Abstract
Accumulating evidence over the past decades has revealed an intricate relationship between dysregulation of cellular metabolism and the progression of atherosclerotic cardiovascular disease. However, an integrated understanding of dysregulated cellular metabolism in atherosclerotic cardiovascular disease and its potential value as a therapeutic target is missing. In this Review, we (1) summarize recent advances concerning the role of metabolic dysregulation during atherosclerosis progression in lesional cells, including endothelial cells, vascular smooth muscle cells, macrophages and T cells; (2) explore the complexity of metabolic cross-talk between these lesional cells; (3) highlight emerging technologies that promise to illuminate unknown aspects of metabolism in atherosclerosis; and (4) suggest strategies for targeting these underexplored metabolic alterations to mitigate atherosclerosis progression and stabilize rupture-prone atheromas with a potential new generation of cardiovascular therapeutics.
Atherosclerotic cardiovascular disease (ASCVD) progressively develops over decades and is characterized by chronic, non-resolving inflammation initiated by the accumulation of ApoB-containing lipoproteins1. Among the leading metabolic pathways driving ASCVD progression are imbalances in lipid metabolism, glucose utilization and amino acid metabolism2–4 (Fig. 1). Dysregulations in these pathways elicit a panoply of pro-atherogenic responses arising from an intricate interplay between endothelial cells (ECs), vascular smooth muscle cells (vSMCs) and innate and adaptive immune cells (for example, monocyte-derived macrophages and lymphocytes)5. As the disease progresses, the lipid-rich plaque expands and eventually becomes prone to rupture or superficial erosion, resulting in thromboembolic events, such as myocardial infarction and stroke6,7. Despite optimal lipid-lowering treatment with statins and PCSK9-blocking antibodies8, ASCVD remains the leading cause of morbidity and mortality worldwide9. Dysregulation of metabolic pathways in lesional cells that augment inflammation and hamper its resolution contribute substantially to this residual risk10–13. This Review explores recently identified cellular and metabolic pathways that contribute to atherosclerosis, their associated molecular mechanisms and the consequences thereof. We further highlight critical gaps in the field being addressed with emerging technologies and newly identified preclinical strategies aimed at optimizing cellular metabolism.
Fig. 1 |. Cellular events leading to the inception, progression and manifestation of cardiovascular events.

Top, Metabolic pathways upregulated in atheroresistant/stable disease areas are in blue, and pathways enriched in atheroprone/unstable disease areas are in red. Bottom, Circulating monocytes bind to adhesion molecules presented on the surface of activated ECs and transmigrate into the vessel wall. Monocytes then mature into macrophages and become foam cells, which can later become apoptotic due to uncontrolled uptake of modified LDL. In early atherosclerosis, these dead cells are efficiently cleared by macrophages. However, as atherosclerosis advances, the capacity for AC removal becomes impaired and promotes necrotic core formation. T cells similarly extravasate into the vessel wall and differentiate into subsets that influence the function of lesional cells. Upon a variety of atherogenic insults, medial vSMCs dedifferentiate and migrate towards the intima, where they initially assemble extracellular matrix in the fibrous cap. In addition, vSMCs can also derive from ECs that have undergone a mesenchymal transition. At later phases of atherosclerosis, vSMCs can adopt multiple cellular phenotypes that can destabilize the fibrous cap and drive the formation of rupture-prone atheromas.
ECs
EC activation as an initiating factor in atherosclerosis
Before the appearance of foam cells and the formation of fatty streaks, disturbed ‘fluid shear stress’ (FSS), characterized by low magnitudes in blood flow with complex changes in direction at curvatures, branch-points and bifurcations, drives EC activation14,15. ECs in these regions exhibit increased inflammatory gene expression (for example, intercellular adhesion molecule 1 (ICAM1) and vascular cell adhesion molecule 1 (VCAM1)), enhanced leukocyte recruitment, and paracellular permeability16 (Fig. 1). ApoB-containing lipoproteins accumulate in areas of paracellular pores or are transported across ECs at sites of disturbed flow via transcytosis17–19. These positively charged lipoproteins are retained within the negatively charged, proteoglycan-rich subendothelial matrix, where they undergo various modifications (for example, oxidation, glycation and aggregation)20,21. Subsequently, recruited monocyte-derived macrophages take up these modified lipoproteins and transform into foam cells (Fig. 1). Notably, foam cell formation compromises their beneficial immune functions, hampers their ability to clear dying cells and enhances their susceptibility to cell death22–24.
EC metabolism in atherosclerosis
The vasculature responds to tissue metabolic demands through homeostatic processes predominantly orchestrated by the endothelium. This regulation determines vascular network formation during embryonic development and adult angiogenesis, vessel diameter and tone, transendothelial transport of macromolecules and micro-molecules, and inflammation25. ECs possess relatively low mitochondrial mass and generate most of their ATP via glycolysis26, converting glucose to lactate (Fig. 2). However, most of these studies were conducted under static conditions, not considering the role of FSS that ECs experience continuously in vivo. Therefore, a comprehensive understanding of EC metabolism requires studying the effects of FSS.
Fig. 2 |. Metabolic pathways in lesional cells relevant to atherosclerosis.

Glucose is transported through the GLUT1 transporter and then proceeds down the glycolysis pathway. This pathway can give rise to lactate or feed into the TCA cycle. Pyruvate, a product of glycolysis, can enter the mitochondria through MPC1 or MPC2 and be converted to acetyl-CoA, which then enters the TCA cycle. Simultaneously, acyl-CoA can be transported into the mitochondria through the carnitine palmitoyltransferase (CPT) 1A and 2 enzymes and enter β-oxidation, generating acetyl-CoA that also feed into the TCA cycle. By-products of these pathways also fuel electron transport chain activity. Glutaminolysis occurs through the enzymes GLS1 and GLUD1, which generate α-ketoglutarate. Additionally, arginine can be converted into NO through the NOS enzymes (eNOS, nNOS or iNOS) or into the polyamine biosynthetic pathway, which also relies on methionine metabolism. Tryptophan is degraded by IDO1 into kynurenine and generates kynurenic acid, 3-hydroxykynurenine, 3-hydroxyanthranilic acid or quinolinic acid. Serine is converted to glycine by SHMT1 or SHMT2 and gives rise the formation of the potent anti-oxidant GSH. GLUT1, glucose transporter 1; HK2, hexokinase 2; GPI, glucose-6-phosphate isomerase; PFKFB, 6-phosphofructo-2-kinase/fructose-2,6-biphosphatases; PFK1, phosphofructokinase-1; TPI1, triosephosphate isomerase 1; ALDO, aldolase; GAPDH, glyceraldehyde 3-phosphate dehydrogenase; PGK1, phosphoglycerate kinase 1; PGM1, phosphoglycerate mutase 1; ENO1, enolase 1; PKM, pyruvate kinase isozymes M1/M2; LDH, lactate dehydrogenase; MPC1/MPC2, mitochondrial pyruvate carriers 1 and 2; SHMT, serine hydroxymethyltransferase; IDO1, indoleamine 2,3-dioxygenase 1; ODC1, ornithine decarboxylase; SRM, spermidine synthase; SRM, spermine synthase; MAT2A, methionine adenosyltransferase 2A; AMD1, adenosylmethionine decarboxylase 1; GLUD1, glutamate dehydrogenase 1.
Mechanical forces and EC metabolism.
Atherosclerosis preferentially forms at sites of disturbed FSS, as noted above. Recent evidence links disturbed FSS with early EC activation and metabolic dysregulation, enhanced oxidative stress, paracellular permeability and impaired nutrient trafficking. Disturbed FSS promotes aerobic glycolysis, whereby glucose is broken down in the presence of oxygen to produce lactate (Fig. 2), modulating mechanosensitive Yes-associated protein (YAP)/transcriptional coactivator with PDZ-binding motif (TAZ) complex activation27 (Fig. 3a), which reciprocally perpetuates glycolysis28. Consistently, inhibiting YAP/TAZ mitigates atherosclerosis progression29. Additionally, disturbed FSS promotes glycolysis through the hypoxia-inducing factor-1α (HIF1α)-mediated upregulation of glycolysis-related genes, contributing to EC activation30 (Fig. 3a). Mechanistically, HIF1α expression is mediated by disturbed FSS-induced AMP-activated protein kinase (AMPK) activation31 (Fig. 3a). Surprisingly, EC deletion of AMPK increases endothelial permeability and inflammatory responses and augments atherosclerosis progression31, underscoring the complexity between mechanical forces and cellular metabolism.
Fig. 3 |. Metabolic pathways in lesional cells and their consequences.

a, Unilaminar FSS induces the expression of transcription factors KLF2 and KLF4, which promote eNOS-mediated NO production, suppressing NF-κB activation and endothelial permeability. Mechanistically, FSS induces KLF2/KLF4 through the MEKK2/MEKK3–MEK5–ERK5 signalling cascade. FSS also decreases glycolysis in a KLF2-dependent manner, supporting mitochondrial metabolism, NADPH production and redox homeostasis. By contrast, disturbed FSS enhances glycolysis through HIF1α, contributing to EC activation. However, some reports indicate a complicated role for glycolysis in EC activation (dashed line). b, Growth factors and cytokines stimulate glucose uptake and promote glycolysis, resulting in dedifferentiation and phenotypic switching. Additionally, the PPP helps to maintain redox balance and inhibits apoptosis, which is counteracted by NOS-mediated NO. Alternatively, arginine metabolism into the polyamine biosynthetic pathway drives proliferation, migration and collagen deposition. IDO1-mediated kynurenine synthesis suppresses osteogenic reprogramming of vSMCs by stimulating the degradation of RUNX2, thereby restraining vascular calcification. c, Glycolysis stimulates pro-inflammatory cytokine secretion. However, lactate simultaneously drives a robust pro-resolving response—stimulating both IL-10 and continual efferocytosis. FAO prevents foam cell formation by suppressing CD36 expression and reducing foam cell formation. GSH synthesis drives continual efferocytosis and lowers superoxide levels. Efferotabolism of AC-derived cargo further stimulates continual efferocytosis, drives TGFβ production, and expands pro-resolving macrophages through the process known as ‘EIMP’. d, mTORC1-mediated glycolysis, serine metabolism into glycine and one-carbon units, and methionine-mediated histone methylation support the differentiation and expansion of Teff cells. Additionally, cholesterol accumulation destabilizes Treg cell differentiation and promotes T cell exhaustion and exTreg cell expansion. OXPHOS stimulates AMPK-mediated repression of glycolysis and drives Treg cell differentiation and stability. Cross-talk pathways driven by metabolites and soluble factors are shown in red.
Conversely, in straight regions of arteries where FSS caused by blood flow is high and unidirectional, termed ‘unilaminar FSS’, pro-inflammatory, prothrombotic and oxidative stress-inducing pathways are suppressed. Unilaminar FSS induces the transcription factors Kruppel-like factor (KLF) 2 and 4, which have a major role in preventing atherosclerosis at these sites32. KLF2 and KLF4 drive the expression of endothelial nitric oxide synthase (eNOS) to generate nitric oxide (NO)32 (Fig. 2), induce thrombomodulin to limit thrombosis, and suppress nuclear factor κ-light-chain-enhanced of activated B cells (NF-κB) activation. Mechanistically, unilaminar FSS induces KLF2/KLF4 through the MEKK2/MEKK3–MEK5–ERK5 signalling cascade that mediates the binding of myocyte enhancer factor 2 (MEF2) to KLF2/KLF4 gene regulatory elements32 (Fig. 3a). Interestingly, transcriptomic profiling revealed that unilaminar FSS reduces the expression of key glycolytic enzymes, such as 6-phosphofructo-2-kinase/fructose-2,6-biphosphatase 3 (PFKFB3), phosphofructokinase-1 (PFKM) and hexokinase 2 (HK2), and inhibits glycolysis in a KLF2-dependent manner33. Unilaminar FSS also stimulates mitochondrial oxidative phosphorylation (OXPHOS), leading to a moderate increase in reactive oxygen species production34. This rise in reactive oxygen species initiates mitophagy, which, in turn, facilitates the assembly of a protein complex that scaffolds the ERK5 signalling axis to amplify KLF2-mediated eNOS expression34 (Fig. 3a). These insights linking the mechanosensitive and metabolic responses of ECs are crucial in developing targeted therapies for atherosclerosis.
Lipids and EC activation.
The endothelium also contributes to atherosclerosis through fatty acid uptake following hydrolysis of circulating lipoproteins by endothelial lipase35. Rate-limiting steps in fatty acid uptake include mitochondrial ATP production and activity of fatty acid transport proteins 3 and 4 (FATP3 and FATP4)36,37. Acyl-CoA synthetase long-chain 1 (ACSL1) subsequently converts long-chain fatty acids into fatty acyl-CoA esters that are either subjected to β-oxidation or stored in lipid droplets38. Low-density lipoprotein (LDL) particles are internalized by the LDL receptor (LDLR), cluster of differentiation 36 (CD36), activin receptor-like kinase 1 (ALK1) and scavenger receptor class B type 1 (SR-B1)19,37,39. Interestingly, uptake via LDLR favours utilization, whereas uptake via ALK1 or SR-B1 is coupled with transcytosis and subendothelial LDL accumulation19,39. Modification of subendothelial LDL promotes EC activation in the oxidative plaque microenvironment40. Mechanistically, bioactive lysophosphatidic acid, a cleavage product of lysophosphatidylcholine derived from oxidized LDL (oxLDL), binds the endothelial receptors LPA1 and LPA3 and drives CXCL1-mediated monocyte recruitment41, further linking metabolic pathways and plaque inflammation.
Metabolism controls EndMT.
The transition of ECs to mesenchymal cells, termed the ‘endothelial-to-mesenchymal transition’ (EndMT), contributes to atherosclerosis progression and plaque instability42,43. Suppression of transforming growth factor-β receptor 1 and 2 (TGFBR1 and TGFBR2) signalling, essential for EndMT (Fig. 3a), not only halts the progression of atherosclerosis but also induces its regression44. Recent studies have demonstrated that metabolic reprogramming drives EndMT45. In this setting, acetate is increased via its atypical production from glucose and then converted into cytosolic acetyl-CoA by ACSS2 (ref. 45). This leads to SMAD2, SMAD4 and ALK5 acetylation, prolonging their half-life and sustaining TGFβ signalling and EndMT45. Importantly, EC-specific deletion of ACSS2 hampers EndMT and reduces atherosclerosis in male and female mice45.
vSMCs
vSMCs in plaque stability
vSMCs have a crucial role in plaque stability by forming the collagen-rich fibrous cap that overlies clinically dangerous necrotic cores46–48 (Fig. 1). Emerging evidence using fate-mapping and single-cell transcriptomics has demonstrated that neointimal vSMCs arise from either ECs undergoing EndMT or the dedifferentiation and selective expansion of medial vSMCs49–51. Dedifferentiation of vSMCs requires the downregulation of myocardin-mediated expression of contractile and other vSMC-related genes (for example, MYH11 and ACTA2)52, which are strongly diminished in the vast majority of vSMCs present in atherosclerotic plaques53,54. However, these dedifferentiated vSMCs can simultaneously acquire genes resembling myofibroblasts55,56 and macrophages53,56,57. Additionally, vSMCs also contribute to vascular calcification due to their transition to chondrocyte-like and osteoblast-like cells58,59. Notably, the presence of vascular calcification has emerged as a superior predictor of cardiovascular events and cardiovascular-related mortality, surpassing all other factors or risk equations to date60,61.
Cellular metabolism guides phenotypic modulation of vSMCs Glucose utilization controls vSMC phenotype.
Recent studies interrogating the complex metabolic interactions within atherosclerotic plaques have uncovered a critical role for glucose uptake and glycolysis in controlling vSMC dedifferentiation62. The soluble mediators platelet-derived growth factor increase GLUT1-mediated glucose uptake and glycolysis in vSMCs63, leading to their dedifferentiation, proliferation and migration (Fig. 3b)63. Enhancing glycolysis via GLUT1 overexpression in vSMCs accelerates atherosclerosis in male and female mice that exhibit features resembling metabolic syndrome in humans64. Furthermore, deletion of pyruvate kinase muscle isozyme M2 (PKM2), which catalyses the final step in glycolysis, inhibited platelet-derived growth factor-induced vSMC proliferation and migration and hampered neointimal hyperplasia65,66. Single-cell RNA sequencing (scRNA-seq) revealed that vSMC dedifferentiation during atherosclerosis requires metabolic reprogramming towards glycolysis67. This was supported by unbiased network preservation analysis using vSMCs isolated from a multi-ethnic cohort of 151 heart transplant donors68. Also, mass spectrometry imaging (MSI) revealed elevated lactate in unstable human atheromas compared to stable plaques69. This is particularly interesting considering that vSMCs residing in a microenvironment rich in lactate show a more synthetic phenotype70. Although these studies point to a critical role for glycolysis in vSMC phenotypic modulation, a shift in glucose utilization towards the hexosamine biosynthesis pathway and the pentose phosphate pathway (PPP) is also observed (Fig. 2)71. The PPP has a crucial role in maintaining cellular redox balance via the production of nicotinamide adenine dinucleotide phosphate (NADPH)72,73. Activation of glucose-6-phosphate dehydrogenase, the rate-limiting enzyme of the PPP (Fig. 2), mitigates vSMC apoptosis by maintaining redox homeostasis72,73 (Fig. 3b). Given that vSMC apoptosis accelerates features of plaque instability74, enhancing glucose-6-phosphate dehydrogenase-mediated flux through the PPP may offer a novel strategy for stabilizing rupture-prone atheromas.
