COVID-19: Proposing a Ketone-Based Metabolic Therapy as a Treatment to Blunt the Cytokine Storm

Patrick C Bradshaw, William A Seeds, Alexandra C Miller, Vikrant R Mahajan, William M Curtis, Patrick C Bradshaw, William A Seeds, Alexandra C Miller, Vikrant R Mahajan, William M Curtis

Abstract

Human SARS-CoV-2 infection is characterized by a high mortality rate due to some patients developing a large innate immune response associated with a cytokine storm and acute respiratory distress syndrome (ARDS). This is characterized at the molecular level by decreased energy metabolism, altered redox state, oxidative damage, and cell death. Therapies that increase levels of (R)-beta-hydroxybutyrate (R-BHB), such as the ketogenic diet or consuming exogenous ketones, should restore altered energy metabolism and redox state. R-BHB activates anti-inflammatory GPR109A signaling and inhibits the NLRP3 inflammasome and histone deacetylases, while a ketogenic diet has been shown to protect mice from influenza virus infection through a protective γδ T cell response and by increasing electron transport chain gene expression to restore energy metabolism. During a virus-induced cytokine storm, metabolic flexibility is compromised due to increased levels of reactive oxygen species (ROS) and reactive nitrogen species (RNS) that damage, downregulate, or inactivate many enzymes of central metabolism including the pyruvate dehydrogenase complex (PDC). This leads to an energy and redox crisis that decreases B and T cell proliferation and results in increased cytokine production and cell death. It is hypothesized that a moderately high-fat diet together with exogenous ketone supplementation at the first signs of respiratory distress will increase mitochondrial metabolism by bypassing the block at PDC. R-BHB-mediated restoration of nucleotide coenzyme ratios and redox state should decrease ROS and RNS to blunt the innate immune response and the associated cytokine storm, allowing the proliferation of cells responsible for adaptive immunity. Limitations of the proposed therapy include the following: it is unknown if human immune and lung cell functions are enhanced by ketosis, the risk of ketoacidosis must be assessed prior to initiating treatment, and permissive dietary fat and carbohydrate levels for exogenous ketones to boost immune function are not yet established. The third limitation could be addressed by studies with influenza-infected mice. A clinical study is warranted where COVID-19 patients consume a permissive diet combined with ketone ester to raise blood ketone levels to 1 to 2 mM with measured outcomes of symptom severity, length of infection, and case fatality rate.

Conflict of interest statement

The authors declare no competing interests. Dr. William Seeds has no current role in the operation and no financial interest in drseeds.com, and his legal separation from that entity is pending.

Copyright © 2020 Patrick C. Bradshaw et al.

Figures

Figure 1
Figure 1
Mechanisms that lead to acute respiratory distress syndrome (ARDS) and mortality following SARS-CoV-2 infection are shown. The cells of the innate immune response secrete increasing amounts of cytokines. The cells that normally protect against a cytokine storm lose this ability leading to a runaway positive feedback loop of cytokine production. Abbreviations: 11β-HSD1 and 11β-HSD2: 11β-hydroxysteroid dehydrogenase types 1 and 2; ACE2: angiotensin-converting enzyme 2; AEC I and AEC II: alveolar epithelial cell types I and II; DCs: dendritic cells; FOXO1: forkhead box O1 transcription factor; FOXO3: forkhead box O3 transcription factor; HIF-1α: hypoxia-inducible factor 1 alpha; IFN: interferon; IL-6: interleukin-6; IRF3: IFN-regulatory factor 3; NAD(H): nicotinamide adenine dinucleotide; NADP(H): nicotinamide adenine dinucleotide phosphate; ONOO−: peroxynitrite; PGC1-α: PPARG coactivator 1-alpha; PDK1 and PDK4: pyruvate dehydrogenase kinases 1 and 4; RLR: retinoic acid-inducible gene I-like receptors; RNS: reactive nitrogen species; ROS: reactive oxygen species; SOD: superoxide dismutase; TGF-β: transforming growth factor-β; TNF-α: tumor necrosis factor-alpha.
Figure 2
Figure 2
Proposed mechanisms and time course of SARS-CoV-2 infection when using a ketone-based metabolic therapy. Abbreviations: HBP: hexosamine biosynthesis pathway.
Figure 3
Figure 3
The major pathways of central metabolism and reactions that alter the ratios of coenzyme couples are shown. The mitochondrial matrix and the cytoplasm have independent coenzyme couple ratios. ATP synthesized in the mitochondrial matrix is exported to the cytoplasm in exchange for ADP by the adenine nucleotide translocase present in the inner mitochondrial membrane. Acetyl-CoA synthesized in the mitochondrial matrix must also be exported to the cytoplasm to provide two carbon units for fatty acid synthesis and protein acetylation. However, acetyl-CoA cannot cross the mitochondrial inner membrane. Therefore, the transfer of acetyl units across the inner mitochondrial membrane is accomplished using the citrate-pyruvate shuttle or the citrate-malate shuttle. These shuttles alter coenzyme levels and use inner membrane carrier proteins for the transport of citrate, pyruvate, and malate. The net result of the citrate-pyruvate shuttle on coenzyme levels is the use of energy from ATP hydrolysis and NADH oxidation to reduce NADP+ to NADPH. There is also a citrate-alpha-ketoglutarate shuttle system that has the net effect of using ATP hydrolysis to increase NADPH in the cytoplasm instead of increasing NADH or NADPH in the mitochondrial matrix. Synthesizing fatty acids in the cytoplasm and reducing antioxidants require NADPH. The malate-aspartate shuttle transfers reducing equivalents from NADH between the cytoplasm and the mitochondrial matrix. The glycerol 3-phosphate shuttle transfers reducing equivalents from cytoplasmic NADH to the mitochondrial ETC. Glucose is catabolized by three pathways including the hexosamine biosynthesis pathway that synthesizes uridine diphosphate (UDP) N-acetylglucosamine, glycolysis that reduces NAD+ to NADH and synthesizes ATP from ADP and Pi, and the pentose phosphate pathway (PPP) that reduces NADP+ to NADPH and synthesizes ribose sugars for nucleotide synthesis. When the pyruvate dehydrogenase complex (PDC) is inhibited, lactate is synthesized from pyruvate to recycle NAD+ from NADH so glycolysis can continue. However, the lactate is exported from the cell and may contribute to lactic acidosis and multiorgan failure [20]. R-BHB decreases the reliance of cells on glycolysis leading to reduced cellular lactate export. Abbreviations: Glut1 and Glut3: glucose transporters 1 and 3; MCT: monocarboxylate transporter; NNT: nicotinamide nucleotide transhydrogenase; PDC: pyruvate dehydrogenase complex.
Figure 4
Figure 4
Phagocyte ROS and RNS metabolism. Most of the major forms of ROS and RNS are derived from superoxide or nitric oxide (NO).
Figure 5
Figure 5
The half-lives and diffusion limits of ROS/RNS.
Figure 6
Figure 6
Restoring PDC activity or metabolizing R-BHB and fatty acids to bypass the inhibited PDC activity that occurs following viral infection is central to beneficial energy reprogramming. Most transcription factor names are taken from the Encyclopedia of Signaling Molecules [97, 98]. Abbreviations: 1,3-BPG: 1,3-bisphosphoglyceric acid; C/EBPβ: CCAAT-enhancer-binding protein β; G3P: glyceraldehyde 3-phosphate; E2F1: E2F transcription factor 1; ERRα and ERRγ: estrogen-related receptor alpha and gamma; GR: glucocorticoid receptor; HNF4α: hepatic nuclear factor 4 alpha; ICER/CREM: inducible cAMP early repressor/cAMP-responsive element modulator; MPC: mitochondrial pyruvate carrier; PDC: pyruvate dehydrogenase complex; PGC1α: PPARG coactivator 1 alpha; PPARα and PPARγ: peroxisome proliferator-activated receptor alpha and gamma; RAR: retinoic acid receptor; RXR: retinoic x receptor; SIRT3: sirtuin 3; STAT5: signal transducer and activator of transcription 5; TR: thyroid hormone receptor.
Figure 7
Figure 7
Maintaining ATP levels during respiratory viral infection is the key to maintain proper ion distribution to avoid edema. The active and passive cotransporters and channels maintain the ion gradients between AEC and extracellular fluids. The apical ligand-gated channels function to maintain appropriate osmotic pressure to provide sufficient fluid in the airway without causing edema. AEC I express superoxide-activated ENaC Na+ channels and NOX2 that regulates them. Abbreviations: 3Na+/2K+-ATPase: 3 sodium 2 potassium ATPase; AE: anion exchange chloride bicarbonate exchanger; BCFTR: basolateral cystic fibrosis transmembrane conductance regulator-like channel; BHBDH: β-hydroxybutyrate dehydrogenase; BIRC: basolateral inward rectifying channel; BKCa: large-conductance Ca2+- and voltage-gated big K+ channel; BORC: basolateral outward rectifying channel; CACC/TMEM: calcium-activated chloride channels/transmembrane protein; CFTR: cystic fibrosis transmembrane conductance regulator; CNG: cyclic nucleotide-gated ion channel; ENaC: epithelial sodium channel; KCa3.1: calcium-activated potassium channel; Kv7.1: voltage-dependent potassium channel; NCX: sodium calcium exchanger; NHE: sodium-hydrogen exchanger; NKCC: sodium potassium chloride cotransporter; NOX2: NADPH oxidase 2; Pit1/2: sodium-dependent phosphate transporters 1 and 2; RYR: ryanodine receptor calcium-induced Ca2+ channel; SERCA: sarcoplasmic/endoplasmic reticulum Ca2+-ATPase; SK4: SK4 calcium-activated potassium channel.
Figure 8
Figure 8
Proinflammatory signaling that occurs during viral infection and mechanisms through which R-BHB inhibits this signaling. (a) The ACE system. (b) The mechanism by which R-BHB inhibits the NLRP3 inflammasome is unknown. Four possible mechanisms are shown. (c) The hexosamine biosynthesis pathway is required to initiate an effective antiviral innate immune response, but its increased activity can also stimulate a cytokine storm. Abbreviations: 6PGDL: 6-phosphonoglocono-D-lactone; 6PG: 6-phosphogluconate; ACE2: angiotensin-converting enzyme 2; ACE: angiotensin-converting enzyme; ANG (1-9): angiotensin (1-9); ANG (1-7): angiotensin (1-7), a vasodilator; ANG II: angiotensin II, a vasoconstrictor; ANG III: angiotensin III, a metabolite of ANG II; BRCC3: Lys-63-specific deubiquitinase BRCC36; CARD: caspase recruitment domain; CLIC: chloride intracellular channel protein; F1,6BP: fructose 1,6-bisphosphate; F6P: fructose 6-phosphate; FADD: fas-associated protein with death domain; GlcNAc: G,N-acetylglucosamine; G3P: glyceraldehyde 3-phosphate; G6P: glucose 6-phosphate; GlcN-6P: glucosamine-6-phosphate; GlcNAc-6P: N-acetyl glucosamine-6-phosphate; GlcNAc-1p: N-acetyl glucosamine-1-phosphate; HBP: hexosamine biosynthesis pathway; IKKε: IκB kinase ε; IL-1R: interleukin-1 receptor; IRAK-1 and IRAK-4: interleukin-1 receptor kinases 1 and 4; IRF3 and IRF5: IFN-regulatory factors 3 and 5; JNK1: c-Jun N-terminal protein kinase 1; K63-Ub: K63-linked polyubiquitin binding; MasR: Mas receptor; MAVS: mitochondrial antiviral signaling protein; MDA5: melanoma differentiation-associated gene 5; MyD88: myeloid differentiation primary response 88; NF-κB: nuclear factor kappa-light-chain-enhancer of activated B cells; NLRP3: NOD-, LRR-, and pyrin domain-containing protein 3; NOD1 and NOD2: nucleotide-binding oligomerization domain-containing proteins 1 and 2; OGT: O-linked N-acetylglucosamine (GlcNAc) transferase; P2X7: P2X purinoceptor 7; PDPK: phosphoinositide-dependent kinase-1; PPP: pentose phosphate pathway; PTMs: posttranslational modifications; R5P: ribose 5-phosphate; RIG-1: retinoic acid-inducible gene I; S7P: sedoheptulose 7-phosphate; SERCA: sarcoplasmic/endoplasmic reticulum Ca2+-ATPase; STAT3: signal transducer and activator of transcription 3; TBK-1: TANK-binding kinase 1; TLR: toll-like receptor; TNFR: tumor necrosis factor receptor; TRAF6: tumor necrosis factor receptor- (TNFR-) associated factor 6; TRIF: TIR domain-containing adapter-inducing interferon β; TRX: thioredoxin; TXNIP: thioredoxin-interacting protein; UDP-GlcNAc: uridine diphosphate N-acetylglucosamine.
