Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2024 Apr 11.
Published in final edited form as: Immunity. 2023 Apr 11;56(4):723–741. doi: 10.1016/j.immuni.2023.03.007

Metabolism in type 2 immune responses

Agnieszka M Kabat 1, Erika L Pearce 1,2, Edward J Pearce 1,3,*
PMCID: PMC10938369  NIHMSID: NIHMS1889367  PMID: 37044062

Abstract

The immune response is tailored to the environment in which it takes place. Immune cells sense and adapt to changes in their surroundings, and it is now appreciated that in addition to cytokines made by stromal and epithelial cells, metabolic cues provide key adaptation signals. Changes in immune cell activation states are linked to changes in cellular metabolism that support function. Furthermore, metabolites themselves can signal between as well as within cells. Here we discuss recent progress in our understanding of how metabolic regulation relates to type 2 immunity firstly by considering specifics of metabolism within type 2 immune cells, and secondly by stressing how type 2 immune cells are integrated more broadly into the metabolism of the organism as a whole.

Introduction

Different functional states of immune cells, for example differentiated, quiescent or activated, are underpinned by metabolic adaptations1. Furthermore, as immune cells circulate, migrating from lymphoid organs to tissues, they adjust metabolically to their new environments2. Complexity is added by the fact that within any given tissue, conditions are not immutable; tissues experiencing inflammation or undergoing repair are likely to have a different metabolic landscape compared to those in homeostasis.

Type 2 immune responses, the focus of this review, are linked to resistance to helminth parasites, allergy and asthma, tissue homeostasis and repair. The notable tissue integration of key elements of type 2 immunity, discussed herein, raises particularly interesting questions about the metabolism of type 2 immune cells, especially since they are engaged in the homeostasis of central metabolic tissues including adipose tissue, pancreas, and liver3, 4, 5, and in responses that occur at barrier sites where metabolic conditions are dynamic. Type 2 immune cells include innate elements such as group 2 innate lymphoid cells (ILC2) and eosinophils, and an adaptive component consisting of type 2 helper T cells (Th2 cells), which all share the ability to make cytokines that are a defining feature of type 2 immunity, including interleukin (IL)-4, IL-13 or IL-5. Type 2 inflammation leads to the formation of Th2 memory, including long-lived resident memory Th2 cells (Th2RM), which have recently emerged as key perpetuators of chronic allergic diseases6, 7, 8. IL-4 or IL-13 play critical roles in directing macrophages and B cells to assume type 2 immunity-specific functions. In macrophages, this involves expression of a set of genes, that define a so-called alternative activation (or “M2”, or M(IL-4)) state, and in B cells it involves a switch to IgE or IgG1 (mouse)/IgG4 (human) production.

In this review we first discuss what is known about cell-intrinsic metabolism linked to cellular fate and function in the major type 2 immune cell populations that have been investigated in this regard. In the second part, we consider how type 2 immune functions are an integral component of the sensory system that responds to and regulates whole-body metabolism. These functions are coordinated by cytokines, hormones, neurotransmitters, nutrients and metabolites and we focus on how type 2 immune cells interpret cues in the gastrointestinal tract and adipose tissue to facilitate host protection and metabolic homeostasis.

Metabolic pathways required for Th2 cell development from priming to tissue residency

The polarization of naïve CD4+ T cells towards the Th2 lineage occurs in the lymph nodes (LN), where they begin to express GATA3 and IL-4. However, full differentiation, which includes the capacity to secrete IL-13 and IL-5, occurs only after cells leave the lymphatics and receive tissue-derived signals IL-33, IL-25 and thymic stromal lymphopoietin (TSLP)9. Furthermore, like ILC2, Th2 cells elicited by parasitic infections can disseminate to distant barrier sites10, 11, 12. Th2RM cells exhibit innate-like properties that mirror ILC2, including antigen-independent cytokine secretion and neuropeptide expression13, 14, 15, 16, 17, 18. Our understanding of Th2RM cell metabolism is fragmented, but recent studies have revealed distinct metabolic adaptations to tissue residency.

Early work on in vitro polarized Th cells showed that activated Th2 cells increase expression of the glucose transporter GLUT1 and have the highest glucose uptake compared to Th1 or Th17 cells19 (see Figure 1 for metabolic pathway information). In contrast to Th1 cells, and similarly to T regulatory (Treg cells), in vitro Th2 cells can differentiate in the absence of glutaminolysis20. Like other Th cells and activated ILC2, Th2 cells require signalling through mTORC1 for the upregulation of glycolysis and cell cycle entry21. However, Th2 cells, more than other Th subtypes, also depend on mTORC2 engagement, through pathways involving GTPase RhoA and AGC kinases22. Since mTORC2 signals promote cell survival, and migratory cytoskeleton rearrangement, mTORC2 could facilitate Th2 maturation and the motile behavior presumably required for tissue residency. Activated CD4+ T cells secrete IL-2, and autocrine signalling through the high affinity IL-2 receptor (IL-2Rα) has broad effects on metabolic remodeling in activated Th cells that depend on mTORC123, but are also directly downstream of STAT5, which drives transcription of multiple genes encoding enzymes in glycolysis and amino acid synthesis pathways24. Interestingly, Th2RM cells in lung, intestine, and adipose tissue have high expression of Il2ra6, 14, 18 and depend on IL-2 signalling for establishing tissue residency6. We speculate that this requirement could be linked to the metabolic changes that are induced through IL-2Rα, which perhaps are essential for supporting the final Th2 differentiation steps that occur within tissues.

Figure 1. Overview of core metabolic pathways used by type 2 immune cells.

Figure 1.

The tricarboxylic acid (TCA) cycle is a central metabolic pathway. Localized to mitochondria and fuelled by glucose, fatty acids and amino acids, it utilizes a series of reactions to generate NADH and FADH which are then oxidized by the electron transport chain to create a proton gradient across the inner mitochondrial membrane. This gradient drives the ATP synthase to generate ATP through OXPHOS. Glucose carbon enters mitochondria as pyruvate, generated by glycolysis, and is there converted to Acetyl-CoA. Intermediates of the glycolysis pathway can be redirected for nucleotide synthesis in the pentose-phosphate pathway, or metabolised to amino sugar uridine diphosphate N-acetylglucosamine (UDP-GlcNac), which is the major sugar donor for glycosylation. Another side branch of glycolysis pathway is the serine synthesis and one carbon metabolism pathway, which generates not only serine but nucleotides and S-adenosyl methionine, the methionine donor for DNA and histone methylation. Glutamine and branched chain amino acids (BCAA) have various points of entry into the TCA cycle. Fatty acids (FA) fuel the TCA cycle after being transported into mitochondria by carnitine palmitoyltransferase 1A (CPT1a) and cleaved in the cycles of β-oxidation to generate Acetyl-CoA (the process of fatty acid oxidation, FAO). FA (and sterols) can be synthesized from citrate (by fatty acid synthesis, FAS), imported from the extracellular space as FA, or generated by lipolysis within lysosomes from triacylglycerols (TAG) taken up from the exterior. FA can be either used for FAO or membrane synthesis, or used for TAG synthesis, for storage along with cholesterol esters in lipid droplets (LD). LD TAG can be reconverted to FA by regulated lipolysis or autophagy. Acetyl-CoA made from citrate exported from mitochondria serves as the acetyl donor for protein (including histone) acetylation. αKG - alpha-ketoglutarate, MCT - Monocarboxylate transporters, LAA – system L amino acids transporters; ASCT – alanine/serine/cysteine transporters; HMG-CoA - β-Hydroxy β-methylglutaryl-CoA; ACC1 - Acetyl-CoA carboxylase1; DGAT1 - Diacylglycerol O-acyltransferase 1. Created with BioRender.com.

Several lipid metabolism pathways regulate Th2 cell effector functions. Fatty acid synthesis (FAS) contributes to Th2 cell differentiation, although Th2 cells are less dependent than Th17 cells on this pathway25. FAS supports membrane biogenesis in rapidly proliferating cells, but might also facilitate T cell activation through increased production of ROS, as FAS limits NADPH availability for antioxidant responses. Consistent with this, deletion or inhibition of ACC1 (the rate limiting enzyme for FAS, Figure 1) in Th cells increases NADPH/NADP+, decreases cellular ROS, and protects from apoptosis26. ACC1 inhibition also increases the abundance of tricarboxylic acid (TCA) cycle intermediates, promotes mitochondrial spare respiratory capacity (SRC), and, consistent with this being a feature of memory T cells27, increases commitment towards Th2 memory26. Thus, decreasing FAS might be an important metabolic switch in transiting towards the memory phenotype. IL-33-driven reactivation of memory Th2 cells is again associated with an increase in FAS: IL-33R+ memory Th2 cells express more ACC1 compared to IL-33R- Th2 cells, which is driven by mTORC1 activation and corelates with increased IL-5 and IL-13 production28.

In addition to its use for FAS, citrate from the TCA cycle is also utilized as a substrate for cholesterol and sterol hormone synthesis in the mevalonate pathway. In helminth infection, Th2 cells that express Cyp11a118, 29, an enzyme that converts cholesterol to pregnenolone, a precursor for steroid synthesis, are able to inhibit Th proliferation in vitro29, and in this way exert regulatory effects that remain to be fully explored. Interestingly, lipid metabolism gene expression is enriched in Th2RM cells compared to circulating memory Th2 cells7. Indeed, in the HDM-induced asthma model, lung Th2 cells express genes associated with FA storage and oxidation7, 13, a signature also observed in Th2 cells from the small intestine and adipose tissue of mice infected with a helminth parasite14, 18 (Figure 2). Th2 cell maintenance in the lungs is negatively affected by the pharmacological inhibition of not only FAS and FA oxidation (FAO), but also glycolysis, suggesting that all three pathways are critical, although glycolysis inhibition has the most potent effect13. This may reflect a need to augment glycolysis to reach full effector potential30. In line with this, during enteric infection with the helminth parasite Heligmosomoides polygyrus, expression of several glycolysis pathway enzymes is increased in Th2 cells in the small intestine compared to Th2 cells in mesenteric adipose tissue, which is a more distal responsive tissue in this infection (Figure 2).

Figure 2. Metabolic adaptation to different tissues of residence.

Figure 2.

Infection with the helminth parasite H. polygyrus is restricted to the small intestine and induces a strong type 2 immune response that protects against reinfection. The mesenteric lymph node (mLN) is the draining LN for this infection, in which the adaptive immune response is initiated. Th2 cells then enter the intestine, where they mediate a protective response that includes promoting alternatively activated macrophages. In this infection, the mesenteric adipose tissue (mAT) also becomes populated by resident memory Th2 cells. For illustrative purposes, fraction dot plot representation is shown for expression of selected genes encoding transcription factors, cytokines, enzymes of glycolytic and lipid metabolism pathways, amino acid (aa) transporters, and genes involved in metabolite sensing in the intestine. These single-cell RNA-sequencing (scRNAseq) data are from an analysis available in Gene Expression Omnibus (GEO) under accession number GSE157314, as previously described14. The schematic was created with BioRender.

PPARγ is a lipid-activated transcription factor that regulates expression of genes in lipid metabolism pathways31. Its role in Th2 responses during helminth infection and lung inflammation has been established in mouse models with T cell-specific PPARγ deletion32, 33. PPARγ function in Th2 cells appears multifaceted. PPARγ expression is downstream of mTORC1 in naïve CD4+ T cells activated in vitro and promotes lipid uptake necessary for proliferation34. Beyond this role in early stages of Th2 differentiation34, 35, PPARγ activity appears essential for Th2RM cell function, since PPARγ-deficient Th2 cells have diminished IL-33R expression and effector cytokine production32, 33, 34. Mice lacking PPARγ in T cells are less protected from re-infection with H. polygyrus28, which is Th2RM cell-dependent11. PPARγ likely links Th2RM cell maturation with lipid metabolism by modulating expression of several genes associated with differentiated Th2 cells (Gata3, Bhlhe40, Il2ra, Il5, Il13), as well as genes linked to lipid uptake (Fabp5, Ldlr, Scarb1)34, 35. In line with this, pharmacologically inhibiting PPARγ decreases FA uptake and proliferation in re-activated memory Th2 cells34. Furthermore, CRISPR/Cas9 screening and chromatin accessibility assays have highlighted a crucial role for cross-regulation between PPARγ and Bhlhe40 in the Th2 differentiation programme35. Bhlhe40 is a transcription factor that is highly expressed during inflammation by Th2RM cells in the lung and small intestine, but not in adipose tissue (Figure 2)13, 18. Bhlhe40 transcriptionally represses most target genes through both histone deacetylase-dependent and independent mechanisms, but in other contexts can also activate transcription36. Bhlhe40 activity in Th2 cells is required for the production of the common β-chain cytokines IL-5 and GM-CSF during H. polygyrus infection and is necessary for protective responses after re-challenge, highlighting its crucial role for functional Th2 memory18. Interestingly, Bhlhe40 is also expressed by CD8+ TRM, but not circulating memory CD8+ T cells, where it regulates the expression of several mitochondrial genes, contributing to mitochondrial fitness37. Bhlhe40 deficiency in CD8+ TRM cells decreases TCA cycle intermediate levels, which might be linked with the impaired gene acetylation observed in these cells37. It remains to be investigated whether Bhlhe40 expression in Th2RM cells similarly contributes to mitochondrial homeostasis.

