1932

Abstract

The global prevalence of metabolic diseases such as type 2 diabetes mellitus, steatohepatitis, myocardial infarction, and stroke has increased dramatically over the past two decades. These obesity-fueled disorders result, in part, from the aberrant accumulation of harmful lipid metabolites in tissues not suited for lipid storage (e.g., the liver, vasculature, heart, and pancreatic beta-cells). Among the numerous lipid subtypes that accumulate, sphingolipids such as ceramides are particularly impactful, as they elicit the selective insulin resistance, dyslipidemia, and ultimately cell death that underlie nearly all metabolic disorders. This review summarizes recent findings on the regulatory pathways controlling ceramide production, the molecular mechanisms linking the lipids to these discrete pathogenic events, and exciting attempts to develop therapeutics to reduce ceramide levels to combat metabolic disease.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-physiol-031620-093815
2021-02-10
2024-03-28
Loading full text...

Full text loading...

/deliver/fulltext/physiol/83/1/annurev-physiol-031620-093815.html?itemId=/content/journals/10.1146/annurev-physiol-031620-093815&mimeType=html&fmt=ahah

Literature Cited

  1. 1. 
    International Diabetes Federation 2020. IDF Diabetes Atlas International Diabetes Federation Brussels: https://www.diabetesatlas.org/en/
  2. 2. 
    Spengler EK, Loomba R. 2015. Recommendations for diagnosis, referral for liver biopsy, and treatment of nonalcoholic fatty liver disease and nonalcoholic steatohepatitis. Mayo Clin. Proc. 90:1233–46
    [Google Scholar]
  3. 3. 
    World Health Organ 2019. Cardovascular diseases (CVDs) Fact Sheet, World Health Organ. Geneva: https://www.who.int/news-room/fact-sheets/detail/cardiovascular-diseases-(cvds)
  4. 4. 
    Randle PJ, Garland PB, Hales CN, Newsholme EA 1963. The glucose fatty-acid cycle. Its role in insulin sensitivity and the metabolic disturbances of diabetes mellitus. Lancet 1:785–89
    [Google Scholar]
  5. 5. 
    Shulman GI. 2000. Cellular mechanisms of insulin resistance. J. Clin. Investig. 106:171–76
    [Google Scholar]
  6. 6. 
    Turinsky J, O'Sullivan DM, Bayly BP 1990. 1,2-Diacylglycerol and ceramide levels in insulin-resistant tissues of the rat in vivo. J. Biol. Chem. 265:16880–85
    [Google Scholar]
  7. 7. 
    Summers SA, Garza LA, Zhou H, Birnbaum MJ 1998. Regulation of insulin-stimulated glucose transporter GLUT4 translocation and Akt kinase activity by ceramide. Mol. Cell. Biol. 18:5457–64
    [Google Scholar]
  8. 8. 
    Chavez JA, Holland WL, Bar J, Sandhoff K, Summers SA 2005. Acid ceramidase overexpression prevents the inhibitory effects of saturated fatty acids on insulin signaling. J. Biol. Chem. 280:20148–53
    [Google Scholar]
  9. 9. 
    Chavez JA, Knotts TA, Wang LP, Li G, Dobrowsky RT et al. 2003. A role for ceramide, but not diacylglycerol, in the antagonism of insulin signal transduction by saturated fatty acids. J. Biol. Chem. 278:10297–303
    [Google Scholar]
  10. 10. 
    Chavez JA, Summers SA. 2003. Characterizing the effects of saturated fatty acids on insulin signaling and ceramide and diacylglycerol accumulation in 3T3-L1 adipocytes and C2C12 myotubes. Arch. Biochem. Biophys. 419:101–9
    [Google Scholar]
  11. 11. 
    Holland WL, Brozinick JT, Wang LP, Hawkins ED, Sargent KM et al. 2007. Inhibition of ceramide synthesis ameliorates glucocorticoid-, saturated-fat-, and obesity-induced insulin resistance. Cell Metab 5:167–79
    [Google Scholar]
  12. 12. 
    Chaurasia B, Tippetts TS, Monibas RM, Liu J, Li Y et al. 2019. Targeting a ceramide double bond improves insulin resistance and hepatic steatosis. Science 365:386–92
    [Google Scholar]
  13. 13. 