Figure 9
Figure 9
The effects of increased R-BHB levels on the activity of the transcription factors FOXO3a, FOXO1, HIF1-α, Nrf2, and PGC-1α. (a) DNA wrapped around histones with deacetylated lysines blocks the access of transcription factors, so HDAC function represses transcription. (b) R-BHB, by inhibiting class I HDACs, allows increased histone acetylation, relaxed chromatin, and FOXO3a expression. (c) Chromatin is also relaxed by β-hydroxybutyrylation at promoters such as that of FOXO1 to increase transcription. (d) HIF-1α is activated by RNS. Peroxynitrite S-nitrosylates HIF-1α to inhibit its von Hippel-Landau factor-induced ubiquitination and proteasomal degradation. This stabilizes HIF-1α leading to increased PDK1 expression. This stabilization is likely inhibited by R-BHB, which increases NADPH to decrease ROS/RNS levels. (e) Nrf2 is activated by oxidized DJ-1, and this is regulated by the redox potential of NADP+/NADPH that controls the cellular antioxidant potential. The DJ-1 chaperone protein, which is activated by moderate levels of hydrogen peroxide, stimulates the release of Nrf2 from KEAP1 allowing Nrf2 to enter the nucleus and induce antioxidant response element gene expression. (f) PGC-1α is the master regulator of mitochondrial biogenesis. It is a coactivator of ERR-α, FOXO1, FOXO3a, PPARs, and nuclear respiratory factor 1 (NRF1). Abbreviations: Ac: acetate; ARE: antioxidant response element; CA9: carbonic anhydrase 9; CCR7: C-C chemokine receptor type 7; ERR-α: estrogen-related receptor alpha; ETC: electron transport chain; K: lysine; HDAC: histone deacetylase; FIH-1: factor inhibiting HIF-1; FOXO1: forkhead box O1; FOXO3: forkhead box O3; FOXP3: forkhead box P3; G6PC: glucose-6-phosphatase; G6PDH: glucose 6-phosphate dehydrogenase; γGCLM: glutamate cysteine ligase modifier subunit; Glut1: glucose transporter 1; GSRX: glutathione reductase; HIF-1α and HIF-1β: hypoxia-inducible factor 1 alpha and beta; HRE: hypoxia response element; IFNG: interferon gamma gene; IL7R: interleukin-7 receptor; iNOS: inducible nitric oxide synthase; IRF7: interferon regulatory factor 7; KEAP1: Kelch-like ECH-associated protein 1; MAF: musculoaponeurotic fibrosarcoma; Nrf1: nuclear respiratory factor 1; Nrf2: nuclear factor erythroid 2-related factor 2 (NFE2L2); RAG1 and RAG2: recombination activating genes 1 and 2; ONOO−: peroxynitrite; ONOOH: peroxynitrous acid; P300/CBP: binding protein 300/CREB-binding protein; PDK1 and PDK4: pyruvate dehydrogenase kinases 1 and 4; PGC1-α: PPARG coactivator 1 alpha; PHD2: prolyl-hydroxylase domain protein 2; pVHL: von Hippel-Landau tumor suppressor protein; SCOT/OXCT1: succinyl-CoA-3-oxaloacid CoA transferase also known as 3-oxoacid CoA-transferase 1; SELL: selectin L; SOD2: superoxide dismutase 2; Ub: ubiquitin; VEGF: vascular endothelial growth factor.

References

    1. Islam M. R., Fischer A. A Transcriptome Analysis Identifies Potential Preventive and Therapeutic Approaches Towards COVID-19. Preprints; 2020.
    1. Kamepalli R., Kamepalli B. How immune T-cell augmentation can help prevent COVID-19: a possible nutritional solution using ketogenic lifestyle. The University of Louisville Journal of Respiratory Infections. 2020;4:7.
    1. Stubbs B., Koutnik A. P., Goldberg E. L., et al. Investigating ketone bodies as immunometabolic countermeasures against respiratory viral infections. Med. 2020 doi: 10.1016/j.medj.2020.06.008.
    1. Hamming I., Timens W., Bulthuis M. L. C., Lely A. T., Navis G. J., van Goor H. Tissue distribution of ACE2 protein, the functional receptor for SARS coronavirus. A first step in understanding SARS pathogenesis. The Journal of Pathology. 2004;203(2):631–637. doi: 10.1002/path.1570.
    1. Fehrenbach H. Alveolar epithelial type II cell: defender of the alveolus revisited. Respiratory Research. 2001;2(1):33–46. doi: 10.1186/rr36.
    1. Su W. Y., Folz R., Chen J. S., Crapo J. D., Chang L. Y. Extracellular superoxide dismutase mRNA expressions in the human lung by in situ hybridization. American Journal of Respiratory Cell and Molecular Biology. 1997;16(2):162–170. doi: 10.1165/ajrcmb.16.2.9032123.
    1. Makris S., Paulsen M., Johansson C. Type I interferons as regulators of lung inflammation. Frontiers in Immunology. 2017;8 doi: 10.3389/fimmu.2017.00259.
    1. Cui H., Xie N., Banerjee S., Ge J., Guo S., Liu G. Impairment of fatty acid oxidation in alveolar epithelial cells mediates acute lung injury. American Journal of Respiratory Cell and Molecular Biology. 2019;60(2):167–178. doi: 10.1165/rcmb.2018-0152OC.
    1. Grandvaux N., Mariani M., Fink K. Lung epithelial NOX/DUOX and respiratory virus infections. Clinical Science. 2015;128(6):337–347. doi: 10.1042/CS20140321.
    1. McAlpine W., Sun L., Wang K. W., et al. Excessive endosomal TLR signaling causes inflammatory disease in mice with defective SMCR8-WDR41-C9ORF72 complex function. Proceedings of the National Academy of Sciences. 2018;115(49):E11523–E11531. doi: 10.1073/pnas.1814753115.
    1. Vaninov N. In the eye of the COVID-19 cytokine storm. Nature Reviews Immunology. 2020;20(5) doi: 10.1038/s41577-020-0305-6.
    1. Huang C., Wang Y., Li X., et al. Clinical features of patients infected with 2019 novel coronavirus in Wuhan, China. The Lancet. 2020;395(10223):497–506. doi: 10.1016/S0140-6736(20)30183-5.
    1. Mehta P., McAuley D., Brown M., et al. COVID-19: consider cytokine storm syndromes and immunosuppression. The Lancet. 2020;395(10229):1033–1034. doi: 10.1016/S0140-6736(20)30628-0.
    1. Veech R. L., Todd King M., Pawlosky R., Kashiwaya Y., Bradshaw P. C., Curtis W. The "great" controlling nucleotide coenzymes. IUBMB Life. 2018;71(5):565–579. doi: 10.1002/iub.1997.
    1. Ghoneum A., Abdulfattah A. Y., Warren B. O., Shu J., Said N. Redox homeostasis and metabolism in cancer: a complex mechanism and potential targeted therapeutics. International Journal of Molecular Sciences. 2020;21(9):p. 3100. doi: 10.3390/ijms21093100.
    1. Pawlosky R. J., Kemper M. F., Kashiwaya Y., King M. T., Mattson M. P., Veech R. L. Effects of a dietary ketone ester on hippocampal glycolytic and tricarboxylic acid cycle intermediates and amino acids in a 3xTgAD mouse model of Alzheimer's disease. Journal of Neurochemistry. 2017;141(2):195–207. doi: 10.1111/jnc.13958.
    1. Yamane K., Indalao I. L., Chida J., Yamamoto Y., Hanawa M., Kido H. Diisopropylamine dichloroacetate, a novel pyruvate dehydrogenase kinase 4 inhibitor, as a potential therapeutic agent for metabolic disorders and multiorgan failure in severe influenza. PLoS One. 2014;9(5, article e98032) doi: 10.1371/journal.pone.0098032.
    1. Sanchez E. L., Lagunoff M. Viral activation of cellular metabolism. Virology. 2015;479-480:609–618. doi: 10.1016/j.virol.2015.02.038.
    1. Trauner D. A., Horvath E., Davis L. E. Inhibition of fatty acid beta oxidation by influenza B virus and salicylic acid in mice: implications for Reye's syndrome. Neurology. 1988;38(2):239–241. doi: 10.1212/WNL.38.2.239.
    1. Effros R. M., Lipchik R. J. Why does lactic acidosis occur in acute lung injury? Chest. 1997;111(5):1157–1158. doi: 10.1378/chest.111.5.1157.
    1. Córdova Martínez A., Martorell Pons M., Sureda Gomila A., Tur Marí J. A., Pons Biescas A. Changes in circulating cytokines and markers of muscle damage in elite cyclists during a multi-stage competition. Clinical Physiology and Functional Imaging. 2015;35(5):351–358. doi: 10.1111/cpf.12170.
    1. Poffé C., Ramaekers M., Thienen R., Hespel P. Ketone ester supplementation blunts overreaching symptoms during endurance training overload. The Journal of Physiology. 2019;597(12):3009–3027. doi: 10.1113/JP277831.
    1. Lin Y., Kim J., Metter E. J., et al. Changes in blood lymphocyte numbers with age in vivo and their association with the levels of cytokines/cytokine receptors. Immunity & Ageing. 2016;13(1) doi: 10.1186/s12979-016-0079-7.
    1. Czesnikiewicz-Guzik M., Lee W.-W., Cui D., et al. T cell subset-specific susceptibility to aging. Clinical Immunology. 2008;127(1):107–118. doi: 10.1016/j.clim.2007.12.002.
    1. Chousterman B. G., Swirski F. K., Weber G. F. Cytokine storm and sepsis disease pathogenesis. Seminars in Immunopathology. 2017;39(5):517–528. doi: 10.1007/s00281-017-0639-8.
    1. Tisoncik J. R., Korth M. J., Simmons C. P., Farrar J., Martin T. R., Katze M. G. Into the eye of the cytokine storm. Microbiology and Molecular Biology Reviews : MMBR. 2012;76(1):16–32. doi: 10.1128/MMBR.05015-11.
    1. Jose R. J., Manuel A. COVID-19 cytokine storm: the interplay between inflammation and coagulation. The Lancet Respiratory Medicine. 2020;8(6):e46–e47. doi: 10.1016/S2213-2600(20)30216-2.
    1. Ye Q., Wang B., Mao J. The pathogenesis and treatment of the ‘cytokine storm’ in COVID-19. The Journal of Infection. 2020;80(6):607–613. doi: 10.1016/j.jinf.2020.03.037.
    1. Miller A. C., Rivas R., McMahon R., et al. Radiation protection and mitigation potential of phenylbutyrate: delivered via oral administration. International Journal of Radiation Biology. 2017;93(9):907–919. doi: 10.1080/09553002.2017.1350301.
    1. Li X., Cui W., Hull L., Smith J. T., Kiang J. G., Xiao M. Effects of low-to-moderate doses of gamma radiation on mouse hematopoietic system. Radiation Research. 2018;190(6):612–622. doi: 10.1667/RR15087.1.
    1. Zhang M., Yin L., Zhang K., et al. Response patterns of cytokines/chemokines in two murine strains after irradiation. Cytokine. 2012;58(2):169–177. doi: 10.1016/j.cyto.2011.12.023.
    1. Kiang J. G., Smith J. T., Hegge S. R., Ossetrova N. I. Circulating cytokine/chemokine concentrations respond to ionizing radiation doses but not radiation dose rates: granulocyte-colony stimulating factor and interleukin-18. Radiation Research. 2018;189(6):634–643. doi: 10.1667/RR14966.1.
    1. Aneva N., Zaharieva E., Katsarska O., et al. Inflammatory profile dysregulation in nuclear workers occupationally exposed to low-dose gamma radiation. Journal of Radiation Research. 2019;60(6):768–779. doi: 10.1093/jrr/rrz059.
    1. Di Maggio F., Minafra L., Forte G., et al. Portrait of inflammatory response to ionizing radiation treatment. Journal of Inflammation. 2015;12(1):p. 14. doi: 10.1186/s12950-015-0058-3.
    1. Singh V. K., Seed T. M. A review of radiation countermeasures focusing on injury-specific medicinals and regulatory approval status: part I. Radiation sub-syndromes, animal models and FDA-approved countermeasures. International Journal of Radiation Biology. 2017;93(9):851–869. doi: 10.1080/09553002.2017.1332438.
    1. Horta Z. P., Case C. M., DiCarlo A. L. Use of growth factors and cytokines to treat injuries resulting from a radiation public health emergency. Radiation Research. 2019;192(1):92–97. doi: 10.1667/RR15383.1.
    1. Curtis W., Kemper M. L., Miller A. L., Pawlosky R., King M. T., Veech R. L. Mitigation of damage from reactive oxygen species and ionizing radiation by ketone body esters. In: Masino S. A., editor. Ketogenic Diet and Metabolic Therapy. Oxford: Oxford University Press; 2017. pp. 254–270.