In atopic individuals, high PPARγ expression in Th2 cells is seen in several settings, including in circulating Th2 cells from individuals with allergy and asthma33, 38, 39, Th2 cells isolated from nasal polyps40, esophagus41, lungs of asthmatic individuals42, and skin of dermatitis patients43. Consistent with the known functions of PPARγ, human PPARγ+ Th2 cells isolated from nasal polyps have increased expression of lipid storage genes compared to other T cell subsets isolated from this tissue40. Human studies also link PPARγ expression with IL-9 production in Th9 cells, which are argued to be a subpopulation of memory Th2 cells present in atopic individuals38, 40, 43, and are crucial for allergic airway recall responses in a mouse model44. CD8+ TRM cells also express PPARγ along with Bhlhe4037, and depend on FAO, which is fuelled by exogenous FA, and promotes their long-term survival in the tissue45. Together with the essential role of PPARγ and Bhlhe40 in tissue resident macrophages46 and of PPARγ in ILC247, 48, 49, 50, PPARγ and Bhlhe40 expression emerges as a shared adaptation for long-lasting metabolic fitness of tissue resident immune cells, that in Th2 cells is also intertwined with the core regulatory network that governs their differentiation and acquisition of effector functions35, 51. In adipocytes, Bhlhe40 expression is induced by hypoxia and downregulates FAO through PPARγ repression52, so it is tempting to speculate that cross-regulation of PPARγ and Bhlhe40 in resident immune cells fine tunes mitochondrial metabolism in hypoxic tissue environments to balance survival with effector function.

Metabolic adaptations necessary for ILC2 differentiation and persistence

ILC2 are the innate counterparts of Th2 cells. They do not express the T cell receptor (TCR) and hence are unable to react in an antigen-specific manner or form adaptive memory, but instead rapidly release type 2 cytokines in response to tissue derived signals, primarily IL-25, IL-33 and TSLP, which can be made by epithelial cells and certain types of stromal cells5. Circulating human ILC2 from healthy donors express a range of system L amino acid (LAA) transporters, which import neutral amino acids, including branched chain amino acids (BCAA). In these cells, arginine and BCAA are the main fuels used to support the TCA cycle and mitochondrial respiration53, which are required for survival (Figure 3). This conclusion is supported by findings in mice, in which the LAA transporters Slc7a5 (LAT1) and Slc7a8 (LAT2) are highly expressed in tissue-resident ILC254, 55 and their deletion results in fewer of these cells in lung, intestine, and adipose tissue, and reduced type 2 responses in models of lung inflammation and helminth infection55, 56. While the requirements for amino acid transport and metabolism have not been comprehensively explored in Th2RM cells, it is worth noting that they have distinct expression patterns of amino acid transporters, including expression of Slc7a818 and Slc38a1, depending on the tissue within which they reside (Figure 2).

Figure 3. Metabolism changes significantly as ILC2 transition between resting and activated states.

Figure 3.

Resting ILC2 predominantly use mitochondrial respiration fuelled by arginine and BCAA taken up through LAAs transporters. Upon activation with IL-33, ILC2 express PPARγ and increase GLUT1 expression and glycolysis, as well as FA uptake and mitochondrial respiration. Further, glutamine transporter expression is upregulated and glutamine becomes a major substrate for mitochondrial respiration, while FA are transiently stored in LD and subsequently used for membrane biogenesis needed for proliferation. The autophagy pathway is induced upon activation and my facilitate utilisation of lipids from LD. Upregulated glycolysis is necessary for proliferation and effector cytokine secretion. BCAA – branch chain amino acids; LAA – system L amino acids transporters; FA – fatty acids; LD – lipid droplets; FAO – fatty acid oxidation. Created with BioRender.com

The use of pathway selective inhibitors indicated that IL-33-driven human ILC2 activation is dependent on glutamine oxidation53. Furthermore, arginine is depleted in these cells53 and in mouse ILC2 during lung inflammation57, possibly for polyamine synthesis. Based on work in macrophages, the important step here may be the use of the polyamine spermidine to hypusinate the eukaryotic translation initiation factor (eIF)5A. Hypusinated eIF5A (eIF5AH) is required for the efficient translation of a subset of proteins that are involved in the TCA cycle and OXPHOS58. Human ILC2 also increase FA uptake upon IL-33 stimulation, but this may not be required to sustain OXPHOS, as inhibition of FAO has no effect on mitochondrial mass or membrane potential53. Mouse ILC2 in the skin, adipose and intestinal tissues also show a propensity for FA uptake59. Furthermore, in a model of lung inflammation, chronically activated ILC2 increase FA uptake, and store these as triacylglycerides (TAG) in lipid droplets (LD), from which they are released to support membrane synthesis49. LD production and cellular activation in the context of IL-33 stimulation are dependent on PPARγ, which regulates many facets of fatty acid metabolism, including the expression of DGAT1 which is important for TAG synthesis. Interestingly, expression of DGAT1 is also increased in small intestinal Th2 cells compared to mesenteric adipose Th2 cells during H. polygyrus infection (Figure 2). The importance of PPARγ is supported by recent studies linking it to ILC2 effector function in lung, adipose tissue, and colorectal cancer47, 48, 49, 50. Utilization of intracellular lipid stores is mediated through organized LD-intrinsic lipolysis, or through autophagy, which is also upregulated in activated ILC260; deletion of Atg5, an essential autophagy gene, results in impaired cytokine production and increased apoptosis in these cells60. This is in contrast to Th2 cells, which are able to expand in the absence of functional autophagy pathways61.

Chronic activation of mouse ILC2 is characterized by increased glucose uptake. In these cells, mTORC1 activation requires the sensing of sufficient glucose, and glucose deprivation consequently also limits chronic ILC2 activation. This may be a therapeutically tractable finding, since mice on a ketogenic diet were shown to develop less severe chronic allergic pulmonary disease associated with diminished ILC2 activation49. Similarly, inhibition of glucose uptake in activated human ILC2 results in decreased proliferation and IL-13 production, although has no significant effect on mitochondrial function53. However, enhanced glycolysis, which results from decreased proteasomal degradation of Hif1α, attenuates mitochondrial respiration, leading to reduced permissive H3K4me3 methylation at the Gata3 and Il5 loci, as well as at the Il1rl1 (IL-33R) locus, which effectively diminishes IL-33 responsiveness and ILC2 numbers in non-lymphoid tissues62. Furthermore, hypoxia-driven upregulation of glycolysis in human ILC2 also results in decreased GATA3 and IL-33R expression53. Thus, while glucose is needed for ILC2 activation, it appears that the ratio between glycolysis and mitochondrial respiration must be tightly regulated for ILC2 tissue residence62.

Overall, a picture of ILC2 differentiation governed by progressive metabolic rewiring is emerging where initial activation increases glycolytic flux and mitochondrial respiration, and the balance between these two metabolic modules is crucial for subsequent maturation. Glycolysis is required for initial proliferation and pro-inflammatory cytokine production, but must be decreased for ILC2 to acquire tissue resident properties and rewire towards a metabolic state that favors longevity.

The metabolism of alternatively activated macrophages

Macrophages sense a broad range of signals within their microenvironment and in response express genes that confer distinct functional attributes cued to enforce normal physiology, but which can cause disease when inappropriately elicited63. There are a range of signals to which macrophages can respond in transcriptionally distinct ways64, 65, 66, but the metabolic changes, which clearly accompany stimulus-induced changes in biology, have been investigated in the context of only a subset of stimuli, including IL-4 and IL-13, which both signal through IL-4Rα and STAT6, to induce alternative activation. Alternative activation is broadly associated with efferocytosis and tissue homeostasis, the control of inflammation, resistance to helminth parasites, and wound healing67, 68, 69, 70.

IL-4 induces expression of genes encoding enzymes in the TCA cycle, and involved in fatty acid metabolism, including PPARγ, which promotes FAO to support the TCA cycle, OXPHOS and mitochondrial biogenesis71, 72, 73, 74, 75; these findings have been confirmed over multiple studies (reviewed in76, 77, 78). An active and complete TCA cycle in IL-4 stimulated macrophages is a striking contrast to the decline in mitochondrial activity, increase in Warburg-type metabolism, and interruption and repurposing of the TCA cycle to produce the metabolite itaconate, observed in macrophages that have become proinflammatory in response to stimulation by IFNγ plus TLR agonists79. The inflammatory metabolic state is resistant to reversal by subsequent IL-4 signalling80, possibly due to the ability of itaconate to post-translationally modify, and in so doing to inhibit, STAT681. In contrast though, alternatively activated macrophages can become inflammatory in response to IFNγ plus TLR agonists80. Integrating RNAseq and metabolomics data revealed an important role for glutamine and an emphasis on UDP-GlcNAC synthesis during alternative activation79. Mechanistically, the need for glutamine reflects a critical role for αKG, produced by glutaminolysis, in the activation of the histone demethylase Jmjd382. Expression of Jmjd3 increases following stimulation with IL-4 and is required for the demethylation of genes associated with alternative activation, such as Retnla, Mrc1, and Arg182. Glutamine is also required for increased FAO in response to IL-482. Further, flux through the UDP-GlcNAc synthesis pathway, which can be promoted by Hedgehog signalling, which can also potentiate alternative activation83, contributes to alternative activation because UDP-GlcNAc is a donor for the O-GlcNAcylation of STAT6, an activating posttranslational modification84.Adequate TCA cycle flux is also required to produce acetyl-CoA from citrate, which is important for histone acetylation and gene expression to support alternative activation85.

Although enhanced FAO is a feature of IL-4 stimulated macrophages, it is not required for alternative activation86, 87 (and discussed in77, 88) because macrophages can utilize disparate substrates to support this pathway77. It is nevertheless the case that inhibition or enhancement of OXPHOS and/or FAO often go hand in hand with the inhibition or enhancement of alternative activation. For example, increased FAO in IL-4 stimulated macrophages is accompanied by the regulated expression of numerous lipid metabolism genes including Cd36, which is important for fatty acid uptake, and Lipa, which encodes the lysosomal TAG lipase, and inhibition of these pathways limits the extent of alternative activation89. Moreover, FAS is also increased in IL-4 stimulated macrophages, in an SREBP1-dependent fashion, and inhibition of this pathway also limits alternative activation90. FAS may be important here because it contributes FA for FAO (in a futile cycle), and/or because it results in the depletion of NADPH, an essential cofactor in lipid synthesis, thereby limiting NADPH availability for antioxidant defences, allowing the accumulation of sufficient ROS to support the alternative activation process91, 92. Despite increased FAS, IL-4 activated macrophages do not store FA as TAG in lipid droplets (LD), an indication that FA are being utilized for other purposes, presumably including FAO. Further, since IL-4 drives macrophage proliferation in vivo93, which requires phospholipids for membrane synthesis, it is possible that induced FAS is needed to support this process. It is of interest that LD accumulate in IL-4-stimulated macrophages when lipase activity is pharmacologically inhibited89, suggesting that FA are fluxing through the TAG pathway in these cells. Increased FAO in alternatively activated macrophages is linked to their ability to mitigate atherosclerosis94, 95, a disease associated with lipid droplet accumulation in macrophages in the arterial intima. In this context it is intriguing that respiration is significantly inhibited by endogenous nitric oxide production in inflammatory macrophages, and that these cells have a high LD content96. In this setting, LD serve a pro-inflammatory role as a platform for prostaglandin production96, and prevention of LD accumulation in alternatively activated macrophages may therefore serve to guard against the activation of this pathway. It is of interest to speculate that the observed association of LD with pathogenic ILC249 might reflect a capacity of these cells to produce inflammatory lipid mediators.

There is great interest in the role of macrophages in tumor growth and metastasis97, 98, and while tumor associated macrophage (TAM) populations are heterogeneous they contain cells that share transcriptional and metabolic features with alternatively activated macrophages99, 100, 101. Indeed, mice in which IL-4 signaling is blocked are more capable of controlling cancer growth and metastasis than are control mice102, 103. Recent findings have described a population of Trem2-positive “lipid associated macrophages (LAM)” in adipose tissue, that are essential for healthy tissue homeostasis and glucose metabolism104. Cells with a similar transcriptional signature have been identified in multiple tumor subtypes100, and these cells share transcriptional similarities, including expression of Lipa, with alternatively activated macrophages. It is possible that FA uptake in LAM promotes FAO and OXPHOS, and that this synergizes with STAT6-dependent processes to drive an alternative activation state. Interestingly, OXPHOS is required for the persistence of tissue resident macrophages in niches such as the alveoli, where they are exposed to extracellular lipids and cholesterol, which in the absence of OXPHOS induce cellular stress and apoptosis105. What might be an analogous situation is seen during efferocytosis, in which macrophages engulf cell corpse-associated lipids through a process that has evolved to not be inflammatory, and consequently synergizes with IL-4 to promote alternative activation106. Metabolically, this synergy may relate to the fact that both stimulation with IL-4 and exposure to the tumor microenvironment promote the PERK arm of the unfolded protein response107. This has two effects. First, PERK promotes mitochondrial function and FAO. Second, downstream of PERK, the induction of ATF-4 leads to increased PSAT1 activation and serine synthesis, which is coupled to a transamination reaction that converts glutamate to αKG. As described above, αKG is critical for Jmjd3-dependent histone demethylation at alternative activation genes and genes related to FAO, and consistent with this, loss of PSAT1 function inhibits IL-4 induced FAO. Moreover, serine synthesis inhibition can increase IGF1 production, which antagonizes STAT6 signaling, and simultaneously promotes classical activation by activating p38-dependent JAK-STAT1 signalling108.