    Chen TC, Lee RA, Tsai SL, Kanamaluru D, Gray NE et al. 2019. An ANGPTL4-ceramide-protein kinase Cζ axis mediates chronic glucocorticoid exposure-induced hepatic steatosis and hypertriglyceridemia in mice. J. Biol. Chem. 294:9213–24
    [Google Scholar]
  14. 14. 
    Correnti JM, Juskeviciute E, Swarup A, Hoek JB 2014. Pharmacological ceramide reduction alleviates alcohol-induced steatosis and hepatomegaly in adiponectin knockout mice. Am. J. Physiol. Gastrointest. Liver Physiol. 306:G959–73
    [Google Scholar]
  15. 15. 
    Glaros EN, Kim WS, Quinn CM, Jessup W, Rye KA, Garner B 2008. Myriocin slows the progression of established atherosclerotic lesions in apolipoprotein E gene knockout mice. J. Lipid Res. 49:324–31
    [Google Scholar]
  16. 16. 
    Glaros EN, Kim WS, Wu BJ, Suarna C, Quinn CM et al. 2007. Inhibition of atherosclerosis by the serine palmitoyl transferase inhibitor myriocin is associated with reduced plasma glycosphingolipid concentration. Biochem. Pharmacol. 73:1340–46
    [Google Scholar]
  17. 17. 
    Hojjati MR, Li Z, Zhou H, Tang S, Huan C et al. 2005. Effect of myriocin on plasma sphingolipid metabolism and atherosclerosis in apoE-deficient mice. J. Biol. Chem. 280:10284–89
    [Google Scholar]
  18. 18. 
    Ji R, Akashi H, Drosatos K, Liao X, Jiang H et al. 2017. Increased de novo ceramide synthesis and accumulation in failing myocardium. JCI Insight 2:e82922
    [Google Scholar]
  19. 19. 
    Kasumov T, Li L, Li M, Gulshan K, Kirwan JP et al. 2015. Ceramide as a mediator of non-alcoholic fatty liver disease and associated atherosclerosis. PLOS ONE 10:e0126910
    [Google Scholar]
  20. 20. 
    Kurek K, Piotrowska DM, Wiesiolek-Kurek P, Łukaszuk B, Chabowski A et al. 2014. Inhibition of ceramide de novo synthesis reduces liver lipid accumulation in rats with nonalcoholic fatty liver disease. Liver Int 34:1074–83
    [Google Scholar]
  21. 21. 
    Park TS, Hu Y, Noh HL, Drosatos K, Okajima K et al. 2008. Ceramide is a cardiotoxin in lipotoxic cardiomyopathy. J. Lipid Res. 49:2101–12
    [Google Scholar]
  22. 22. 
    Park TS, Panek RL, Mueller SB, Hanselman JC, Rosebury WS et al. 2004. Inhibition of sphingomyelin synthesis reduces atherogenesis in apolipoprotein E-knockout mice. Circulation 110:3465–71
    [Google Scholar]
  23. 23. 
    Park TS, Panek RL, Rekhter MD, Mueller SB, Rosebury WS et al. 2006. Modulation of lipoprotein metabolism by inhibition of sphingomyelin synthesis in ApoE knockout mice. Atherosclerosis 189:264–72
    [Google Scholar]
  24. 24. 
    Park TS, Rosebury W, Kindt EK, Kowala MC, Panek RL 2008. Serine palmitoyltransferase inhibitor myriocin induces the regression of atherosclerotic plaques in hyperlipidemic ApoE-deficient mice. Pharmacol. Res. 58:45–51
    [Google Scholar]
  25. 25. 
    Raichur S, Wang ST, Chan PW, Li Y, Ching J et al. 2014. CerS2 haploinsufficiency inhibits β-oxidation and confers susceptibility to diet-induced steatohepatitis and insulin resistance. Cell Metab 20:687–95
    [Google Scholar]
  26. 26. 
    Turpin SM, Nicholls HT, Willmes DM, Mourier A, Brodesser S et al. 2014. Obesity-induced CerS6-dependent C16:0 ceramide production promotes weight gain and glucose intolerance. Cell Metab 20:678–86
    [Google Scholar]
  27. 27. 
    Summers SA. 2018. Could ceramides become the new cholesterol?. Cell Metab 27:276–80
    [Google Scholar]
  28. 28. 
    Hannun YA, Obeid LM. 2018. Sphingolipids and their metabolism in physiology and disease. Nat. Rev. Mol. Cell Biol. 19:175–91
    [Google Scholar]
  29. 29. 