    1. Schaue D., Kachikwu E. L., McBride W. H. Cytokines in radiobiological responses: a review. Radiation Research. 2012;178(6):505–523. doi: 10.1667/RR3031.1.
    1. Zhang C., Wu Z., Li J. W., Zhao H., Wang G. Q. Cytokine release syndrome in severe COVID-19: interleukin-6 receptor antagonist tocilizumab may be the key to reduce mortality. International Journal of Antimicrobial Agents. 2020;55(5):p. 105954. doi: 10.1016/j.ijantimicag.2020.105954.
    1. Cavalli G., de Luca G., Campochiaro C., et al. Interleukin-1 blockade with high-dose anakinra in patients with COVID-19, acute respiratory distress syndrome, and hyperinflammation: a retrospective cohort study. The Lancet Rheumatology. 2020;2(6):e325–e331. doi: 10.1016/s2665-9913(20)30127-2.
    1. Mantlo E., Bukreyeva N., Maruyama J., Paessler S., Huang C. Antiviral activities of type I interferons to SARS-CoV-2 infection. Antiviral Research. 2020;179, article 104811 doi: 10.1016/j.antiviral.2020.104811.
    1. Jamilloux Y., Henry T., Belot A., et al. Should we stimulate or suppress immune responses in COVID-19? Cytokine and anti-cytokine interventions. Autoimmunity Reviews. 2020;19(7, article 102567) doi: 10.1016/j.autrev.2020.102567.
    1. Djafarzadeh S., Vuda M., Takala J., Ochs M., Jakob S. M. Toll-like receptor-3-induced mitochondrial dysfunction in cultured human hepatocytes. Mitochondrion. 2011;11(1):83–88. doi: 10.1016/j.mito.2010.07.010.
    1. Garaude J., Acín-Pérez R., Martínez-Cano S., et al. Mitochondrial respiratory-chain adaptations in macrophages contribute to antibacterial host defense. Nature Immunology. 2016;17(9):1037–1045. doi: 10.1038/ni.3509.
    1. Yang C. S., Kim J. J., Lee S. J., et al. TLR3-triggered reactive oxygen species contribute to inflammatory responses by activating signal transducer and activator of transcription-1. The Journal of Immunology. 2013;190(12):6368–6377. doi: 10.4049/jimmunol.1202574.
    1. Soucy-Faulkner A., Mukawera E., Fink K., et al. Requirement of NOX2 and reactive oxygen species for efficient RIG-I-mediated antiviral response through regulation of MAVS expression. PLoS Pathogens. 2010;6(6) doi: 10.1371/journal.ppat.1000930.
    1. To E. E., Broughton B. R. S., Hendricks K. S., Vlahos R., Selemidis S. Influenza A virus and TLR7 activation potentiate NOX2 oxidase-dependent ROS production in macrophages. Free Radical Research. 2014;48(8):940–947. doi: 10.3109/10715762.2014.927579.
    1. Khomich O., Kochetkov S., Bartosch B., Ivanov A. Redox biology of respiratory viral infections. Viruses. 2018;10(8):p. 392. doi: 10.3390/v10080392.
    1. Naik E., Dixit V. M. Mitochondrial reactive oxygen species drive proinflammatory cytokine production. The Journal of Experimental Medicine. 2011;208(3):417–420. doi: 10.1084/jem.20110367.
    1. Czerkies M., Korwek Z., Prus W., et al. Cell fate in antiviral response arises in the crosstalk of IRF, NF-κB and JAK/STAT pathways. Nature Communications. 2018;9(1) doi: 10.1038/s41467-017-02640-8.
    1. Salminen A., Ojala J., Kaarniranta K., Kauppinen A. Mitochondrial dysfunction and oxidative stress activate inflammasomes: impact on the aging process and age-related diseases. Cellular and Molecular Life Sciences. 2012;69(18):2999–3013. doi: 10.1007/s00018-012-0962-0.
    1. Kelley N., Jeltema D., Duan Y., He Y. The NLRP3 inflammasome: an overview of mechanisms of activation and regulation. International Journal of Molecular Sciences. 2019;20(13):p. 3328. doi: 10.3390/ijms20133328.
    1. Farioli-Vecchioli S., Nardacci R., Falciatori I., Stefanini S. Catalase immunocytochemistry allows automatic detection of lung type II alveolar cells. Histochemistry and Cell Biology. 2001;115(4):333–339. doi: 10.1007/s004180100259.
    1. Folz R. J., Guan J., Seldin M. F., Oury T. D., Enghild J. J., Crapo J. D. Mouse extracellular superoxide dismutase: primary structure, tissue-specific gene expression, chromosomal localization, and lung in situ hybridization. American Journal of Respiratory Cell and Molecular Biology. 1997;17(4):393–403. doi: 10.1165/ajrcmb.17.4.2826.
    1. Ateş N. A., Yildirim Ö., Tamer L., et al. Plasma catalase activity and malondialdehyde level in patients with cataract. Eye. 2004;18(8):785–788. doi: 10.1038/sj.eye.6700718.
    1. Komuro I., Keicho N., Iwamoto A., Akagawa K. S. Human alveolar macrophages and granulocyte-macrophage colony-stimulating factor-induced monocyte-derived macrophages are resistant to H2O2 via their high basal and inducible levels of catalase activity. Journal of Biological Chemistry. 2001;276(26):24360–24364. doi: 10.1074/jbc.M102081200.
    1. He Z., Sun X., Mei G., Yu S., Li N. Nonclassical secretion of human catalase on the surface of CHO cells is more efficient than classical secretion. Cell Biology International. 2008;32(4):367–373. doi: 10.1016/j.cellbi.2007.12.003.
    1. Nikpouraghdam M., Jalali Farahani A., Alishiri G. H., et al. Epidemiological characteristics of coronavirus disease 2019 (COVID-19) patients in IRAN: a single center study. Journal of Clinical Virology. 2020;127, article 104378 doi: 10.1016/j.jcv.2020.104378.
    1. Abouhashem A. S., Singh K., Azzazy H. M. E., Sen C. K. Is low alveolar type II cell SOD3in the lungs of elderly linked to the observed severity of COVID-19? Antioxidants & Redox Signaling. 2020;33(2):59–65. doi: 10.1089/ars.2020.8111.
    1. Goodson P., Kumar A., Jain L., et al. Nadph oxidase regulates alveolar epithelial sodium channel activity and lung fluid balance in vivo via O2− signaling. American Journal of Physiology-Lung Cellular and Molecular Physiology. 2012;302(4):L410–L419. doi: 10.1152/ajplung.00260.2011.
    1. Takemura Y., Goodson P., Bao H. F., Jain L., Helms M. N. Rac1-mediated NADPH oxidase release of O2- regulates epithelial sodium channel activity in the alveolar epithelium. American Journal of Physiology-Lung Cellular and Molecular Physiology. 2010;298(4):L509–L520. doi: 10.1152/ajplung.00230.2009.
    1. Ganguly K., Depner M., Fattman C., et al. Superoxide dismutase 3, extracellular (SOD3) variants and lung function. Physiological Genomics. 2009;37(3):260–267. doi: 10.1152/physiolgenomics.90363.2008.
    1. Ansar M., Ivanciuc T., Garofalo R. P., Casola A. Increased lung catalase activity confers protection against experimental RSV infection. Scientific Reports. 2020;10(1):p. 3653. doi: 10.1038/s41598-020-60443-2.
    1. Kakimoto Y., Seto Y., Ochiai E., et al. Cytokine elevation in sudden death with respiratory syncytial virus: a case report of 2 children. Pediatrics. 2016;138(6, article e20161293) doi: 10.1542/peds.2016-1293.
    1. Ghosh S., Janocha A. J., Aronica M. A., et al. Nitrotyrosine proteome survey in asthma identifies oxidative mechanism of catalase inactivation. Journal of Immunology. 2006;176(9):5587–5597. doi: 10.4049/jimmunol.176.9.5587.
    1. Guo X. J., Thomas P. G. New fronts emerge in the influenza cytokine storm. Seminars in Immunopathology. 2017;39(5):541–550. doi: 10.1007/s00281-017-0636-y.
    1. To E. E., Erlich J. R., Liong F., et al. Mitochondrial reactive oxygen species contribute to pathological inflammation during influenza A virus infection in mice. Antioxidants & Redox Signaling. 2020;32(13):929–942. doi: 10.1089/ars.2019.7727.
    1. To E. E., Luong R., Diao J., et al. Novel endosomal NOX2 oxidase inhibitor ameliorates pandemic influenza A virus-induced lung inflammation in mice. Respirology. 2019;24(10):1011–1017. doi: 10.1111/resp.13524.
    1. Vlahos R., Stambas J., Bozinovski S., Broughton B. R. S., Drummond G. R., Selemidis S. Inhibition of Nox2 oxidase activity ameliorates influenza A virus-induced lung inflammation. PLoS Pathogens. 2011;7(2, article e1001271) doi: 10.1371/journal.ppat.1001271.
    1. Cronin S. J. F., Woolf C. J., Weiss G., Penninger J. M. The role of iron regulation in immunometabolism and immune-related disease. Frontiers in Molecular Biosciences. 2019;6 doi: 10.3389/fmolb.2019.00116.
    1. Pacher P., Beckman J. S., Liaudet L. Nitric oxide and peroxynitrite in health and disease. Physiological Reviews. 2007;87(1):315–424. doi: 10.1152/physrev.00029.2006.
    1. Ascenzi P., Bocedi A., Visca P., Minetti M., Clementi E. Does CO2 modulate peroxynitrite specificity? IUBMB Life. 2006;58(10):611–613. doi: 10.1080/15216540600746344.
    1. Budzińska M., Gałgańska H., Karachitos A., Wojtkowska M., Kmita H. The TOM complex is involved in the release of superoxide anion from mitochondria. Journal of Bioenergetics and Biomembranes. 2009;41(4):361–367. doi: 10.1007/s10863-009-9231-9.
    1. Lustgarten M. S., Bhattacharya A., Muller F. L., et al. Complex I generated, mitochondrial matrix-directed superoxide is released from the mitochondria through voltage dependent anion channels. Biochemical and Biophysical Research Communications. 2012;422(3):515–521. doi: 10.1016/j.bbrc.2012.05.055.
    1. Guzzi P. H., Mercatelli D., Ceraolo C., Giorgi F. M. Master regulator analysis of the SARS-CoV-2/human interactome. Journal of Clinical Medicine. 2020;9(4):p. 982. doi: 10.3390/jcm9040982.
    1. Pantel A., Teixeira A., Haddad E., Wood E. G., Steinman R. M., Longhi M. P. Direct type I IFN but not MDA5/TLR3 activation of dendritic cells is required for maturation and metabolic shift to glycolysis after poly IC stimulation. PLoS Biology. 2014;12(1, article e1001759) doi: 10.1371/journal.pbio.1001759.
    1. Villena J. A. New insights into PGC-1 coactivators: redefining their role in the regulation of mitochondrial function and beyond. The FEBS Journal. 2015;282(4):647–672. doi: 10.1111/febs.13175.
    1. Olmos Y., Valle I., Borniquel S., et al. Mutual dependence of Foxo3a and PGC-1α in the induction of oxidative stress genes. Journal of Biological Chemistry. 2009;284(21):14476–14484. doi: 10.1074/jbc.M807397200.
    1. Shao H., Lan D., Duan Z., et al. Upregulation of mitochondrial gene expression in PBMC from convalescent SARS patients. Journal of Clinical Immunology. 2006;26(6):546–554. doi: 10.1007/s10875-006-9046-y.
    1. Singh K. K., Chaubey G., Chen J. Y., Suravajhala P. Decoding SARS-CoV-2 hijacking of host mitochondria in COVID-19 pathogenesis. American Journal of Physiology-Cell Physiology. 2020;319(2):C258–C267. doi: 10.1152/ajpcell.00224.2020.
    1. Wu K., Zou J., Chang H. Y. RNA-GPS predicts SARS-CoV-2 RNA localization to host mitochondria and nucleolus. bioRxiv; 2020.
    1. Channappanavar R., Fehr A. R., Vijay R., et al. Dysregulated type I interferon and inflammatory monocyte-macrophage responses cause lethal pneumonia in SARS-CoV-infected mice. Cell Host & Microbe. 2016;19(2):181–193. doi: 10.1016/j.chom.2016.01.007.
    1. Kindler E., Thiel V., Weber F. Interaction of SARS and MERS coronaviruses with the antiviral interferon response. Advances in Virus Research. 2016;96:219–243. doi: 10.1016/bs.aivir.2016.08.006.
    1. Furuyama T., Kitayama K., Yamashita H., Mori N. Forkhead transcription factor FOXO1 (FKHR)-dependent induction of PDK4 gene expression in skeletal muscle during energy deprivation. The Biochemical Journal. 2003;375(2):365–371. doi: 10.1042/bj20030022.