Other metabolic factors can also synergize with IL-4 to potentiate alternative activation. For example, extracellular adenosine signaling through AB2 adenosine receptor (A2BAR)109, and the TCA cycle intermediate succinate, sensed through the G protein coupled succinate receptor SUCNR1110. During helminth infection, adenosine generated by the CD39-mediated breakdown of ATP in the extracellular environment, presumably due to cell damage, plays a role in broadly promoting type 2 immunity by stimulating epithelial cells, via AB2AR, to release IL-33 and IL-25. This process is critical for the subsequent development of Th2RM cells and immunity to secondary infection111. Moreover, through SUCNR1, succinate from intestinal microbes or intestinal helminth parasites activates Tuft cells in the intestinal epithelium to produce IL-25, which in turn activates local ILC2 (reviewed in112). Additionally, in settings of strong type 2 immune response-mediated inflammation, the successful establishment of recruited monocytes into tissue-resident macrophages, which assume an alternatively activated state, is vitamin A-dependent113. These data indicate broad links between the cellular sensing of succinate (in contrast to intracellular succinate, which can be pro-inflammatory), adenosine and vitamin A, with effective type 2 immunity.

Another interesting feature of a lipid rich-environment is that, at least within tumors, it induces the activation of PI3Kγ in macrophages. This is important because targeted inhibition of macrophage PI3Kγ has shown promise in cancer therapy114. Mechanistically, this reflects the role of PI3kγ in the activation of Akt and mTORC2 to induce the C/EBPβ-dependent expression of anti-inflammatory genes. Intriguingly, the induction of Arg1 by IL-4 is regulated by a response element containing STAT6 and C/EBP binding sites and C/EBPβ has been directly implicated in IL-4 induced alternative activation115, raising the possibility that PI3Kγ synergizes with IL-4 in the development of tumor promoting TAM. There is a further link here with TAM lipid metabolism, in that membrane cholesterol efflux through ABC transporters, and associated lipid raft depletion, potentiate IL-4-induced alternative activation through increased STAT6 signaling and the PI3K-mTORC2-Akt pathway103; mTORC2 is critical for alternative activation116, 117. The importance of cholesterol efflux for full alternative activation is supported by the fact that TAM from mice with a myeloid deficiency in ABC cholesterol efflux transporters exhibit diminished expression of IL-4-inducible genes and enrichment of tumoricidal genes, and in this way they phenocopy PI3K-deficient TAM.

The expression of Arg1 by IL-4 activated macrophages results in these cells being able to diminish extracellular arginine levels, thereby depriving adjacent cells of this amino acid, with negative effects on their biology118. This is associated with increased tumor growth (e.g.119), most probably because of T cells’ sensitivity to arginine depletion120. Sensing arginine sufficiency is one of the permissive cues for mTORC1 activation121, a process required for the type of anabolic metabolism that typifies fast-proliferating T lymphocytes118. However, arginine depletion by macrophages expressing Arg1 can also serve protective functions as it can limit immunopathology122, and play direct roles in the killing of parasitic helminths123. Arginase is important not only as an effector enzyme, but also in the production of ornithine, a precursor for polyamine synthesis and thereby eIF5A hypusination, which is critical for enhanced TCA cycle activity and OXPHO associated with alternative activation, as discussed above58. Thus, arginase is required for both the establishment and effector functions of alternatively activated macrophages (Figure 4).

Figure 4. Alternatively activated macrophages manipulate amino acid metabolism to impact neighbouring cells.

Figure 4.

Secretion of IL-4 and IL-13 by ILC2 and Th2 cells induces a state of alternative activation in macrophages through IL-4Rα. This enhances expression of amino acid metabolising enzymes IL4i1 and Arg1. IL4i1 converts tryptophan into indole-3-pyruvate, which is a ligand for nuclear receptor Ahr. Signalling through Ahr induces exhaustion in effector T (Teff) cells, while promoting Treg cell development. IL4i1 also converts tyrosine into 4-hydroxy-phenylpuryvate. Both indole-3-pyruvate and 4-hydroxy-phenylpuryvate act as ROS scavengers (ROS is shown outside the cells but is actually intracellular). Depletion of arginine by Arg1 inhibits Teff cell activation. Arginine sufficiency is one of the permissive signals for mTORC activation. In macrophages, arginine is converted to ornithine and further used for polyamine synthesis, which is essential for alternative activation. Created with BioRender.com

The expression of distinct amino acid metabolizing enzymes is a feature of differentially activated macrophages124. Arginase is one of these, but so too are inducible NO synthase (iNOS), which generates NO from arginine, and IDO1 and IDO2, which degrade tryptophan125. These three enzymes are expressed in IFNγ−exposed macrophages. In contrast, IL-4 induces the expression of IL4i1, a tryptophan, phenylalanine, and tyrosine catabolizing secreted enzyme that was first identified as being expressed in IL-4-stimulated B cells126, and which like IDO1 and IDO2 has been implicated in the suppression of T cell proliferation127. Kynurenic acid, and in particular derivatives of indole-3-pyruvic acid produced in tryptophan breakdown by IL4i1, drive AhR dependent cancer cell motility and adaptive immune response suppression, promoting cancer progression128. AhR-dependent suppressive effects of IL4i1 reflect enhanced Treg cell development by AhR signaling129 and the induction of an exhausted phenotype in effector T cells (Figure 4). The broad significance of IL4i1 expression and tryptophan catabolism in cancer is supported by the finding that Il4i1 expressing monocyte-derived TAM, also expressing IDO1 and PD-L1, are common to cancers represented in multiple single cell RNAseq datasets across different human tissues100.

While Il4i1 was originally defined as an IL-4-inducible gene, IL4i1+ TAM express genes indicative of stimulation by IFNγ and CD40L100, which is consistent with reports that expression of Il4i1 is driven more by inflammatory signals in macrophages than by IL-4130. However, Il4i1 expression is highest in myeloid cells131 and Il4i1 expression in macrophages is induced by IL-4, and indeed there is evidence that it promotes IL-4-driven alternative activation132. This is of interest from the context of a second function of IL4i1-its role in ferroptosis prevention. Specifically, 4-hydroxy-phenylpyruvate from tyrosine catabolism, as well as indole-3-pyruvic acid from tryptophan, are radical scavengers which activate protective pathways that suppress ferroptosis133, an oxidative cell death program characterized by iron-dependent lipid peroxidation. One of the genes induced in this context is Slc7a11, which encodes a component of the cysteine importer Xc- (xCT). Cysteine import is critical for glutathione synthesis, and in this way the activity of GPX4, which catalyzes the reduction of lipid peroxides to mitigate against ferroptosis and cell death in IL-4 activated macrophages, although not in resting or inflammatory macrophages, suggesting that regulation of this pathway is of particular importance for alternatively activated macrophage survival134. It is conceivable that the tumor promoting properties of IL4i1 may reflect a role in protecting both alternatively activated TAM and tumor cells from ferroptosis. It is intriguing that in uniform manifold approximation and projection plots, Il4i1-expressing TAM cluster closely to TAM expressing genes more typical of LAM and IL-4-induced alternatively activated macrophages, including Trem2, ApoE, Fabp5 and Lipa100, raising the possibility that IL4i1+ TAM preferentially support the survival of adjacent alternatively activated macrophages in the tumor environment. This would be expected to promote resilience in the anti-inflammatory macrophage TAM compartment.

In addition to acting as a substrate for Arg1 and iNOS, and a precursor for polyamine synthesis, arginine is required for creatine synthesis. Glycine amidinotransferase (Gatm), the rate limiting enzyme in this pathway, is expressed in alternatively activated macrophages, and Gatm deletion in macrophages diminishes the IL-4 induced expression of alternative activation genes, but has no effect on inflammatory activation135. Moreover, addition of exogeneous creatine significantly inhibits STAT1 signaling and IFNγ responsiveness, but promotes alternative activation whereas deletion or inhibition of the creatine transporter Slc6a8 diminishes the IL-4 induced expression of certain alternative activation genes136. Phosphocreatine donates phosphate for rapid regeneration of ATP from ADP and data indicate that creatine promotes expression of STAT6 target genes through ATP-dependent SWI-SNF-mediated chromatin remodeling136.

In summary, increased OXPHOS accompanied by FAO is a feature of alternatively activated macrophages related to their pro-homeostatic function in catabolizing lipids in lipid-rich environments and during efferocytosis. Alternatively activated macrophages are also specialized to break down several amino acids through their expression of distinct amino acid degrading enzymes. Amongst these, arginase is also expressed by ILC257, 137 and Th2 cells14, suggesting that type 2 immune cells exhibit functional overlap related to arginine depletion from the microenvironment to limit the metabolic potential and therefore functions of other cells. In the context of the regulation of type 1 immunity, this may underlie the effects of alternatively activated macrophages and type 2 immunity that are related to wound healing and tumor progression. Interestingly, alternative activation is enhanced by metabolites such as adenosine in the extracellular environment, suggesting a mechanism by which the appearance of damage-associated signals in the environment reinforce the development of macrophages associated with the regulation of inflammation and tissue repair.

Metabolic regulation of IgE production

Antibody production by plasma cells is an important part of adaptive type 2 responses and IgE production in particular contributes to protection from certain parasites, venoms and toxins138, 139, 140. In healthy individuals, IgE is the least abundant antibody class in the circulation, with most of IgE being bound to high affinity receptor FcεRI on the surface of mucosa or skin residing mast cells. Class switching to IgE requires IL-4141 and high serum IgE levels are associated with many allergic disorders. Cross-linking of FcεRI, mediated by antigen binding, leads to mast cell degranulation and release of inflammatory mediators. In sensitized individuals these inflammatory signals trigger hypersensitivity reactions, including anaphylaxis139. The role of IgE in many type 2 disorders makes IgE an attractive therapeutic target; anti-IgE antibodies are currently used for the treatment of asthma and chronic spontaneous urticaria. However, they require prolonged dosing to reach a desired reduction in circulating IgE levels142. A better understanding of the metabolic control of IgE+ humoral responses could contribute to the development of novel and more effective therapies.

In order to become antibody secreting plasma cells (PC), B cells transit through several stages of differentiation, each marked by rapid metabolic change. While short lived PC (SLPC) produce low affinity antibodies and are generated early in the humoral response through the extrafollicular pathway, long-lived plasma cells (LLPC) are mainly high affinity and are generated in a T cell dependent manner. Substantial progress has been made in understanding metabolic regulation of humoral responses143, 144. Antigen-driven exit of B cells from a resting state results in a rapid increase in cell mass and proliferation. These changes are supported by anabolic reprogramming mediated by PKCβ and mTORC1 activity, and include increased glucose uptake, paralleled by mitochondrial remodeling and a progressive increase in mitochondrial respiration145, 146, 147. Activated B cells responding to protein antigens subsequently form germinal canters (GC), where they undergo cycles of proliferation, somatic hypermutation, and affinity selection. Preferential use of OXPHOS is paramount for GC B cells148, 149, 150, a metabolic requirement that can be compromised in bacterial infections, leading to decreased antibody production151. IL-4 signaling is important for B cell reprograming towards OXPHOS through epigenetic regulation, which in turn contributes to increased expression of Bcl6 through epigenetic remodeling of the Bcl6 locus150. BCL6 is a central transcription factor for GC B cells, and its expression has previously been linked to glycolysis repression152. Indeed, a distinctive feature of GC B cell energetics is the reliance on mitochondrial and peroxisomal FAO and a minimal requirement for glucose149. Overall, these data indicate that as GC B cells differentiate, the fuel choice is switched from glucose to FA. This raises an interesting possibility that GC B cells and Tfh cells utilize distinct metabolic pathways as a strategy to avoid nutrient competition, since Tfh cells use glucose and glutamine to fuel mitochondrial respiration153. Such distinctive metabolic adaptations could facilitate targeted therapeutic manipulation of either GC B cells or Tfh cells in antibody mediated diseases. In this context, it is of interest that a subset of Tfh cells that selectively promotes anaphylactic IgE responses has recently been identified154. Whether these cells are metabolically distinct from other Tfh cell subsets remains to be investigated, but it seems plausible, since metabolic differences have been shown between autoreactive and infection-driven Tfh cells155.