    Merrill AH Jr., Schmelz EM, Dillehay DL, Spiegel S, Shayman JA et al. 1997. Sphingolipids—the enigmatic lipid class: biochemistry, physiology, and pathophysiology. Toxicol. Appl. Pharmacol. 142:208–25
    [Google Scholar]
  30. 30. 
    Hannun YA, Obeid LM. 1995. Ceramide: an intracellular signal for apoptosis. Trends Biochem. Sci. 20:73–77
    [Google Scholar]
  31. 31. 
    Summers SA, Chaurasia B, Holland WL 2019. Metabolic messengers: ceramides. Nat. Metab. 1:1051–58
    [Google Scholar]
  32. 32. 
    Merrill AH Jr 2002. De novo sphingolipid biosynthesis: a necessary, but dangerous, pathway. J. Biol. Chem. 277:25843–46
    [Google Scholar]
  33. 33. 
    Han G, Gupta SD, Gable K, Niranjanakumari S, Moitra P et al. 2009. Identification of small subunits of mammalian serine palmitoyltransferase that confer distinct acyl-CoA substrate specificities. PNAS 106:8186–91
    [Google Scholar]
  34. 34. 
    Davis D, Kannan M, Wattenberg B 2018. Orm/ORMDL proteins: gate guardians and master regulators. Adv. Biol. Regul. 70:3–18
    [Google Scholar]
  35. 35. 
    Lone MA, Santos T, Alecu I, Silva LC, Hornemann T 2019. 1-Deoxysphingolipids. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 1864:512–21
    [Google Scholar]
  36. 36. 
    Bertea M, Rutti MF, Othman A, Marti-Jaun J, Hersberger M et al. 2010. Deoxysphingoid bases as plasma markers in diabetes mellitus. Lipids Health Dis 9:84
    [Google Scholar]
  37. 37. 
    Brozinick JT, Hawkins E, Hoang Bui H, Kuo MS, Tan B et al. 2013. Plasma sphingolipids are biomarkers of metabolic syndrome in non-human primates maintained on a Western-style diet. Int. J. Obes. 37:1064–70
    [Google Scholar]
  38. 38. 
    Gai Z, Gui T, Alecu I, Lone MA, Hornemann T et al. 2019. Farnesoid X receptor activation induces the degradation of hepatotoxic 1-deoxysphingolipids in non-alcoholic fatty liver disease. Liver Int 40:844–59
    [Google Scholar]
  39. 39. 
    Othman A, Bianchi R, Alecu I, Wei Y, Porretta-Serapiglia C et al. 2015. Lowering plasma 1-deoxysphingolipids improves neuropathy in diabetic rats. Diabetes 64:1035–45
    [Google Scholar]
  40. 40. 
    Othman A, Rutti MF, Ernst D, Saely CH, Rein P et al. 2012. Plasma deoxysphingolipids: a novel class of biomarkers for the metabolic syndrome. Diabetologia 55:421–31
    [Google Scholar]
  41. 41. 
    Othman A, Saely CH, Muendlein A, Vonbank A, Drexel H et al. 2015. Plasma 1-deoxysphingolipids are predictive biomarkers for type 2 diabetes mellitus. BMJ Open Diabetes Res. Care 3:e000073
    [Google Scholar]
  42. 42. 
    Zelnik ID, Rozman B, Rosenfeld-Gur E, Ben-Dor S, Futerman AH 2019. A stroll down the CerS lane. Adv. Exp. Med. Biol. 1159:49–63
    [Google Scholar]
  43. 43. 
    Sociale M, Wulf AL, Breiden B, Klee K, Thielisch M et al. 2018. Ceramide synthase schlank is a transcriptional regulator adapting gene expression to energy requirements. Cell Rep 22:967–78
    [Google Scholar]
  44. 44. 
    Siddique MM, Li Y, Chaurasia B, Kaddai VA, Summers SA 2015. Dihydroceramides: from bit players to lead actors. J. Biol. Chem. 290:15371–79
    [Google Scholar]
  45. 45. 
    Mizutani Y, Kihara A, Igarashi Y 2004. Identification of the human sphingolipid C4-hydroxylase, hDES2, and its up-regulation during keratinocyte differentiation. FEBS Lett 563:93–97
    [Google Scholar]
  46. 46. 
    Omae F, Miyazaki M, Enomoto A, Suzuki M, Suzuki Y, Suzuki A 2004. DES2 protein is responsible for phytoceramide biosynthesis in the mouse small intestine. Biochem. J. 379:687–95
    [Google Scholar]
  47. 47. 