    1. Wu P., Peters J. M., Harris R. A. Adaptive increase in pyruvate dehydrogenase kinase 4 during starvation is mediated by peroxisome proliferator-activated receptor α. Biochemical and Biophysical Research Communications. 2001;287(2):391–396. doi: 10.1006/bbrc.2001.5608.
    1. Kwon H. S., Huang B., Unterman T. G., Harris R. A. Protein kinase B-α inhibits human pyruvate dehydrogenase kinase-4 gene induction by dexamethasone through inactivation of FOXO transcription factors. Diabetes. 2004;53(4):899–910. doi: 10.2337/diabetes.53.4.899.
    1. Mariappan N., Elks C. M., Fink B., Francis J. TNF-induced mitochondrial damage: a link between mitochondrial complex I activity and left ventricular dysfunction. Free Radical Biology & Medicine. 2009;46(4):462–470. doi: 10.1016/j.freeradbiomed.2008.10.049.
    1. Gerriets V. A., Kishton R. J., Nichols A. G., et al. Metabolic programming and PDHK1 control CD4+ T cell subsets and inflammation. The Journal of Clinical Investigation. 2015;125(1):194–207. doi: 10.1172/JCI76012.
    1. Wang A., Huen S. C., Luan H. H., et al. Opposing effects of fasting metabolism on tissue tolerance in bacterial and viral inflammation. Cell. 2016;166(6):1512–1525.e12. doi: 10.1016/j.cell.2016.07.026.
    1. Moseley C. E., Webster R. G., Aldridge J. R. Peroxisome proliferator-activated receptor and AMP-activated protein kinase agonists protect against lethal influenza virus challenge in mice. Influenza and Other Respiratory Viruses. 2010;4(5):307–311. doi: 10.1111/j.1750-2659.2010.00155.x.
    1. Way J. M., Harrington W. W., Brown K. K., et al. Comprehensive messenger ribonucleic acid profiling reveals that peroxisome proliferator-activated receptor γ activation has coordinate effects on gene expression in multiple insulin-sensitive tissues. Endocrinology. 2001;142(3):1269–1277. doi: 10.1210/endo.142.3.8037.
    1. Klotz L. O., Sánchez-Ramos C., Prieto-Arroyo I., Urbánek P., Steinbrenner H., Monsalve M. Redox regulation of FoxO transcription factors. Redox Biology. 2015;6:51–72. doi: 10.1016/j.redox.2015.06.019.
    1. Ciavarella C., Motta I., Valente S., Pasquinelli G. Pharmacological (or synthetic) and nutritional agonists of PPAR-γ as candidates for cytokine storm modulation in COVID-19 disease. Molecules. 2020;25(9):p. 2076. doi: 10.3390/molecules25092076.
    1. Srivastava S., Kashiwaya Y., King M. T., et al. Mitochondrial biogenesis and increased uncoupling protein 1 in brown adipose tissue of mice fed a ketone ester diet. The FASEB Journal. 2012;26(6):2351–2362. doi: 10.1096/fj.11-200410.
    1. Hasan-Olive M. M., Lauritzen K. H., Ali M., Rasmussen L. J., Storm-Mathisen J., Bergersen L. H. A ketogenic diet improves mitochondrial biogenesis and bioenergetics via the PGC1α-SIRT3-UCP2 axis. Neurochemical Research. 2019;44(1):22–37. doi: 10.1007/s11064-018-2588-6.
    1. Parker B., Walton C., Carr S., et al. β-Hydroxybutyrate elicits favorable mitochondrial changes in skeletal muscle. International Journal of Molecular Sciences. 2018;19(8):p. 2247. doi: 10.3390/ijms19082247.
    1. Harris R. A., Kuntz M. J. PDK. In: Choi S., editor. Encyclopedia of Signaling Molecules. Springer International Publishing; 2018.
    1. Kono M., Yoshida N., Maeda K., et al. Pyruvate dehydrogenase phosphatase catalytic subunit 2 limits Th17 differentiation. Proceedings of the National Academy of Sciences. 2018;115(37):9288–9293. doi: 10.1073/pnas.1805717115.
    1. Veech R. L., King M. T., Pawlosky R., Bradshaw P. C., Curtis W. Relationship between inorganic ion distribution, resting membrane potential, and the ΔG′ of ATP hydrolysis: a new paradigm. The FASEB Journal. 2019;33(12):13126–13130. doi: 10.1096/fj.201901942R.
    1. Lai A. L., Millet J. K., Daniel S., Freed J. H., Whittaker G. R. The SARS-CoV fusion peptide forms an extended bipartite fusion platform that perturbs membrane order in a calcium-dependent manner. Journal of Molecular Biology. 2017;429(24):3875–3892. doi: 10.1016/j.jmb.2017.10.017.
    1. Liu M., Yang Y., Gu C., et al. Spike protein of SARS-CoV stimulates cyclooxygenase-2 expression via both calcium-dependent and calcium-independent protein kinase C pathways. The FASEB Journal. 2007;21(7):1586–1596. doi: 10.1096/fj.06-6589com.
    1. Nieto-Torres J. L., Verdiá-Báguena C., Jimenez-Guardeño J. M., et al. Severe acute respiratory syndrome coronavirus E protein transports calcium ions and activates the NLRP3 inflammasome. Virology. 2015;485:330–339. doi: 10.1016/j.virol.2015.08.010.
    1. Williamson C. D., DeBiasi R. L., Colberg-Poley A. M. Viral product trafficking to mitochondria, mechanisms and roles in pathogenesis. Infectious Disorders Drug Targets. 2012;12(1):18–37. doi: 10.2174/187152612798994948.
    1. Fink K., Martin L., Mukawera E., et al. IFNβ/TNFα synergism induces a non-canonical STAT2/IRF9-dependent pathway triggering a novel DUOX2 NADPH oxidase-mediated airway antiviral response. Cell Research. 2013;23(5):673–690. doi: 10.1038/cr.2013.47.
    1. Strengert M., Jennings R., Davanture S., Hayes P., Gabriel G., Knaus U. G. Mucosal reactive oxygen species are required for antiviral response: role of Duox in influenza A virus infection. Antioxidants & Redox Signaling. 2014;20(17):2695–2709. doi: 10.1089/ars.2013.5353.
    1. Huang K. P. The mechanism of protein kinase C activation. Trends in Neurosciences. 1989;12(11):425–432. doi: 10.1016/0166-2236(89)90091-X.
    1. Webb B. L. J., Hirst S. J., Giembycz M. A. Protein kinase C isoenzymes: a review of their structure, regulation and role in regulating airways smooth muscle tone and mitogenesis. British Journal of Pharmacology. 2000;130(7):1433–1452. doi: 10.1038/sj.bjp.0703452.
    1. Brueggemann L. I., Kakad P. P., Love R. B., et al. Kv7 potassium channels in airway smooth muscle cells: signal transduction intermediates and pharmacological targets for bronchodilator therapy. American Journal of Physiology-Lung Cellular and Molecular Physiology. 2012;302(1):L120–L132. doi: 10.1152/ajplung.00194.2011.
    1. Haick J. M., Brueggemann L. I., Cribbs L. L., Denning M. F., Schwartz J., Byron K. L. PKC-dependent regulation of Kv7.5 channels by the bronchoconstrictor histamine in human airway smooth muscle cells. American Journal of Physiology-Lung Cellular and Molecular Physiology. 2017;312(6):L822–L834. doi: 10.1152/ajplung.00567.2016.
    1. Abd el Sabour Faramawy M., Abd Allah A., el Batrawy S., Amer H. Impact of high fat low carbohydrate enteral feeding on weaning from mechanical ventilation. Egyptian Journal of Chest Diseases and Tuberculosis. 2014;63(4):931–938. doi: 10.1016/j.ejcdt.2014.07.004.
    1. Wilson M. R., Petrie J. E., Shaw M. W., et al. High-fat feeding protects mice from ventilator-induced lung injury, via neutrophil-independent mechanisms. Critical Care Medicine. 2017;45(8):e831–e839. doi: 10.1097/CCM.0000000000002403.
    1. Pontes-Arruda A., Aragão A. M. A., Albuquerque J. D. Effects of enteral feeding with eicosapentaenoic acid, γ-linolenic acid, and antioxidants in mechanically ventilated patients with severe sepsis and septic shock. Critical Care Medicine. 2006;34(9):2325–2333. doi: 10.1097/01.CCM.0000234033.65657.B6.
    1. Goldberg E. L., Molony R. D., Kudo E., et al. Ketogenic diet activates protective γδ T cell responses against influenza virus infection. Science Immunology. 2019;4(41, article eaav2026) doi: 10.1126/sciimmunol.aav2026.
    1. Goldberg E. L., Shchukina I., Asher J. L., Sidorov S., Artyomov M. N., Dixit V. D. Ketogenesis activates metabolically protective γδ T cells in visceral adipose tissue. Nature Metabolism. 2020;2(1):50–61. doi: 10.1038/s42255-019-0160-6.
    1. Goldberg E. L., Iwasaki A., Dixit V. D. Authors’ response to reply: ketogenic diet activates protective γδ T cell responses against influenza virus infection. NutriXiv. 2020 doi: 10.31232/.
    1. Bass K. N., Weiss E., Klatt K. C. Reply: ketogenic diet activates protective Γδ T cell responses against influenza virus infection. NutriXiv. 2020 doi: 10.31232/.
    1. Kennedy A. R., Pissios P., Otu H., et al. A high-fat, ketogenic diet induces a unique metabolic state in mice. American Journal of Physiology. Endocrinology and Metabolism. 2007;292(6):E1724–E1739. doi: 10.1152/ajpendo.00717.2006.
    1. Taggart A. K. P., Kero J., Gan X., et al. (D)-β-Hydroxybutyrate inhibits adipocyte lipolysis via the nicotinic acid receptor PUMA-G. The Journal of Biological Chemistry. 2005;280(29):26649–26652. doi: 10.1074/jbc.C500213200.
    1. Pailla K., El-Mir M.-Y., Cynober l., Blonde-Cynober F. Cytokine-mediated inhibition of ketogenesis is unrelated to nitric oxide or protein synthesis. Clinical Nutrition. 2001;20(4):313–317. doi: 10.1054/clnu.2001.0421.
    1. Bostock E. C. S., Kirkby K. C., Taylor B. V., Hawrelak J. A. Consumer reports of "keto flu" associated with the ketogenic diet. Frontiers in Nutrition. 2020;7:p. 20. doi: 10.3389/fnut.2020.00020.
    1. Harvey C. J. d. C., Schofield G. M., Williden M., McQuillan J. A. The effect of medium chain triglycerides on time to nutritional ketosis and symptoms of keto-induction in healthy adults: a randomised controlled clinical trial. Journal of Nutrition and Metabolism. 2018;2018:9. doi: 10.1155/2018/2630565.2630565
    1. Harvey C. J. d. C., Schofield G. M., Williden M. The use of nutritional supplements to induce ketosis and reduce symptoms associated with keto-induction: a narrative review. PeerJ. 2018;6, article e4488 doi: 10.7717/peerj.4488.
    1. Li T., Li X., Attri K. S., et al. O-GlcNAc transferase links glucose metabolism to MAVS-mediated antiviral innate immunity. Cell Host & Microbe. 2018;24(6):791–803.e6. doi: 10.1016/j.chom.2018.11.001.
    1. Wang Q., Fang P., He R., et al. O-GlcNAc transferase promotes influenza A virus–induced cytokine storm by targeting interferon regulatory factor–5. Science Advances. 2020;6(16, article eaaz7086) doi: 10.1126/sciadv.aaz7086.
    1. Kim S. R., Lee S. G., Kim S. H., et al. SGLT2 inhibition modulates NLRP3 inflammasome activity via ketones and insulin in diabetes with cardiovascular disease. Nature Communications. 2020;11(1) doi: 10.1038/s41467-020-15983-6.
    1. Langhans W., Hrupka B. J. Cytokines and appetite. In: Kronfol Z., editor. Cytokines and Mental Health. Boston, MA: Springer; 2003. pp. 167–209.
    1. Pan L., Mu M., Yang P., et al. Clinical characteristics of COVID-19 patients with digestive symptoms in Hubei, China: a descriptive, cross-sectional, multicenter study. The American Journal of Gastroenterology. 2020;115(5):766–773. doi: 10.14309/ajg.0000000000000620.
    1. Lei Y., Moore C. B., Liesman R. M., et al. MAVS-mediated apoptosis and its inhibition by viral proteins. PLoS One. 2009;4(5, article e5466) doi: 10.1371/journal.pone.0005466.
    1. Jiang H., Shi H., Sun M., et al. PFKFB3-driven macrophage glycolytic metabolism is a crucial component of innate antiviral defense. Journal of Immunology. 2016;197(7):2880–2890. doi: 10.4049/jimmunol.1600474.
    1. Angelidi A. M., Belanger M. J., Mantzoros C. S. Commentary: COVID-19 and diabetes mellitus: what we know, how our patients should be treated now, and what should happen next. Metabolism. 2020;107, article 154245 doi: 10.1016/j.metabol.2020.154245.