B cells that successfully emerge from the GC reaction provide long lasting humoral immunity by becoming either memory B cells or LLPC. Continuous antibody production and remarkable longevity in this population are supported by a specific set of metabolic adaptations, including increased mitochondrial SRC, amino acid uptake, and activation of autophagy pathways143. Moreover, in contrast to SLPC, LLPC have an increased requirement for glucose, which is at least in part mediated by CD28 signalling156, reminiscent of the CD28 co-stimulation requirement in T cells for the upregulation of glycolysis and mitochondrial SRC157. Interestingly, the majority of glucose acquired by LLPC is not used to fuel energetic demands, which are met by FAO and glutaminolysis158, but is instead diverted into the hexosamine pathway to produce UDP-GlcNAc, a sugar donor required for antibody glycosylation159, 160. N-glycosylation of the heavy chain is essential for antibody folding and, depending on the glycosylation pattern, modulates its interactions with Fc receptors, diversifying antibody functionality161. Alterations in antibody glycosylation have been observed in infections and autoimmune diseases. Recent characterization of IgE glycosylation revealed a disease-specific pattern in individuals with peanut allergy, marked by increased sialylation162. It would be of interest to investigate whether metabolic alterations in the hexosamine pathway in LLPC underlie allergy-associated modifications in glycosylation.

The majority of LLPC reside in the specialized bone marrow (BM) niches, where they continuously receive pro-survival signals from supporting cells, while SLPC localize to secondary lymphoid organs (SLO). With regard to IgE serological memory, it has been unclear whether IgE+ LLPC exist and if so, where they reside. The potentially life-threatening consequences of systemic IgE reactions necessitate tight control that limits longevity and anatomic localization of these responses. Previous research showed that IgE+ PC that are generated in physiological circumstances, such as in helminth infections, localize to SLO and are short lived163. The limited lifespan of IgE+ PC was linked to mitochondrial apoptosis that is driven by persistent calcium signaling downstream of the BCR164, suggesting that mitochondria remodeling that leads to reduced calcium buffering might control IgE+ PC longevity. These regulatory mechanisms are altered in individuals with chronic allergies, where clinical findings indicate the existence of long-lived IgE responses165. In line with this, recent studies in allergic individuals and in mice chronically exposed to antigens revealed that PC producing high affinity IgE are found outside of the SLO, including in the BM and gastrointestinal tissues166, 167. Understanding whether metabolic adaptations support the persistence of IgE+ PC in the BM and mucosal niches in allergic individuals could be of clinical relevance.

Sensing of the outside world – type 2 immunity in the intestine

As mentioned above, type 2 immunity plays an important role at barrier surfaces. This raises interesting questions regarding the intestine, where the need for continuous immune surveillance and reinforcement of barrier function must be balanced with the maintenance of adequate nutrient absorption. In homeostasis, the intestinal mucosa is dominated by type 1 and type 3 immune responses with critical contributions from Treg and myeloid cells that ensure tolerance towards food and commensal antigens. The homeostatic development and functions of mucosal immune cells are promoted by dietary- and microbiota-derived metabolites168. In contrast, expansion of intestinal type 2 immune cells is often associated with metabolic perturbations, such as those caused by helminth infections or malnutrition169, 170, 171.

It has been postulated that type 2 responses are engaged in food quality monitoring, protecting the host from ingesting noxious substances, for instance by changing absorption or promoting expulsion172. From this perspective, dysregulated protective mechanisms are proposed to underlie the development of food allergies and abdominal pain172, 173. Part of such control might be to sense malnutrition, which is evolutionarily linked with persistent presence of intestinal helminths, prototypic examples of type 2 immune response inducers174. Indeed, helminth parasite infections are associated with decreased blood glucose levels, decreased essential amino acids, including BCAA, in blood and tissues, and increased circulating lipids along with alterations in ketone body metabolism170, 175, 176. Longitudinal metabolomic studies indicate that the chronic stage of helminth infection is characterized by the establishment of a new metabolic homeostatic setpoint, in which circulating metabolites (with the exception of lipids) are brought to pre-infection levels, but the metabolic state of several tissues remains altered176.

SCFA (acetate, butyrate and propionate) are produced by anaerobic fermentation of complex carbohydrates by intestinal microbiota. Alterations in microbiota-derived metabolites, including SCFA and bile acid derivatives accompany helminth infections175, 177. Moreover, acetate is produced by parasites such as Tritrichomonas and H. polygyrus and can promote parasite invasion through disruption of the epithelial barrier178, 179. SCFA bind to GPR41, GPR109a, and GPR43168, the latter of which is highly expressed by SI Th2 cells, but not mesenteric adipose tissue Th2 cells in H. polygyrus infected mice (Figure 2). SCFA signaling enhances human Th2 effector functions; GPR41 and IL-5 expression are strongly correlated in Th2 cells isolated from tissue biopsies of patients with eosinophilic esophagitis, and in vitro stimulation with butyrate enhances Th2 cytokine production41. Thus, alterations in SCFA levels could act as a danger signal that boosts intestinal type 2 responses during parasitic infection.

Bile acids are produced from cholesterol, where they are conjugated to taurine or glycine to increases their solubility180. In addition to their role in the absorption of dietary lipids and fat-soluble vitamins, bile acids and secondary bile acids (produced when microbiota metabolize bile acids in the intestinal lumen) act as ligands for several GPR and nuclear receptors, including FXR, PXR, LXR, Rorγt, VDR, and TGR5180. Bile acids regulate metabolic pathways in many tissues, and have emerged as modulators of innate and adaptive immune cells181. For example, microbiota-derived secondary bile acids directly regulate Th17 and Treg cell differentiation downstream of Rorγt or VDR181. Intestinal Th2 cells express VDR and LXR during H. polygyrus infection18, but whether they are also subjected to similar regulation remains to be tested.

Imbalance in secondary bile acid metabolism might be sensed as a danger signal that induces inflammation182, 183. For example, a diet rich in inulin alters the intestinal microbiota with a resultant increase in the levels of circulating unconjugated bile acids. This in turn activates intestinal stromal cells, in an FXR-dependent manner, to secrete IL-33 that activates ILC2 and promotes eosinophilia, responses that are able to mediate resistance to an intestinal helminth183. Taken together, changes in the abundance or composition of bile acid metabolites might act as a sensitive readout of metabolic perturbations that are associated with intestinal infection or malnutrition. Such cues could be interpreted by type 2 immune cells to mount the appropriate response. Alterations in the levels of micronutrients may synergize in such pathways. For example, ratios of conjugated to unconjugated bile acids are altered during acute vitamin A deficiency, which may contribute to observed microbiota shifts in this model184. Consistent with this idea, deficiency in retinoic acid, a derivative of vitamin A, has been previously linked to enhanced IL-13+ ILC2 response in the intestine59, 169.

Eosinophils are a prominent component of type 2 immunity as a result of their dependence on IL-5, and play roles in tissue homeostasis and protection against helminth parasites, but also in a variety of immunopathologic conditions associated with type 2 immunity185. Eosinophils are numerous in the GI tract and in the steady state contribute to an anti-inflammatory environment by inhibiting Th17 and Th1 responses186, 187, and moreover contribute to the development of the architecture of the small intestine in a microbiota- and IL-33-dependent manner188. Mice lacking eosinophils have altered extracellular matrix (ECM) turnover and maladaptive changes characterised by villous atrophy, altered muscle contractility, and decreased intestinal epithelial cell (IEC) migration along the basement membrane, which together results in impaired lipid absorption from the lumen, underscoring the homeostatic role of type 2 immunity and resident eosinophils in interpreting microbial signals to adjust intestinal absorptive functions188. Signaling through AhR is needed to acquire the tissue resident phenotype by intestinal eosinophils189. Within the GI tract, AhR in thought to be activated by diet or microbiota-derived tryptophan metabolites. A role of AhR signalling in promoting mucosal IEC, ILC3, and Th17 responses is well documented190. In eosinophils, AhR activity promotes transcriptional reprograming needed to adapt to tissue residency, including induction of genes that regulate ECM interactions and remodeling, which affected eosinophil adhesion189. In contrast, AhR negatively regulates intestinal ILC2191. AhR facilitates its own gene chromatin accessibility and transcription, in cooperation with the transcription factor Gfi1. This presumably sequesters Gfi1, which is a positive regulator of ILC2 development through its binding to Il1rl1 locus. AhR-deficient ILC2 have increased IL-33 expression and effector functions, including being able to mediate better protection against H. polygyrus infection191. It remains unclear which environmental signals regulate this self-enhancing mechanism of AhR expression, as germ free mice or mice on a diet deprived of AhR ligands had comparable chromatin remodeling around the Ahr locus191. One possibility would be that the activity of IL4i1 secreted by alternatively activated macrophages provides a local source of such ligands, as discussed above. Whether Ahr regulates intestinal Th2 cells is not known, but it is interesting that AhR expression in mouse SI Th2 cells (Figure 2), and in human Th2 cells from nasal polyps or endoscopic biopsies correlates with the expression of effector cytokines and the SCFA receptors40, 41. Together, environmental sensing through AhR could affect type 2 immune cells in the intestine, favoring homeostatic functions, while inhibiting excessive inflammatory type 2 responses.

Intestinal type 2 immune cells participate in the monitoring of metabolic homeostasis and appear well equipped to respond to alterations related to malnutrition and parasitic infections. Their intracellular metabolic adaptability might enable them to mount a prolonged protective response in the presence of metabolic stress. As dietary interventions are being proposed for the treatment of chronic inflammatory disorders, including asthma192, a better understanding of the relationships between nutrients, microbiota, and mucosal type 2 immune responses could inform such therapeutical strategies.

Type 2 immune cells in adipose tissue and systemic metabolism

Adipose tissue plays a central role in regulating systemic metabolism. Composition and function of this highly plastic tissue differs depending on the type and anatomical location. The primary task of white adipose tissue (WAT) is energy storage, insulation and organ protection, with the main depots being visceral (VAT) and subcutaneous adipose tissues (scWAT). Brown adipose tissue (BAT), which is less abundant and developmentally distinct from WAT, is able to generate heat upon cold exposure through adaptive thermogenesis. Beige adipose tissue is developmentally related to WAT, but shares characteristics of thermogenic BAT, including high mitochondrial content and expression of uncoupling protein 1 (UCP-1), which permits dissipation of the mitochondrial proton gradient in a manner that generates heat193. Beyond its well-recognized roles in lipid storage and body temperature regulation, adipose tissue participates in immune surveillance, host defense, and tissue regeneration. In addition to adipocytes, it hosts heterogeneous populations of immune cells, stromal cells in various stages of commitment towards mature adipocytes, as well as neurons, and endothelial cells. Cross-talk between these cell types impacts tissue homeostasis and function194.

Metabolically healthy WAT responds to dynamically fluctuating levels of nutrients and contributes to circulating glucose homeostasis. Elevated levels of glucose and FA activate insulin production in the pancreas, to which adipocytes respond by increasing glucose uptake, TAG synthesis, and storage. However, chronic overnutrition leads to adipose tissue inflammation and expansion, and inflammatory signals, such as CCL2 and TNF, further recruit and polarize immune cells, e.g. macrophages, towards inflammatory phenotypes195. The paradigm in which inflammatory macrophages promote obesity, while alternatively activated macrophages maintain healthy adipose tissue through phagocytosis of dead adipocytes and anti-inflammatory cytokine secretion, has provided a framework for conceptualizing how the balance between type 1 and type 2 inflammation contributes to adipose tissue physiology. However, our understanding of the versatile roles of macrophages in adipose tissues has expanded greatly in recent years, providing a more complex picture. Several macrophage populations, positioned in distinct anatomical niches, have been identified194, 196. Understanding their roles in metabolic homeostasis is an important task for ongoing and future research.

Initial inflammatory responses associated with expanding adipose tissue are considered beneficial in that they can help the tissue to cope with prolonged nutritional stress by, for instance, enhancing angiogenesis to ameliorate hypoxia in the growing tissue197. Gradually however, chronic inflammation contributes to resistance to anabolic signals provided by insulin, as well as catabolic signals provided by catecholamines and leptin, leaving adipocytes metabolically inflexible. In this state, both energy storage and expenditure are dysregulated. Ultimately, such disfunction contributes to the development of metabolic disorders including type 2 diabetes mellitus, non-alcoholic fatty liver disease (NAFLD) and cardiovascular disease195.