The mammalian target of rapamycin complex 1 (mTORC1) is a pivotal regulator of nutrient sensing and metabolism and has a central role in vSMC phenotypic modulation75. This complex comprises the core catalytic subunit, mTOR, and the regulatory associated subunit, Raptor75. Conditional deletion of Raptor in vSMCs, essential for mTORC1 activation, impairs relaxation and contractility in the aorta accompanied by alterations in autophagic signalling76. mTORC1 activation is suppressed by the tuberous sclerosis complex (TSC), and deletion of TSC1 leads to chronic mTORC1 hyperactivation. Interestingly, chronic mTORC1 activation in vSMCs achieved by vSMC-specific TSC1 deletion diminished vSMC contractile genes and led to a degradative vSMC phenotype that caused extracellular matrix proteolysis and progressive aortic disease77. These studies highlight the critical role of mTORC1 in vascular health and disease.
vSMC amino acid metabolism in plaque stability.
Arginine metabolism has a central role in vSMC phenotypic modulation and plaque stability. Through the NOS family of enzymes, including nNOS, iNOS and eNOS, arginine metabolism leads to the generation of citrulline and NO78 (Fig. 2). EC-generated NO stimulates vasodilation via cyclic GMP-mediated activation of soluble guanylate cyclase in vSMCs, and vSMC-generated NO operates as a free radical that inhibits proliferation and migration and promotes apoptosis79,80 (Fig. 3b). NOS enzymes and arginases compete for available arginine. Thus, flux through arginase 1 (ARG1) limits NO production and promotes survival, proliferation and collagen synthesis while inhibiting lipopolysaccharide (LPS)-induced inflammation79–81 (Figs. 2 and 3b). Furthermore, ARG1 overexpression in vivo, which substantially increases lesional polyamines (small, positively charged molecules integral to crucial cellular processes), enhances vSMC proliferation, drives the production of pro-resolving cytokines and promotes features associated with plaque stability82. Treatment with the polyamine spermidine mitigates necrotic core expansion by stimulating vSMC autophagy and preventing lipid accumulation, supporting polyamine synthesis as an important mediator83.
In addition to altered arginine metabolism, dysregulated tryptophan metabolism in vSMCs affects vascular calcification and atherosclerosis (Fig. 3b). Importantly, individuals with coronary artery calcification show evidence of altered tryptophan metabolism84. Because lesional vSMCs adopt osteoblast-like properties during atherosclerosis, dysregulation in tryptophan metabolism probably contributes to vSMC-mediated vascular calcification. Indeed, vSMC-specific deletion of the tryptophan-catabolizing enzyme, indoleamine 2,3-dioxygenase 1 (IDO1), increases runt-related transcription factor 2 (RUNX2) expression and exacerbates vascular calcification, which can be reversed by administering the IDO1 product kynurenine84 (Fig. 3b). By activating the aryl hydrocarbon receptor, kynurenine limits the transition of vSMCs to chondrocytes and preserves the integrity of the fibrous cap85. However, observational studies in humans have demonstrated inconsistent correlations between the tryptophan–kynurenine pathway to ASCVD and stroke86–88. Additionally, the tryptophan metabolite 3-hydroxyanthranilic acid activates NF-κB and increases MMP2 expression in vSMCs89. Furthermore, 3-hydroxyanthranilic acid treatment in mice augments the formation of abdominal aortic aneurysms89. Although these insights into the metabolic reprogramming of vSMCs highlight their role in atherosclerosis, the exact implications of tryptophan metabolism remain to be fully resolved.
Macrophage immunometabolism
Macrophages show remarkable plasticity in their function and occupy a range of phenotypes that extend from pro-inflammatory to pro-resolving states90,91. This plasticity is essential as certain phenotypes are adapted for combating acute infections, and others aid in tissue repair. A plethora of studies have established a link between the metabolic activities of macrophages and their phenotypic states. For instance, pro-inflammatory macrophages display increased aerobic glycolysis, flux through the PPP, and iNOS-mediated NO production92–95 (Fig. 2). By contrast, pro-resolving macrophages are characterized by sustained tricarboxylic acid (TCA) cycle flux, elevated fatty acid β-oxidation (FAO), increased oxygen consumption, enhanced glutaminolysis and polyamine biosynthesis92,93,96 (Fig. 2). Importantly, the balance between pro-inflammatory and pro-resolving macrophages in atherosclerotic plaques has important consequences in their clinical manifestations, as they can either drive plaque instability or stabilize rupture-prone atheromas. As discussed below, metabolic reprogramming of lesional macrophages is highly dynamic, and emerging studies suggest that correcting dysregulated macrophage metabolism is a promising therapeutic approach.
Cholesterol metabolism in foam cell formation
Dysregulated cholesterol metabolism in macrophages is intricately linked to the pathogenesis of atherosclerosis. The cell-surface lipid-sensing triggering receptor expressed on myeloid cells 2 (TREM2), identified in foamy plaque macrophages97, upregulates the expression of the scavenger receptor CD36 (ref. 98). Uptake of modified LDLs through scavenger receptors, such as CD36 or SR-B1, bypasses the regulatory feedback that normally limits cholesterol accumulation. Internalized cholesterol is esterified by acyl-CoA:cholesterol acyltransferase (ACAT) to store cholesterol as cholesterol esters in lipid droplets, transforming macrophages into foam cells99. However, dysregulations in sterol metabolism prevent the safe disposition of cholesterol into lipid droplets and promote the formation of membrane-damaging cholesterol crystals100,101. This is further exacerbated as normal cholesterol efflux through the transporters ABCA1 and ABCG1 becomes compromised during atherosclerosis. Although these foam cells are not inflammatory, their diminished migratory capability and the constant barrage of atherogenic stimuli eventually lead to their death through apoptosis102–106. If these apoptotic foam cells are not rapidly cleared, they further proceed into post-apoptotic necrosis, spilling their contents into the plaque microenvironment102,103,107. The uptake of cholesterol crystals activates the NLRP3 inflammasome108,109. Activation of this multiprotein complex leads to caspase-1-mediated secretion of the pro-inflammatory and pyrogenic cytokines, interleukin (IL)-1β and IL-18 (refs. 110,111). Notably, suppression of the inflammasome via enhancing cholesterol efflux or by inhibiting NLRP3 assembly blunts atherosclerosis progression and enhances features of plaque stability101,112–114.
Glycolysis and OXPHOS in inflammation and its resolution
Inflammatory stimuli have been shown to enhance the expression of the key glycolytic enzymes GLUT1 (encoded by SLC2A1)115, hexokinase 3 (HK3)116, 6-phosphofructo-2-kinase/fructose-2,6-biphosphatase 3 (PFKFB3)116,117 and PKM2 (refs. 118–120). This glycolytic shift is further driven by atherogenic factors, such as cytokines, oxLDL and cholesterol crystals, which increases their polarization towards an inflammatory phenotype121. The concept that pro-inflammatory lesional macrophages readily take up glucose has been leveraged in positron emission tomography using the glucose analogue [18F]-fluoro-2-deoxy-d glucose (18F-FDG) to detect advancing atheromas in humans122,123.
Enhanced glycolysis in pro-inflammatory macrophages breaks the TCA cycle, leading to the accumulation of the TCA cycle intermediates, succinate and citrate124,125. These intermediates drive reactive oxygen species-mediated IL-1β production and tumour necrosis factor (TNF) translation124,125 (Fig. 3c). Moreover, pro-inflammatory macrophages exposed to bioactive components of modified LDL increase citrate production. This is subsequently converted into oxaloacetate by ATP citrate lyase (ACLY), promoting HIF1α-mediated IL-1β production126. Although likely due to modulation of fatty acid and cholesterol metabolism rather than glycolysis, deleting ACLY in myeloid cells enhances fibrous cap formation and decreases necrotic core size127. Despite these clear connections between glycolysis and pro-inflammatory responses, the role of glycolysis in macrophages during atherosclerosis is complex. For instance, mice heterozygous for the Slc2a1 gene in their haematopoietic cells showed reductions in monocyte recruitment to plaques and attenuations in atherosclerosis128. By contrast, myeloid-specific GLUT1 deletion augmented features of plaque instability129. Also, transgenic overexpression of GLUT1 in CD68+ cells enhanced glucose uptake and increased glycolysis, but atherosclerosis progression or features of plaque stability were unaffected116. These data suggest a difference between limiting versus abolishing glucose flux in macrophages. As discussed later, this may be partly due to the recently discovered effects of lactate stimulating inflammation resolution.
Although pro-resolving macrophages show enhanced FAO-mediated oxygen consumption, FAO is dispensable for their polarization130,131. Interestingly, inhibiting macrophage FAO by deleting CPT1α and CPT2 enhanced CD36 expression and increased foam cell formation (Fig. 3c), ultimately driving atherosclerosis progression132. This process is influenced by oxLDL/CD36 signalling, which stimulates the NLRP3 inflammasome and cytokine production, shifting the metabolic balance in macrophages from OXPHOS to glycolysis133. The potent anti-inflammatory cytokine IL-10 suppresses LPS-induced glycolysis and drives OXPHOS, which blunts inflammasome-mediated IL-1β secretion134. The establishment of FAO driving IL-10 and, reciprocally, IL-10 promoting FAO, highlights the existence of a feed-forward loop that augments resolution135. Peroxisome proliferator-activated receptor gamma (PPARγ) critically regulates this pathway as its deletion reduces oxygen consumption and exacerbates atherosclerosis136,137. PPARγ deletion lowers the expression of isocitrate dehydrogenase 1 (IDH1) and increases the expression of immune-responsive gene 1 (IRG1, encoded by ACOD1), resulting in itaconate accumulation136,138. This metabolite inhibits succinate dehydrogenase and subsequently leads to increased succinate levels, which block macrophage polarization towards a pro-resolving phenotype119,138,139 (Fig. 3c). Although manipulating these pathways is an attractive therapeutic strategy, the risk of unintended consequences, such as heightened susceptibility to infections and tumorigenesis140, cannot be ignored. This delicate interplay between metabolism, inflammation and resolution in atherosclerosis highlights the need for precise and carefully considered therapeutic interventions.
Amino acid metabolism in macrophages
Accumulating evidence over the past two decades has increasingly highlighted a critical role for amino acid metabolism in macrophages. The amino acids glutamine, arginine, tryptophan, serine and glycine have been shown to have important roles in macrophage function, particularly during atherosclerosis. As an example, glutamine can be transaminated by glutaminase 1 (GLS1) into glutamate141 (Fig. 2). This is then channelled by the malate–aspartate shuttle via aspartate aminotransferase (GOT; Fig. 3c). Macrophages deficient in GLS1 show deficiencies in OXPHOS and are unable to meet the high-energy demand for cytoskeletal rearrangement required for efferocytosis. Mice with myeloid-specific deletion of GLS1 manifest large necrotic cores during atherosclerosis141. As another example, macrophages can metabolize arginine through the NO pathway, in which iNOS converts arginine to NO, or the polyamine pathway, in which arginine is converted to ornithine and then into polyamines (Fig. 2). The temporal features of arginine metabolism through these routes have been demonstrated in models of inflammation and its resolution. iNOS-mediated NO production is high during acute inflammation142, which is useful when eliminating pathogens. At later phases, arginase mediated ornithine formation and polyamine biosynthesis dominates142. As mentioned earlier, macrophages exist among a spectrum of phenotypes and often express iNOS and ARG1 simultaneously in vivo. However, the metabolism of arginine by one pathway often suppresses the other. For instance, NO-mediated nitrosylation on ornithine decarboxylase (ODC1) suppresses its activity and prevents polyamine synthesis143. Conversely, NO production is restricted by ARG1-mediated polyamine synthesis, as polyamines suppress iNOS translation144–146. Interestingly, low expression of ARG1 in macrophages is associated with susceptibility to atherosclerosis147, and deletion of ARG1 in myeloid cells blunts atherosclerosis regression in male mice, where ARG1+ macrophages are abundant148.
The degradation of tryptophan into kynurenine by IDO1 (Fig. 2) contributes to the non-inflammatory property of efferocytosis149. Furthermore, kynurenine possesses diagnostic and prognostic value for atherosclerosis, and a deviation in this pathway, particularly a downregulation of kynurenic acid branch enzymes, is associated with unstable atherosclerosis150,151. Tryptophan degradation through this pathway exerts anti-inflammatory effects due to the subsequent accumulation of uncharged tRNAs that activate the serine/threonine kinase general control nonderepressible 2 (GCN2). Consequently, GCN2 activation modulates ribosome assembly and induces a repertoire of transcription factors that promote Il10 and Tgfb expression while simultaneously suppressing Il12 production149. Treating macrophages with the tryptophan metabolite 3-hydroxyanthralinic acid blocks inflammasome activation, lowers oxLDL uptake and decreases atherosclerosis progression152,153. These pathways probably account for the finding that deleting or inhibiting IDO1 in apolipoprotein-E-deficient (Apoe−/−) male mice enhances atherosclerosis, increases necrotic core area and augments the expression of pro-inflammatory cytokines154,155. Additionally, treating atheroprone male mice with the IDO1 inhibitor 1-methyl-dl-tryptophan (1-MT) blocked atherosclerosis regression elicited by the oral administration of eicosapentaenoic acid156. Furthermore, preclinical and clinical studies demonstrate that the synthetic tryptophan metabolite 3,4-dimethoxycinnamoyl anthranilic acid (3,4-DAA, also known as ‘Tranilast’) reduced atherosclerosis progression in mice, lowered the restenosis rate in patients with ASCVD after transluminal angioplasty and lowered myocardial infarction rates in humans in the Prevention of REStenosis with Tranilast and its Outcomes (PRESTO) trial154,157–159. However, other studies suggest a more complicated role of IDO1 in atherosclerosis. For instance, inhibiting IDO1 using 1-MT or deleting IDO1 in haematopoietic cells enhanced high-density lipoprotein (HDL) and increased IL-10 production160,161. Moreover, MSI studies revealed that the tryptophan metabolite 5-hydroxyindoleacetic acid, associated with metabolic syndrome and inflammation162, is enriched in the fibrous cap of unstable human plaques69. Additionally, tryptophan metabolism, as assessed by the ratio of kynurenine to tryptophan, was positively correlated with stroke severity163.
Serine supports a complex network of metabolic pathways that contribute to cell proliferation and synthesis of nucleotides, NADPH and S-adenosyl-methionine (SAM) through one-carbon metabolism164,165 (Fig. 2). Additionally, serine has a crucial role in maintaining redox homeostasis by its conversion to glycine via serine hydroxymethyltransferases 1 and 2 and subsequently into the potent anti-oxidant glutathione (GSH; Figs. 2 and 3c). Despite the role of GSH in suppressing oxidant stress and inflammation, reducing GSH synthesis by serine deprivation decreases LPS-induced IL-1β production166. Another study corroborated this finding yet suggested that serine has a role in IL-1β production through SAM-mediated epigenetic reprogramming by histone methylation rather than by reducing GSH167. Nonetheless, GSH is among the metabolites most decreased in pro-inflammatory macrophages, and in vivo strategies that enhance GSH production decrease atherosclerosis168–170. Furthermore, glycine, a GSH precursor, is negatively associated with ASCVD in humans171,172, and glycine-based treatments reduce atherosclerosis through de novo GSH biosynthesis in macrophages173.
Mechanistic insights into efferotabolism
Macrophage-mediated clearance of apoptotic cells (ACs), termed ‘efferocytosis’, is vital to tissue homeostasis107. Monocyte-derived macrophages in early atherosclerotic lesions display intact AC clearance107 (Fig. 1). However, as atherosclerosis advances, efferocytosis becomes defective, which drives the formation of large necrotic cores that are intimately linked to clinically dangerous unstable atheromas23,174–177. Preclinical studies aimed at restoring efferocytosis in mouse models of atherosclerosis demonstrated that enhancing the clearance of ACs diminishes necrotic core area and increases the size of the protective fibrous cap178–181. Emerging evidence indicates that the efficient clearance of ACs in vivo requires successive rounds of phagocytosis148,182,183, which is metabolically distinct from single phagocytic events. After the internalization of an AC, phagocytes degrade and process the AC-derived macromolecules through a process requiring microtubule-associated protein 1A/1B light chain 3 (refs. 184–186). This metabolically demanding process, which we have termed ‘efferotabolism’187, describes the process by which phagocytes break down, metabolize and respond to AC-derived products (Fig. 4). Efferotabolism is necessary to trigger a resolution response and has been associated with features of plaque stability. Conversely, impaired processing of AC-derived macromolecules sustains inflammation and is now considered a hallmark of advanced atherosclerosis188.
Fig. 4 |. Current understanding of efferocytosis, the metabolism of AC-derived cargo and consequences thereof.

Binding of an AC to a macrophage activates cell-surface receptors that drive ERK1/ERK2 activation and DRP1-mediated mitochondrial fission. Free cholesterol from a degraded AC is either esterified by ACAT in the ER or exported by the cholesterol transporters ABCA1 and ABCG1. Methionine from a degraded AC is converted to SAM, which is used for DNA methylation that subsequently suppresses DUSP4 expression. This lifts repression of ERK1/ERK2 activation. ERK1/ERK2 signalling drives MYC expression, PGE2-mediated TGFβ production and AP1-mediated IL-10 secretion. AC-derived arginine is converted into putrescine and triggers cytoskeleton remodelling and continual efferocytosis. Recycling of AC-derived nucleotides mediates EIMP through a DNA-PK–mTORC2 signalling cascade. Fatty acids derived from ACs are used for FAO that promotes NAD+–SIRT1–PBX1 signalling that drives IL-10 expression. Concurrently, GLUT1-mediated glucose uptake promotes glycolysis and lactate production. Lactate is secreted from the macrophage via MCT1 to prime a pro-resolving microenvironment. Guided by mitochondrial fission, lactate is utilized for histone lactylation, which drives an epigenetic programme that promotes ARG1 expression.