    1. Myette-Côté É., Caldwell H. G., Ainslie P. N., Clarke K., Little J. P. A ketone monoester drink reduces the glycemic response to an oral glucose challenge in individuals with obesity: a randomized trial. The American Journal of Clinical Nutrition. 2019;110(6):1491–1501. doi: 10.1093/ajcn/nqz232.
    1. Stubbs B. J., Cox P. J., Evans R. D., et al. On the metabolism of exogenous ketones in humans. Frontiers in Physiology. 2017;8:848–848. doi: 10.3389/fphys.2017.00848.
    1. Lottes R. G., Newton D. A., Spyropoulos D. D., Baatz J. E. Lactate as substrate for mitochondrial respiration in alveolar epithelial type II cells. American Journal of Physiology-Lung Cellular and Molecular Physiology. 2015;308(9):L953–L961. doi: 10.1152/ajplung.00335.2014.
    1. Uhlén M., Fagerberg L., Hallström B. M., et al. Proteomics. Tissue-based map of the human proteome. Science. 2015;347(6220, article 1260419) doi: 10.1126/science.1260419.
    1. Youm Y. H., Nguyen K. Y., Grant R. W., et al. The ketone metabolite β-hydroxybutyrate blocks NLRP3 inflammasome-mediated inflammatory disease. Nature Medicine. 2015;21(3):263–269. doi: 10.1038/nm.3804.
    1. Scambler T., Jarosz-Griffiths H. H., Lara-Reyna S., et al. ENaC-mediated sodium influx exacerbates NLRP3-dependent inflammation in cystic fibrosis. eLife. 2019;8 doi: 10.7554/eLife.49248.
    1. Neudorf H., Durrer C., Myette-Cote E., Makins C., O'Malley T., Little J. P. Oral ketone supplementation acutely increases markers of NLRP3 inflammasome activation in human monocytes. Molecular Nutrition & Food Research. 2019;63(11, article 1801171) doi: 10.1002/mnfr.201801171.
    1. Liu X., Zhang X., Ding Y., et al. Nuclear factor E2-related factor-2 negatively regulates NLRP3 inflammasome activity by inhibiting reactive oxygen species-induced NLRP3 priming. Antioxidants & Redox Signaling. 2017;26(1):28–43. doi: 10.1089/ars.2015.6615.
    1. Neudorf H., Myette-Côté É., Little J. P. The impact of acute ingestion of a ketone monoester drink on LPS-stimulated NLRP3 activation in humans with obesity. Nutrients. 2020;12(3):p. 854. doi: 10.3390/nu12030854.
    1. Shaw D. M., Merien F., Braakhuis A., Keaney L., Dulson D. K. Acute hyperketonaemia alters T-cell-related cytokine gene expression within stimulated peripheral blood mononuclear cells following prolonged exercise. European Journal of Applied Physiology. 2020;120(1):191–202. doi: 10.1007/s00421-019-04263-x.
    1. Li J., Wang X., Chen J., Zuo X., Zhang H., Deng A. COVID-19 infection may cause ketosis and ketoacidosis. Diabetes, Obesity & Metabolism. 2020;22 doi: 10.1111/dom.14057.
    1. Chamorro-Pareja N., Parthasarathy S., Annam J., Hoffman J., Coyle C., Kishore P. Letter to the editor: unexpected high mortality in COVID-19 and diabetic ketoacidosis. Metabolism. 2020;110, article 154301 doi: 10.1016/j.metabol.2020.154301.
    1. Brahma M. K., Ha C. M., Pepin M. E., et al. Increased glucose availability attenuates myocardial ketone body utilization. Journal of the American Heart Association. 2020;9(15, article e013039) doi: 10.1161/JAHA.119.013039.
    1. Kraut J. A., Madias N. E. Treatment of acute metabolic acidosis: a pathophysiologic approach. Nature Reviews Nephrology. 2012;8(10):589–601. doi: 10.1038/nrneph.2012.186.
    1. Bolla A. M., Caretto A., Laurenzi A., Scavini M., Piemonti L. Low-carb and ketogenic diets in type 1 and type 2 diabetes. Nutrients. 2019;11(5):p. 962. doi: 10.3390/nu11050962.
    1. Poffé C., Ramaekers M., Bogaerts S., Hespel P. Bicarbonate unlocks the ergogenic action of ketone monoester intake in endurance exercise. Medicine and Science in Sports and Exercise. 2020 doi: 10.1249/MSS.0000000000002467.
    1. Nyenwe E. A., Kitabchi A. E. The evolution of diabetic ketoacidosis: an update of its etiology, pathogenesis and management. Metabolism. 2016;65(4):507–521. doi: 10.1016/j.metabol.2015.12.007.
    1. Traba J., Kwarteng-Siaw M., Okoli T. C., et al. Fasting and refeeding differentially regulate NLRP3 inflammasome activation in human subjects. The Journal of Clinical Investigation. 2015;125(12):4592–4600. doi: 10.1172/JCI83260.
    1. Goedecke J. H., Christie C., Wilson G., et al. Metabolic adaptations to a high-fat diet in endurance cyclists. Metabolism. 1999;48(12):1509–1517. doi: 10.1016/S0026-0495(99)90238-X.
    1. Sherrier M., Li H. The impact of keto-adaptation on exercise performance and the role of metabolic-regulating cytokines. The American Journal of Clinical Nutrition. 2019;110(3):562–573. doi: 10.1093/ajcn/nqz145.
    1. Shimazu T., Hirschey M. D., Newman J., et al. Suppression of oxidative stress by β-hydroxybutyrate, an endogenous histone deacetylase inhibitor. Science. 2013;339(6116):211–214. doi: 10.1126/science.1227166.
    1. Thio C. L., Chi P. Y., Lai A. C. Y., Chang Y. J. Regulation of type 2 innate lymphoid cell-dependent airway hyperreactivity by butyrate. The Journal of Allergy and Clinical Immunology. 2018;142(6):1867–1883.e12. doi: 10.1016/j.jaci.2018.02.032.
    1. Sutendra G., Kinnaird A., Dromparis P., et al. A nuclear pyruvate dehydrogenase complex is important for the generation of acetyl-CoA and histone acetylation. Cell. 2014;158(1):84–97. doi: 10.1016/j.cell.2014.04.046.
    1. Kikani C. K., Verona E. V., Ryu J., et al. Proliferative and antiapoptotic signaling stimulated by nuclear-localized PDK1 results in oncogenesis. Science Signaling. 2012;5(249) doi: 10.1126/scisignal.2003065.
    1. Chang P. V., Hao L., Offermanns S., Medzhitov R. The microbial metabolite butyrate regulates intestinal macrophage function via histone deacetylase inhibition. Proceedings of the National Academy of Sciences of the United States of America. 2014;111(6):2247–2252. doi: 10.1073/pnas.1322269111.
    1. Ohira H., Fujioka Y., Katagiri C., et al. Butyrate attenuates inflammation and lipolysis generated by the interaction of adipocytes and macrophages. Journal of Atherosclerosis and Thrombosis. 2013;20(5):425–442. doi: 10.5551/jat.15065.
    1. Larsen L., Tonnesen M., Ronn S. G., et al. Inhibition of histone deacetylases prevents cytokine-induced toxicity in beta cells. Diabetologia. 2007;50(4):779–789. doi: 10.1007/s00125-006-0562-3.
    1. Chriett S., Dąbek A., Wojtala M., Vidal H., Balcerczyk A., Pirola L. Prominent action of butyrate over β-hydroxybutyrate as histone deacetylase inhibitor, transcriptional modulator and anti-inflammatory molecule. Scientific Reports. 2019;9(1):p. 742. doi: 10.1038/s41598-018-36941-9.
    1. Cavaleri F., Bashar E. Potential synergies of β-hydroxybutyrate and butyrate on the modulation of metabolism, inflammation, cognition, and general health. Journal of Nutrition and Metabolism. 2018;2018:13. doi: 10.1155/2018/7195760.7195760
    1. Schulthess J., Pandey S., Capitani M., et al. The short chain fatty acid butyrate imprints an antimicrobial program in macrophages. Immunity. 2019;50(2):432–445.e7. doi: 10.1016/j.immuni.2018.12.018.
    1. Kasotakis G., Kintsurashvili E., Galvan M. D., et al. Histone deacetylase 7 inhibition in a murine model of gram-negative pneumonia-induced acute lung injury. Shock. 2020;53(3):344–351. doi: 10.1097/SHK.0000000000001372.
    1. Chakraborty K., Raundhal M., Chen B. B., et al. The mito-DAMP cardiolipin blocks IL-10 production causing persistent inflammation during bacterial pneumonia. Nature Communications. 2017;8(1) doi: 10.1038/ncomms13944.
    1. Imai Y., Kuba K., Neely G. G., et al. Identification of oxidative stress and Toll-like receptor 4 signaling as a key pathway of acute lung injury. Cell. 2008;133(2):235–249. doi: 10.1016/j.cell.2008.02.043.
    1. Offermanns S. Hydroxy-carboxylic acid receptor actions in metabolism. Trends in Endocrinology & Metabolism. 2017;28(3):227–236. doi: 10.1016/j.tem.2016.11.007.
    1. Docampo M. D., Stein-Thoeringer C. K., Lazrak A., Burgos da Silva M. D., Cross J., van den Brink M. R. M. Expression of the butyrate/niacin receptor, GPR109a on T cells plays an important role in a mouse model of graft versus host disease. The Journal of Immunology. 2018;132(1 Supplement) doi: 10.1182/blood-2018-99-118783.
    1. Ma X., Luo X., Zhou S., et al. Hydroxycarboxylic acid receptor 2 is a Zika virus restriction factor that can be induced by Zika virus infection through the IRE1-XBP1 pathway. Frontiers in Cellular and Infection Microbiology. 2020;9:p. 480. doi: 10.3389/fcimb.2019.00480.
    1. Thangaraju M., Cresci G. A., Liu K., et al. GPR109A is a G-protein-coupled receptor for the bacterial fermentation product butyrate and functions as a tumor suppressor in colon. Cancer Research. 2009;69(7):2826–2832. doi: 10.1158/0008-5472.CAN-08-4466.
    1. Zandi-Nejad K., Takakura A., Jurewicz M., et al. The role of HCA2 (GPR109A) in regulating macrophage function. The FASEB Journal. 2013;27(11):4366–4374. doi: 10.1096/fj.12-223933.
    1. Ye L., Cao Z., Lai X., et al. Niacin fine-tunes energy homeostasis through canonical GPR109A signaling. The FASEB Journal. 2019;33(4):4765–4779. doi: 10.1096/fj.201801951R.
    1. Singh N., Gurav A., Sivaprakasam S., et al. Activation of Gpr109a, receptor for niacin and the commensal metabolite butyrate, suppresses colonic inflammation and carcinogenesis. Immunity. 2014;40(1):128–139. doi: 10.1016/j.immuni.2013.12.007.
    1. Guo W., Liu J., Sun J., et al. Butyrate alleviates oxidative stress by regulating NRF2 nuclear accumulation and H3K9/14 acetylation via GPR109A in bovine mammary epithelial cells and mammary glands. Free Radical Biology & Medicine. 2020;152:728–742. doi: 10.1016/j.freeradbiomed.2020.01.016.
    1. Chen G., Huang B., Fu S., et al. G protein-coupled receptor 109A and host microbiota modulate intestinal epithelial integrity during sepsis. Frontiers in Immunology. 2018;9:p. 2079. doi: 10.3389/fimmu.2018.02079.
    1. Ahmed A., Siman-Tov G., Hall G., Bhalla N., Narayanan A. Human antimicrobial peptides as therapeutics for viral infections. Viruses. 2019;11(8):p. 704. doi: 10.3390/v11080704.
    1. Takiguchi T., Morizane S., Yamamoto T., Kajita A., Ikeda K., Iwatsuki K. Cathelicidin antimicrobial peptide LL-37 augments interferon‐β expression and antiviral activity induced by double-stranded RNA in keratinocytes. The British Journal of Dermatology. 2014;171(3):492–498. doi: 10.1111/bjd.12942.
    1. McHugh B. J., Wang R., Li H. N., et al. Cathelicidin is a "fire alarm", generating protective NLRP3-dependent airway epithelial cell inflammatory responses during infection with Pseudomonas aeruginosa. PLoS Pathogens. 2019;15(4, article e1007694) doi: 10.1371/journal.ppat.1007694.
    1. Hsieh I. N., Hartshorn K. L. The role of antimicrobial peptides in influenza virus infection and their potential as antiviral and immunomodulatory therapy. Pharmaceuticals. 2016;9(3):p. 53. doi: 10.3390/ph9030053.