Healthy WAT is enriched in type 2 immune cells, which participate in sensing and tuning nutritional and thermoregulatory signals. Mice deficient in components of type 2 immunity, including IL-13198, 199 and IL-33200, 201 have increased body weight compared to control animals. Catabolic processes that promote energy expenditure within adipose tissue are under control of the sympathetic nervous system, with the main neurotransmitters being catecholamines, including epinephrine and norepinephrine. These signal via β-adrenergic receptors to promote lipid mobilization and thermogenesis193, 202. Research over the past decade has established a link between adipose tissue type 2 immune cells and sympathetic signals, and a major role for ILC2, which coordinate eosinophil and macrophage responses203, 204. IL-5 expression by ILC2 maintains resident eosinophils, which are the main source of IL-4 in the adipose tissue. In turn, IL-4, together with ILC2-derived IL-13, skews adipose macrophages towards an alternatively activated phenotype, improving insulin sensitivity and resistance to HFD-induced obesity205, 206. The eosinophil – macrophage axis is also thought to underlie beneficial effects on glucose homeostasis that are observed when type 2 inflammation is induced by helminth infections207. A consequence of helminth infection is also an establishment of an adipose Th2RM population, which becomes a major source of type 2 cytokines and maintains eosinophils in the adipose tissue14. However, whether these cells take over homeostatic roles of ILC2 in promoting long-term metabolic homeostasis remains to be tested. Decreased frequencies of WAT eosinophils are observed in aged mice and humans, which may contribute to deteriorating metabolic health associated with aging208. Transfer of eosinophils from young animals decreases systemic low-grade inflammation and adipose hypertrophy observed in aged mice, partially in an IL-4 dependent manner208.

Signals derived from adipose stromal cells are essential in providing the niche for type 2 immune cells. Eosinophil recruitment to adipose tissue is influenced by the chemokine CCL11, which is secreted by stromal cells in response to IL-4 and IL-13209, or by mature adipocytes following sympathetic stimulation210. Adipose tissue resident ILC2 also respond to stroma-derived signals, particularly IL-33206, 211. A DPP4+ multipotent progenitor cell stromal population, analogous to universal fibroblast progenitors212, is the main source of IL-33 in adipose tissue14, 201, 209, 213, 214, 215. These DPP4+ cells comprise the stem compartment for renewal of adipocytes within adipose tissues216. In adipose, but also in other organs, these cells are enriched in interstitial niches, where they are in close proximity to blood vessels and lymphatics, positioning that may facilitate sensing of metabolic cues, and interactions with vasculature-associated macrophages, ILC2, and Th2 cells14, 217, 218, 219, 220. Signals that promote IL-33 production in adipose tissue are not well understood, but TNF, IL-17A, as well as sympathetic stimulation, have been implicated214, 221. Mesothelial cells also express IL-33 in adipose tissue, providing an additional source of this cytokine during inflammatory stress201 or in aged animals222. The switch in IL-33 source in aged adipose tissue is associated with a senescence-like phenotype of ILC2, indicating that the cellular context of IL-33 production might impact ILC2 functionality and contribute to age-related metabolic dysfunction222. Glial-derived neurotrophic factor (GDNF) is another stromal-derived factor that maintains ILC2 in VAT223. Hypothalamus-derived signals via sympathetic neurons induce GDNF production in stromal cells, and this in turn supports ILC2 numbers and cytokine production through the tyrosine kinase receptor RET. Mice harbouring RET-deficient ILC2 have increased body weights and impaired glucose tolerance compared with control animals223. Stromal cells can also support ILC2 in a contact-dependent manner, where binding of LFA-1 expressed by ILC2 with ICAM-1 expressed by stromal cells induces ILC2 proliferation and IL-5 production209.

Type 2 inflammation expands DPP4+ multipotent progenitor cells in the VAT14, while signalling through IL-4Rα increases proliferation of adipocyte progenitors (broadly defined as PDGFRα+ Sca1+ stromal cells)224. On the other hand, adipose eosinophils, macrophages, and Th2 cells can produce TGFβ114, 225,226 - a cytokine that is essential for the maintenance of multipotent progenitor cells, as it inhibits their differentiation towards committed adipocytes216. Therefore, the balance between different type 2 cytokines might regulate the pool of highly plastic progenitor cells, priming the tissue for rapid adaptation to changing conditions, i.e. by allowing for hypertrophic tissue expansion in response to calorie excess, and promoting beige adipocyte development when increased energy expenditure is required.

Early studies indicated that mice deficient in eosinophils or IL-5 show decreased oxygen consumption and heat production, indicating that these cells play a role in enhancing oxidative metabolism and energy expenditure206. Increased energy expenditure in the adipose tissue is promoted by the activity of brown and beige adipocytes. Signals derived from type 2 immune cells aid in this metabolic regulation mainly by facilitating beiging of WAT in the subcutaneous compartments200, 224, 227, which is also known to be more prone to beiging compared to visceral WAT228. In particular, signaling through IL-4Rα on stromal cells, but not on mature adipocytes, promotes beiging of scWAT, which contributes to increased energy expenditure in a UCP-1 dependent manner224. Both ILC2- and eosinophil-derived IL-4 and IL-13 promote beiging of scWAT224. Eosinophils might contribute to beiging through factors beyond IL-4, as eosinophils deficient in the transcriptional repressor KLF3 express more of the beige fat-inducing factor meteorin-like229, 230 (Metrn), and mice with KLF3-deficiency have enhanced beiging in scWAT230. Several studies have implicated alternatively activated macrophages in beiging227, 229, 231, 232, 233, however precisely how these cells perform this function is not well understood234, 235. Additionally, ILC2 induce beiging at the later stages of adipocyte differentiation through the production of methionine-enkephalin (Met-Enk) from the opioid peptide proenkephalin A200.

The physiological importance of type 2 immune cells in the activation of thermogenesis within BAT of adult animals has not been convincingly demonstrated235. This suggests that the main impact of type 2 signals is on WAT beiging, which might serve functions beyond classical heat generation. Beige adipocytes, with high mitochondrial content compared to white adipocytes, are more efficient in utilizing glucose and lipids, likely contributing to improved glucose homeostasis that is associated with BAT activity236. Additionally, brown and beige adipocytes have endocrine functions distinct from WAT. Indeed, in recent years several brown adipokines (batokines), which regulate immune cell recruitment, as well as liver, pancreas, and muscle metabolism, have been identified237. For example, neuregulin 4 (Nrg4) expression is increased during cold acclimation in BAT and scWAT and acts on liver to downregulate lipogenesis. Nrg4-deficient mice have increased insulin resistance and develop hepatic steatosis when fed HFD238. Thus, through such mechanisms, beiging induced by type 2 signals might contribute to metabolic regulation in the homeostatic state, but the same pathways could also be utilized during prolonged type 2 inflammation, when perhaps a new energy homeostasis setpoint needs to be established.

While WAT is a central metabolic hub, other organs including pancreas, liver, and muscle participate in the maintenance of whole-body metabolic homeostasis. Mediators of type 2 immunity are emerging as important metabolic regulators also in these tissues3, 4, 239, 240, indicating that much remains to be discovered about the roles that type 2 immune cells play in the regulation of metabolic health.

Concluding Remarks

Here we have discussed how type 2 immune cells adopt specific metabolic pathways to support their functional needs. Decisions about cellular differentiation and activation are significantly impacted by signals from cytokines along with intersecting signals from changes in nutrient and metabolite levels in the environment, which in turn drive emphasis on particular metabolic pathways that support or limit cellular function. This remains a rich and diverse area for investigation.

While advances have been made in understanding the roles of metabolic pathways in the biology of type 2 immune cells in an in vitro context, it is clear that these conditions do not fully capture the conditions in vivo, where cells are undoubtedly exposed to more complex stimuli, with consequences for their metabolic profiles. Moreover, scRNAseq has revealed the heterogeneity of activation states in vivo, and while transcriptional signatures can be informative about the metabolic status of individual cells, single cell metabolomics, ideally with matched scRNAseq, will be required for high-definition characterization. These approaches will hopefully become available in the near future. This would not only open the door to understanding metabolic reprogramming at the single cell level, but also to the metabolic analysis of cells that are integral to type 2 immunity, but which currently remain largely uncharacterized metabolically, such as eosinophils, cDC2, basophils and mast cells.

There are certain metabolic features that have been found to be common to multiple Type 2 immune cells. For example, tissue ILC2, Th2 cells and alternatively activated macrophages share increased expression of genes involved in fatty acid metabolism downstream of PPARγ activation. PPARγ binds medium to long-chain fatty acids and certain eicosanoids, and it is of particular interest in this light that type 2 immunity is so heavily implicated in adipose tissue homeostasis.

As a final point, it is noteworthy that rapid progress has been made in biologic therapies for allergic diseases and asthma, with multiple antibody therapies that target critical components of type 2 immunity, such as IL-4, IL-5, IL-33, and TSLP either already on the market or in trials241. Many of these drugs have remarkably beneficial effects. An inevitable outcome of their extended use will be clarification of whether conclusions drawn from the study of specific aspects of type 2 immunity, often in mice, hold true in human biology over the long term. For example, how important are IL-4, IL-5 or IL-33 for adipose tissue homeostasis and consequently is weight gain or insulin sensitivity a concern in people treated with these biologics? Likewise, does IL-4 signaling inhibition, with the attendant loss of expression of genes such as Arg1 in alternatively activated TAMs, confer advantage in cancer? Insights from the use of these drugs thus promise to be incredibly informative and valuable.

Key elements of Type 2 immune responses are integrated within tissues and are part of tissue homeostasis and repair. Pearce and colleagues review how metabolic regulation relates to type 2 immunity, discussing both specifics of metabolism and metabolic adaptation within type 2 immune cells and how type 2 immune cells are integrated more broadly into the metabolism of the organism as a whole.