After AC degradation in phagolysosomes, intracellular cholesterol levels rise, especially when digesting cholesterol-rich apoptotic foam cells189. To manage this, macrophages direct AC-derived cholesterol to ACAT in the endoplasmic reticulum (ER)190 (Fig. 4). Concurrently, AC-derived sterols stimulate LXRs and PPARs, enhancing the expression of cholesterol transporters ABCA1 and ABCG1 (refs. 191–193; Fig. 4). Cholesterol efflux through these transporters has a vital role in macrophage survival after efferocytosis189,194. Furthermore, TREM2 deficiency leads to a failure of LXR-mediated cholesterol efflux and efferocytosis as well as increasing ER stress responses. Despite these pro-atherogenic responses, TREM2 deletion in monocytes reduced atherosclerosis195. However, this initial benefit diminishes as the disease progresses196. In addition to cholesterol, intracellular arginine levels also increase after efferocytosis148. Pro-resolving macrophages expressing high levels of ARG1 hydrolyse arginine into ornithine, which is subsequently decarboxylated by ODC1 to generate putrescine. Putrescine biosynthesis leads to Rac1-mediated cytoskeletal remodelling that permits the internalization of multiple ACs (Figs. 3c and 4). Deletion of ARG1 from myeloid cells or silencing ODC1 in lesional macrophages using macrophage-targeting nanoparticles leads to a selective defect in continual efferocytosis, lowers IL-10 production and blunts atherosclerosis regression in male mice148,197.
The binding of an AC with a macrophage triggers CD36-dependent ERK1/ERK2 activation, leading to prostaglandin E2 biosynthesis and TGFβ1 expression198. However, CD36-mediated ERK1/ERK2 activation is insufficient to surmount the repression imposed by dual specificity phosphatase 4 (DUSP4). This regulatory feedback is overcome through a methionine salvage pathway originating from the phagolysosomal degradation of an AC198. AC-derived methionine is enzymatically converted to SAM by methionine adenosyltransferase 2A198. SAM is then utilized by DNA methyltransferase 3A (DNMT3A) to repress DUSP4 expression via DNA methylation, permitting ERK1/ERK2–PGE2-mediated TGFβ1 upregulation198. Haematopoietic deletion of DNMT3A impairs this ERK1/ERK2–PGE2–TGFβ1 signalling axis in mice and drives features of plaque instability198 (Figs. 3c and 4). In parallel, nucleotides salvaged from AC degradation potentiate macrophage proliferation via a process termed ‘efferocytosis-induced macrophage proliferation’ (EIMP)199 (Figs. 3c and 4). Similarly to CD36, AC binding with a macrophage activates MerTK to stimulate ERK1/ERK2. Simultaneously, mTORC2 is stimulated by DNA-PK-mediated Rictor activation. These pathways converge on the transcriptional repressor BHLEH40, which represses c-Maf-mediated suppression of cell cycle progression, a process contingent on Myc. Importantly, mTORC2 deletion in macrophages lowered IL-1β production during atherosclerosis progression and impaired proliferation of efferocytosis-competent macrophages during regression199,200.
Interestingly, macrophage ingestion of an AC stimulates both glycolysis and OXPHOS135,182. Unbiased metabolomics revealed that macrophages use fatty acids from ACs to fuel mitochondrial β-oxidation, increasing NAD+ levels135. Inhibiting complex III of the electron transport chain limited IL-10 production, which can be restored by supplying NAD+ precursors. Additionally, RNA-seq of efferocytes revealed upregulation of GLUT1, enhancing glucose uptake and stimulating a transient increase in glycolysis dependent on PFKFB2 activation182,201. Efferocytosis-mediated glycolysis promotes cytoskeleton remodelling and enhances the expression of AC receptors to drive continual efferocytosis182,201 (Fig. 4). Furthermore, lactate generated in macrophages after efferocytosis is exported via MCT1 into the surrounding microenvironment182,202. Lactate also promotes inflammation resolution through epigenetic modifications whereby lactate moieties are covalently attached to lysine residues on histone proteins203. Histone lactylation in macrophages enhances ARG1 expression, increases their phagocytic capacity and redirects macrophage polarization towards a pro-resolving phenotype204,205. Interestingly, histone lactylation is tightly coupled to mitochondrial dynamics as mitochondrial fission, an early event in efferocytosis206, drives histone lactylation205 (Fig. 4).
A recent integrative study combining transcriptomic and proteomic approaches identified two distinct states in macrophages under chronic physiological hypoxia (~1% oxygen)—the ‘primed’ and ‘poised’ states207—that are associated with enhanced conversion of glucose into a noncanonical PPP loop. This increases the production of NADPH that drives phagolysosomal acidification and maintains cellular redox homeostasis, which are necessary for continual efferocytosis. These findings are particularly interesting because advanced human atheromas that are stable show signs of inflammation resolution and may involve metabolic reprogramming of lesional macrophages208. Gaining a comprehensive understanding of the interplay between various metabolic pathways in macrophages during efferocytosis will provide valuable insights into the progression of ASCVD and offers potential avenues for therapeutic interventions.
T cell metabolism in atherosclerosis
T cell function in atherosclerosis
T cells are critical modulators of atherogenesis. Regulatory T (Treg) cells, which express the transcription factor FOXP3 and the high-affinity IL-2 receptor CD25, exert anti-inflammatory and pro-resolving effects209. By enhancing efferocytosis and stimulating the production of pro-resolving mediators that promote atherosclerosis regression, Treg cells have been repeatedly shown to be atheroprotective210–212. Alternatively, specific CD4+ helper T cells are associated with pro-atherogenic responses. This is most clear for the TH1 subset of helper T cells210. Type 2 helper T (TH2) cells, although probably having a minor impact on atherosclerosis210, warrant further investigation. Moreover, the intricate interplay between T cell metabolism and features of autoimmunity adds another layer of complexity that needs to be elucidated for a more comprehensive understanding of atherosclerosis pathogenesis.
T cell-mediated autoimmunity in atherosclerosis
Clonal expansion of T cells and B cells in atherosclerosis provides increasing evidence for the essential role of autoimmunity in atherosclerosis213. Individuals with autoimmune disorders, such as systemic lupus erythematosus, rheumatoid arthritis, inflammatory bowel disease and systemic sclerosis, exhibit an increased risk for ASCVD214. Cancer immunotherapies increase the risk of ASCVD, which is assumed to be mediated by disinhibition of autoreactive T cells215. Accordingly, growing evidence has revealed the critical role of CD4+ T cell-mediated autoimmunity in atherosclerosis216–221. Recent single-cell T cell antigen receptor sequencing has revealed a break in tolerance in atherosclerosis, whereby adaptive immune cells no longer distinguish self-antigens from non-self-antigens218,219. Treg cells and effector T (Teff) cells recognize epitopes in plaque-associated autoantigens210. The most studied atherosclerosis-related autoantigen is ApoB, the core protein of LDL, chylomicrons and other lipoproteins. In the 1990s, Hansson and colleagues isolated CD4+ T cells responding to oxLDL from human atherosclerotic plaques222. Whereas this T cell response was suggested to rely on the formation of LDL-neoepitopes through oxidation222, subsequent work identified CD4+ T cells that recognize epitopes in the native ApoB protein, restricted by binding to major histocompatibility complex (MHC). CD4+ T cells recognize peptides bound to MHC class II, and CD8+ T cells recognize peptides bound to MHC class I. Many MHC class II-restricted epitopes have been discovered in the native ApoB protein in mice216,220,223 and humans217,220, and vaccine-based strategies using ApoB peptides substantially decrease atherosclerosis in animal models224. Phenotyping by flow cytometry and RNA-seq has revealed that ApoB-reactive (ApoB+) CD4+ T cells predominantly exhibit Treg cell-like signatures in healthy mice and humans216,221.
Treg cells can become unstable during atherosclerosis and convert into exTreg cells216, described as TH1-like, TFH-like, TH17-like or cytotoxic216,220,225–227. ApoB+ T cells are particularly susceptible to adopting TH1/TH17 effector cell characteristics with minimal Treg cell signatures and to switching into exTreg cells216,221. Conversely, exTreg cells upregulate Teff cell markers216. Notably, exTreg cells are highly pro-inflammatory and pro-atherogenic221, and recent lineage tracing studies have shown that some exTreg cells are a cytotoxic type of CD4+ T cells228. The switch from antigen-specific Treg cells to exTreg cells may participate in the break in tolerance to self during atherogenesis. To date, the mechanisms responsible for this switch are unknown210,229. Cellular metabolic changes might have a critical role in governing this process as the conversion from Treg cells to exTreg cells is accelerated in mice fed a Western diet216.
Metabolic reprogramming determines the function and fate of T cells
Accumulating evidence suggests that T cell responses are dynamically regulated by metabolic signals230. Following activation, quiescent and naive CD4+ T cells reprogramme their cellular metabolism to meet the increased energy demand required for proliferation and differentiation230. Specific metabolic programmes are required for differentiation into pro-inflammatory effector CD4+ T cells (for example, TH1, TH2 and TH17 cells) or anti-inflammatory Treg cells. Whereas Teff cells are highly glycolytic, Treg cells predominantly utilize mitochondrial OXPHOS of fatty acids or pyruvate for energy production231,232. The PI3K–Akt–mTORC1 signalling pathway is one of the central regulators of pro-glycolytic metabolism in CD4+ T cells, which is characterized by GLUT1 upregulation233. GLUT1 expression is essential for the activation and proliferation of Teff cells but not required for the functionality of Treg cells 233. Additionally, pyruvate dehydrogenase kinase 1 (PDK1)232, which inhibits the conversion of pyruvate into acetyl-CoA and thereby prevents subsequent OXPHOS, or the glycolysis-promoting transcription factor HIF1α234, are both highly expressed in Teff cells but low in Treg cells. CD4+ T cell-specific deletion and/or pharmacological depletion of GLUT1, GLUT3, PDK1 or HIF1α all protect mice from experimental autoimmune diseases by attenuating Teff cell differentiation and activation232–235. In Apoe−/− mice, PDK1 inhibition by dichloroacetate reduced plaque infiltration by CD4+ T cells, induced a shift from pro-inflammatory to anti-inflammatory T cell responses, and conferred atheroprotection236. By contrast, Treg cell-specific deletion of the metabolic sensor liver kinase B1, critical for maintaining OXPHOS, reduces Treg cell number and impairs their function, thereby inducing a fatal autoimmune disorder in mice237.
Besides affecting T cell differentiation and functions, metabolic alterations can shift pre-existing lineage decisions. Specifically, uncontrolled glycolysis impairs Treg cell stability (Fig. 3d) and induces spontaneous autoimmunity in mice238–240. Treg cells lacking the phosphatase PTEN, which reduces the pro-glycolytic activity of mTORC2 by inhibiting PI3K–Akt signalling, lose their suppressive capacity as well as their expression of FOXP3 and give rise to autoimmunity in mice238,239. Accordingly, Treg cells expressing constitutively active GLUT1 have diminished suppressive capacity and lose FOXP3 expression240. Although the effects of energy metabolism and intermediate metabolites on T cells in the context of atherosclerosis remain largely unexplored, it is likely that metabolic reprogramming contributes to these responses.
A recent study demonstrated that CD4+ T cells are susceptible to metabolic exhaustion in advanced atherosclerosis. Prolonged Western diet feeding to Apoe−/− male mice reduced glycolysis in naive CD4+ T cells, which led to increased apoptosis and a diminished proliferative capacity241 (Fig. 3d). Moreover, compelling evidence suggests that hypercholesterolaemia also contributes to the instability of Treg cells during atherogenesis (Fig. 3d). A shift from high to normal cholesterol intake reduces the loss of lesional Treg cells and plaque progression in male and female Ldlr−/− mice242. In a related study, treating Apoe−/− mice with ApoAI, the core protein of HDL, diminished Western diet-related intracellular cholesterol accumulation in Treg cells and prevented their conversion into exTreg cells 227. By contrast, exaggerated cholesterol accumulation in T cells induced apoptosis and impaired T cell functionality (Fig. 3d), which was associated with reduced atherosclerosis in aged Ldlr−/− mice243. Because unrestrained glycolysis renders Treg cells unstable in autoimmune disease models238–240, hypercholesterolaemia-driven Treg cell instability might involve a similar mechanism. However, the specific impact of hypercholesterolaemia on metabolic reprogramming of T cells in the context of atherosclerosis must be determined by future studies.
Nutrients modulate T cell responses in autoimmunity
The typical Western diet of the 21st century, characterized by an excess intake of refined sugars, saturated fats, animal protein, salt and cholesterol, is known to increase the risk of ASCVD244,245. Although the effects of diet on T cell-mediated immunity in atherosclerosis remain largely unexplored, emerging evidence strongly suggests that specific nutrients, mainly glucose and amino acids, directly influence T cell fate and function. For example, glucose restriction leads to AMPK-mediated inhibition of mTORC1 signalling, suppressing Teff cell responses and promoting Treg cell differentiation246 (Fig. 3d). Consistently, high glucose intake exacerbates experimental autoimmunity by induction of TH17 cells247. Notably, increased mTORC1 signalling has also been implicated in Treg cell lineage instability. Leucine-induced mTORC1 activation drives CD4+ T cell differentiation into TH1 and TH17 cells, but Treg cell differentiation remains unaffected248. α-Ketoglutarate, a cell-permeable metabolite of glutamine, impairs Treg cell differentiation and induces the generation of pro-inflammatory T cells249. In line with this, glutamine deprivation shifts the differentiation of naive CD4+ T cells under TH1 polarizing conditions towards a Treg cell phenotype by attenuating mTORC1 activation250. Interestingly, transient inhibition of glutaminase, which catalyses the first step of glutaminolysis, induces differentiation towards a TH1 phenotype by increasing histone methylation and thereby enhancing mTORC1 signalling251.
In addition to leucine, other amino acids also regulate T cell responses. Serine fuels proliferating T cells with glycine and one-carbon units for de novo purine nucleotide synthesis (Figs. 2 and 3d), independent of glucose252. Methionine, found in meat, fish and dairy products, is a critical substrate for the methyl donor SAM (Fig. 2), which is needed for histone methylation in activated T cells, and thus regulates the expression of genes involved in Teff cell proliferation and cytokine production253. Furthermore, dietary methionine restriction reduces the expansion of pathogenic TH17 cells253. Tryptophan metabolism also has a notable role in T cell behaviour. IDO1-mediated tryptophan depletion and biosynthesis of kynurenine and its downstream metabolites drives Treg cell differentiation and blunts inflammatory responses254,255. Consistently, IDO1 deletion in dendritic cells stimulates IL-6 production and differentiation of T cells towards TH17 cells256. IDO1 expression by vSMCs has been shown to inhibit T cell infiltration into the arterial media, a phenomenon that is called the ‘medial immunoprivilege’257. Tregcells induce IDO1 expression in vSMCs, ECs and macrophages through their co-inhibitory surface receptor CTLA4. Notably, the induction of the ‘Treg–IDO axis’ in the vessel wall is atheroprotective258.
Future research investigating the following aspects will further elucidate the role of dysregulated T cell metabolism in atherosclerosis: (1) whether and how T cells, especially those responding to disease-mediating autoantigens such as ApoB, are metabolically reprogrammed during atherogenesis; (2) the impact of metabolic adaptations on lineage stability of Treg cells; and (3) whether atheroprotective immunity can be enhanced by experimental targeting of distinct metabolic pathways. Ultimately, understanding the causes and consequences of metabolic adaptations in T cells during atherosclerosis could be pivotal in development of novel immunotherapies.
Intercellular metabolic cross-talk in atherosclerosis
Studies exploring the molecular mechanisms of dysregulated metabolism in atherosclerosis often either use purified cell types in vitro or involve the deletion of critical metabolic enzymes in specific cells in vivo. However, these approaches may overlook vital intercellular communications that involve various cell types and multiple enzymes, potentially missing important connections that occur during atherosclerosis in humans. Additionally, metabolites can act as communicative factors between lesional cells. The soluble factors they produce can also influence the metabolism in distant cells, governing their fate and function. This interplay between ECs, vSMCs and immune cells can either accelerate atherosclerosis and enhance plaque instability or attenuate atherosclerosis progression and drive regression or stabilization. Additionally, alterations in metabolic pathways may be pro-atherogenic in one cell type or anti-atherogenic in others. Furthermore, due to the epidemic of obesity, dyslipidaemia and insulin resistance, the common use of medications treating these diseases is creating a new population of individuals experiencing long-term metabolic alterations. This warrants careful consideration in the design of preclinical models of atherosclerosis and the development of novel therapies thereof. The following sections highlight the mechanisms driving pathological metabolic cross-talk during atherosclerosis and cover the pleiotropic effects of commonly used metabolism-based drugs.
Intercellular metabolic communication between lesional cells
Glycolysis promotes atherosclerosis by driving EC activation and vSMC dedifferentiation while limiting the differentiation of CD4+ T cells towards Treg cells30,67,70 (Fig. 3). ECs in regions of disturbed FSS probably experience a persistent state of glycolysis30, suggesting that low levels of lactate may be released into the subendothelial space. This scenario merits exploration owing to its potential implications in vSMC dedifferentiation, proliferation, migration and calcification70,259. However, glycolysis simultaneously exerts anti-atherogenic properties by promoting efferocytosis and stimulating the production of pro-resolving mediators by macrophages182,201.
The processing of AC-derived cargo in the context of efferotabolism is metabolically demanding as evidenced by the simultaneous stimulation of glycolysis, FAO glutaminolysis and the PPP135,141,182,201,260. These changes in cellular metabolism following efferocytosis have a key role in their cross-talk with vSMCs261. Additionally, the ACs themselves release polyamines and nucleotides that directly serve as tissue messengers262. Secreted metabolites from ACs or macrophages following efferocytosis act in an autocrine or paracrine manner to further increase the successive clearance of dying cells and promote the expression of IL-10 and TGFβ261,262. However, disparate responses by other cell types may antagonize these atheroma-stabilizing properties. For example, TGFβ secreted from efferocytes may also promote EndMT, which has a central role in plaque instability43. In addition, efferocytes secrete the potent angiogenic factor vascular endothelial growth factor A, which may inadvertently enhance intraplaque angiogenesis263, which has been shown to permit haemorrhage and promote plaque instability264–266.