    1. White M. R., Tripathi S., Verma A., et al. Collectins, H-ficolin and LL-37 reduce influence viral replication in human monocytes and modulate virus-induced cytokine production. Innate Immunity. 2017;23(1):77–88. doi: 10.1177/1753425916678470.
    1. Pinheiro da Silva F., Machado M. C. The dual role of cathelicidins in systemic inflammation. Immunology Letters. 2017;182:57–60. doi: 10.1016/j.imlet.2017.01.004.
    1. Casanova V., Sousa F. H., Shakamuri P., et al. Citrullination alters the antiviral and immunomodulatory activities of the human cathelicidin LL-37 during rhinovirus infection. Frontiers in Immunology. 2020;11 doi: 10.3389/fimmu.2020.00085.
    1. Damgaard D., Bjørn M. E., Jensen P. Ø., Nielsen C. H. Reactive oxygen species inhibit catalytic activity of peptidylarginine deiminase. Journal of Enzyme Inhibition and Medicinal Chemistry. 2017;32(1):1203–1208. doi: 10.1080/14756366.2017.1368505.
    1. Fuchs T. A., Abed U., Goosmann C., et al. Novel cell death program leads to neutrophil extracellular traps. The Journal of Cell Biology. 2007;176(2):231–241. doi: 10.1083/jcb.200606027.
    1. Liu Q., Liu J., Roschmann K. I. L., et al. Histone deacetylase inhibitors up-regulate LL-37 expression independent of toll-like receptor mediated signalling in airway epithelial cells. Journal of Inflammation. 2013;10(1):p. 15. doi: 10.1186/1476-9255-10-15.
    1. Zhang K., Hussain T., Wang J., et al. Sodium butyrate abrogates the growth and pathogenesis of Mycobacterium bovis via regulation of cathelicidin (LL37) expression and NF-κB signaling. Frontiers in Microbiology. 2020;11:p. 433. doi: 10.3389/fmicb.2020.00433.
    1. Langfort J., Pilis W., Zarzeczny R., Nazar K., Kaciuba-Uściłko H. Effect of low-carbohydrate-ketogenic diet on metabolic and hormonal responses to graded exercise in men. Journal of Physiology and Pharmacology. 1996;47(2):361–371.
    1. Volek J. S., Sharman M. J., Love D. M., et al. Body composition and hormonal responses to a carbohydrate-restricted diet. Metabolism. 2002;51(7):864–870. doi: 10.1053/meta.2002.32037.
    1. Nakamura Y., Walker B. R., Ikuta T. Systematic review and meta-analysis reveals acutely elevated plasma cortisol following fasting but not less severe calorie restriction. Stress. 2016;19(2):151–157. doi: 10.3109/10253890.2015.1121984.
    1. Cox P. J., Kirk T., Ashmore T., et al. Nutritional ketosis alters fuel preference and thereby endurance performance in athletes. Cell Metabolism. 2016;24(2):256–268. doi: 10.1016/j.cmet.2016.07.010.
    1. Coutinho A. E., Chapman K. E. The anti-inflammatory and immunosuppressive effects of glucocorticoids, recent developments and mechanistic insights. Molecular and Cellular Endocrinology. 2011;335(1):2–13. doi: 10.1016/j.mce.2010.04.005.
    1. Lansbury L., Rodrigo C., Leonardi-Bee J., Nguyen-van-Tam J., Lim W. S., Cochrane Acute Respiratory Infections Group Corticosteroids as adjunctive therapy in the treatment of influenza. Cochrane Database of Systematic Reviews. 2019;(2) doi: 10.1002/14651858.CD010406.pub3.CD010406
    1. Selvaraj V., Dapaah-Afriyie K., Finn A., Flanigan T. P. Short-term dexamethasone in Sars-CoV-2 patients. Rhode Island medical journal (2013) 2020;103(6):39–43.
    1. Horby P., Lim W. S., Emberson J., et al. Effect of dexamethasone in hospitalized patients with COVID-19: preliminary report. medRxiv. 2020 doi: 10.1101/2020.06.22.20137273.
    1. Li W., Moore M. J., Vasilieva N., et al. Angiotensin-converting enzyme 2 is a functional receptor for the SARS coronavirus. Nature. 2003;426(6965):450–454. doi: 10.1038/nature02145.
    1. Bernstein K. E., Khan Z., Giani J. F., Cao D. Y., Bernstein E. A., Shen X. Z. Angiotensin-converting enzyme in innate and adaptive immunity. Nature Reviews. Nephrology. 2018;14(5):325–336. doi: 10.1038/nrneph.2018.15.
    1. Wan Y., Shang J., Graham R., Baric R. S., Li F. Receptor recognition by the novel coronavirus from Wuhan: an analysis based on decade-long structural studies of SARS coronavirus. Journal of Virology. 2020;94(7) doi: 10.1128/JVI.00127-20.
    1. Bourgonje A. R., Abdulle A. E., Timens W., et al. Angiotensin-converting enzyme 2 (ACE2),SARS-CoV-2 and the pathophysiology of coronavirus disease 2019 (COVID-19) The Journal of Pathology. 2020;251(3):228–248. doi: 10.1002/path.5471.
    1. Kuba K., Imai Y., Rao S., et al. A crucial role of angiotensin converting enzyme 2 (ACE2) in SARS coronavirus-induced lung injury. Nature Medicine. 2005;11(8):875–879. doi: 10.1038/nm1267.
    1. Sungnak W., Network H. C. A. L. B., Huang N., et al. SARS-CoV-2 entry factors are highly expressed in nasal epithelial cells together with innate immune genes. Nature Medicine. 2020;26(5):681–687. doi: 10.1038/s41591-020-0868-6.
    1. Gurwitz D. Angiotensin receptor blockers as tentative SARS-CoV-2 therapeutics. Drug Development Research. 2020;81(5):537–540. doi: 10.1002/ddr.21656.
    1. Passos-Silva D. G., Brandan E., Santos R. A. S. Angiotensins as therapeutic targets beyond heart disease. Trends in Pharmacological Sciences. 2015;36(5):310–320. doi: 10.1016/j.tips.2015.03.001.
    1. Carvajal G., Rodríguez-Vita J., Rodrigues-Díez R., et al. Angiotensin II activates the Smad pathway during epithelial mesenchymal transdifferentiation. Kidney International. 2008;74(5):585–595. doi: 10.1038/ki.2008.213.
    1. Komai T., Inoue M., Okamura T., et al. Transforming growth factor-β and interleukin-10 synergistically regulate humoral immunity via modulating metabolic signals. Frontiers in Immunology. 2018;9 doi: 10.3389/fimmu.2018.01364.
    1. Zhang L., du J., Hu Z., et al. IL-6 and serum amyloid A synergy mediates angiotensin II-induced muscle wasting. Journal of the American Society of Nephrology. 2009;20(3):604–612. doi: 10.1681/ASN.2008060628.
    1. Ma L. J., Corsa B. A., Zhou J., et al. Angiotensin type 1 receptor modulates macrophage polarization and renal injury in obesity. American Journal of Physiology-Renal Physiology. 2011;300(5):F1203–F1213. doi: 10.1152/ajprenal.00468.2010.
    1. Huss D. J., Winger R. C., Cox G. M., et al. TGF-β signaling via Smad4 drives IL-10 production in effector Th1 cells and reduces T-cell trafficking in EAE. European Journal of Immunology. 2011;41(10):2987–2996. doi: 10.1002/eji.201141666.
    1. Choi J., Park S., Kwon T. K., Sohn S. I., Park K. M., Kim J. I. Role of the histone deacetylase inhibitor valproic acid in high-fat diet-induced hypertension via inhibition of HDAC1/angiotensin II axis. International Journal of Obesity. 2017;41(11):1702–1709. doi: 10.1038/ijo.2017.166.
    1. Wang L., Zhu Q., Lu A., et al. Sodium butyrate suppresses angiotensin II-induced hypertension by inhibition of renal (pro)renin receptor and intrarenal renin-angiotensin system. Journal of Hypertension. 2017;35(9):1899–1908. doi: 10.1097/HJH.0000000000001378.
    1. Swartz T. H., Mamedova L. K., Bradford B. J. Beta-hydroxybutyrate alters the mRNA cytokine profile from mouse macrophages challenged with Streptococcus uberis. Kansas Agricultural Experiment Station Research Reports. 2019;5(9) doi: 10.4148/2378-5977.7874.
    1. Miyachi Y., Tsuchiya K., Shiba K., et al. A reduced M1-like/M2-like ratio of macrophages in healthy adipose tissue expansion during SGLT2 inhibition. Scientific Reports. 2018;8(1) doi: 10.1038/s41598-018-34305-x.
    1. Shi X., Li X., Li D., et al. ß-Hydroxybutyrate activates the NF-κB signaling pathway to promote the expression of pro-inflammatory factors in calf hepatocytes. Cellular Physiology and Biochemistry. 2014;33(4):920–932. doi: 10.1159/000358664.
    1. Shen T., Li X., Loor J. J., et al. Hepatic nuclear factor kappa B signaling pathway and NLR family pyrin domain containing 3 inflammasome is over-activated in ketotic dairy cows. Journal of Dairy Science. 2019;102(11):10554–10563. doi: 10.3168/jds.2019-16706.
    1. Fu S. P., Li S. N., Wang J. F., et al. BHBA suppresses LPS-induced inflammation in BV-2 cells by inhibiting NF-κB activation. Mediators of Inflammation. 2014;2014:12. doi: 10.1155/2014/983401.983401
    1. Kajitani N., Iwata M., Miura A., et al. Prefrontal cortex infusion of beta-hydroxybutyrate, an endogenous NLRP3 inflammasome inhibitor, produces antidepressant-like effects in a rodent model of depression. Neuropsychopharmacology Reports. 2020;40(2):157–165. doi: 10.1002/npr2.12099.
    1. Yamanashi T., Iwata M., Kamiya N., et al. Beta-hydroxybutyrate, an endogenic NLRP3 inflammasome inhibitor, attenuates stress-induced behavioral and inflammatory responses. Scientific Reports. 2017;7(1) doi: 10.1038/s41598-017-08055-1.
    1. Zdzisińska B., Filar J., Paduch R., Kaczor J., Lokaj I., Kandefer-Szerszeń M. The influence of ketone bodies and glucose on interferon, tumor necrosis factor production and NO release in bovine aorta endothelial cells. Veterinary Immunology and Immunopathology. 2000;74(3-4):237–247. doi: 10.1016/S0165-2427(00)00175-6.
    1. Kim D. H., Park M. H., Ha S., et al. Anti-inflammatory action of β-hydroxybutyrate via modulation of PGC-1α and FoxO1, mimicking calorie restriction. Aging. 2019;11(4):1283–1304. doi: 10.18632/aging.101838.
    1. Miyauchi T., Uchida Y., Kadono K., et al. Up-regulation of FOXO1 and reduced inflammation by β-hydroxybutyric acid are essential diet restriction benefits against liver injury. Proceedings of the National Academy of Sciences of the United States of America. 2019;116(27):13533–13542. doi: 10.1073/pnas.1820282116.
    1. Cowell R. M., Talati P., Blake K. R., Meador-Woodruff J. H., Russell J. W. Identification of novel targets for PGC-1alpha and histone deacetylase inhibitors in neuroblastoma cells. Biochemical and Biophysical Research Communications. 2009;379(2):578–582. doi: 10.1016/j.bbrc.2008.12.109.
    1. Cho H. M., Seok Y., Lee H., Song M., Kim I. K. Repression of transcriptional activity of forkhead box O1 by histone deacetylase inhibitors ameliorates hyperglycemia in type 2 diabetic rats. International Journal of Molecular Sciences. 2018;19(11):p. 3539. doi: 10.3390/ijms19113539.
    1. Summermatter S., Baum O., Santos G., Hoppeler H., Handschin C. Peroxisome proliferator-activated receptor γ coactivator 1α (PGC-1α) promotes skeletal muscle lipid refueling in vivo by activating de novo lipogenesis and the pentose phosphate pathway. Journal of Biological Chemistry. 2010;285(43):32793–32800. doi: 10.1074/jbc.M110.145995.
    1. Yeo H., Lyssiotis C. A., Zhang Y., et al. FoxO3 coordinates metabolic pathways to maintain redox balance in neural stem cells. The EMBO Journal. 2013;32(19):2589–2602. doi: 10.1038/emboj.2013.186.
    1. Greer E. L., Dowlatshahi D., Banko M. R., et al. An AMPK-FOXO pathway mediates longevity induced by a novel method of dietary restriction in C. elegans. Current Biology. 2007;17(19):1646–1656. doi: 10.1016/j.cub.2007.08.047.
    1. Hariharan N., Maejima Y., Nakae J., Paik J., DePinho R. A., Sadoshima J. Deacetylation of FoxO by Sirt1 plays an essential role in mediating starvation-induced autophagy in cardiac myocytes. Circulation Research. 2010;107(12):1470–1482. doi: 10.1161/CIRCRESAHA.110.227371.