Acknowledgments

We thank members of the EJP and ELP laboratories, and especially Drs. Anna Kania, Katarzyna Grzes and David Sanin for helpful discussions.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Pearce EL & Pearce EJ Metabolic pathways in immune cell activation and quiescence. Immunity 38, 633–643 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Caputa G, Castoldi A. & Pearce EJ Metabolic adaptations of tissue-resident immune cells. Nat Immunol 20, 793–801 (2019). [DOI] [PubMed] [Google Scholar]
  • 3.Dalmas E. et al. Interleukin-33-Activated Islet-Resident Innate Lymphoid Cells Promote Insulin Secretion through Myeloid Cell Retinoic Acid Production. Immunity 47, 928–942 e927 (2017). [DOI] [PubMed] [Google Scholar]
  • 4.Fujimoto M. et al. Liver group 2 innate lymphoid cells regulate blood glucose levels through IL-13 signaling and suppression of gluconeogenesis. Nat Commun 13, 5408 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Spits H. & Mjosberg J. Heterogeneity of type 2 innate lymphoid cells. Nat. Rev. Immunol (2022). [DOI] [PMC free article] [PubMed]
  • 6.Hondowicz BD et al. Interleukin-2-Dependent Allergen-Specific Tissue-Resident Memory Cells Drive Asthma. Immunity 44, 155–166 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Rahimi RA, Nepal K, Cetinbas M, Sadreyev RI & Luster AD Distinct functions of tissue-resident and circulating memory Th2 cells in allergic airway disease. J. Exp. Med 217 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Turner DL et al. Biased Generation and In Situ Activation of Lung Tissue-Resident Memory CD4 T Cells in the Pathogenesis of Allergic Asthma. J. Immunol 200, 1561–1569 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Van Dyken SJ et al. A tissue checkpoint regulates type 2 immunity. Nat Immunol 17, 1381–1387 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Filbey KJ et al. Intestinal helminth infection promotes IL-5- and CD4(+) T cell-dependent immunity in the lung against migrating parasites. Mucosal Immunol. 12, 352–362 (2019). [DOI] [PubMed] [Google Scholar]
  • 11.Classon CH et al. Intestinal helminth infection transforms the CD4(+) T cell composition of the skin. Mucosal Immunol 15, 257–267 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Ricardo-Gonzalez RR, Molofsky AB & Locksley RM ILC2s - development, divergence, dispersal. Curr. Opin. Immunol 75, 102168 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Tibbitt CA et al. Single-Cell RNA Sequencing of the T Helper Cell Response to House Dust Mites Defines a Distinct Gene Expression Signature in Airway Th2 Cells. Immunity 51, 169–184 e165 (2019). [DOI] [PubMed] [Google Scholar]
  • 14.Kabat AM et al. Resident TH2 cells orchestrate adipose tissue remodeling at a site adjacent to infection. Sci Immunol 7, eadd3263 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Guo L. et al. Innate immunological function of TH2 cells in vivo. Nat Immunol 16, 1051–1059 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Minutti CM et al. Epidermal Growth Factor Receptor Expression Licenses Type-2 Helper T Cells to Function in a T Cell Receptor-Independent Fashion. Immunity 47, 710–722 e716 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Okano M. et al. Interleukin-33-activated neuropeptide CGRP-producing memory Th2 cells cooperate with somatosensory neurons to induce conjunctival itch. Immunity (2022). [DOI] [PubMed]
  • 18.Jarjour NN et al. BHLHE40 Promotes TH2 Cell-Mediated Antihelminth Immunity and Reveals Cooperative CSF2RB Family Cytokines. J. Immunol 204, 923–932 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.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]
  • 20.Klysz D. et al. Glutamine-dependent alpha-ketoglutarate production regulates the balance between T helper 1 cell and regulatory T cell generation. Sci Signal 8, ra97 (2015). [DOI] [PubMed] [Google Scholar]
  • 21.Yang K. et al. T cell exit from quiescence and differentiation into Th2 cells depend on Raptor-mTORC1-mediated metabolic reprogramming. Immunity 39, 1043–1056 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Walker JA & McKenzie ANJ TH2 cell development and function. Nat. Rev. Immunol 18, 121–133 (2018). [DOI] [PubMed] [Google Scholar]
  • 23.Ross SH & Cantrell DA Signaling and Function of Interleukin-2 in T Lymphocytes. Annu. Rev. Immunol 36, 411–433 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Villarino AV et al. A central role for STAT5 in the transcriptional programing of T helper cell metabolism. Sci Immunol 7, eabl9467 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Berod L. et al. De novo fatty acid synthesis controls the fate between regulatory T and T helper 17 cells. Nat. Med 20, 1327–1333 (2014). [DOI] [PubMed] [Google Scholar]
  • 26.Endo Y. et al. ACC1 determines memory potential of individual CD4(+) T cells by regulating de novo fatty acid biosynthesis. Nat Metab 1, 261–275 (2019). [DOI] [PubMed] [Google Scholar]
  • 27.van der Windt GJ et al. Mitochondrial respiratory capacity is a critical regulator of CD8+ T cell memory development. Immunity 36, 68–78 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Nakajima T. et al. ACC1-expressing pathogenic T helper 2 cell populations facilitate lung and skin inflammation in mice. J. Exp. Med 218 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Mahata B. et al. Single-cell RNA sequencing reveals T helper cells synthesizing steroids de novo to contribute to immune homeostasis. Cell Rep 7, 1130–1142 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Huang SCC et al. Metabolic Reprogramming Mediated by the mTORC2-IRF4 Signaling Axis Is Essential for Macrophage Alternative Activation. Immunity 45, 817–830 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Wang YX PPARs: diverse regulators in energy metabolism and metabolic diseases. Cell Res. 20, 124–137 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Chen T. et al. PPAR-gamma promotes type 2 immune responses in allergy and nematode infection. Sci Immunol 2 (2017). [DOI] [PubMed] [Google Scholar]
  • 33.Nobs SP et al. PPARgamma in dendritic cells and T cells drives pathogenic type-2 effector responses in lung inflammation. J Exp Med 214, 3015–3035 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Angela M. et al. Fatty acid metabolic reprogramming via mTOR-mediated inductions of PPARgamma directs early activation of T cells. Nat Commun 7, 13683 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Henriksson J. et al. Genome-wide CRISPR Screens in T Helper Cells Reveal Pervasive Crosstalk between Activation and Differentiation. Cell 176, 882–896 e818 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Cook ME, Jarjour NN, Lin CC & Edelson BT Transcription Factor Bhlhe40 in Immunity and Autoimmunity. Trends Immunol. 41, 1023–1036 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Li C. et al. The Transcription Factor Bhlhe40 Programs Mitochondrial Regulation of Resident CD8(+) T Cell Fitness and Functionality. Immunity 51, 491–507 e497 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Wambre E. et al. A phenotypically and functionally distinct human TH2 cell subpopulation is associated with allergic disorders. Sci. Transl. Med 9 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Seumois G. et al. Single-cell transcriptomic analysis of allergen-specific T cells in allergy and asthma. Sci Immunol 5 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Ma J. et al. Single-cell analysis pinpoints distinct populations of cytotoxic CD4(+) T cells and an IL-10(+)CD109(+) TH2 cell population in nasal polyps. Sci Immunol 6 (2021). [DOI] [PubMed] [Google Scholar]
  • 41.Wen T. et al. Single-cell RNA sequencing identifies inflammatory tissue T cells in eosinophilic esophagitis. J. Clin. Invest 129, 2014–2028 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Vieira Braga FA et al. A cellular census of human lungs identifies novel cell states in health and in asthma. Nat. Med 25, 1153–1163 (2019). [DOI] [PubMed] [Google Scholar]
  • 43.Micosse C. et al. Human “TH9” cells are a subpopulation of PPAR-gamma(+) TH2 cells. Sci Immunol 4 (2019). [DOI] [PubMed] [Google Scholar]
  • 44.Ulrich BJ et al. Allergic airway recall responses require IL-9 from resident memory CD4(+) T cells. Sci Immunol 7, eabg9296 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Pan Y. et al. Survival of tissue-resident memory T cells requires exogenous lipid uptake and metabolism. Nature 543, 252–256 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Jarjour NN et al. Bhlhe40 mediates tissue-specific control of macrophage proliferation in homeostasis and type 2 immunity. Nat. Immunol 20, 687–700 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Ercolano G. et al. PPAR drives IL-33-dependent ILC2 pro-tumoral functions. Nat Commun 12, 2538 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Fali T. et al. Metabolic regulation by PPARgamma is required for IL-33-mediated activation of ILC2s in lung and adipose tissue. Mucosal Immunol. 14, 585–593 (2021). [DOI] [PubMed] [Google Scholar]
  • 49.Karagiannis F. et al. Lipid-Droplet Formation Drives Pathogenic Group 2 Innate Lymphoid Cells in Airway Inflammation. Immunity 52, 620–634 e626 (2020). [DOI] [PubMed] [Google Scholar]
  • 50.Xiao Q. et al. PPARgamma enhances ILC2 function during allergic airway inflammation via transcription regulation of ST2. Mucosal Immunol. 14, 468–478 (2021). [DOI] [PubMed] [Google Scholar]
  • 51.Proserpio V. et al. Single-cell analysis of CD4+ T-cell differentiation reveals three major cell states and progressive acceleration of proliferation. Genome Biol. 17, 103 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Yun Z, Maecker HL, Johnson RS & Giaccia AJ Inhibition of PPAR gamma 2 gene expression by the HIF-1-regulated gene DEC1/Stra13: a mechanism for regulation of adipogenesis by hypoxia. Dev. Cell 2, 331–341 (2002). [DOI] [PubMed] [Google Scholar]
  • 53.Surace L. et al. Dichotomous metabolic networks govern human ILC2 proliferation and function. Nat. Immunol 22, 1367–1374 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Robinette ML et al. Transcriptional programs define molecular characteristics of innate lymphoid cell classes and subsets. Nat. Immunol 16, 306–317 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Hodge SH et al. Amino acid availability acts as a metabolic rheostat to determine the magnitude of ILC2 responses. J. Exp. Med 220 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Panda SK et al. SLC7A8 is a key amino acids supplier for the metabolic programs that sustain homeostasis and activation of type 2 innate lymphoid cells. Proc. Natl. Acad. Sci. U. S. A 119, e2215528119 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Monticelli LA et al. Arginase 1 is an innate lymphoid-cell-intrinsic metabolic checkpoint controlling type 2 inflammation. Nat. Immunol 17, 656–665 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Puleston DJ et al. Polyamines and eIF5A Hypusination Modulate Mitochondrial Respiration and Macrophage Activation. Cell Metab 30, 352–363 e358 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Wilhelm C. et al. Critical role of fatty acid metabolism in ILC2-mediated barrier protection during malnutrition and helminth infection. J. Exp. Med 213, 1409–1418 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Galle-Treger L. et al. Autophagy is critical for group 2 innate lymphoid cell metabolic homeostasis and effector function. J. Allergy Clin. Immunol 145, 502–517 e505 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Kabat AM et al. The autophagy gene Atg16l1 differentially regulates Treg and TH2 cells to control intestinal inflammation. Elife 5, e12444 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Li Q. et al. E3 Ligase VHL Promotes Group 2 Innate Lymphoid Cell Maturation and Function via Glycolysis Inhibition and Induction of Interleukin-33 Receptor. Immunity 48, 258–270 e255 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Murray PJ & Wynn TA Protective and pathogenic functions of macrophage subsets. Nat Rev Immunol 11, 723–737 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Sica A. & Mantovani A. Macrophage plasticity and polarization: in vivo veritas. J Clin Invest 122, 787–795 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Xue J. et al. Transcriptome-based network analysis reveals a spectrum model of human macrophage activation. Immunity 40, 274–288 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Murray PJ Macrophage Polarization. Annu Rev Physiol 79, 541–566 (2017). [DOI] [PubMed] [Google Scholar]
  • 67.Gause WC, Wynn TA & Allen JE Type 2 immunity and wound healing: evolutionary refinement of adaptive immunity by helminths. Nat. Rev. Immunol 13, 607–614 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Wynn TA, Chawla A. & Pollard JW Macrophage biology in development, homeostasis and disease. Nature 496, 445–455 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Eming SA, Murray PJ & Pearce EJ Metabolic orchestration of the wound healing response. Cell Metab 33, 1726–1743 (2021). [DOI] [PubMed] [Google Scholar]
  • 70.Lechner A, Bohnacker S. & Esser-von Bieren J. Macrophage regulation & function in helminth infection. Semin. Immunol 53, 101526 (2021). [DOI] [PubMed] [Google Scholar]
  • 71.Odegaard JI et al. Macrophage-specific PPARgamma controls alternative activation and improves insulin resistance. Nature 447, 1116–1120 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Vats D. et al. Oxidative metabolism and PGC-1beta attenuate macrophage-mediated inflammation. Cell Metab 4, 13–24 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Thomas GD et al. The biology of nematode- and IL4Ralpha-dependent murine macrophage polarization in vivo as defined by RNA-Seq and targeted lipidomics. Blood 120, e93–e104 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Szanto A. et al. STAT6 transcription factor is a facilitator of the nuclear receptor PPARgamma-regulated gene expression in macrophages and dendritic cells. Immunity 33, 699–712 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Daniel B. et al. The Nuclear Receptor PPARgamma Controls Progressive Macrophage Polarization as a Ligand-Insensitive Epigenomic Ratchet of Transcriptional Memory. Immunity 49, 615–626 e616 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Yan J. & Horng T. Lipid Metabolism in Regulation of Macrophage Functions. Trends Cell Biol 30, 979–989 (2020). [DOI] [PubMed] [Google Scholar]
  • 77.van Teijlingen Bakker N. & Pearce EJ Cell-intrinsic metabolic regulation of mononuclear phagocyte activation: Findings from the tip of the iceberg. Immunol Rev 295, 54–67 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Wang Y, Li N, Zhang X. & Horng T. Mitochondrial metabolism regulates macrophage biology. J Biol Chem 297, 100904 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Jha AK et al. Network integration of parallel metabolic and transcriptional data reveals metabolic modules that regulate macrophage polarization. Immunity 42, 419–430 (2015). [DOI] [PubMed] [Google Scholar]