In addition, the mechanisms necessary to process the metabolic burden following efferocytosis are only partially understood but may involve novel and unexpected pathways. Mitochondria from an ingested AC may escape phagolysosomal degradation and fuse with the phagocyte mitochondria to increase mitochondrial mass and enhance respiration, as was seen for endosymbiosis. The exchange of respiring mitochondria between adipocytes and macrophages has also been observed267–269. Transfer of endocytic organelles via nanotubes may also occur, an emerging and exciting field of research that raises questions about how cells cope with metabolic stress270. Organelle transfer becomes particularly intriguing when considering a hypothetical scenario where non-phagocytic cells residing in atheromas could share the metabolic burden of processing ingested ACs.
The modulation of T cell metabolic reprogramming and regulation of immune responses hinge on the actions of EC-secreted metabolites271. NO and sphingosine-1-phosphate primarily shape T cell metabolism and its subsets (Fig. 3). NO, a potent messenger molecule derived from ECs, influences various physiological and pathological conditions by stimulating glycolysis and concurrently impairing mitochondrial reserve capacity272. Furthermore, NO-mediated nitrosylation of very long-chain acyl-CoA dehydrogenase273, which catalyses the first step in FAO, potentially alters T cell subset differentiation. This implies its critical role in dictating lipid metabolism and mitochondrial respiration and, thus, indirectly forming T cell subsets. By contrast, sphingosine-1-phosphate, an important EC metabolite, promotes the survival of naive T cells by maintaining mitochondrial content274,275, adding a new facet to the dynamic regulation of T cell differentiation and survival. In addition, IDO1 manages tryptophan catabolism and curbs excessive immune responses, with elevations in IDO1 activity by ECs indirectly determining immune cell apoptosis, thereby steering T cell polarization and overall inflammation276,277. These intersecting metabolic pathways form a complex network of EC–T cell cross-talk extending beyond basic cell function to dictate broader immune and inflammatory responses that could be relevant to ASCVD.
In the complex microenvironment of atherosclerotic plaques, a variety of metabolite-sensing G-protein-coupled receptors (GPCRs) have a critical role in orchestrating intercellular communication278. These GPCRs enable cross-talk between different cell types, with metabolites produced by one cell acting as extracellular signals for receptors on others. As an example, succinate, often released by hypoxic cells within atherosclerotic plaques, activates SUCNR1/GPR91 on ECs and immune cells279,280. This interaction triggers Gq/11, leading to phospholipase C-mediated signalling, calcium mobilization, NF-κB activation and inflammasome-mediated IL-1β secretion236,279,280. GPR35, responsive to kynurenic acid from the kynurenine pathway, acts conversely by dampening inflammatory responses281. However, consistent with the currently unresolved role of tryptophan metabolism in ASCVD, deleting the GPR35 receptor in haematopoietic cells does not affect atherosclerosis282. Binding of medium-chain fatty acids to GPR84 on macrophages promotes pro-inflammatory cytokine production283. Additionally, short-chain fatty acids, such as acetate and propionate, activate OLFR78 on vSMCs, altering vascular tone and blood pressure284,285. These GPCRs highlight the dynamic and multifaceted role of metabolites in governing intercellular communication within atherosclerotic lesions.
Effects of metabolism-altering therapeutics on lesional cells Statins.
Statins are best known for lowering cholesterol but exhibit atheroprotective pleiotropic effects even in normocholesterolaemic settings286. One such benefit is the inhibition of leukocyte–endothelial interactions due to decreased adhesion molecule presentation on ECs287. Additionally, inhibiting HMG-CoA reductase by statins prevents isoprenylation, lowers membrane levels of small GTPases, increases NO production by ECs and enhances vascular reactivity288,289. These effects probably involve statin-induced increases in KLF2 driving the expression of anti-inflammatory genes, including eNOS, which decreases pro-inflammatory signalling by inhibiting NF-κB and YAP/TAZ29,290–292. Despite these benefits, nearly 50% of individuals treated with statins still experience recurrent cardiovascular events293. Notably, the rate of statin use among the global population is increasing, coinciding with a shift in the prevalence of acute clinical events towards superficial plaque erosion rather than plaque rupture294,295. Plaque erosion has been mechanistically linked to the formation of neutrophil extracellular traps (NETs)296. EC-targeting nanoparticles carrying inhibitors to peptidyl arginine deiminase-4, which mediates histone citrullination necessary for NETosis297, mitigate the formation of NETs and preserves endothelial integrity298.
ACLY inhibitors.
ACLY has emerged as a crucial enzyme linking carbohydrate and lipid metabolism. By converting citrate to acetyl-CoA, ACLY has a vital role in fatty acid and cholesterol biosynthesis299, processes central to the pathophysiology of ASCVD. Bempedoic acid (ETC-1002) is a first-in-class inhibitor of ACLY, offering a unique approach to modulating lipid metabolism300. By inhibiting ACLY, bempedoic acid increases hepatic LDLR expression and lowers circulating cholesterol300. In statin-intolerant patients, bempedoic acid reduced the risk of major adverse cardiovascular events301. Interestingly, the use of bempedoic acid extends beyond its primary application as a lipid-lowering agent, potentially exerting anti-atherogenic effects by modulating macrophage-mediated inflammation within atherosclerotic lesions. As an example, ACLY is activated in inflammatory macrophages and its deletion from macrophages enhances collagen cap thickness and decreases necrotic core area127.
Thiazolidinediones.
Thiazolidinediones have been lauded for their ability to lower blood glucose in type 2 diabetes, primarily by enhancing insulin sensitivity in peripheral tissues, leading to improved glucose utilization and reduced insulin resistance302. Additionally, pioglitazone has demonstrated beneficial effects on lipid profiles, increasing HDL cholesterol and reducing triglyceride levels303, with evidence of reduced atherosclerosis304. However, pioglitazone also results in weight gain, a bona fide risk factor for atherosclerosis305. Its tendency to cause fluid retention leading to oedema could also exacerbate heart failure in susceptible patients. Moreover, while enhancing macrophage apoptosis in a PPARγ-independent manner, pioglitazone and rosiglitazone increase AC clearance by macrophages, albeit with a net effect of increasing ACs and plaque necrosis306. Thus, the above risks highlight the importance of individualizing therapy, weighing the beneficial glucose-lowering and possible cardiovascular benefits against potential complications.
Metformin.
Metformin, primarily used for type 2 diabetes management, benefits ASCVD beyond its glucose-lowering capability307,308. For instance, metformin increases eNOS-mediated NO production in ECs309–311. Through a similar mechanism, metformin also suppresses vascular calcification and vSMC senescence312,313. As another example, metformin drives macrophage polarization towards a pro-resolving phenotype via regulating an AMPK–mTOR–NLRP3 signalling axis that accelerates wound healing314. As discussed below, metformin, in combination with drug-eluting stents (DESs), delays re-endothelialization and, potentially, heightens the risk for in-stent thrombosis315.
Insulin.
Whereas metformin exerts multiple beneficial effects beyond its glucose-lowering capability, insulin therapy has a more intricate relationship with respect to ASCVD. Insulin is an irreplaceable component in managing type 1 diabetes and frequently becomes necessary in the later stages of type 2 diabetes316. Its potent glucose-lowering abilities improve glycaemic control, which limits microvascular complications such as nephropathy, retinopathy and neuropathy317. Nonetheless, insulin therapy carries challenges. Using exogenous insulin can lead to weight gain, potentially exacerbating insulin resistance and the risk of atherosclerosis318. High-dose insulin therapy has also been implicated in cardiovascular concerns, promoting sodium retention, vasoconstriction and vSMC proliferation, migration and inflammation, factors that contribute to atherosclerosis319–321. The relationship between insulin therapy and atherosclerosis is complex and influenced by factors such as the patient’s underlying metabolic status and concurrent therapies322.
mTOR inhibitors.
The advent of first-generation DESs coated with the potent mTOR inhibitor rapamycin (for example, sirolimus) marked an important advance in interventional cardiology. This was primarily due to its superior ability to maintain vessel patency compared to its bare-metal counterparts323. Acting as a central regulator of nutrient sensing and metabolism, mTOR has a pivotal role in the proliferation and migration of vSMCs. Inhibiting the mTOR pathway using rapamycin induces CNN1, ACTA2 and TAGLN gene expression, all associated with a quiescent vSMC phenotype324–326. This inhibition also blocks nutrient sensing, impeding cell cycle progression and subsequent proliferation. This, in turn, reduces restenosis, offering a compelling advantage over bare-metal stents323. However, these first-generation DESs also delay re-endothelialization, thereby heightening the risk of stent thrombosis327–329. As a result, patients receiving first-generation DESs require an extended period of anticoagulation therapy. Interestingly, newer-generation DESs, which still contain rapamycin or its derivatives, are overcoming this hurdle due to improved drug release mechanisms, polymer coatings and stent design330–332. Thus, the success of mTOR inhibitors in DESs underscores the potential of exploiting metabolic pathways for therapeutic purposes.
Conclusions and future directions
Our understanding of ASCVD has evolved beyond the simplistic notion of lipid accumulation, embracing the intricate interplay of dysregulated cellular metabolism and complex interactions between cell types that drive disease progression. To optimize cellular metabolism and address the persistent cardiovascular risk despite existing therapies, we must prioritize critical areas for exploration. Understanding the complex links between metabolic pathways, cellular activities and cell differentiation within the plaque microenvironment is essential. This knowledge will drive the development of targeted therapeutics that address specific metabolic pathways. In this regard, the explosion of new metabolomic technologies, such as single-cell metabolomics, MSI, spatial transcriptomics and cytometry by time of flight (CyTOF), has begun to reveal the complexity of cell metabolism in atherosclerosis (Box 1).
Box 1. Single-cell technologies reveal novel insights into atherosclerosis.
Unbiased single-cell technologies, such as scRNA-seq, single-cell assay for transposase-accessible chromatin with high-throughput sequencing, CITE sequencing and single-cell mass cytometry, are powerful platforms that have ushered in a new age for understanding cellular dynamics while also revealing molecular signatures of newly identified cell types in atherosclerotic plaques335–339. Specifically, these technologies have permitted cellular indexing of transcriptomes (scRNA-seq), allowed for the analysis of genetically open regions of chromatin that are available for active transcription (single-cell assay for transposase-accessible chromatin with high-throughput sequencing), elucidated cell-surface markers and epitopes (CITE sequencing) and profiled single-cell proteomics in heterogeneous samples using metal isotopes (CyTOF). Furthermore, advancements in computational modelling have allowed for the deconstruction of plaques to reveal cell–cell communications through specific ligand–receptor interactions. These approaches have revealed inflammatory Il1b+Nlrp3+ macrophages in plaques97. The clinical relevance of these findings was highlighted by the CANTOS trial where blocking IL-1β signalling reduced cardiovascular events12. CyTOF uncovered the expansion of monocytes, plasmacytoid dendritic cells and CD11c+ immune cells with a simultaneous diminishment of CD206+CD169+ macrophages and type 2 conventional cells during atherosclerosis340. In the non-immune cell compartment, single-cell omics platforms confirmed the plasticity of vSMCs and revealed a new vSMC-derived population, termed the ‘SEM’, which can either maintain vSMC properties or differentiate into macrophage-like or fibrochondrocyte-like cells56. Additionally, spatial transcriptomics using MERFISH identified a new population of dedifferentiated vSMCs in males dominated by the abundance of CARTPT341. Profiling the transcriptional landscape of human atheromas revealed genes in vSMCs commonly associated with bone mineralization, vascular calcification and matrix remodelling in the necrotic cores of plaques. Lineage tracing coupled to scRNA-seq further identified Tcf21 as a critical component of vSMC dedifferentiation and their trajectory resembling other cell types55. Interestingly, stem cell antigen 1+ vSMCs, which show a loss of expression in differentiated vSMC markers50, were found in the core of atherosclerotic plaques but were absent in the fibrous cap53.
Although these technologies have provided invaluable insights into atherosclerosis, they also open new avenues for exploration and highlight gaps in our understanding. The complexity of cellular interactions within plaques, particularly in metabolic processes, remains a challenge. Advances in single-cell metabolomics, such as SCENITH and SpaceM333,334, promises to unravel this complexity. As an example, a recent study combining RNA-seq and MSI in human atheromas uncovered previously unrecognized metabolic pathways involved in plaque instability69. Future explorations integrating MSI with mass-tagged antibodies could offer profound insights by merging single-cell, spatial metabolomics with spatial transcriptomics. Nonetheless, integration of data from genomics, transcriptomics, proteomics and metabolomics in a broader approach will require continued advances in computational methods, machine learning and imaging techniques.
Additionally, methods to target metabolic pathways for treating ASCVD, including the exploration of cell-type-specific, nanoparticle-based strategies for delivering inhibitors, mRNA and RNAi, hold great promise. Although isolated metabolic pathways are relatively well understood, further exploration is needed to comprehensively characterize the contribution of noncanonical pathways, dysregulation within these pathways and the mechanisms underlying intercellular cross-talk. Innovation in these areas is set to provide a more comprehensive understanding of atherosclerosis at both the molecular and cellular levels, ultimately revealing potential biomarkers and novel therapeutic strategies to treat ASCVD.
One area deserving further investigation is the protective role of efferotabolism in atherosclerosis, which remains understudied. Strategies that increase lesional efferocytosis have been repeatedly shown to mitigate atherosclerosis progression and even enhance its regression. However, macrophages must undergo metabolic reprogramming to effectively take up and process multiple ACs. Thus, gaining a deeper understanding of the interplay among various metabolic processes in efferotabolism will not only shed light on the progression of ASCVD but also reveal strategies for therapeutic intervention. Moreover, because most research interrogating EC metabolism has been predominantly conducted under static conditions, the metabolism of ECs under pathophysiological conditions in the setting of FSS warrants attention. A comprehensive understanding of EC metabolism is critical, given their role on leukocyte adhesion and phenotypic transition to mesenchymal cells. Similarly, insights from T cell studies in autoimmune diseases could deepen our understanding of their metabolism in atherosclerosis. This is particularly important as recent studies have clearly demonstrated that alterations in T cell specification rely on metabolic reprogramming. Bridging knowledge gaps between interconnected metabolic pathways within lesional cells could open new avenues for therapeutic manipulation. By focusing on these underexplored aspects of dysregulated cellular metabolism and their interdependencies, we can unlock a new era of effective interventions to mitigate ASCVD (Boxes 1 and 2).
Box 2. From lipids to proteins and amino acids: a paradigm shift in nutrient metabolism and atherosclerosis.
The critical role of lipid/lipoprotein metabolism in ASCVD has been extensively studied, leading to the development of effective therapeutics that reduce LDL cholesterol and cardiovascular risk342,343. Despite this, numerous individuals with normal LDL cholesterol levels still suffer from major cardiovascular events344,345, emphasizing the importance of other metabolic pathways. Over the past decades, a growing body of evidence uncovered essential roles for protein and amino acids in ASCVD pathogenesis. In Apoe−/− mice, high-protein diets enhance the size and complexity of atherosclerotic plaques346,347. In humans, however, high-protein intake can either increase or decrease the cardiovascular risk348–350, highlighting the importance of understanding the roles of specific amino acids, rather than total protein intake.
Among 79,838 participants, a higher intake of essential amino acids (for example, branched-chained amino acids (BCAAs)) was associated with increased cardiovascular mortality, whereas a higher intake of non-essential amino acids (for example, glycine) was associated with a lower risk351. Individuals with ASCVD show a distinct pattern of circulating amino acids, where BCAAs are increased, and glycine is decreased in association with disease severity171–173,352,353. Although it was shown that increased BCAAs drive glycine depletion through activation of the pyruvate–alanine cycle354, recent isotope tracing studies revealed that increased synthesis of serine through reversed serine hydroxymethyltransferase activity drives glycine depletion independent of BCAAs355. Consistent with its lower levels in coronary artery disease, glycine depletion enhances, whereas glycine-based treatment reduces, atherosclerosis in Apoe−/− mice through lipid-lowering effects and inducing de novo GSH biosynthesis in macrophages173,356. The role of BCAAs, however, is controversial357,358. Supplementation of BCAAs or leucine alone lowers atherosclerosis in Apoe−/− mice by reducing LDL cholesterol and pro-atherogenic/inflammatory chemokines359,360. In lipid-laden macrophages, leucine reduces intracellular cholesterol and triglycerides by increasing cholesterol efflux and inhibiting very-low-density lipoprotein uptake361,362. By contrast, leucine activates mTORC1 in macrophages and synergizes with atherogenic lipids to enhance mitochondrial dysfunction and apoptosis347. Thus, although higher circulating BCAAs are linked with ASCVD, dysregulated BCAA metabolism and the effects of specific BCAAs on different cell types of the atherosclerotic plaque warrant further research.
Acknowledgements
We acknowledge funding from an American Heart Association Postdoctoral Fellowship (23POST1026505) to A.C.F.; a Louisiana State University Health Sciences Center, Shreveport and Center for Cardiovascular Diseases and Sciences Postdoctoral Fellowship to A.C.F.; a Louisiana State University Health Sciences Center, Shreveport, CIRP, and Chancellor’s Pathways Research Award to A.Y.J. and O.R.; Deutsche Forschungsgemeinschaft (NE 2574/1-1) to F.S.N.; National Institutes of Health R01 HL135582 to M.S.; P01 HL136275 and R35 HL145241 to K.L.; R00 HL150233, R01 DK134011 and R01 DK136685 to O.R.; R00 HL145131 and R01 HL167758 to A.Y.J.
Footnotes
Competing interests
K.L. is founder and co-owner of Atherovax and receives no compensation from Atherovax. No Atherovax funds were used in this study. O.R. is a scientific advisor at Diapin Therapeutics and receives no compensation from Diapin Therapeutics. No Diapin Therapeutics funds were used in this study. All other authors declare no competing interests.
Additional information
Peer review information Nature Metabolism thanks Daniel Ketelhuth, Laurent Yvan-Charvet and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Primary Handling Editor: Christoph Schmitt, in collaboration with the Nature Metabolism team.
Reprints and permissions information is available at www.nature.com/reprints.