    Senkal CE, Salama MF, Snider AJ, Allopenna JJ, Rana NA et al. 2017. Ceramide is metabolized to acylceramide and stored in lipid droplets. Cell Metab 25:686–97
    [Google Scholar]
  48. 48. 
    Coant N, Sakamoto W, Mao C, Hannun YA 2017. Ceramidases, roles in sphingolipid metabolism and in health and disease. Adv. Biol. Regul. 63:122–31
    [Google Scholar]
  49. 49. 
    Holland WL, Miller RA, Wang ZV, Sun K, Barth BM et al. 2011. Receptor-mediated activation of ceramidase activity initiates the pleiotropic actions of adiponectin. Nat. Med. 17:55–63
    [Google Scholar]
  50. 50. 
    Straub LG, Scherer PE. 2019. Metabolic messengers: adiponectin. Nat. Metab. 1:334–39
    [Google Scholar]
  51. 51. 
    Vasiliauskaite-Brooks I, Sounier R, Rochaix P, Bellot G, Fortier M et al. 2017. Structural insights into adiponectin receptors suggest ceramidase activity. Nature 544:120–23
    [Google Scholar]
  52. 52. 
    Cartier A, Hla T. 2019. Sphingosine 1-phosphate: lipid signaling in pathology and therapy. Science 366:eaar5551
    [Google Scholar]
  53. 53. 
    Boulgaropoulos B, Amenitsch H, Laggner P, Pabst G 2010. Implication of sphingomyelin/ceramide molar ratio on the biological activity of sphingomyelinase. Biophys. J. 99:499–506
    [Google Scholar]
  54. 54. 
    Claus RA, Dorer MJ, Bunck AC, Deigner HP 2009. Inhibition of sphingomyelin hydrolysis: targeting the lipid mediator ceramide as a key regulator of cellular fate. Curr. Med. Chem. 16:1978–2000
    [Google Scholar]
  55. 55. 
    Kitatani K, Idkowiak-Baldys J, Hannun YA 2008. The sphingolipid salvage pathway in ceramide metabolism and signaling. Cell Signal 20:1010–18
    [Google Scholar]
  56. 56. 
    Kitatani K, Sheldon K, Rajagopalan V, Anelli V, Jenkins RW et al. 2009. Involvement of acid β-glucosidase 1 in the salvage pathway of ceramide formation. J. Biol. Chem. 284:12972–78
    [Google Scholar]
  57. 57. 
    Jensen PN, Fretts AM, Yu C, Hoofnagle AN, Umans JG et al. 2019. Circulating sphingolipids, fasting glucose, and impaired fasting glucose: The Strong Heart Family Study. EBioMedicine 41:44–49
    [Google Scholar]
  58. 58. 
    Lemaitre RN, Yu C, Hoofnagle A, Hari N, Jensen PN et al. 2018. Circulating sphingolipids, insulin, HOMA-IR, and HOMA-B: The Strong Heart Family Study. Diabetes 67:1663–72
    [Google Scholar]
  59. 59. 
    Thorens B, Rodriguez A, Cruciani-Guglielmacci C, Wigger L, Ibberson M, Magnan C 2019. Use of preclinical models to identify markers of type 2 diabetes susceptibility and novel regulators of insulin secretion—a step towards precision medicine. Mol. Metab. 27S:S147–54
    [Google Scholar]
  60. 60. 
    Wigger L, Cruciani-Guglielmacci C, Nicolas A, Denom J, Fernandez N et al. 2017. Plasma dihydroceramides are diabetes susceptibility biomarker candidates in mice and humans. Cell Rep 18:2269–79
    [Google Scholar]
  61. 61. 
    de Carvalho LP, Tan SH, Ow GS, Tang Z, Ching J et al. 2018. Plasma ceramides as prognostic biomarkers and their arterial and myocardial tissue correlates in acute myocardial infarction. JACC Basic Transl. Sci. 3:163–75
    [Google Scholar]
  62. 62. 
    Hilvo M, Meikle PJ, Pedersen ER, Tell GS, Dhar I et al. 2019. Development and validation of a ceramide- and phospholipid-based cardiovascular risk estimation score for coronary artery disease patients. Eur. Heart J. 41:371–80
    [Google Scholar]
  63. 63. 
    Laaksonen R, Ekroos K, Sysi-Aho M, Hilvo M, Vihervaara T et al. 2016. Plasma ceramides predict cardiovascular death in patients with stable coronary artery disease and acute coronary syndromes beyond LDL-cholesterol. Eur. Heart J. 37:1967–76
    [Google Scholar]
  64. 64. 