  • 80.Van den Bossche J. et al. Mitochondrial Dysfunction Prevents Repolarization of Inflammatory Macrophages. Cell Rep 17, 684–696 (2016). [DOI] [PubMed] [Google Scholar]
  • 81.Runtsch MC et al. Itaconate and itaconate derivatives target JAK1 to suppress alternative activation of macrophages. Cell Metab 34, 487–501 e488 (2022). [DOI] [PubMed] [Google Scholar]
  • 82.Liu PS et al. alpha-ketoglutarate orchestrates macrophage activation through metabolic and epigenetic reprogramming. Nat Immunol 18, 985–994 (2017). [DOI] [PubMed] [Google Scholar]
  • 83.Petty AJ et al. Hedgehog signaling promotes tumor-associated macrophage polarization to suppress intratumoral CD8+ T cell recruitment. J Clin Invest 129, 5151–5162 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Hinshaw DC et al. Hedgehog Signaling Regulates Metabolism and Polarization of Mammary Tumor-Associated Macrophages. Cancer Res 81, 5425–5437 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Covarrubias AJ et al. Akt-mTORC1 signaling regulates Acly to integrate metabolic input to control of macrophage activation. Elife 5 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Nomura M. et al. Fatty acid oxidation in macrophage polarization. Nat Immunol 17, 216–217 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Divakaruni AS et al. Etomoxir Inhibits Macrophage Polarization by Disrupting CoA Homeostasis. Cell Metab 28, 490–503 e497 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Namgaladze D. & Brune B. Macrophage fatty acid oxidation and its roles in macrophage polarization and fatty acid-induced inflammation. Biochim Biophys Acta 1861, 1796–1807 (2016). [DOI] [PubMed] [Google Scholar]
  • 89.Huang SC et al. Cell-intrinsic lysosomal lipolysis is essential for alternative activation of macrophages. Nat Immunol 15, 846–855 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Bidault G. et al. SREBP1-induced fatty acid synthesis depletes macrophages antioxidant defences to promote their alternative activation. Nat Metab 3, 1150–1162 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Zhang Y. et al. ROS play a critical role in the differentiation of alternatively activated macrophages and the occurrence of tumor-associated macrophages. Cell Res 23, 898–914 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Griess B, Mir S, Datta K. & Teoh-Fitzgerald M. Scavenging reactive oxygen species selectively inhibits M2 macrophage polarization and their pro-tumorigenic function in part, via Stat3 suppression. Free Radic Biol Med 147, 48–60 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Jenkins SJ et al. Local macrophage proliferation, rather than recruitment from the blood, is a signature of TH2 inflammation. Science 332, 1284–1288 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.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]
  • 95.Rahman K. et al. Inflammatory Ly6Chi monocytes and their conversion to M2 macrophages drive atherosclerosis regression. J. Clin. Invest 127, 2904–2915 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Castoldi A. et al. Triacylglycerol synthesis enhances macrophage inflammatory function. Nat Commun 11, 4107 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Bingle L, Brown NJ & Lewis CE The role of tumour-associated macrophages in tumour progression: implications for new anticancer therapies. J Pathol 196, 254–265 (2002). [DOI] [PubMed] [Google Scholar]
  • 98.DeNardo DG & Ruffell B. Macrophages as regulators of tumour immunity and immunotherapy. Nat Rev Immunol 19, 369–382 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Zilionis R. et al. Single-Cell Transcriptomics of Human and Mouse Lung Cancers Reveals Conserved Myeloid Populations across Individuals and Species. Immunity 50, 1317–1334 e1310 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Mulder K. et al. Cross-tissue single-cell landscape of human monocytes and macrophages in health and disease. Immunity 54, 1883–1900 e1885 (2021). [DOI] [PubMed] [Google Scholar]
  • 101.Geeraerts X. et al. Macrophages are metabolically heterogeneous within the tumor microenvironment. Cell Rep 37, 110171 (2021). [DOI] [PubMed] [Google Scholar]
  • 102.Rodriguez-Tirado C. et al. Interleukin 4 Controls the Pro-Tumoral Role of Macrophages in Mammary Cancer Pulmonary Metastasis in Mice. Cancers (Basel) 14 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Goossens P. et al. Membrane Cholesterol Efflux Drives Tumor-Associated Macrophage Reprogramming and Tumor Progression. Cell Metab. 29, 1376–1389 e1374 (2019). [DOI] [PubMed] [Google Scholar]
  • 104.Jaitin DA et al. Lipid-Associated Macrophages Control Metabolic Homeostasis in a Trem2-Dependent Manner. Cell 178, 686–698 e614 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Wculek SK et al. Oxidative phosphorylation selectively orchestrates tissue macrophage homeostasis. Immunity (2023). [DOI] [PubMed]
  • 106.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 e445 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Raines LN et al. PERK is a critical metabolic hub for immunosuppressive function in macrophages. Nat Immunol 23, 431–445 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Shan X. et al. Serine metabolism orchestrates macrophage polarization by regulating the IGF1-p38 axis. Cell Mol Immunol 19, 1263–1278 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Patel N. et al. A2B adenosine receptor induces protective antihelminth type 2 immune responses. Cell Host Microbe 15, 339–350 (2014). [DOI] [PubMed] [Google Scholar]
  • 110.Keiran N. et al. SUCNR1 controls an anti-inflammatory program in macrophages to regulate the metabolic response to obesity. Nat. Immunol 20, 581–592 (2019). [DOI] [PubMed] [Google Scholar]
  • 111.El-Naccache DW et al. Adenosine metabolized from extracellular ATP promotes type 2 immunity through triggering A(2B)AR signaling in intestinal epithelial cells. Cell Rep 40, 111150 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Schneider C, O’Leary CE & Locksley RM Regulation of immune responses by tuft cells. Nat. Rev. Immunol 19, 584–593 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Gundra UM et al. Vitamin A mediates conversion of monocyte-derived macrophages into tissue-resident macrophages during alternative activation. Nat Immunol 18, 642–653 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Kaneda MM et al. PI3Kgamma is a molecular switch that controls immune suppression. Nature 539, 437–442 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Pauleau AL et al. Enhancer-mediated control of macrophage-specific arginase I expression. J Immunol 172, 7565–7573 (2004). [DOI] [PubMed] [Google Scholar]
  • 116.Huang SC et al. Metabolic Reprogramming Mediated by the mTORC2-IRF4 Signaling Axis Is Essential for Macrophage Alternative Activation. Immunity 45, 817–830 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Hallowell RW et al. mTORC2 signalling regulates M2 macrophage differentiation in response to helminth infection and adaptive thermogenesis. Nat Commun 8, 14208 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Murray PJ Amino acid auxotrophy as a system of immunological control nodes. Nat Immunol 17, 132–139 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Van de Velde LA et al. Neuroblastoma Formation Requires Unconventional CD4 T Cells and Arginase-1-Dependent Myeloid Cells. Cancer Res 81, 5047–5059 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Bronte V. et al. IL-4-induced arginase 1 suppresses alloreactive T cells in tumor-bearing mice. J. Immunol 170, 270–278 (2003). [DOI] [PubMed] [Google Scholar]
  • 121.Chantranupong L. et al. The CASTOR Proteins Are Arginine Sensors for the mTORC1 Pathway. Cell 165, 153–164 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Pesce JT et al. Arginase-1-expressing macrophages suppress Th2 cytokine-driven inflammation and fibrosis. PLoS Pathog 5, e1000371 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Anthony RM et al. Memory T(H)2 cells induce alternatively activated macrophages to mediate protection against nematode parasites. Nat Med 12, 955–960 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Kieler M, Hofmann M. & Schabbauer G. More than just protein building blocks: how amino acids and related metabolic pathways fuel macrophage polarization. FEBS J 288, 3694–3714 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Fiore A. & Murray PJ Tryptophan and indole metabolism in immune regulation. Curr Opin Immunol 70, 7–14 (2021). [DOI] [PubMed] [Google Scholar]
  • 126.Chu CC & Paul WE Fig1, an interleukin 4-induced mouse B cell gene isolated by cDNA representational difference analysis. Proc Natl Acad Sci U S A 94, 2507–2512 (1997). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Boulland ML et al. Human IL4I1 is a secreted L-phenylalanine oxidase expressed by mature dendritic cells that inhibits T-lymphocyte proliferation. Blood 110, 220–227 (2007). [DOI] [PubMed] [Google Scholar]
  • 128.Sadik A. et al. IL4I1 Is a Metabolic Immune Checkpoint that Activates the AHR and Promotes Tumor Progression. Cell 182, 1252–1270 e1234 (2020). [DOI] [PubMed] [Google Scholar]
  • 129.Quintana FJ et al. Control of T(reg) and T(H)17 cell differentiation by the aryl hydrocarbon receptor. Nature 453, 65–71 (2008). [DOI] [PubMed] [Google Scholar]
  • 130.Marquet J. et al. Dichotomy between factors inducing the immunosuppressive enzyme IL-4-induced gene 1 (IL4I1) in B lymphocytes and mononuclear phagocytes. Eur. J. Immunol 40, 2557–2568 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Rieckmann JC et al. Social network architecture of human immune cells unveiled by quantitative proteomics. Nat Immunol 18, 583–593 (2017). [DOI] [PubMed] [Google Scholar]
  • 132.Yue Y. et al. IL4I1 Is a Novel Regulator of M2 Macrophage Polarization That Can Inhibit T Cell Activation via L-Tryptophan and Arginine Depletion and IL-10 Production. PLoS One 10, e0142979 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Zeitler L. et al. Anti-ferroptotic mechanism of IL4i1-mediated amino acid metabolism. Elife 10 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Piattini F. et al. Differential sensitivity of inflammatory macrophages and alternatively activated macrophages to ferroptosis. Eur J Immunol 51, 2417–2429 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Yu L. et al. Reprogramming alternative macrophage polarization by GATM-mediated endogenous creatine synthesis: A potential target for HDM-induced asthma treatment. Front. Immunol 13, 937331 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Ji L. et al. Slc6a8-Mediated Creatine Uptake and Accumulation Reprogram Macrophage Polarization via Regulating Cytokine Responses. Immunity 51, 272–284 e277 (2019). [DOI] [PubMed] [Google Scholar]
  • 137.Bando JK, Nussbaum JC, Liang HE & Locksley RM Type 2 innate lymphoid cells constitutively express arginase-I in the naive and inflamed lung. J. Leukoc. Biol 94, 877–884 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Galli SJ, Starkl P, Marichal T. & Tsai M. Mast cells and IgE in defense against venoms: Possible “good side” of allergy? Allergol Int 65, 3–15 (2016). [DOI] [PubMed] [Google Scholar]
  • 139.Gould HJ & Sutton BJ IgE in allergy and asthma today. Nat. Rev. Immunol 8, 205–217 (2008). [DOI] [PubMed] [Google Scholar]
  • 140.Fitzsimmons CM, Falcone FH & Dunne DW Helminth Allergens, Parasite-Specific IgE, and Its Protective Role in Human Immunity. Front. Immunol 5, 61 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Haase P. & Voehringer D. Regulation of the humoral type 2 immune response against allergens and helminths. Eur. J. Immunol 51, 273–279 (2021). [DOI] [PubMed] [Google Scholar]
  • 142.Hu J. et al. Anti-IgE therapy for IgE-mediated allergic diseases: from neutralizing IgE antibodies to eliminating IgE(+) B cells. Clin Transl Allergy 8, 27 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Egawa T. & Bhattacharya D. Regulation of metabolic supply and demand during B cell activation and subsequent differentiation. Curr. Opin. Immunol 57, 8–14 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Vivas-Garcia Y. & Efeyan A. The metabolic plasticity of B cells. Front Mol Biosci 9, 991188 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Waters LR, Ahsan FM, Wolf DM, Shirihai O. & Teitell MA Initial B Cell Activation Induces Metabolic Reprogramming and Mitochondrial Remodeling. iScience 5, 99–109 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Price MJ, Patterson DG, Scharer CD & Boss JM Progressive Upregulation of Oxidative Metabolism Facilitates Plasmablast Differentiation to a T-Independent Antigen. Cell Rep 23, 3152–3159 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Tsui C. et al. Protein Kinase C-beta Dictates B Cell Fate by Regulating Mitochondrial Remodeling, Metabolic Reprogramming, and Heme Biosynthesis. Immunity 48, 1144–1159 e1145 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Chen D. et al. Coupled analysis of transcriptome and BCR mutations reveals role of OXPHOS in affinity maturation. Nat. Immunol 22, 904–913 (2021). [DOI] [PubMed] [Google Scholar]
  • 149.Weisel FJ et al. Germinal center B cells selectively oxidize fatty acids for energy while conducting minimal glycolysis. Nat Immunol 21, 331–342 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Haniuda K, Fukao S. & Kitamura D. Metabolic Reprogramming Induces Germinal Center B Cell Differentiation through Bcl6 Locus Remodeling. Cell Rep 33, 108333 (2020). [DOI] [PubMed] [Google Scholar]
  • 151.Biram A. et al. Bacterial infection disrupts established germinal center reactions through monocyte recruitment and impaired metabolic adaptation. Immunity 55, 442–458 e448 (2022). [DOI] [PubMed] [Google Scholar]
  • 152.Oestreich KJ et al. Bcl-6 directly represses the gene program of the glycolysis pathway. Nat. Immunol 15, 957–964 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Ribeiro F, Perucha E. & Graca L. T follicular cells: The regulators of germinal center homeostasis. Immunol. Lett 244, 1–11 (2022). [DOI] [PubMed] [Google Scholar]
  • 154.Gowthaman U. et al. Identification of a T follicular helper cell subset that drives anaphylactic IgE. Science 365 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Choi SC et al. Inhibition of glucose metabolism selectively targets autoreactive follicular helper T cells. Nat Commun 9, 4369 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Utley A. et al. CD28 Regulates Metabolic Fitness for Long-Lived Plasma Cell Survival. Cell Rep 31, 107815 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Klein Geltink RI et al. Mitochondrial Priming by CD28. Cell 171, 385–397 e311 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Lam WY & Bhattacharya D. Metabolic Links between Plasma Cell Survival, Secretion, and Stress. Trends Immunol. 39, 19–27 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Lam WY et al. Mitochondrial Pyruvate Import Promotes Long-Term Survival of Antibody-Secreting Plasma Cells. Immunity 45, 60–73 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Lam WY et al. Metabolic and Transcriptional Modules Independently Diversify Plasma Cell Lifespan and Function. Cell Rep 24, 2479–2492 e2476 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Jennewein MF & Alter G. The Immunoregulatory Roles of Antibody Glycosylation. Trends Immunol. 38, 358–372 (2017). [DOI] [PubMed] [Google Scholar]