References
- 1.Libby P et al. Atherosclerosis. Nat. Rev. Dis. Primers 5, 56 (2019). [DOI] [PubMed] [Google Scholar]
- 2.Rahman MS, Murphy AJ & Woollard KJ Effects of dyslipidaemia on monocyte production and function in cardiovascular disease. Nat. Rev. Cardiol. 14, 387–400 (2017). [DOI] [PubMed] [Google Scholar]
- 3.Lim GB Hyperglycaemia-induced trained immunity promotes atherosclerosis. Nat. Rev. Cardiol. 18, 687 (2021). [DOI] [PubMed] [Google Scholar]
- 4.Nitz K, Lacy M & Atzler D Amino acids and their metabolism in atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 39, 319–330 (2019). [DOI] [PubMed] [Google Scholar]
- 5.Yurdagul A Jr., Finney AC, Woolard MD & Orr AW The arterial microenvironment: the where and why of atherosclerosis. Biochem. J. 473, 1281–1295 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Hansson GK, Libby P & Tabas I Inflammation and plaque vulnerability. J. Intern. Med. 278, 483–493 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Partida RA, Libby P, Crea F & Jang IK Plaque erosion: a new in vivo diagnosis and a potential major shift in the management of patients with acute coronary syndromes. Eur. Heart J. 39, 2070–2076 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Michos ED, McEvoy JW & Blumenthal RS Lipid management for the prevention of atherosclerotic cardiovascular disease. N. Engl. J. Med. 381, 1557–1567 (2019). [DOI] [PubMed] [Google Scholar]
- 9.Tsao CW et al. Heart Disease and Stroke Statistics–2023 Update: a report from the American Heart Association. Circulation 147, e93–e621 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Back M, Yurdagul A Jr., Tabas I, Oorni K & Kovanen PT Inflammation and its resolution in atherosclerosis: mediators and therapeutic opportunities. Nat. Rev. Cardiol. 16, 389–406 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Doran AC Inflammation resolution: implications for atherosclerosis. Circ. Res. 130, 130–148 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Ridker PM et al. Antiinflammatory therapy with canakinumab for atherosclerotic disease. N. Engl. J. Med. 377, 1119–1131 (2017). [DOI] [PubMed] [Google Scholar]
- 13.Tardif JC et al. Efficacy and safety of low-dose colchicine after myocardial infarction. N. Engl. J. Med. 381, 2497–2505 (2019). [DOI] [PubMed] [Google Scholar]
- 14.Hahn C & Schwartz MA Mechanotransduction in vascular physiology and atherogenesis. Nat. Rev. Mol. Cell Biol. 10, 53–62 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Mullick AE et al. Increased endothelial expression of Toll-like receptor 2 at sites of disturbed blood flow exacerbates early atherogenic events. J. Exp. Med. 205, 373–383 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Humphrey JD, Dufresne ER & Schwartz MA Mechanotransduction and extracellular matrix homeostasis. Nat. Rev. Mol. Cell Biol. 15, 802–812 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Boren J & Williams KJ The central role of arterial retention of cholesterol-rich apolipoprotein-B-containing lipoproteins in the pathogenesis of atherosclerosis: a triumph of simplicity. Curr. Opin. Lipidol. 27, 473–483 (2016). [DOI] [PubMed] [Google Scholar]
- 18.Ramirez CM et al. Caveolin-1 regulates atherogenesis by attenuating low-density lipoprotein transcytosis and vascular inflammation independently of endothelial nitric oxide synthase activation. Circulation 140, 225–239 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Huang L et al. SR-B1 drives endothelial cell LDL transcytosis via DOCK4 to promote atherosclerosis. Nature 569, 565–569 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Camejo G, Olofsson SO, Lopez F, Carlsson P & Bondjers G Identification of Apo B-100 segments mediating the interaction of low density lipoproteins with arterial proteoglycans. Arteriosclerosis 8, 368–377 (1988). [DOI] [PubMed] [Google Scholar]
- 21.Tsiantoulas D et al. APRIL limits atherosclerosis by binding to heparan sulfate proteoglycans. Nature 597, 92–96 (2021). [DOI] [PubMed] [Google Scholar]
- 22.Li Y et al. Cholesterol-induced apoptotic macrophages elicit an inflammatory response in phagocytes, which is partially attenuated by the Mer receptor. J. Biol. Chem. 281, 6707–6717 (2006). [DOI] [PubMed] [Google Scholar]
- 23.Schrijvers DM, De Meyer GR, Kockx MM, Herman AG & Martinet W Phagocytosis of apoptotic cells by macrophages is impaired in atherosclerosis. Arterioscler Thromb. Vasc. Biol. 25, 1256–1261 (2005). [DOI] [PubMed] [Google Scholar]
- 24.Kojima Y, Weissman IL & Leeper NJ The role of efferocytosis in atherosclerosis. Circulation 135, 476–489 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Humphrey JD & Schwartz MA Vascular mechanobiology: homeostasis, adaptation, and disease. Annu. Rev. Biomed. Eng. 23, 1–27 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Rohlenova K, Veys K, Miranda-Santos I, De Bock K & Carmeliet P Endothelial cell metabolism in health and disease. Trends Cell Biol. 28, 224–236 (2018). [DOI] [PubMed] [Google Scholar]
- 27.Enzo E et al. Aerobic glycolysis tunes YAP/TAZ transcriptional activity. EMBO J. 34, 1349–1370 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.White SM et al. YAP/TAZ inhibition induces metabolic and signaling rewiring resulting in targetable vulnerabilities in NF2-deficient tumor cells. Dev. Cell 49, 425–443 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Wang KC et al. Flow-dependent YAP/TAZ activities regulate endothelial phenotypes and atherosclerosis. Proc. Natl Acad. Sci. USA 113, 11525–11530 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Wu D et al. HIF-1alpha is required for disturbed flow-induced metabolic reprogramming in human and porcine vascular endothelium. Elife 10.7554/eLife.25217 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Yang Q et al. PRKAA1/AMPKα1-driven glycolysis in endothelial cells exposed to disturbed flow protects against atherosclerosis. Nat. Commun. 9, 4667 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Atkins GB & Jain MK Role of Kruppel-like transcription factors in endothelial biology. Circ. Res. 100, 1686–1695 (2007). [DOI] [PubMed] [Google Scholar]
- 33.Doddaballapur A et al. Laminar shear stress inhibits endothelial cell metabolism via KLF2-mediated repression of PFKFB3. Arterioscler. Thromb. Vasc. Biol. 35, 137–145 (2015). [DOI] [PubMed] [Google Scholar]
- 34.Coon BG et al. A mitochondrial contribution to anti-inflammatory shear stress signaling in vascular endothelial cells. J. Cell Biol. 10.1083/jcb.202109144 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Goldberg IJ et al. Lipolytic enzymes and free fatty acids at the endothelial interface. Atherosclerosis 329, 1–8 (2021). [DOI] [PubMed] [Google Scholar]
- 36.Mehrotra D, Wu J, Papangeli I & Chun HJ Endothelium as a gatekeeper of fatty acid transport. Trends Endocrinol. Metab. 25, 99–106 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Abumrad NA et al. Endothelial cell receptors in tissue lipid uptake and metabolism. Circ. Res. 128, 433–450 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Ibrahim A, Yucel N, Kim B & Arany Z Local mitochondrial ATP production regulates endothelial fatty acid uptake and transport. Cell Metab. 32, 309–319 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Kraehling JR et al. Genome-wide RNAi screen reveals ALK1 mediates LDL uptake and transcytosis in endothelial cells. Nat. Commun. 7, 13516 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Libby P, Ridker PM & Hansson GK Progress and challenges in translating the biology of atherosclerosis. Nature 473, 317–325 (2011). [DOI] [PubMed] [Google Scholar]
- 41.Zhou Z et al. Lipoprotein-derived lysophosphatidic acid promotes atherosclerosis by releasing CXCL1 from the endothelium. Cell Metab. 13, 592–600 (2011). [DOI] [PubMed] [Google Scholar]
- 42.Chen PY et al. Endothelial-to-mesenchymal transition drives atherosclerosis progression. J. Clin. Invest. 125, 4514–4528 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Evrard SM et al. Endothelial to mesenchymal transition is common in atherosclerotic lesions and is associated with plaque instability. Nat. Commun. 7, 11853 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Chen PY et al. Endothelial TGF-beta signalling drives vascular inflammation and atherosclerosis. Nat. Metab. 1, 912–926 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Zhu X et al. Acetate controls endothelial-to-mesenchymal transition. Cell Metab. 10.1016/j.cmet.2023.05.010 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Bennett MR, Sinha S & Owens GK Vascular smooth muscle cells in atherosclerosis. Circ. Res. 118, 692–702 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Sorokin V et al. Role of vascular smooth muscle cell plasticity and interactions in vessel wall inflammation. Front Immunol. 11, 599415 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Worssam MD & Jorgensen HF Mechanisms of vascular smooth muscle cell investment and phenotypic diversification in vascular diseases. Biochem. Soc. Trans. 49, 2101–2111 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Liu M & Gomez D Smooth muscle cell phenotypic diversity. Arterioscler. Thromb. Vasc. Biol. 39, 1715–1723 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Dobnikar L et al. Disease-relevant transcriptional signatures identified in individual smooth muscle cells from healthy mouse vessels. Nat. Commun. 9, 4567 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Kovacic JC et al. Endothelial to mesenchymal transition in cardiovascular disease: JACC state-of-the-art review. J. Am. Coll. Cardiol. 73, 190–209 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Chen J, Kitchen CM, Streb JW & Miano JM Myocardin: a component of a molecular switch for smooth muscle differentiation. J. Mol. Cell. Cardiol. 34, 1345–1356 (2002). [DOI] [PubMed] [Google Scholar]
- 53.Shankman LS et al. KLF4-dependent phenotypic modulation of smooth muscle cells has a key role in atherosclerotic plaque pathogenesis. Nat. Med. 21, 628–637 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Yap C, Mieremet A, de Vries CJM, Micha D & de Waard V Six shades of vascular smooth muscle cells illuminated by KLF4 (kruppel-like factor 4). Arterioscler. Thromb. Vasc. Biol. 41, 2693–2707 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Wirka RC et al. Atheroprotective roles of smooth muscle cell phenotypic modulation and the TCF21 disease gene as revealed by single-cell analysis. Nat. Med. 25, 1280–1289 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Pan H et al. Single-cell genomics reveals a novel cell state during smooth muscle cell phenotypic switching and potential therapeutic targets for atherosclerosis in mouse and human. Circulation 142, 2060–2075 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Vengrenyuk Y et al. Cholesterol loading reprograms the microRNA-143/145-myocardin axis to convert aortic smooth muscle cells to a dysfunctional macrophage-like phenotype. Arterioscler. Thromb. Vasc. Biol. 35, 535–546 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Alencar GF et al. Stem cell pluripotency genes Klf4 and Oct4 regulate complex SMC phenotypic changes critical in late-stage atherosclerotic lesion pathogenesis. Circulation 142, 2045–2059 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Malhotra R et al. HDAC9 is implicated in atherosclerotic aortic calcification and affects vascular smooth muscle cell phenotype. Nat. Genet. 51, 1580–1587 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Sangiorgi G et al. Arterial calcification and not lumen stenosis is highly correlated with atherosclerotic plaque burden in humans: a histologic study of 723 coronary artery segments using nondecalcifying methodology. J. Am. Coll. Cardiol. 31, 126–133 (1998). [DOI] [PubMed] [Google Scholar]
- 61.Bos D et al. Comparison of atherosclerotic calcification in major vessel beds on the risk of all-cause and cause-specific mortality: The Rotterdam Study. Circ. Cardiovasc. Imaging 10.1161/CIRCIMAGING.115.003843 (2015). [DOI] [PubMed] [Google Scholar]
- 62.Guillermier C et al. Imaging mass spectrometry reveals heterogeneity of proliferation and metabolism in atherosclerosis. JCI Insight 10.1172/jci.insight.128528 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Perez J, Hill BG, Benavides GA, Dranka BP & Darley-Usmar VM Role of cellular bioenergetics in smooth muscle cell proliferation induced by platelet-derived growth factor. Biochem. J. 428, 255–267 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Wall VZ et al. Smooth muscle glucose metabolism promotes monocyte recruitment and atherosclerosis in a mouse model of metabolic syndrome. JCI Insight 3, e96544 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Jain M et al. Smooth muscle cell-specific PKM2 (pyruvate kinase muscle 2) promotes smooth muscle cell phenotypic switching and neointimal hyperplasia. Arterioscler. Thromb. Vasc. Biol. 41, 1724–1737 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Zhao X et al. PKM2-dependent glycolysis promotes the proliferation and migration of vascular smooth muscle cells during atherosclerosis. Acta Biochim. Biophys. Sin. 52, 9–17 (2020). [DOI] [PubMed] [Google Scholar]
- 67.Newman AAC et al. Multiple cell types contribute to the atherosclerotic lesion fibrous cap by PDGFRbeta and bioenergetic mechanisms. Nat. Metab. 3, 166–181 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Perry RN, Albarracin D, Aherrahrou R & Civelek M Network preservation analysis reveals dysregulated metabolic pathways in human vascular smooth muscle cell phenotypic switching. Circ. Genom. Precis. Med. 16, e003781 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Seeley EH et al. Spatially resolved metabolites in stable and unstable human atherosclerotic plaques identified by mass spectrometry imaging. Arterioscler. Thromb. Vasc. Biol. 10.1161/ATVBAHA.122.318684 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Yang L et al. Lactate promotes synthetic phenotype in vascular smooth muscle cells. Circ. Res. 121, 1251–1262 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Shi J, Yang Y, Cheng A, Xu G & He F Metabolism of vascular smooth muscle cells in vascular diseases. Am. J. Physiol. Heart Circ. Physiol. 319, H613–H631 (2020). [DOI] [PubMed] [Google Scholar]
- 72.Peiro C et al. Inflammation, glucose, and vascular cell damage: the role of the pentose phosphate pathway. Cardiovasc. Diabetol. 15, 82 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Dong LH et al. TRAF6-mediated SM22alpha K21 ubiquitination promotes G6PD activation and NADPH production, contributing to GSH homeostasis and VSMC survival in vitro and in vivo. Circ. Res. 117, 684–694 (2015). [DOI] [PubMed] [Google Scholar]
- 74.Clarke MC et al. Apoptosis of vascular smooth muscle cells induces features of plaque vulnerability in atherosclerosis. Nat. Med. 12, 1075–1080 (2006). [DOI] [PubMed] [Google Scholar]
- 75.Liu GY & Sabatini DM mTOR at the nexus of nutrition, growth, ageing and disease. Nat. Rev. Mol. Cell Biol. 21, 183–203 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Reho JJ, Guo DF, Morgan DA & Rahmouni K mTORC1 (mechanistic target of rapamycin complex 1) signaling in endothelial and smooth muscle cells is required for vascular function. Hypertension 77, 594–604 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Li G et al. Chronic mTOR activation induces a degradative smooth muscle cell phenotype. J. Clin. Invest. 130, 1233–1251 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Zhao Y, Vanhoutte PM & Leung SW Vascular nitric oxide: beyond eNOS. J. Pharm. Sci. 129, 83–94 (2015). [DOI] [PubMed] [Google Scholar]
- 79.Tzeng HP, Lan KC, Yang TH, Chung MN & Liu SH Benzo[a]pyrene activates interleukin-6 induction and suppresses nitric oxide-induced apoptosis in rat vascular smooth muscle cells. PLoS ONE 12, e0178063 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Ignarro LJ et al. Role of the arginine-nitric oxide pathway in the regulation of vascular smooth muscle cell proliferation. Proc. Natl Acad. Sci. USA 98, 4202–4208 (2001). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Grossi M et al. Inhibition of polyamine formation antagonizes vascular smooth muscle cell proliferation and preserves the contractile phenotype. Basic Clin. Pharmacol. Toxicol. 115, 379–388 (2014). [DOI] [PubMed] [Google Scholar]
- 82.Wang XP et al. Arginase I enhances atherosclerotic plaque stabilization by inhibiting inflammation and promoting smooth muscle cell proliferation. Eur. Heart J. 35, 911–919 (2014). [DOI] [PubMed] [Google Scholar]
- 83.Michiels CF, Kurdi A, Timmermans JP, De Meyer GRY & Martinet W Spermidine reduces lipid accumulation and necrotic core formation in atherosclerotic plaques via induction of autophagy. Atherosclerosis 251, 319–327 (2016). [DOI] [PubMed] [Google Scholar]
- 84.Ouyang L et al. Indoleamine 2,3-dioxygenase 1 deletion-mediated kynurenine insufficiency in vascular smooth muscle cells exacerbates arterial calcification. Circulation 145, 1784–1798 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Kim JB et al. Environment-sensing aryl hydrocarbon receptor inhibits the chondrogenic fate of modulated smooth muscle cells in atherosclerotic lesions. Circulation 142, 575–590 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Li M, Kwok MK, Fong SSM & Schooling CM Indoleamine 2,3-dioxygenase and ischemic heart disease: a Mendelian randomization study. Sci. Rep. 9, 8491 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Niinisalo P et al. Indoleamine 2,3-dioxygenase activity associates with cardiovascular risk factors: the Health 2000 study. Scand. J. Clin. Lab. Invest. 68, 767–770 (2008). [DOI] [PubMed] [Google Scholar]
- 88.Brouns R et al. The role of tryptophan catabolism along the kynurenine pathway in acute ischemic stroke. Neurochem. Res. 35, 1315–1322 (2010). [DOI] [PubMed] [Google Scholar]
- 89.Wang Q et al. Tryptophan-derived 3-hydroxyanthranilic acid contributes to angiotensin II-induced abdominal aortic aneurysm formation in mice in vivo. Circulation 136, 2271–2283 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Moore KJ & Tabas I Macrophages in the pathogenesis of atherosclerosis. Cell 145, 341–355 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Galvan-Pena S & O’Neill LA Metabolic reprograming in macrophage polarization. Front. Immunol. 5, 420 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Viola A, Munari F, Sanchez-Rodriguez R, Scolaro T & Castegna A The metabolic signature of macrophage responses. Front. Immunol. 10, 1462 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Wculek SK, Dunphy G, Heras-Murillo I, Mastrangelo A & Sancho D Metabolism of tissue macrophages in homeostasis and pathology. Cell Mol. Immunol. 19, 384–408 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Sarrazy V et al. Maintenance of macrophage redox status by ChREBP limits inflammation and apoptosis and protects against advanced atherosclerotic lesion formation. Cell Rep. 13, 132–144 (2015). [DOI] [PubMed] [Google Scholar]