    1. Fan W., Morinaga H., Kim J. J., et al. FoxO1 regulates Tlr4 inflammatory pathway signalling in macrophages. The EMBO Journal. 2010;29(24):4223–4236. doi: 10.1038/emboj.2010.268.
    1. Chung S., Lee T. J., Reader B. F., et al. FoxO1 regulates allergic asthmatic inflammation through regulating polarization of the macrophage inflammatory phenotype. Oncotarget. 2016;7(14):17532–17546. doi: 10.18632/oncotarget.8162.
    1. Yang J. B., Zhao Z. B., Liu Q. Z., et al. FoxO1 is a regulator of MHC-II expression and anti-tumor effect of tumor-associated macrophages. Oncogene. 2018;37(9):1192–1204. doi: 10.1038/s41388-017-0048-4.
    1. Wang Y. P., Zhou L. S., Zhao Y. Z., et al. Regulation of G6PD acetylation by SIRT2 and KAT9 modulates NADPH homeostasis and cell survival during oxidative stress. The EMBO Journal. 2014;33(12):1304–1320. doi: 10.1002/embj.201387224.
    1. He M., Chiang H. H., Luo H., et al. An acetylation switch of the NLRP3 inflammasome regulates aging-associated chronic inflammation and insulin resistance. Cell Metabolism. 2020;31(3):580–591.e5. doi: 10.1016/j.cmet.2020.01.009.
    1. Nemoto S., Fergusson M. M., Finkel T. SIRT1 functionally interacts with the metabolic regulator and transcriptional coactivator PGC-1α. The Journal of Biological Chemistry. 2005;280(16):16456–16460. doi: 10.1074/jbc.M501485200.
    1. Qiu X., Brown K., Hirschey M. D., Verdin E., Chen D. Calorie restriction reduces oxidative stress by SIRT3-mediated SOD2 activation. Cell Metabolism. 2010;12(6):662–667. doi: 10.1016/j.cmet.2010.11.015.
    1. Someya S., Yu W., Hallows W. C., et al. Sirt3 mediates reduction of oxidative damage and prevention of age-related hearing loss under caloric restriction. Cell. 2010;143(5):802–812. doi: 10.1016/j.cell.2010.10.002.
    1. Ahn B. H., Kim H. S., Song S., et al. A role for the mitochondrial deacetylase Sirt3 in regulating energy homeostasis. Proceedings of the National Academy of Sciences. 2008;105(38):14447–14452. doi: 10.1073/pnas.0803790105.
    1. Cortés-Rojo C., Vargas-Vargas M. A., Olmos-Orizaba B. E., Rodríguez-Orozco A. R., Calderón-Cortés E. Interplay between NADH oxidation by complex I, glutathione redox state and sirtuin-3, and its role in the development of insulin resistance. Biochimica et Biophysica Acta (BBA) - Molecular Basis of Disease. 2020;1866(8, article 165801) doi: 10.1016/j.bbadis.2020.165801.
    1. Rouillard A. D., Gundersen G. W., Fernandez N. F., et al. The harmonizome: a collection of processed datasets gathered to serve and mine knowledge about genes and proteins. Database. 2016;2016 doi: 10.1093/database/baw100.
    1. Ranhotra H. S. Estrogen-related receptor alpha and mitochondria: tale of the titans. Journal of Receptor and Signal Transduction. 2015;35(5):386–390. doi: 10.3109/10799893.2014.959592.
    1. Svensson K., Albert V., Cardel B., Salatino S., Handschin C. Skeletal muscle PGC-1α modulates systemic ketone body homeostasis and ameliorates diabetic hyperketonemia in mice. The FASEB Journal. 2016;30(5):1976–1986. doi: 10.1096/fj.201500128.
    1. Villena J. A., Hock M. B., Chang W. Y., Barcas J. E., Giguere V., Kralli A. Orphan nuclear receptor estrogen-related receptor is essential for adaptive thermogenesis. Proceedings of the National Academy of Sciences. 2007;104(4):1418–1423. doi: 10.1073/pnas.0607696104.
    1. Giguere V. Transcriptional control of energy homeostasis by the estrogen-related receptors. Endocrine Reviews. 2008;29(6):677–696. doi: 10.1210/er.2008-0017.
    1. Tripathi M., Yen P. M., Singh B. K. Estrogen-related receptor alpha: an under-appreciated potential target for the treatment of metabolic diseases. International Journal of Molecular Sciences. 2020;21(5):p. 1645. doi: 10.3390/ijms21051645.
    1. Schnyder S., Svensson K., Cardel B., Handschin C. Muscle PGC-1α is required for long-term systemic and local adaptations to a ketogenic diet in mice. American Journal of Physiology. Endocrinology and Metabolism. 2017;312(5):E437–E446. doi: 10.1152/ajpendo.00361.2016.
    1. Srivastava S., Baxa U., Niu G., Chen X., Veech R. L. A ketogenic diet increases brown adipose tissue mitochondrial proteins and UCP1 levels in mice. IUBMB Life. 2013;65(1):58–66. doi: 10.1002/iub.1102.
    1. Marcondes S., Turko I. V., Murad F. Nitration of succinyl-CoA:3-oxoacid CoA-transferase in rats after endotoxin administration. Proceedings of the National Academy of Sciences of the United States of America. 2001;98(13):7146–7151. doi: 10.1073/pnas.141222598.
    1. Rebrin I., Brégère C., Kamzalov S., Gallaher T. K., Sohal R. S. Nitration of tryptophan 372 in succinyl-CoA:3-ketoacid CoA transferase during aging in rat heart mitochondria†. Biochemistry. 2007;46(35):10130–11044. doi: 10.1021/bi7001482.
    1. Dittenhafer-Reed K. E., Richards A. L., Fan J., et al. SIRT3 mediates multi-tissue coupling for metabolic fuel switching. Cell Metabolism. 2015;21(4):637–646. doi: 10.1016/j.cmet.2015.03.007.
    1. Barroso W. A., Victorino V. J., Jeremias I. C., et al. High-fat diet inhibits PGC-1α suppressive effect on NFκB signaling in hepatocytes. European Journal of Nutrition. 2018;57(5):1891–1900. doi: 10.1007/s00394-017-1472-5.
    1. Veech R. L. Ketone esters increase brown fat in mice and overcome insulin resistance in other tissues in the rat. Annals of the New York Academy of Sciences. 2013;1302(1):42–48. doi: 10.1111/nyas.12222.
    1. McCormick M., Chen K., Ramaswamy P., Kenyon C. New genes that extend Caenorhabditis elegans' lifespan in response to reproductive signals. Aging Cell. 2012;11(2):192–202. doi: 10.1111/j.1474-9726.2011.00768.x.
    1. Shaw W. M., Luo S., Landis J., Ashraf J., Murphy C. T. The C. elegans TGF-β Dauer pathway regulates longevity via insulin signaling. Current Biology. 2007;17(19):1635–1645. doi: 10.1016/j.cub.2007.08.058.
    1. Balsa E., Perry E. A., Bennett C. F., et al. Defective NADPH production in mitochondrial disease complex I causes inflammation and cell death. Nature Communications. 2020;11(1) doi: 10.1038/s41467-020-16423-1.
    1. Chen Q., Kirk K., Shurubor Y. I., et al. Rewiring of glutamine metabolism is a bioenergetic adaptation of human cells with mitochondrial DNA mutations. Cell Metabolism. 2018;27(5):1007–1025.e5. doi: 10.1016/j.cmet.2018.03.002.
    1. Gaude E., Schmidt C., Gammage P. A., et al. NADH shuttling couples cytosolic reductive carboxylation of glutamine with glycolysis in cells with mitochondrial dysfunction. Molecular Cell. 2018;69(4):581–593.e7. doi: 10.1016/j.molcel.2018.01.034.
    1. Parmeggiani A., Bowman R. H. Regulation of phosphofructokinase activity by citrate in normal and diabetic muscle. Biochemical and Biophysical Research Communications. 1963;12(4):268–273. doi: 10.1016/0006-291X(63)90294-8.
    1. Weber G., Lea M. A., Stamm N. B. Inhibition of pyruvate kinase and glucokinase by acetyl CoA and inhibition of glucokinase by phosphoenolpyruvate. Life Sciences. 1967;6(22):2441–2452. doi: 10.1016/0024-3205(67)90071-9.
    1. Infantino V., Pierri C. L., Iacobazzi V. Metabolic routes in inflammation: the citrate pathway and its potential as therapeutic target. Current Medicinal Chemistry. 2020;26(40):7104–7116. doi: 10.2174/0929867325666180510124558.
    1. Jha A. K., Huang S. C. C., Sergushichev A., et al. Network integration of parallel metabolic and transcriptional data reveals metabolic modules that regulate macrophage polarization. Immunity. 2015;42(3):419–430. doi: 10.1016/j.immuni.2015.02.005.
    1. De Santa F., Narang V., Yap Z. H., et al. Jmjd3 contributes to the control of gene expression in LPS-activated macrophages. The EMBO Journal. 2009;28(21):3341–3352. doi: 10.1038/emboj.2009.271.
    1. Stender J. D., Pascual G., Liu W., et al. Control of proinflammatory gene programs by regulated trimethylation and demethylation of histone H4K20. Molecular Cell. 2012;48(1):28–38. doi: 10.1016/j.molcel.2012.07.020.
    1. Kang M. K., Mehrazarin S., Park N. H., Wang C. Y. Epigenetic gene regulation by histone demethylases: emerging role in oncogenesis and inflammation. Oral Diseases. 2017;23(6):709–720. doi: 10.1111/odi.12569.
    1. Lambert A. J., Brand M. D. Superoxide production by NADH:ubiquinone oxidoreductase (complex I) depends on the pH gradient across the mitochondrial inner membrane. Biochemical Journal. 2004;382(2):511–517. doi: 10.1042/BJ20040485.
    1. Murphy M. P. How mitochondria produce reactive oxygen species. Biochemical Journal. 2009;417(1):1–13. doi: 10.1042/BJ20081386.
    1. Sullivan P. G., Rippy N. A., Dorenbos K., Concepcion R. C., Agarwal A. K., Rho J. M. The ketogenic diet increases mitochondrial uncoupling protein levels and activity. Annals of Neurology. 2004;55(4):576–580. doi: 10.1002/ana.20062.
    1. Lu Y., Yang Y. Y., Zhou M. W., et al. Ketogenic diet attenuates oxidative stress and inflammation after spinal cord injury by activating Nrf2 and suppressing the NF-κB signaling pathways. Neuroscience Letters. 2018;683:13–18. doi: 10.1016/j.neulet.2018.06.016.
    1. Wei T., Tian W., Liu F., Xie G. Protective effects of exogenous β-hydroxybutyrate on paraquat toxicity in rat kidney. Biochemical and Biophysical Research Communications. 2014;447(4):666–671. doi: 10.1016/j.bbrc.2014.04.074.
    1. Sun J., Ren X., Simpkins J. W. Sequential upregulation of superoxide dismutase 2 and heme oxygenase 1 by tert-butylhydroquinone protects mitochondria during oxidative stress. Molecular Pharmacology. 2015;88(3):437–449. doi: 10.1124/mol.115.098269.
    1. Mann G. E., Niehueser-Saran J., Watson A., et al. Nrf2/ARE regulated antioxidant gene expression in endothelial and smooth muscle cells in oxidative stress: implications for atherosclerosis and preeclampsia. Sheng Li Xue Bao. 2007;59(2):117–127.
    1. Mitsuishi Y., Taguchi K., Kawatani Y., et al. Nrf2 redirects glucose and glutamine into anabolic pathways in metabolic reprogramming. Cancer Cell. 2012;22(1):66–79. doi: 10.1016/j.ccr.2012.05.016.
    1. Clements C. M., McNally R. S., Conti B. J., Mak T. W., Ting J. P. Y. DJ-1, a cancer- and Parkinson's disease-associated protein, stabilizes the antioxidant transcriptional master regulator Nrf2. Proceedings of the National Academy of Sciences. 2006;103(41):15091–15096. doi: 10.1073/pnas.0607260103.
    1. Park Y.-K., Ahn D. R., Oh M., et al. Nitric oxide donor, (±)-S-nitroso-N-acetylpenicillamine, stabilizes transactive hypoxia-inducible factor-1α by inhibiting von Hippel-Lindau recruitment and asparagine hydroxylation. Molecular Pharmacology. 2008;74(1):236–245. doi: 10.1124/mol.108.045278.
    1. Walmsley S. R., Print C., Farahi N., et al. Hypoxia-induced neutrophil survival is mediated by HIF-1α–dependent NF-κB activity. Journal of Experimental Medicine. 2005;201(1):105–115. doi: 10.1084/jem.20040624.
    1. Lu H., Forbes R. A., Verma A. Hypoxia-inducible factor 1 activation by aerobic glycolysis implicates the Warburg effect in carcinogenesis. Journal of Biological Chemistry. 2002;277(26):23111–23115. doi: 10.1074/jbc.M202487200.