  • 162.Shade KC et al. Sialylation of immunoglobulin E is a determinant of allergic pathogenicity. Nature 582, 265–270 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Yang Z, Sullivan BM & Allen CD Fluorescent in vivo detection reveals that IgE(+) B cells are restrained by an intrinsic cell fate predisposition. Immunity 36, 857–872 (2012). [DOI] [PubMed] [Google Scholar]
  • 164.Newman R. & Tolar P. Chronic calcium signaling in IgE(+) B cells limits plasma cell differentiation and survival. Immunity 54, 2756–2771 e2710 (2021). [DOI] [PubMed] [Google Scholar]
  • 165.Luger EO et al. Allergy for a lifetime? Allergol Int 59, 1–8 (2010). [DOI] [PubMed] [Google Scholar]
  • 166.Asrat S. et al. Chronic allergen exposure drives accumulation of long-lived IgE plasma cells in the bone marrow, giving rise to serological memory. Sci Immunol 5 (2020). [DOI] [PubMed] [Google Scholar]
  • 167.Hoh RA et al. Origins and clonal convergence of gastrointestinal IgE(+) B cells in human peanut allergy. Sci Immunol 5 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Rooks MG & Garrett WS Gut microbiota, metabolites and host immunity. Nat. Rev. Immunol 16, 341–352 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Spencer SP et al. Adaptation of innate lymphoid cells to a micronutrient deficiency promotes type 2 barrier immunity. Science 343, 432–437 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Shea-Donohue T, Qin B. & Smith A. Parasites, nutrition, immune responses and biology of metabolic tissues. Parasite Immunol. 39 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Rytter MJ, Kolte L, Briend A, Friis H. & Christensen VB The immune system in children with malnutrition--a systematic review. PLoS One 9, e105017 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Florsheim EB, Sullivan ZA, Khoury-Hanold W. & Medzhitov R. Food allergy as a biological food quality control system. Cell 184, 1440–1454 (2021). [DOI] [PubMed] [Google Scholar]
  • 173.Aguilera-Lizarraga J. et al. Local immune response to food antigens drives meal-induced abdominal pain. Nature 590, 151–156 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Sorobetea D, Svensson-Frej M. & Grencis R. Immunity to gastrointestinal nematode infections. Mucosal Immunol. 11, 304–315 (2018). [DOI] [PubMed] [Google Scholar]
  • 175.Kokova D. & Mayboroda OA Twenty Years on: Metabolomics in Helminth Research. Trends Parasitol 35, 282–288 (2019). [DOI] [PubMed] [Google Scholar]
  • 176.Kokova D. et al. Metabolic Homeostasis in Chronic Helminth Infection Is Sustained by Organ-Specific Metabolic Rewiring. ACS Infect Dis 7, 906–916 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Houlden A. et al. Chronic Trichuris muris Infection in C57BL/6 Mice Causes Significant Changes in Host Microbiota and Metabolome: Effects Reversed by Pathogen Clearance. PLoS One 10, e0125945 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Schneider C. et al. A Metabolite-Triggered Tuft Cell-ILC2 Circuit Drives Small Intestinal Remodeling. Cell 174, 271–284 e214 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Schalter F. et al. Acetate, a metabolic product of Heligmosomoides polygyrus, facilitates intestinal epithelial barrier breakdown in a FFAR2-dependent manner. Int. J. Parasitol 52, 591–601 (2022). [DOI] [PubMed] [Google Scholar]
  • 180.de Aguiar Vallim TQ, Tarling EJ & Edwards PA Pleiotropic roles of bile acids in metabolism. Cell Metab. 17, 657–669 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Godlewska U, Bulanda E. & Wypych TP Bile acids in immunity: Bidirectional mediators between the host and the microbiota. Front. Immunol 13, 949033 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Mohammed AD et al. Defective humoral immunity disrupts bile acid homeostasis which promotes inflammatory disease of the small bowel. Nat Commun 13, 525 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Arifuzzaman M. et al. Inulin fibre promotes microbiota-derived bile acids and type 2 inflammation. Nature 611, 578–584 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Hibberd MC et al. The effects of micronutrient deficiencies on bacterial species from the human gut microbiota. Sci. Transl. Med 9 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Constantine GM & Klion AD Recent advances in understanding the role of eosinophils. Fac Rev 11, 26 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Arnold IC et al. Eosinophils suppress Th1 responses and restrict bacterially induced gastrointestinal inflammation. J. Exp. Med 215, 2055–2072 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Sugawara R. et al. Small intestinal eosinophils regulate Th17 cells by producing IL-1 receptor antagonist. J. Exp. Med 213, 555–567 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Ignacio A. et al. Small intestinal resident eosinophils maintain gut homeostasis following microbial colonization. Immunity 55, 1250–1267 e1212 (2022). [DOI] [PubMed] [Google Scholar]
  • 189.Diny NL et al. The aryl hydrocarbon receptor contributes to tissue adaptation of intestinal eosinophils in mice. J. Exp. Med 219 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Stockinger B, Shah K. & Wincent E. AHR in the intestinal microenvironment: safeguarding barrier function. Nat. Rev. Gastroenterol. Hepatol 18, 559–570 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Li S. et al. Aryl Hydrocarbon Receptor Signaling Cell Intrinsically Inhibits Intestinal Group 2 Innate Lymphoid Cell Function. Immunity 49, 915–928 e915 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Wilhelm C, Surendar J. & Karagiannis F. Enemy or ally? Fasting as an essential regulator of immune responses. Trends Immunol. 42, 389–400 (2021). [DOI] [PubMed] [Google Scholar]
  • 193.Shapira SN & Seale P. Transcriptional Control of Brown and Beige Fat Development and Function. Obesity (Silver Spring) 27, 13–21 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Trim WV & Lynch L. Immune and non-immune functions of adipose tissue leukocytes. Nat Rev Immunol 22, 371–386 (2022). [DOI] [PubMed] [Google Scholar]
  • 195.Reilly SM & Saltiel AR Adapting to obesity with adipose tissue inflammation. Nat. Rev. Endocrinol 13, 633–643 (2017). [DOI] [PubMed] [Google Scholar]
  • 196.Russo L. & Lumeng CN Properties and functions of adipose tissue macrophages in obesity. Immunology 155, 407–417 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Ghaben AL & Scherer PE Adipogenesis and metabolic health. Nat. Rev. Mol. Cell Biol 20, 242–258 (2019). [DOI] [PubMed] [Google Scholar]
  • 198.Hams E, Locksley RM, McKenzie AN & Fallon PG Cutting edge: IL-25 elicits innate lymphoid type 2 and type II NKT cells that regulate obesity in mice. J. Immunol 191, 5349–5353 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Stanya KJ et al. Direct control of hepatic glucose production by interleukin-13 in mice. J. Clin. Invest 123, 261–271 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Brestoff JR et al. Group 2 innate lymphoid cells promote beiging of white adipose tissue and limit obesity. Nature 519, 242–246 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Mahlakoiv T. et al. Stromal cells maintain immune cell homeostasis in adipose tissue via production of interleukin-33. Sci Immunol 4 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Zeng W. et al. Sympathetic neuro-adipose connections mediate leptin-driven lipolysis. Cell 163, 84–94 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Rankin LC & Artis D. Beyond Host Defense: Emerging Functions of the Immune System in Regulating Complex Tissue Physiology. Cell 173, 554–567 (2018). [DOI] [PubMed] [Google Scholar]
  • 204.Zhou L, Lin Q. & Sonnenberg GF Metabolic control of innate lymphoid cells in health and disease. Nat Metab 4, 1650–1659 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205.Wu D. et al. Eosinophils sustain adipose alternatively activated macrophages associated with glucose homeostasis. Science 332, 243–247 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Molofsky AB et al. Innate lymphoid type 2 cells sustain visceral adipose tissue eosinophils and alternatively activated macrophages. J. Exp. Med 210, 535–549 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Moyat M, Coakley G. & Harris NL The interplay of type 2 immunity, helminth infection and the microbiota in regulating metabolism. Clin Transl Immunology 8, e01089 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Brigger D. et al. Eosinophils regulate adipose tissue inflammation and sustain physical and immunological fitness in old age. Nat Metab 2, 688–702 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Rana BMJ et al. A stromal cell niche sustains ILC2-mediated type-2 conditioning in adipose tissue. J. Exp. Med 216, 1999–2009 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Huang Z. et al. The FGF21-CCL11 Axis Mediates Beiging of White Adipose Tissues by Coupling Sympathetic Nervous System to Type 2 Immunity. Cell Metab 26, 493–508 e494 (2017). [DOI] [PubMed] [Google Scholar]
  • 211.Molofsky AB et al. Interleukin-33 and Interferon-gamma Counter-Regulate Group 2 Innate Lymphoid Cell Activation during Immune Perturbation. Immunity 43, 161–174 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Buechler MB et al. Cross-tissue organization of the fibroblast lineage. Nature (2021). [DOI] [PubMed]
  • 213.Spallanzani RG et al. Distinct immunocyte-promoting and adipocyte-generating stromal components coordinate adipose tissue immune and metabolic tenors. Sci Immunol 4 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214.Kohlgruber AC et al. gammadelta T cells producing interleukin-17A regulate adipose regulatory T cell homeostasis and thermogenesis. Nat. Immunol 19, 464–474 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Vasanthakumar A. et al. Sex-specific adipose tissue imprinting of regulatory T cells. Nature 579, 581–585 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Merrick D. et al. Identification of a mesenchymal progenitor cell hierarchy in adipose tissue. Science 364 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Sbierski-Kind J, Mroz N. & Molofsky AB Perivascular stromal cells: Directors of tissue immune niches. Immunol. Rev 302, 10–31 (2021). [DOI] [PubMed] [Google Scholar]
  • 218.Silva HM et al. Vasculature-associated fat macrophages readily adapt to inflammatory and metabolic challenges. J. Exp. Med 216, 786–806 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219.Moura Silva H. et al. c-MAF-dependent perivascular macrophages regulate diet-induced metabolic syndrome. Sci Immunol 6, eabg7506 (2021). [DOI] [PubMed] [Google Scholar]
  • 220.Chakarov S. et al. Two distinct interstitial macrophage populations coexist across tissues in specific subtissular niches. Science 363 (2019). [DOI] [PubMed] [Google Scholar]
  • 221.Ding X. et al. IL-33-driven ILC2/eosinophil axis in fat is induced by sympathetic tone and suppressed by obesity. J. Endocrinol 231, 35–48 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Goldberg EL et al. IL-33 causes thermogenic failure in aging by expanding dysfunctional adipose ILC2. Cell Metab. 33, 2277–2287 e2275 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Cardoso F. et al. Neuro-mesenchymal units control ILC2 and obesity via a brain-adipose circuit. Nature 597, 410–414 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Lee MW et al. Activated type 2 innate lymphoid cells regulate beige fat biogenesis. Cell 160, 74–87 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Jimenez MT, Michieletto MF & Henao-Mejia J. A new perspective on mesenchymal-immune interactions in adipose tissue. Trends Immunol. 42, 375–388 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 226.Nawaz A. et al. CD206(+) M2-like macrophages regulate systemic glucose metabolism by inhibiting proliferation of adipocyte progenitors. Nat Commun 8, 286 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Qiu Y. et al. Eosinophils and type 2 cytokine signaling in macrophages orchestrate development of functional beige fat. Cell 157, 1292–1308 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228.Wang W. & Seale P. Control of brown and beige fat development. Nat. Rev. Mol. Cell Biol 17, 691–702 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Rao RR et al. Meteorin-like is a hormone that regulates immune-adipose interactions to increase beige fat thermogenesis. Cell 157, 1279–1291 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230.Knights AJ et al. Eosinophil function in adipose tissue is regulated by Kruppel-like factor 3 (KLF3). Nat Commun 11, 2922 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 231.Fabbiano S. et al. Caloric Restriction Leads to Browning of White Adipose Tissue through Type 2 Immune Signaling. Cell Metab. 24, 434–446 (2016). [DOI] [PubMed] [Google Scholar]
  • 232.Nguyen KD et al. Alternatively activated macrophages produce catecholamines to sustain adaptive thermogenesis. Nature 480, 104–108 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233.Hui X. et al. Adiponectin Enhances Cold-Induced Browning of Subcutaneous Adipose Tissue via Promoting M2 Macrophage Proliferation. Cell Metab 22, 279–290 (2015). [DOI] [PubMed] [Google Scholar]
  • 234.Villarroya F, Cereijo R, Villarroya J, Gavalda-Navarro A. & Giralt M. Toward an Understanding of How Immune Cells Control Brown and Beige Adipobiology. Cell Metab. 27, 954–961 (2018). [DOI] [PubMed] [Google Scholar]
  • 235.Fischer K. et al. Alternatively activated macrophages do not synthesize catecholamines or contribute to adipose tissue adaptive thermogenesis. Nat. Med 23, 623–630 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 236.Kajimura S, Spiegelman BM & Seale P. Brown and Beige Fat: Physiological Roles beyond Heat Generation. Cell Metab. 22, 546–559 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.Ahmad B. et al. Brown/Beige adipose tissues and the emerging role of their secretory factors in improving metabolic health: The batokines. Biochimie 184, 26–39 (2021). [DOI] [PubMed] [Google Scholar]
  • 238.Wang GX et al. The brown fat-enriched secreted factor Nrg4 preserves metabolic homeostasis through attenuation of hepatic lipogenesis. Nat. Med 20, 1436–1443 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 239.Knudsen NH et al. Interleukin-13 drives metabolic conditioning of muscle to endurance exercise. Science 368 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 240.Yang X. et al. Very-low-density lipoprotein receptor-enhanced lipid metabolism in pancreatic stellate cells promotes pancreatic fibrosis. Immunity 55, 1185–1199 e1188 (2022). [DOI] [PubMed] [Google Scholar]
  • 241.Saco T, Ugalde IC, Cardet JC & Casale TB Strategies for choosing a biologic for your patient with allergy or asthma. Ann. Allergy. Asthma. Immunol 127, 627–637 (2021). [DOI] [PubMed] [Google Scholar]

RESOURCES