- 95.Haschemi A et al. The sedoheptulose kinase CARKL directs macrophage polarization through control of glucose metabolism. Cell Metab. 15, 813–826 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Puleston DJ et al. Polyamines and eIF5A hypusination modulate mitochondrial respiration and macrophage activation. Cell Metab. 30, 352–363 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Cochain C et al. Single-cell RNA-seq reveals the transcriptional landscape and heterogeneity of aortic macrophages in murine atherosclerosis. Circ. Res. 122, 1661–1674 (2018). [DOI] [PubMed] [Google Scholar]
- 98.Guo X et al. TREM2 promotes cholesterol uptake and foam cell formation in atherosclerosis. Cell. Mol. Life Sci. 80, 137 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Wang N & Westerterp M ABC transporters, cholesterol efflux, and implications for cardiovascular diseases. Adv. Exp. Med. Biol. 1276, 67–83 (2020). [DOI] [PubMed] [Google Scholar]
- 100.Sheedy FJ et al. CD36 coordinates NLRP3 inflammasome activation by facilitating intracellular nucleation of soluble ligands into particulate ligands in sterile inflammation. Nat. Immunol. 14, 812–820 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Duewell P et al. NLRP3 inflammasomes are required for atherogenesis and activated by cholesterol crystals. Nature 464, 1357–1361 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Kim K et al. Transcriptome analysis reveals nonfoamy rather than foamy plaque macrophages are proinflammatory in atherosclerotic murine models. Circ. Res. 123, 1127–1142 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Park YM, Febbraio M & Silverstein RL CD36 modulates migration of mouse and human macrophages in response to oxidized LDL and may contribute to macrophage trapping in the arterial intima. J. Clin. Invest. 119, 136–145 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.van Gils JM et al. The neuroimmune guidance cue netrin-1 promotes atherosclerosis by inhibiting the emigration of macrophages from plaques. Nat. Immunol. 13, 136–143 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Hou P et al. Macrophage polarization and metabolism in atherosclerosis. Cell Death Dis. 14, 691 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Spann NJ et al. Regulated accumulation of desmosterol integrates macrophage lipid metabolism and inflammatory responses. Cell 151, 138–152 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Doran AC, Yurdagul A Jr. & Tabas I Efferocytosis in health and disease. Nat. Rev. Immunol. 20, 254–267 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Dang EV & Cyster JG Loss of sterol metabolic homeostasis triggers inflammasomes - how and why. Curr. Opin. Immunol. 56, 1–9 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.de la Roche M et al. Trafficking of cholesterol to the ER is required for NLRP3 inflammasome activation. J. Cell Biol. 217, 3560–3576 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Tall AR & Westerterp M Inflammasomes, neutrophil extracellular traps, and cholesterol. J. Lipid Res. 60, 721–727 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Grebe A, Hoss F & Latz E NLRP3 inflammasome and the IL-1 pathway in atherosclerosis. Circ. Res. 122, 1722–1740 (2018). [DOI] [PubMed] [Google Scholar]
- 112.Soehnlein O & Libby P Targeting inflammation in atherosclerosis - from experimental insights to the clinic. Nat. Rev. Drug Discov. 20, 589–610 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Hendrikx T et al. Bone marrow-specific caspase-1/11 deficiency inhibits atherosclerosis development in Ldlr−/− mice. FEBS J. 282, 2327–2338 (2015). [DOI] [PubMed] [Google Scholar]
- 114.Zeng W et al. The selective NLRP3 inhibitor MCC950 hinders atherosclerosis development by attenuating inflammation and pyroptosis in macrophages. Sci. Rep. 11, 19305 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Fukuzumi M, Shinomiya H, Shimizu Y, Ohishi K & Utsumi S Endotoxin-induced enhancement of glucose influx into murine peritoneal macrophages via GLUT1. Infect. Immun. 64, 108–112 (1996). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Nishizawa T et al. Testing the role of myeloid cell glucose flux in inflammation and atherosclerosis. Cell Rep. 7, 356–365 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Ruiz-Garcia A et al. Cooperation of adenosine with macrophage Toll-4 receptor agonists leads to increased glycolytic flux through the enhanced expression of PFKFB3 gene. J. Biol. Chem. 286, 19247–19258 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Palsson-McDermott EM et al. Pyruvate kinase M2 regulates Hif-1α activity and IL-1β induction and is a critical determinant of the warburg effect in LPS-activated macrophages. Cell Metab. 21, 65–80 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Tannahill GM et al. Succinate is an inflammatory signal that induces IL-1β through HIF-1α. Nature 496, 238–242 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Van den Bossche J, O’Neill LA & Menon D Macrophage immunometabolism: where are we (going)? Trends Immunol. 38, 395–406 (2017). [DOI] [PubMed] [Google Scholar]
- 121.O’Rourke SA et al. Cholesterol crystals drive metabolic reprogramming and M1 macrophage polarisation in primary human macrophages. Atherosclerosis 352, 35–45 (2022). [DOI] [PubMed] [Google Scholar]
- 122.Rogers IS et al. Feasibility of FDG imaging of the coronary arteries: comparison between acute coronary syndrome and stable angina. JACC Cardiovasc. Imaging 3, 388–397 (2010). [DOI] [PubMed] [Google Scholar]
- 123.Rudd JH et al. Atherosclerosis inflammation imaging with 18F-FDG PET: carotid, iliac, and femoral uptake reproducibility, quantification methods, and recommendations. J. Nucl. Med. 49, 871–878 (2008). [DOI] [PubMed] [Google Scholar]
- 124.Mills EL et al. Succinate dehydrogenase supports metabolic repurposing of mitochondria to drive inflammatory macrophages. Cell 167, 457–470 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Galvan-Pena S et al. Malonylation of GAPDH is an inflammatory signal in macrophages. Nat. Commun. 10, 338 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Di Gioia M et al. Endogenous oxidized phospholipids reprogram cellular metabolism and boost hyperinflammation. Nat. Immunol. 21, 42–53 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Baardman J et al. Macrophage ATP citrate lyase deficiency stabilizes atherosclerotic plaques. Nat. Commun. 11, 6296 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Sarrazy V et al. Disruption of Glut1 in hematopoietic stem cells prevents myelopoiesis and enhanced glucose flux in atheromatous plaques of ApoE−/− mice. Circ. Res. 118, 1062–1077 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Freemerman AJ et al. Myeloid Slc2a1-deficient murine model revealed macrophage activation and metabolic phenotype are fueled by GLUT1. J. Immunol. 202, 1265–1286 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Divakaruni AS et al. Etomoxir inhibits macrophage polarization by disrupting CoA homeostasis. Cell Metab. 28, 490–503 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Nomura M et al. Fatty acid oxidation in macrophage polarization. Nat. Immunol. 17, 216–217 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Nomura M et al. Macrophage fatty acid oxidation inhibits atherosclerosis progression. J. Mol. Cell. Cardiol. 127, 270–276 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Chen Y et al. Mitochondrial metabolic reprogramming by CD36 signaling drives macrophage inflammatory responses. Circ. Res. 125, 1087–1102 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Ip WKE, Hoshi N, Shouval DS, Snapper S & Medzhitov R Anti-inflammatory effect of IL-10 mediated by metabolic reprogramming of macrophages. Science 356, 513–519 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Zhang S et al. Efferocytosis fuels requirements of fatty acid oxidation and the electron transport chain to polarize macrophages for tissue repair. Cell Metab. 29, 443–456 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136.Nelson VL et al. PPARγ is a nexus controlling alternative activation of macrophages via glutamine metabolism. Genes Dev. 32, 1035–1044 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Babaev VR et al. Conditional knockout of macrophage PPARgamma increases atherosclerosis in C57BL/6 and low-density lipoprotein receptor-deficient mice. Arterioscler Thromb. Vasc. Biol. 25, 1647–1653 (2005). [DOI] [PubMed] [Google Scholar]
- 138.Cordes T et al. Immunoresponsive gene 1 and itaconate inhibit succinate dehydrogenase to modulate intracellular succinate levels. J. Biol. Chem. 291, 14274–14284 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139.Ganta VC et al. A MicroRNA93-interferon regulatory factor-9-immunoresponsive gene-1-itaconic acid pathway modulates m2-like macrophage polarization to revascularize ischemic muscle. Circulation 135, 2403–2425 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.O’Neill LAJ & Artyomov MN Itaconate: the poster child of metabolic reprogramming in macrophage function. Nat. Rev. Immunol. 19, 273–281 (2019). [DOI] [PubMed] [Google Scholar]
- 141.Merlin J et al. Non-canonical glutamine transamination sustains efferocytosis by coupling redox buffering to oxidative phosphorylation. Nat. Metab. 3, 1313–1326 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Albina JE, Mills CD, Henry WL Jr. & Caldwell MD Temporal expression of different pathways of 1-arginine metabolism in healing wounds. J. Immunol. 144, 3877–3880 (1990). [PubMed] [Google Scholar]
- 143.Bauer PM, Buga GM, Fukuto JM, Pegg AE & Ignarro LJ Nitric oxide inhibits ornithine decarboxylase via S-nitrosylation of cysteine 360 in the active site of the enzyme. J. Biol. Chem. 276, 34458–34464 (2001). [DOI] [PubMed] [Google Scholar]
- 144.Lee J, Ryu H, Ferrante RJ, Morris SM Jr. & Ratan RR Translational control of inducible nitric oxide synthase expression by arginine can explain the arginine paradox. Proc. Natl Acad. Sci. USA 100, 4843–4848 (2003). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.El-Gayar S, Thuring-Nahler H, Pfeilschifter J, Rollinghoff M & Bogdan C Translational control of inducible nitric oxide synthase by IL-13 and arginine availability in inflammatory macrophages. J. Immunol. 171, 4561–4568 (2003). [DOI] [PubMed] [Google Scholar]
- 146.Bussiere FI et al. Spermine causes loss of innate immune response to Helicobacter pylori by inhibition of inducible nitric-oxide synthase translation. J. Biol. Chem. 280, 2409–2412 (2005). [DOI] [PubMed] [Google Scholar]
- 147.Teupser D et al. Identification of macrophage arginase I as a new candidate gene of atherosclerosis resistance. Arterioscler. Thromb. Vasc. Biol. 26, 365–371 (2006). [DOI] [PubMed] [Google Scholar]
- 148.Yurdagul A Jr. et al. Macrophage metabolism of apoptotic cell-derived arginine promotes continual efferocytosis and resolution of injury. Cell Metab. 31, 518–533 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149.Ravishankar B et al. The amino acid sensor GCN2 inhibits inflammatory responses to apoptotic cells promoting tolerance and suppressing systemic autoimmunity. Proc. Natl Acad. Sci. USA 112, 10774–10779 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Baumgartner R et al. Evidence that a deviation in the kynurenine pathway aggravates atherosclerotic disease in humans. J. Intern. Med. 289, 53–68 (2021). [DOI] [PubMed] [Google Scholar]
- 151.Pedersen ER et al. Associations of plasma kynurenines with risk of acute myocardial infarction in patients with stable angina pectoris. Arterioscler. Thromb. Vasc. Biol. 35, 455–462 (2015). [DOI] [PubMed] [Google Scholar]
- 152.Berg M et al. 3-Hydroxyanthralinic acid metabolism controls the hepatic SREBP/lipoprotein axis, inhibits inflammasome activation in macrophages, and decreases atherosclerosis in Ldlr−/− mice. Cardiovasc. Res. 116, 1948–1957 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Zhang L et al. The tryptophan metabolite 3-hydroxyanthranilic acid lowers plasma lipids and decreases atherosclerosis in hypercholesterolaemic mice. Eur. Heart J. 33, 2025–2034 (2012). [DOI] [PubMed] [Google Scholar]
- 154.Cole JE et al. Indoleamine 2,3-dioxygenase-1 is protective in atherosclerosis and its metabolites provide new opportunities for drug development. Proc. Natl Acad. Sci. USA 112, 13033–13038 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Polyzos KA et al. Inhibition of indoleamine 2,3-dioxygenase promotes vascular inflammation and increases atherosclerosis in Apoe−/− mice. Cardiovasc. Res. 106, 295–302 (2015). [DOI] [PubMed] [Google Scholar]
- 156.Nakajima K et al. Orally administered eicosapentaenoic acid induces rapid regression of atherosclerosis via modulating the phenotype of dendritic cells in LDL receptor-deficient mice. Arterioscler. Thromb. Vasc. Biol. 31, 1963–1972 (2011). [DOI] [PubMed] [Google Scholar]
- 157.Holmes DR Jr. et al. Results of prevention of restenosis with tranilast and its outcomes (PRESTO) trial. Circulation 106, 1243–1250 (2002). [DOI] [PubMed] [Google Scholar]
- 158.Tamai H et al. Impact of tranilast on restenosis after coronary angioplasty: tranilast restenosis following angioplasty trial (TREAT). Am. Heart J. 138, 968–975 (1999). [DOI] [PubMed] [Google Scholar]
- 159.Chen S et al. Novel role for tranilast in regulating NLRP3 ubiquitination, vascular inflammation, and atherosclerosis. J. Am. Heart Assoc. 9, e015513 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160.Liang H et al. The proatherosclerotic function of indoleamine 2, 3-dioxygenase 1 in the developmental stage of atherosclerosis. Signal. Transduct. Target. Ther. 4, 23 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161.Metghalchi S et al. Indoleamine 2,3-dioxygenase fine-tunes immune homeostasis in atherosclerosis and colitis through repression of interleukin-10 production. Cell Metab. 22, 460–471 (2015). [DOI] [PubMed] [Google Scholar]
- 162.Afarideh M et al. Association of peripheral 5-hydroxyindole-3-acetic acid, a serotonin derivative, with metabolic syndrome and low-grade inflammation. Endocr. Pract. 21, 711–718 (2015). [DOI] [PubMed] [Google Scholar]
- 163.Colpo GD, Venna VR, McCullough LD & Teixeira AL Systematic review on the involvement of the kynurenine pathway in stroke: pre-clinical and clinical evidence. Front Neurol. 10, 778 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.Yang M & Vousden KH Serine and one-carbon metabolism in cancer. Nat. Rev. Cancer 16, 650–662 (2016). [DOI] [PubMed] [Google Scholar]
- 165.Locasale JW Serine, glycine and one-carbon units: cancer metabolism in full circle. Nat. Rev. Cancer 13, 572–583 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Rodriguez AE et al. Serine metabolism supports macrophage IL-1β production. Cell Metab. 29, 1003–1011 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167.Yu W et al. One-carbon metabolism supports S-adenosylmethionine and histone methylation to drive inflammatory macrophages. Mol. Cell 75, 1147–1160 (2019). [DOI] [PubMed] [Google Scholar]
- 168.Rosenblat M, Coleman R & Aviram M Increased macrophage glutathione content reduces cell-mediated oxidation of LDL and atherosclerosis in apolipoprotein E-deficient mice. Atherosclerosis 163, 17–28 (2002). [DOI] [PubMed] [Google Scholar]
- 169.Callegari A et al. Gain and loss of function for glutathione synthesis: impact on advanced atherosclerosis in apolipoprotein E-deficient mice. Arterioscler. Thromb. Vasc. Biol. 31, 2473–2482 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170.Hitzel J et al. Oxidized phospholipids regulate amino acid metabolism through MTHFD2 to facilitate nucleotide release in endothelial cells. Nat. Commun. 9, 2292 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 171.Ding Y et al. Plasma glycine and risk of acute myocardial infarction in patients with suspected stable angina pectoris. J. Am. Heart Assoc. 10.1161/JAHA.115.002621 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 172.Wittemans LBL et al. Assessing the causal association of glycine with risk of cardio-metabolic diseases. Nat. Commun. 10, 1060 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Rom O et al. Induction of glutathione biosynthesis by glycine-based treatment mitigates atherosclerosis. Redox Biol. 52, 102313 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 174.Tabas I Consequences and therapeutic implications of macrophage apoptosis in atherosclerosis: the importance of lesion stage and phagocytic efficiency. Arterioscler. Thromb. Vasc. Biol. 25, 2255–2264 (2005). [DOI] [PubMed] [Google Scholar]
- 175.Tabas I Macrophage death and defective inflammation resolution in atherosclerosis. Nat. Rev. Immunol. 10, 36–46 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176.Linton MF et al. Macrophage apoptosis and efferocytosis in the pathogenesis of atherosclerosis. Circ. J. 80, 2259–2268 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177.Thorp E, Cui D, Schrijvers DM, Kuriakose G & Tabas I Mertk receptor mutation reduces efferocytosis efficiency and promotes apoptotic cell accumulation and plaque necrosis in atherosclerotic lesions of Apoe−/− mice. Arterioscler. Thromb. Vasc. Biol. 28, 1421–1428 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178.Kojima Y et al. CD47-blocking antibodies restore phagocytosis and prevent atherosclerosis. Nature 536, 86–90 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 179.Tao W et al. siRNA nanoparticles targeting CaMKIIgamma in lesional macrophages improve atherosclerotic plaque stability in mice. Sci. Transl. Med. 12, eaay1063 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180.Cai B et al. MerTK receptor cleavage promotes plaque necrosis and defective resolution in atherosclerosis. J. Clin. Invest. 127, 564–568 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181.Flores AM et al. Pro-efferocytic nanoparticles are specifically taken up by lesional macrophages and prevent atherosclerosis. Nat. Nanotechnol. 15, 154–161 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 182.Morioka S et al. Efferocytosis induces a novel SLC program to promote glucose uptake and lactate release. Nature 563, 714–718 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183.Park D et al. Continued clearance of apoptotic cells critically depends on the phagocyte Ucp2 protein. Nature 477, 220–224 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 184.Trzeciak A, Wang YT & Perry JSA First we eat, then we do everything else: the dynamic metabolic regulation of efferocytosis. Cell Metab. 33, 2126–2141 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185.Martinez J et al. Molecular characterization of LC3-associated phagocytosis reveals distinct roles for Rubicon, NOX2 and autophagy proteins. Nat. Cell Biol. 17, 893–906 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