    1. Tannahill G. M., Curtis A. M., Adamik J., et al. Succinate is an inflammatory signal that induces IL-1β through HIF-1α. Nature. 2013;496(7444):238–242. doi: 10.1038/nature11986.
    1. Anderson G., Reiter R. J. Melatonin: roles in influenza, Covid-19, and other viral infections. Reviews in Medical Virology. 2020;30(3, article e2109) doi: 10.1002/rmv.2109.
    1. Alexander R. K., Liou Y. H., Knudsen N. H., et al. Bmal1 integrates mitochondrial metabolism and macrophage activation. Elife. 2020;9 doi: 10.7554/eLife.54090.
    1. Delbrel E., Soumare A., Naguez A., et al. HIF-1α triggers ER stress and CHOP-mediated apoptosis in alveolar epithelial cells, a key event in pulmonary fibrosis. Scientific Reports. 2018;8(1, article 17939) doi: 10.1038/s41598-018-36063-2.
    1. Guo X., Zhu Z., Zhang W., et al. Nuclear translocation of HIF-1α induced by influenza A (H1N1) infection is critical to the production of proinflammatory cytokines. Emerging Microbes & Infections. 2017;6(5, article e39) doi: 10.1038/emi.2017.21.
    1. Meng X., Guo X., Zhu Y., Xie H., Wang R. Mechanism of nuclear translocation of hypoxia-inducible factor-1α in influenza A (H1N1) virus infected-alveolar epithelial cells. Zhonghua Wei Zhong Bing Ji Jiu Yi Xue. 2020;32(1):8–13. doi: 10.3760/cma.j.cn121430-20191023-00002.
    1. Zhao C., Chen J., Cheng L., Xu K., Yang Y., Su X. Deficiency of HIF-1α enhances influenza A virus replication by promoting autophagy in alveolar type II epithelial cells. Emerging Microbes & Infections. 2020;9(1):691–706. doi: 10.1080/22221751.2020.1742585.
    1. Kilani M. M., Mohammed K. A., Nasreen N., Tepper R. S., Antony V. B. RSV causes HIF-1α stabilization via NO release in primary bronchial epithelial cells. Inflammation. 2004;28(5):245–251. doi: 10.1007/s10753-004-6047-y.
    1. Masaki T., Kojima T., Okabayashi T., et al. A nuclear factor-κB signaling pathway via protein kinase C δ regulates replication of respiratory syncytial virus in polarized normal human nasal epithelial cells. Molecular Biology of The Cell. 2011;22(13):2144–2156. doi: 10.1091/mbc.e10-11-0875.
    1. Yoboua F., Martel A., Duval A., Mukawera E.´., Grandvaux N. Respiratory syncytial virus-mediated NF-κB p65 phosphorylation at serine 536 is dependent on RIG-I, TRAF6, and IKKβ. Journal of Virology. 2010;84(14):7267–7277. doi: 10.1128/JVI.00142-10.
    1. Bradshaw P. C. Cytoplasmic and mitochondrial NADPH-coupled redox systems in the regulation of aging. Nutrients. 2019;11(3):p. 504. doi: 10.3390/nu11030504.
    1. Nisimoto Y., Jackson H. M., Ogawa H., Kawahara T., Lambeth J. D. Constitutive NADPH-dependent electron transferase activity of the Nox4 dehydrogenase domain. Biochemistry. 2010;49(11):2433–2442. doi: 10.1021/bi9022285.
    1. Worthington D. J., Rosemeyer M. A. Glutathione reductase from human erythrocytes. Catalytic properties and aggregation. European Journal of Biochemistry. 1976;67(1):231–238. doi: 10.1111/j.1432-1033.1976.tb10654.x.
    1. Lin H. R., Wu Y. H., Yen W. C., Yang C. M., Chiu D. T. Y. Diminished COX-2/PGE2-mediated antiviral response due to impaired NOX/MAPK signaling in G6PD-knockdown lung epithelial cells. PLoS One. 2016;11(4, article e0153462) doi: 10.1371/journal.pone.0153462.
    1. Nadeem A., al-Harbi N. O., Ahmad S. F., Ibrahim K. E., Siddiqui N., al-Harbi M. M. Glucose-6-phosphate dehydrogenase inhibition attenuates acute lung injury through reduction in NADPH oxidase-derived reactive oxygen species. Clinical & Experimental Immunology. 2018;191(3):279–287. doi: 10.1111/cei.13097.
    1. Wu Y. H., Chiu D., Lin H. R., Tang H. Y., Cheng M. L., Ho H. Y. Glucose-6-phosphate dehydrogenase enhances antiviral response through downregulation of NADPH sensor HSCARG and upregulation of NF-κB signaling. Viruses. 2015;7(12):6689–6706. doi: 10.3390/v7122966.
    1. Kim H. K., Han J. Tetrahydrobiopterin in energy metabolism and metabolic diseases. Pharmacological Research. 2020;157, article 104827 doi: 10.1016/j.phrs.2020.104827.
    1. Wever R. M., van Dam T., van Rijn H. J. M., de Groot F., Rabelink T. J. Tetrahydrobiopterin regulates superoxide and nitric oxide generation by recombinant endothelial nitric oxide synthase. Biochemical and Biophysical Research Communications. 1997;237(2):340–344. doi: 10.1006/bbrc.1997.7069.
    1. Arora S., Dev K., Agarwal B., Das P., Syed M. A. Macrophages: their role, activation and polarization in pulmonary diseases. Immunobiology. 2018;223(4-5):383–396. doi: 10.1016/j.imbio.2017.11.001.
    1. O’Neill L. A. J. A broken krebs cycle in macrophages. Immunity. 2015;42(3):393–394. doi: 10.1016/j.immuni.2015.02.017.
    1. Palmieri E. M., Gonzalez-Cotto M., Baseler W. A., et al. Nitric oxide orchestrates metabolic rewiring in M1 macrophages by targeting aconitase 2 and pyruvate dehydrogenase. Nature Communications. 2020;11(1) doi: 10.1038/s41467-020-14433-7.
    1. Infantino V., Iacobazzi V., Menga A., Avantaggiati M. L., Palmieri F. A key role of the mitochondrial citrate carrier (SLC25A1) in TNFα- and IFNγ-triggered inflammation. Biochimica et Biophysica Acta (BBA) - Gene Regulatory Mechanisms. 2014;1839(11):1217–1225. doi: 10.1016/j.bbagrm.2014.07.013.
    1. Mills E. L., O'Neill L. A. Reprogramming mitochondrial metabolism in macrophages as an anti-inflammatory signal. European Journal of Immunology. 2016;46(1):13–21. doi: 10.1002/eji.201445427.
    1. Zhang S., Weinberg S., DeBerge M., et al. Efferocytosis fuels requirements of fatty acid oxidation and the electron transport chain to polarize macrophages for tissue repair. Cell metabolism. 2019;29(2):443–456.e5. doi: 10.1016/j.cmet.2018.12.004.
    1. Puchalska P., Martin S. E., Huang X., et al. Hepatocyte-macrophage acetoacetate shuttle protects against tissue fibrosis. Cell Metabolism. 2019;29(2):383–398.e7. doi: 10.1016/j.cmet.2018.10.015.
    1. Krebs H. A., Wallace P. G., Hems R., Freedland R. A. Rates of ketone-body formation in the perfused rat liver. Biochemical Journal. 1969;112(5):595–600. doi: 10.1042/bj1120595.
    1. Karagiannis F., Masouleh S. K., Wunderling K., et al. Lipid-droplet formation drives pathogenic group 2 innate lymphoid cells in airway inflammation. Immunity. 2020;52(4):620–634.e6. doi: 10.1016/j.immuni.2020.03.003.
    1. Guo X.-Z. J., Dash P., Crawford J. C., et al. Lung γδ T cells mediate protective responses during neonatal influenza infection that are associated with type 2 immunity. Immunity. 2018;49(3):531–544.e6. doi: 10.1016/j.immuni.2018.07.011.
    1. Chakhtoura M., Chain R. W., Sato P. Y., et al. Ethyl pyruvate modulates murine dendritic cell activation and survival through their immunometabolism. Frontiers in Immunology. 2019;10 doi: 10.3389/fimmu.2019.00030.
    1. Everts B., Amiel E., Huang S. C. C., et al. TLR-driven early glycolytic reprogramming via the kinases TBK1-IKKɛ supports the anabolic demands of dendritic cell activation. Nature Immunology. 2014;15(4):323–332. doi: 10.1038/ni.2833.
    1. Bajwa G., DeBerardinis R. J., Shao B., Hall B., Farrar J. D., Gill M. A. Cutting edge: critical role of glycolysis in human plasmacytoid dendritic cell antiviral responses. J Immunol. 2016;196(5):2004–2009. doi: 10.4049/jimmunol.1501557.
    1. Nastasi C., Candela M., Bonefeld C. M., et al. The effect of short-chain fatty acids on human monocyte-derived dendritic cells. Scientific Reports. 2015;5(1) doi: 10.1038/srep16148.
    1. Kaisar M. M. M., Pelgrom L. R., van der Ham A. J., Yazdanbakhsh M., Everts B. Butyrate conditions human dendritic cells to prime type 1 regulatory T cells via both histone deacetylase inhibition and G protein-coupled receptor 109A signaling. Frontiers in Immunology. 2017;8 doi: 10.3389/fimmu.2017.01429.
    1. Singh N., Thangaraju M., Prasad P. D., et al. Blockade of dendritic cell development by bacterial fermentation products butyrate and propionate through a transporter (Slc5a8)-dependent inhibition of histone deacetylases. Journal of Biological Chemistry. 2010;285(36):27601–27608. doi: 10.1074/jbc.M110.102947.
    1. Das U. N. Pyruvate is an endogenous anti-inflammatory and anti-oxidant molecule. Medical Science Monitor : International Medical Journal of Experimental and Clinical Research. 2006;12(5):RA79–RA84.
    1. Kohlgruber A. C., Gal-Oz S. T., LaMarche N. M., et al. γδ T cells producing interleukin-17A regulate adipose regulatory T cell homeostasis and thermogenesis. Nature Immunology. 2018;19(5):464–474. doi: 10.1038/s41590-018-0094-2.
    1. Ni F.-F., Li C. R., Liao J. X., et al. The effects of ketogenic diet on the Th17/Treg cells imbalance in patients with intractable childhood epilepsy. Seizure. 2016;38:17–22. doi: 10.1016/j.seizure.2016.03.006.
    1. Berod L., Friedrich C., Nandan A., et al. De novo fatty acid synthesis controls the fate between regulatory T and T helper 17 cells. Nature Medicine. 2014;20(11):1327–1333. doi: 10.1038/nm.3704.
    1. McPherson P. A., McEneny J. The biochemistry of ketogenesis and its role in weight management, neurological disease and oxidative stress. Journal of Physiology and Biochemistry. 2012;68(1):141–151. doi: 10.1007/s13105-011-0112-4.
    1. Cignarella F., Cantoni C., Ghezzi L., et al. Intermittent fasting confers protection in CNS autoimmunity by altering the gut microbiota. Cell Metabolism. 2018;27(6):1222–1235.e6. doi: 10.1016/j.cmet.2018.05.006.
    1. Arpaia N., Campbell C., Fan X., et al. Metabolites produced by commensal bacteria promote peripheral regulatory T-cell generation. Nature. 2013;504(7480):451–455. doi: 10.1038/nature12726.
    1. Zhang J., Jin H., Xu Y., Shan J. Rapamycin modulate Treg/Th17 balance via regulating metabolic pathways: a study in mice. Transplantation Proceedings. 2019;51(6):2136–2140. doi: 10.1016/j.transproceed.2019.04.067.
    1. Kespohl M., Vachharajani N., Luu M., et al. The microbial metabolite butyrate induces expression of Th1-associated factors in CD4(+) T cells. Frontiers in immunology. 2017;8 doi: 10.3389/fimmu.2017.01036.
    1. Zhang H., Tang K., Ma J., et al. Ketogenesis-generated β-hydroxybutyrate is an epigenetic regulator of CD8+ T-cell memory development. Nature Cell Biology. 2020;22(1):18–25. doi: 10.1038/s41556-019-0440-0.
    1. Ma R., Ji T., Zhang H., et al. A Pck1-directed glycogen metabolic program regulates formation and maintenance of memory CD8+ T cells. Nature Cell Biology. 2018;20(1):21–27. doi: 10.1038/s41556-017-0002-2.
    1. Lam M. H.-B., Wing Y.-K., Yu M. W.-M., et al. Mental morbidities and chronic fatigue in severe acute respiratory syndrome survivors: long-term follow-up. Archives of Internal Medicine. 2009;169(22):2142–2147. doi: 10.1001/archinternmed.2009.384.

Source: PubMed

3
Tilaa