- 186.Sanjuan MA et al. Toll-like receptor signalling in macrophages links the autophagy pathway to phagocytosis. Nature 450, 1253–1257 (2007). [DOI] [PubMed] [Google Scholar]
- 187.Yurdagul A Jr. Metabolic consequences of efferocytosis and its impact on atherosclerosis. Immunometabolism 3, e210017 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188.Kumar D, Pandit R & Yurdagul A Jr. Mechanisms of continual efferocytosis by macrophages and its role in mitigating atherosclerosis. Immunometabolism 5, e00017 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189.Viaud M et al. Lysosomal cholesterol hydrolysis couples efferocytosis to anti-inflammatory oxysterol production. Circ. Res. 122, 1369–1384 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190.Cui D et al. Pivotal advance: macrophages become resistant to cholesterol-induced death after phagocytosis of apoptotic cells. J. Leukoc. Biol. 82, 1040–1050 (2007). [DOI] [PubMed] [Google Scholar]
- 191.Fond AM, Lee CS, Schulman IG, Kiss RS & Ravichandran KS Apoptotic cells trigger a membrane-initiated pathway to increase ABCA1. J. Clin. Invest. 125, 2748–2758 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192.Kiss RS, Elliott MR, Ma Z, Marcel YL & Ravichandran KS Apoptotic cells induce a phosphatidylserine-dependent homeostatic response from phagocytes. Curr. Biol. 16, 2252–2258 (2006). [DOI] [PubMed] [Google Scholar]
- 193.A-Gonzalez N et al. Apoptotic cells promote their own clearance and immune tolerance through activation of the nuclear receptor LXR. Immunity 31, 245–258 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 194.Yvan-Charvet L et al. ABCA1 and ABCG1 protect against oxidative stress-induced macrophage apoptosis during efferocytosis. Circ. Res. 106, 1861–1869 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 195.Patterson MT, et al. Trem2 promotes foamy macrophage lipid uptake and survival in atherosclerosis. Nat. Cardiovasc. Res. 10.1038/s44161-023-00354-3 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196.Piollet M et al. TREM2 limits necrotic core formation during atherogenesis by controlling macrophage survival and efferocytosis. Preprint at bioRxiv 10.1101/2023.05.15.539977 (2023). [DOI] [Google Scholar]
- 197.Yurdagul A Jr. et al. ODC (ornithine decarboxylase)-dependent putrescine synthesis maintains MerTK (MER tyrosine-protein kinase) expression to drive resolution. Arterioscler. Thromb. Vasc. Biol. 41, e144–e159 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 198.Ampomah PB et al. Macrophages use apoptotic cell-derived methionine and DNMT3A during efferocytosis to promote tissue resolution. Nat. Metab. 4, 444–457 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 199.Gerlach BD et al. Efferocytosis induces macrophage proliferation to help resolve tissue injury. Cell Metab. 33, 2445–2463 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200.Zhang X et al. Loss of macrophage mTORC2 drives atherosclerosis via FoxO1 and IL-1beta signaling. Circ. Res. 133, 200–219 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 201.Schilperoort M, Ngai D, Katerelos M, Power DA & Tabas I PFKFB2-mediated glycolysis promotes lactate-driven continual efferocytosis by macrophages. Nat. Metab. 10.1038/s42255-023-00736-8 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202.Jha MK et al. Macrophage monocarboxylate transporter 1 promotes peripheral nerve regeneration after injury in mice. J. Clin. Invest. 131, e141964 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203.Tabas I & Bornfeldt KE Intracellular and Intercellular aspects of macrophage immunometabolism in atherosclerosis. Circ. Res. 126, 1209–1227 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204.Zhang D et al. Metabolic regulation of gene expression by histone lactylation. Nature 574, 575–580 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 205.Susser LI et al. Mitochondrial fragmentation promotes inflammation resolution responses in macrophages via histone lactylation. Mol. Cell. Biol. 43, 531–546 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 206.Wang Y et al. Mitochondrial fission promotes the continued clearance of apoptotic cells by macrophages. Cell 171, 331–345 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207.Wang YT et al. Metabolic adaptation supports enhanced macrophage efferocytosis in limited-oxygen environments. Cell Metab. 35, 316–331 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 208.Fredman G et al. An imbalance between specialized pro-resolving lipid mediators and pro-inflammatory leukotrienes promotes instability of atherosclerotic plaques. Nat. Commun. 7, 12859 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 209.Sakaguchi S Naturally arising Foxp3-expressing CD25+CD4+ regulatory T cells in immunological tolerance to self and non-self. Nat. Immunol. 6, 345–352 (2005). [DOI] [PubMed] [Google Scholar]
- 210.Saigusa R, Winkels H & Ley K T cell subsets and functions in atherosclerosis. Nat. Rev. Cardiol. 17, 387–401 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 211.Proto JD et al. Regulatory T cells promote macrophage efferocytosis during inflammation resolution. Immunity 49, 666–677 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 212.Sharma M et al. Regulatory T cells license macrophage pro-resolving functions during atherosclerosis regression. Circ. Res. 127, 335–353 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 213.Wolf D & Ley K Immunity and inflammation in atherosclerosis. Circ. Res. 124, 315–327 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 214.Conrad N et al. Autoimmune diseases and cardiovascular risk: a population-based study on 19 autoimmune diseases and 12 cardiovascular diseases in 22 million individuals in the UK. Lancet 400, 733–743 (2022). [DOI] [PubMed] [Google Scholar]
- 215.Nettersheim FS, Picard FSR, Hoyer FF & Winkels H Immunotherapeutic strategies in cancer and atherosclerosis—two sides of the same coin. Front. Cardiovas. Med. 8, 812702 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 216.Wolf D et al. Pathogenic autoimmunity in atherosclerosis evolves from initially protective ApoB100-reactive CD4+ T-regulatory cells. Circulation 10.1161/CIRCULATIONAHA.119.042863 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 217.Roy P et al. Immunodominant MHC-II (major histocompatibility complex II) restricted epitopes in human apolipoprotein B. Circ. Res. 131, 258–276 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 218.Depuydt MAC et al. Single-cell T cell receptor sequencing of paired human atherosclerotic plaques and blood reveals autoimmune-like features of expanded effector T cells. Nat. Cardiovasc. Res. 2, 112–125 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 219.Wang Z et al. Pairing of single-cell RNA analysis and T cell antigen receptor profiling indicates breakdown of T cell tolerance checkpoints in atherosclerosis. Nat. Cardiovasc. Res. 2, 290–306 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 220.Kimura T et al. Regulatory CD4+ T cells recognize major histocompatibility complex class II molecule-restricted peptide epitopes of apolipoprotein B. Circulation 138, 1130–1143 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 221.Saigusa R et al. Single cell transcriptomics and TCR reconstruction reveal CD4 T cell response to MHC-II-restricted APOB epitope in human cardiovascular disease. Nat. Cardiovas. Res. 1, 462–475 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 222.Stemme S et al. T lymphocytes from human atherosclerotic plaques recognize oxidized low density lipoprotein. Proc. Natl Acad. Sci USA 92, 3893–3897 (1995). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 223.Kimura T et al. Atheroprotective vaccination with MHC-II-restricted ApoB peptides induces peritoneal IL-10-producing CD4 T cells. Am. J. Physiol. Heart Circ. Physiol. 312, H781–H790 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 224.Tse K et al. Atheroprotective vaccination with MHC-II restricted peptides from ApoB-100. Front. Immunol. 4, 493 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 225.Li J et al. CCR5+T-bet+FoxP3+ effector CD4 T cells drive atherosclerosis. Circ. Res. 118, 1540–1552 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 226.Freuchet A et al. Identification of human exTreg cells as CD16+CD56+ cytotoxic CD4+ T cells. Nat. Immunol. 24, 1748–1761 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 227.Gaddis DE et al. Apolipoprotein AI prevents regulatory to follicular helper T cell switching during atherosclerosis. Nat. Commun. 9, 1095 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 228.Freuchet A et al. Identification of human exTreg cells as CD16+CD56+ cytotoxic CD4+ T cells. Nat. Immunol. 24, 1748–1761 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 229.Saxena V, Lakhan R, Iyyathurai J & Bromberg JS Mechanisms of exTreg induction. Eur. J. Immunol. 51, 1956–1967 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 230.Chapman NM, Boothby MR & Chi H Metabolic coordination of T cell quiescence and activation. Nat. Rev. Immunol. 20, 55–70 (2019). [DOI] [PubMed] [Google Scholar]
- 231.Michalek RD et al. Cutting edge: distinct glycolytic and lipid oxidative metabolic programs are essential for effector and regulatory CD4+ T cell subsets. J. Immunol. 186, 3299–3303 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 232.Gerriets VA et al. Metabolic programming and PDHK1 control CD4+ T cell subsets and inflammation. J. Clin. Invest. 125, 194–207 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 233.Macintyre AN et al. The glucose transporter Glut1 is selectively essential for CD4 T cell activation and effector function. Cell Metab. 20, 61–72 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 234.Shi LZ et al. HIF1α-dependent glycolytic pathway orchestrates a metabolic checkpoint for the differentiation of TH17 and Treg cells. J. Exp. Med. 208, 1367–1376 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 235.Hochrein SM et al. The glucose transporter GLUT3 controls T helper 17 cell responses through glycolytic-epigenetic reprogramming. Cell Metab. 34, 516–532 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 236.Forteza MJ et al. Pyruvate dehydrogenase kinase regulates vascular inflammation in atherosclerosis and increases cardiovascular risk. Cardiovasc. Res. 119, 1524–1536 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 237.He N et al. Metabolic control of regulatory T cell (Treg) survival and function by Lkb1. Proc. Natl Acad. Sci. USA 114, 12542–12547 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 238.Shrestha S et al. Treg cells require the phosphatase PTEN to restrain TH1 and TFH cell responses. Nat. Immunol. 16, 178–187 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 239.Huynh A et al. Control of PI3 kinase in Treg cells maintains homeostasis and lineage stability. Nat. Immunol. 16, 188–196 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 240.Gerriets VA et al. Foxp3 and Toll-like receptor signaling balance Treg cell anabolic metabolism for suppression. Nat. Immunol. 17, 1459–1466 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 241.Gaddis DE et al. Atherosclerosis impairs naive CD4 T-cell responses via disruption of glycolysis. Arterioscler. Thromb. Vasc. Biol. 41, 2387–2398 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 242.Maganto-García E, Tarrio ML, Grabie N, Bu DX & Lichtman AH Dynamic changes in regulatory T cells are linked to levels of diet-induced hypercholesterolemia. Circulation 124, 185–195 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 243.Bazioti V et al. T cell cholesterol efflux suppresses apoptosis and senescence and increases atherosclerosis in middle aged mice. Nat. Commun. 13, 3799 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 244.Iqbal R et al. Dietary patterns and the risk of acute myocardial infarction in 52 countries: results of the INTERHEART study. Circulation 118, 1929–1937 (2008). [DOI] [PubMed] [Google Scholar]
- 245.Kelly RK et al. Associations between types and sources of dietary carbohydrates and cardiovascular disease risk: a prospective cohort study of UK Biobank participants. BMC Med. 21, 34 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 246.Kedia-Mehta N & Finlay DK Competition for nutrients and its role in controlling immune responses. Nat. Commun. 10, 2123 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 247.Zhang D et al. High glucose intake exacerbates autoimmunity through reactive-oxygen-species-mediated TGF-β cytokine activation. Immunity 51, 671–681 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 248.Sinclair LV et al. Control of amino-acid transport by antigen receptors coordinates the metabolic reprogramming essential for T cell differentiation. Nat. Immunol. 14, 500–508 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 249.Matias MI et al. Regulatory T cell differentiation is controlled by aKG- induced alterations in mitochondrial metabolism and lipid homeostasis. Cell Rep. 10.1016/j.celrep.2021.109911 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 250.Klysz D et al. Glutamine-dependent α-ketoglutarate production regulates the balance between T helper 1 cell and regulatory T cell generation. Sci. Signal. 8, ra97 (2015). [DOI] [PubMed] [Google Scholar]
- 251.Johnson MO et al. Distinct regulation of Th17 and Th1 cell differentiation by glutaminase-dependent metabolism. Cell 175, 1780–1795 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 252.Ma EH et al. Serine is an essential metabolite for effector T cell expansion. Cell Metab. 25, 345–357 (2017). [DOI] [PubMed] [Google Scholar]
- 253.Roy DG et al. Methionine metabolism shapes T helper cell responses through regulation of epigenetic reprogramming. Cell Metab. 31, 250–266 (2020). [DOI] [PubMed] [Google Scholar]
- 254.Mezrich JD et al. An interaction between kynurenine and the aryl hydrocarbon receptor can generate regulatory T cells. J. Immunol. 185, 3190–3198 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 255.Yan Y et al. IDO upregulates regulatory T cells via tryptophan catabolite and suppresses encephalitogenic T cell responses in experimental autoimmune encephalomyelitis. J. Immunol. 185, 5953–5961 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 256.Sharma MD et al. Indoleamine 2,3-dioxygenase controls conversion of Foxp3+ Tregs to TH17-like cells in tumor-draining lymph nodes. Blood 113, 6102–6111 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 257.Cuffy MC et al. Induction of Indoleamine 2,3-dioxygenase in vascular smooth muscle cells by interferon-γ contributes to medial immunoprivilege1. J. Immunol. 179, 5246–5254 (2007). [DOI] [PubMed] [Google Scholar]
- 258.Forteza MJ et al. Activation of the regulatory T-cell/indoleamine 2,3-dioxygenase axis reduces vascular inflammation and atherosclerosis in hyperlipidemic mice. Front. Immunol. 10.3389/fimmu.2018.00950 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 259.Zhu Y et al. Lactate accelerates calcification in VSMCs through suppression of BNIP3-mediated mitophagy. Cell Signal. 58, 53–64 (2019). [DOI] [PubMed] [Google Scholar]
- 260.Tsai TL et al. Multiomics reveal the central role of pentose phosphate pathway in resident thymic macrophages to cope with efferocytosis-associated stress. Cell Rep. 40, 111065 (2022). [DOI] [PubMed] [Google Scholar]
- 261.Yurdagul A Jr. Crosstalk between macrophages and vascular smooth muscle cells in atherosclerotic plaque stability. Arterioscler. Thromb. Vasc. Biol. 42, 372–380 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 262.Medina CB et al. Metabolites released from apoptotic cells act as tissue messengers. Nature 580, 130–135 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 263.Howangyin KY et al. Myeloid-epithelial-reproductive receptor tyrosine kinase and milk fat globule epidermal growth factor 8 coordinately improve remodeling after myocardial infarction via local delivery of vascular endothelial growth factor. Circulation 133, 826–839 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 264.Virmani R et al. Atherosclerotic plaque progression and vulnerability to rupture: angiogenesis as a source of intraplaque hemorrhage. Arterioscler. Thromb. Vasc. Biol. 25, 2054–2061 (2005). [DOI] [PubMed] [Google Scholar]
- 265.de Vries MR et al. Blockade of vascular endothelial growth factor receptor 2 inhibits intraplaque haemorrhage by normalization of plaque neovessels. J. Intern. Med. 285, 59–74 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 266.Michel JB, Virmani R, Arbustini E & Pasterkamp G Intraplaque haemorrhages as the trigger of plaque vulnerability. Eur. Heart J. 32, 1977–1985 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 267.Rustom A, Saffrich R, Markovic I, Walther P & Gerdes HH Nanotubular highways for intercellular organelle transport. Science 303, 1007–1010 (2004). [DOI] [PubMed] [Google Scholar]
- 268.Borcherding N et al. Dietary lipids inhibit mitochondria transfer to macrophages to divert adipocyte-derived mitochondria into the blood. Cell Metab. 34, 1499–1513 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 269.Brestoff JR et al. Intercellular mitochondria transfer to macrophages regulates white adipose tissue homeostasis and is impaired in obesity. Cell Metab. 33, 270–282 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 270.Gurke S et al. Tunneling nanotube (TNT)-like structures facilitate a constitutive, actomyosin-dependent exchange of endocytic organelles between normal rat kidney cells. Exp. Cell. Res. 314, 3669–3683 (2008). [DOI] [PubMed] [Google Scholar]
- 271.Certo M et al. Endothelial cell and T-cell crosstalk: targeting metabolism as a therapeutic approach in chronic inflammation. Br. J. Pharmacol. 178, 2041–2059 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 272.Lukacs-Kornek V et al. Regulated release of nitric oxide by nonhematopoietic stroma controls expansion of the activated T cell pool in lymph nodes. Nat. Immunol. 12, 1096–1104 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 273.Doulias PT, Tenopoulou M, Greene JL, Raju K & Ischiropoulos H Nitric oxide regulates mitochondrial fatty acid metabolism through reversible protein S-nitrosylation. Sci. Signal. 6, rs1 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 274.Mendoza A et al. Lymphatic endothelial S1P promotes mitochondrial function and survival in naive T cells. Nature 546, 158–161 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 275.Lu Z et al. GPR40/FFA1 and neutral sphingomyelinase are involved in palmitate-boosted inflammatory response of microvascular endothelial cells to LPS. Atherosclerosis 240, 163–173 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 276.Gerriets VA & Rathmell JC Metabolic pathways in T cell fate and function. Trends Immunol. 33, 168–173 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 277.Wu H, Gong J & Liu Y Indoleamine 2, 3-dioxygenase regulation of immune response (Review). Mol. Med. Rep. 17, 4867–4873 (2018). [DOI] [PubMed] [Google Scholar]
- 278.Husted AS, Trauelsen M, Rudenko O, Hjorth SA & Schwartz TW GPCR-mediated signaling of metabolites. Cell Metab. 25, 777–796 (2017). [DOI] [PubMed] [Google Scholar]
- 279.He W et al. Citric acid cycle intermediates as ligands for orphan G-protein-coupled receptors. Nature 429, 188–193 (2004). [DOI] [PubMed] [Google Scholar]
- 280.Xu J et al. Succinate/IL-1beta signaling axis promotes the inflammatory progression of endothelial and exacerbates atherosclerosis. Front. Immunol. 13, 817572 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 281.Divorty N, Mackenzie AE, Nicklin SA & Milligan G G protein-coupled receptor 35: an emerging target in inflammatory and cardiovascular disease. Front. Pharmacol. 6, 41 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 282.Baumgartner R et al. Disruption of GPR35 signaling in bone marrow-derived cells does not influence vascular inflammation and atherosclerosis in hyperlipidemic mice. Metabolites 11, 411 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 283.Recio C et al. Activation of the immune-metabolic receptor GPR84 enhances inflammation and phagocytosis in macrophages. Front. Immunol. 9, 1419 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 284.Natarajan N et al. Microbial short chain fatty acid metabolites lower blood pressure via endothelial G protein-coupled receptor 41. Physiol. Genomics 48, 826–834 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 285.Pluznick JL et al. Olfactory receptor responding to gut microbiota-derived signals plays a role in renin secretion and blood pressure regulation. Proc. Natl Acad. Sci. USA 110, 4410–4415 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 286.McFarlane SI, Muniyappa R, Francisco R & Sowers JR Clinical review 145: pleiotropic effects of statins: lipid reduction and beyond. J. Clin. Endocrinol. Metab. 87, 1451–1458 (2002). [DOI] [PubMed] [Google Scholar]
- 287.Pruefer D, Scalia R & Lefer AM Simvastatin inhibits leukocyte-endothelial cell interactions and protects against inflammatory processes in normocholesterolemic rats. Arterioscler. Thromb. Vasc. Biol. 19, 2894–2900 (1999). [DOI] [PubMed] [Google Scholar]
- 288.Perticone F et al. Effects of atorvastatin and vitamin C on endothelial function of hypercholesterolemic patients. Atherosclerosis 152, 511–518 (2000). [DOI] [PubMed] [Google Scholar]
- 289.Mital S et al. Simvastatin upregulates coronary vascular endothelial nitric oxide production in conscious dogs. Am. J. Physiol. Heart Circ. Physiol. 279, H2649–H2657 (2000). [DOI] [PubMed] [Google Scholar]
- 290.Parmar KM et al. Statins exert endothelial atheroprotective effects via the KLF2 transcription factor. J. Biol. Chem. 280, 26714–26719 (2005). [DOI] [PubMed] [Google Scholar]
- 291.Bu DX et al. Statin-induced Kruppel-like factor 2 expression in human and mouse T cells reduces inflammatory and pathogenic responses. J. Clin. Invest. 120, 1961–1970 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 292.Koushki K et al. Anti-inflammatory action of statins in cardiovascular disease: the role of inflammasome and Toll-like receptor pathways. Clin. Rev. Allergy Immunol. 60, 175–199 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 293.Aday AW & Ridker PM Targeting residual inflammatory risk: a shifting paradigm for atherosclerotic disease. Front. Cardiovasc. Med. 6, 16 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 294.Nilsson J Atherosclerotic plaque vulnerability in the statin era. Eur. Heart J. 38, 1638–1644 (2017). [DOI] [PubMed] [Google Scholar]
- 295.Libby P, Pasterkamp G, Crea F & Jang IK Reassessing the mechanisms of acute coronary syndromes. Circ. Res. 124, 150–160 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 296.Chow OA et al. Statins enhance formation of phagocyte extracellular traps. Cell Host Microbe 8, 445–454 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 297.Tsourouktsoglou TD et al. Histones, DNA, and citrullination promote neutrophil extracellular trap inflammation by regulating the localization and activation of TLR4. Cell Rep. 31, 107602 (2020). [DOI] [PubMed] [Google Scholar]
- 298.Molinaro R et al. Targeted delivery of protein arginine deiminase-4 inhibitors to limit arterial intimal NETosis and preserve endothelial integrity. Cardiovasc. Res. 117, 2652–2663 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 299.Feng X, Zhang L, Xu S & Shen AZ ATP-citrate lyase (ACLY) in lipid metabolism and atherosclerosis: an updated review. Prog. Lipid Res. 77, 101006 (2020). [DOI] [PubMed] [Google Scholar]
- 300.Pinkosky SL et al. AMP-activated protein kinase and ATP-citrate lyase are two distinct molecular targets for ETC-1002, a novel small molecule regulator of lipid and carbohydrate metabolism. J. Lipid Res. 54, 134–151 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 301.Nissen SE et al. Bempedoic acid and cardiovascular outcomes in statin-intolerant patients. N. Engl. J. Med. 388, 1353–1364 (2023). [DOI] [PubMed] [Google Scholar]
- 302.DeFronzo RA Pharmacologic therapy for type 2 diabetes mellitus. Ann. Intern. Med. 131, 281–303 (1999). [DOI] [PubMed] [Google Scholar]
- 303.Staels B PPARgamma and atherosclerosis. Curr. Med. Res. Opin. 21, S13–S20 (2005). [DOI] [PubMed] [Google Scholar]
- 304.Loke YK, Kwok CS & Singh S Comparative cardiovascular effects of thiazolidinediones: systematic review and meta-analysis of observational studies. Brit. Med. J. 342, d1309 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 305.Kahn SE et al. Glycemic durability of rosiglitazone, metformin, or glyburide monotherapy. N. Engl. J. Med. 355, 2427–2443 (2006). [DOI] [PubMed] [Google Scholar]
- 306.Thorp E, Kuriakose G, Shah YM, Gonzalez FJ & Tabas I Pioglitazone increases macrophage apoptosis and plaque necrosis in advanced atherosclerotic lesions of nondiabetic low-density lipoprotein receptor-null mice. Circulation 116, 2182–2190 (2007). [DOI] [PubMed] [Google Scholar]
- 307.Saisho Y Metformin and inflammation: its potential beyond glucose-lowering effect. Endocr. Metab. Immune Disord. Drug Targets 15, 196–205 (2015). [DOI] [PubMed] [Google Scholar]
- 308.Cameron AR et al. Anti-inflammatory effects of metformin irrespective of diabetes status. Circ. Res. 119, 652–665 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 309.Davis BJ, Xie Z, Viollet B & Zou MH Activation of the AMP-activated kinase by antidiabetes drug metformin stimulates nitric oxide synthesis in vivo by promoting the association of heat shock protein 90 and endothelial nitric oxide synthase. Diabetes 55, 496–505 (2006). [DOI] [PubMed] [Google Scholar]
- 310.Katakam PV, Ujhelyi MR, Hoenig M & Miller AW Metformin improves vascular function in insulin-resistant rats. Hypertension 35, 108–112 (2000). [DOI] [PubMed] [Google Scholar]
- 311.Vitale C et al. Metformin improves endothelial function in patients with metabolic syndrome. J. Intern. Med. 258, 250–256 (2005). [DOI] [PubMed] [Google Scholar]
- 312.Tai S et al. Metformin suppresses vascular smooth muscle cell senescence by promoting autophagic flux. J. Adv. Res. 41, 205–218 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 313.Cao X et al. Metformin inhibits vascular calcification in female rat aortic smooth muscle cells via the AMPK–eNOS–NO pathway. Endocrinology 154, 3680–3689 (2013). [DOI] [PubMed] [Google Scholar]
- 314.Qing L et al. Metformin induces the M2 macrophage polarization to accelerate the wound healing via regulating AMPK/mTOR/NLRP3 inflammasome signaling pathway. Am. J. Transl. Res 11, 655–668 (2019). [PMC free article] [PubMed] [Google Scholar]
- 315.Habib A et al. Metformin impairs vascular endothelial recovery after stent placement in the setting of locally eluted mammalian target of rapamycin inhibitors via S6 kinase-dependent inhibition of cell proliferation. J. Am. Coll. Cardiol. 61, 971–980 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 316.Diabetes C et al. The effect of intensive treatment of diabetes on the development and progression of long-term complications in insulin-dependent diabetes mellitus. N. Engl. J. Med. 329, 977–986 (1993). [DOI] [PubMed] [Google Scholar]
- 317.Adler AI et al. Association of systolic blood pressure with macrovascular and microvascular complications of type 2 diabetes (UKPDS 36): prospective observational study. BMJ 321, 412–419 (2000). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 318.UK Prospective Diabetes Study (UKPDS) Group. Effect of intensive blood-glucose control with metformin on complications in overweight patients with type 2 diabetes (UKPDS 34). Lancet 352, 854–865 (1998). [PubMed] [Google Scholar]
- 319.Wang CC, Gurevich I & Draznin B Insulin affects vascular smooth muscle cell phenotype and migration via distinct signaling pathways. Diabetes 52, 2562–2569 (2003). [DOI] [PubMed] [Google Scholar]
- 320.Zoungas S et al. Severe hypoglycemia and risks of vascular events and death. N. Engl. J. Med. 363, 1410–1418 (2010). [DOI] [PubMed] [Google Scholar]
- 321.Cersosimo E, Xu X & Musi N Potential role of insulin signaling on vascular smooth muscle cell migration, proliferation, and inflammation pathways. Am. J. Physiol. Cell Physiol. 302, C652–C657 (2012). [DOI] [PubMed] [Google Scholar]
- 322.Calles-Escandon J & Cipolla M Diabetes and endothelial dysfunction: a clinical perspective. Endocr. Rev. 22, 36–52 (2001). [DOI] [PubMed] [Google Scholar]
- 323.Moses JW et al. Sirolimus-eluting stents versus standard stents in patients with stenosis in a native coronary artery. N. Engl. J. Med. 349, 1315–1323 (2003). [DOI] [PubMed] [Google Scholar]
- 324.Hegner B et al. mTOR regulates vascular smooth muscle cell differentiation from human bone marrow-derived mesenchymal progenitors. Arterioscler. Thromb. Vasc. Biol. 29, 232–238 (2009). [DOI] [PubMed] [Google Scholar]
- 325.Martin KA et al. The mTOR/p70 S6K1 pathway regulates vascular smooth muscle cell differentiation. Am. J. Physiol. Cell Physiol. 286, C507–C517 (2004). [DOI] [PubMed] [Google Scholar]
- 326.Martin KA et al. Rapamycin promotes vascular smooth muscle cell differentiation through insulin receptor substrate-1/phosphatidylinositol 3-kinase/Akt2 feedback signaling. J. Biol. Chem. 282, 36112–36120 (2007). [DOI] [PubMed] [Google Scholar]
- 327.McFadden EP et al. Late thrombosis in drug-eluting coronary stents after discontinuation of antiplatelet therapy. Lancet 364, 1519–1521 (2004). [DOI] [PubMed] [Google Scholar]
- 328.Meier P et al. Coronary collateral function long after drug-eluting stent implantation. J. Am. Coll. Cardiol. 49, 15–20 (2007). [DOI] [PubMed] [Google Scholar]
- 329.Vorpahl M et al. Pathobiology of stent thrombosis after drug-eluting stent implantation. Curr. Pharm. Des. 16, 4064–4071 (2010). [DOI] [PubMed] [Google Scholar]
- 330.Stefanini GG & Holmes DR Jr. Drug-eluting coronary-artery stents. N. Engl. J. Med. 368, 254–265 (2013). [DOI] [PubMed] [Google Scholar]
- 331.Byrne RA et al. Report of a European society of cardiology-European association of percutaneous cardiovascular interventions task force on the evaluation of coronary stents in Europe: executive summary. Eur. Heart J. 36, 2608–2620 (2015). [DOI] [PubMed] [Google Scholar]
- 332.O’Brien B, Zafar H, Ibrahim A, Zafar J & Sharif F Coronary stent materials and coatings: a technology and performance update. Ann. Biomed. Eng. 44, 523–535 (2016). [DOI] [PubMed] [Google Scholar]
- 333.Arguello RJ et al. SCENITH: a flow cytometry-based method to functionally profile energy metabolism with single-cell resolution. Cell Metab. 32, 1063–1075 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 334.Rappez L et al. SpaceM reveals metabolic states of single cells. Nat. Methods 18, 799–805 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 335.Eberhardt N & Giannarelli C How single-cell technologies have provided new insights into atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 42, 243–252 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 336.Ord T et al. Single-cell epigenomics and functional fine-mapping of atherosclerosis GWAS loci. Circ. Res. 129, 240–258 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 337.Williams JW et al. Single cell RNA sequencing in atherosclerosis research. Circ. Res. 126, 1112–1126 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 338.Kott KA et al. Single-cell immune profiling in coronary artery disease: the role of state-of-the-art immunophenotyping with mass cytometry in the diagnosis of atherosclerosis. J. Am. Heart Assoc. 9, e017759 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 339.Winkels H et al. Atlas of the immune cell repertoire in mouse atherosclerosis defined by Single-Cell RNA-sequencing and mass cytometry. Circ. Res. 122, 1675–1688 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 340.Fernandez DM et al. Single-cell immune landscape of human atherosclerotic plaques. Nat. Med. 25, 1576–1588 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 341.Mizrak D et al. Single-molecule spatial transcriptomics of human thoracic aortic aneurysms uncovers calcification-related CARTPT-expressing smooth muscle cells. Arterioscler. Thromb. Vasc. Biol. 10.1161/ATVBAHA.123.319329 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 342.Baigent C et al. Efficacy and safety of cholesterol-lowering treatment: prospective meta-analysis of data from 90,056 participants in 14 randomised trials of statins. Lancet 366, 1267–1278 (2005). [DOI] [PubMed] [Google Scholar]
- 343.Sabatine MS et al. Evolocumab and clinical outcomes in patients with cardiovascular disease. N. Engl. J. Med. 376, 1713–1722 (2017). [DOI] [PubMed] [Google Scholar]
- 344.Sachdeva A et al. Lipid levels in patients hospitalized with coronary artery disease: an analysis of 136,905 hospitalizations in Get With The Guidelines. Am. Heart J. 157, 111–117 (2009). [DOI] [PubMed] [Google Scholar]
- 345.Sampson UK, Fazio S & Linton MF Residual cardiovascular risk despite optimal LDL cholesterol reduction with statins: the evidence, etiology, and therapeutic challenges. Curr. Atheroscler. Rep. 14, 1–10 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 346.Foo SY et al. Vascular effects of a low-carbohydrate high-protein diet. Proc. Natl Acad. Sci. USA 106, 15418–15423 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 347.Zhang X et al. High-protein diets increase cardiovascular risk by activating macrophage mTOR to suppress mitophagy. Nat. Metab. 2, 110–125 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 348.Lagiou P et al. Low carbohydrate-high protein diet and incidence of cardiovascular diseases in Swedish women: prospective cohort study. BMJ 344, e4026 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 349.Hu FB et al. Dietary protein and risk of ischemic heart disease in women. Am. J. Clin. Nutr. 70, 221–227 (1999). [DOI] [PubMed] [Google Scholar]
- 350.Vogtschmidt YD et al. Is protein the forgotten ingredient: effects of higher compared to lower protein diets on cardiometabolic risk factors. A systematic review and meta-analysis of randomised controlled trials. Atherosclerosis 328, 124–135 (2021). [DOI] [PubMed] [Google Scholar]
- 351.Tharrey M et al. Patterns of amino acid intake are strongly associated with cardiovascular mortality, independently of the sources of protein. Int. J. Epidemiol. 49, 312–321 (2020). [DOI] [PubMed] [Google Scholar]
- 352.Ruiz-Canela M et al. Plasma branched-chain amino acids and incident cardiovascular disease in the PREDIMED trial. Clin. Chem. 62, 582–592 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 353.Liu Y et al. Dysregulated oxalate metabolism is a driver and therapeutic target in atherosclerosis. Cell Rep. 36, 109420 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 354.White PJ et al. Muscle-liver trafficking of BCAA-derived nitrogen underlies obesity-related glycine depletion. Cell Rep. 33, 108375 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 355.Ghrayeb A et al. Serine synthesis via reversed SHMT2 activity drives glycine depletion and acetaminophen hepatotoxicity in MASLD. Cell Metab. 36, 116–129.e7 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 356.Rom O et al. Glycine-based treatment ameliorates NAFLD by modulating fatty acid oxidation, glutathione synthesis, and the gut microbiome. Sci. Transl. Med. 12, eaa2841 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 357.Grajeda-Iglesias C, Rom O & Aviram M Branched-chain amino acids and atherosclerosis: friends or foes? Curr. Opin. Lipidol. 29, 166–169 (2018). [DOI] [PubMed] [Google Scholar]
- 358.McGarrah RW & White PJ Branched-chain amino acids in cardiovascular disease. Nat. Rev. Cardiol. 20, 77–89 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 359.Li Z et al. Oral administration of branched-chain amino acids attenuates atherosclerosis by inhibiting the inflammatory response and regulating the gut microbiota in ApoE-deficient mice. Nutrients 14, 5065 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 360.Zhao Y et al. Leucine supplementation via drinking water reduces atherosclerotic lesions in apoE null mice. Acta Pharmacol. Sin. 37, 196–203 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 361.Rom O et al. Atherogenicity of amino acids in the lipid-laden macrophage model system in vitro and in atherosclerotic mice: a key role for triglyceride metabolism. J. Nutr. Biochem. 45, 24–38 (2017). [DOI] [PubMed] [Google Scholar]
- 362.Grajeda-Iglesias C et al. Leucine supplementation attenuates macrophage foam-cell formation: studies in humans, mice, and cultured macrophages. Biofactors 44, 245–262 (2018). [DOI] [PubMed] [Google Scholar]