    Peterson LR, Xanthakis V, Duncan MS, Gross S, Friedrich N et al. 2018. Ceramide remodeling and risk of cardiovascular events and mortality. J. Am. Heart Assoc. 7: https://doi.org/10.1161/JAHA.117.007931
    [Crossref] [Google Scholar]
  65. 65. 
    Havulinna AS, Sysi-Aho M, Hilvo M, Kauhanen D, Hurme R et al. 2016. Circulating ceramides predict cardiovascular outcomes in the population-based FINRISK 2002 cohort. Arterioscler. Thromb. Vasc. Biol. 36:2424–30
    [Google Scholar]
  66. 66. 
    Poss AM, Maschek JA, Cox JE, Hauner BJ, Hopkins PN et al. 2020. Machine learning reveals serum sphingolipids as cholesterol-independent biomarkers of coronary artery disease. J. Clin. Investig. 130:1363–76
    [Google Scholar]
  67. 67. 
    Huynh K, Barlow CK, Jayawardana KS, Weir JM, Mellett NA et al. 2019. High-throughput plasma lipidomics: detailed mapping of the associations with cardiometabolic risk factors. Cell Chem. Biol. 26:71–84.e4
    [Google Scholar]
  68. 68. 
    Luukkonen PK, Sadevirta S, Zhou Y, Kayser B, Ali A et al. 2018. Saturated fat is more metabolically harmful for the human liver than unsaturated fat or simple sugars. Diabetes Care 41:1732–39
    [Google Scholar]
  69. 69. 
    Luukkonen PK, Zhou Y, Sadevirta S, Leivonen M, Arola J et al. 2016. Hepatic ceramides dissociate steatosis and insulin resistance in patients with non-alcoholic fatty liver disease. J. Hepatol. 64:1167–75
    [Google Scholar]
  70. 70. 
    Kolak M, Westerbacka J, Velagapudi VR, Wagsater D, Yetukuri L et al. 2007. Adipose tissue inflammation and increased ceramide content characterize subjects with high liver fat content independent of obesity. Diabetes 56:1960–68
    [Google Scholar]
  71. 71. 
    Coen PM, Dube JJ, Amati F, Stefanovic-Racic M, Ferrell RE et al. 2010. Insulin resistance is associated with higher intramyocellular triglycerides in type I but not type II myocytes concomitant with higher ceramide content. Diabetes 59:80–88
    [Google Scholar]
  72. 72. 
    Coen PM, Goodpaster BH. 2012. Role of intramyocelluar lipids in human health. Trends Endocrinol. Metab. 23:391–98
    [Google Scholar]
  73. 73. 
    Coen PM, Hames KC, Leachman EM, DeLany JP, Ritov VB et al. 2013. Reduced skeletal muscle oxidative capacity and elevated ceramide but not diacylglycerol content in severe obesity. Obesity 21:2362–71
    [Google Scholar]
  74. 74. 
    Dube JJ, Amati F, Toledo FG, Stefanovic-Racic M, Rossi A et al. 2011. Effects of weight loss and exercise on insulin resistance, and intramyocellular triacylglycerol, diacylglycerol and ceramide. Diabetologia 54:1147–56
    [Google Scholar]
  75. 75. 
    Adams JM 2nd, Pratipanawatr T, Berria R, Wang E, DeFronzo RA et al. 2004. Ceramide content is increased in skeletal muscle from obese insulin-resistant humans. Diabetes 53:25–31
    [Google Scholar]
  76. 76. 
    Mantovani A, Bonapace S, Lunardi G, Canali G, Dugo C et al. 2020. Associations between specific plasma ceramides and severity of coronary-artery stenosis assessed by coronary angiography. Diabetes Metab 46:150–57
    [Google Scholar]
  77. 77. 
    Karjalainen JP, Mononen N, Hutri-Kahonen N, Lehtimaki M, Hilvo M et al. 2019. New evidence from plasma ceramides links apoE polymorphism to greater risk of coronary artery disease in Finnish adults. J. Lipid Res. 60:1622–29
    [Google Scholar]
  78. 78. 
    Hilvo M, Meikle PJ, Pedersen ER, Tell GS, Dhar I et al. 2020. Development and validation of a ceramide- and phospholipid-based cardiovascular risk estimation score for coronary artery disease patients. Eur. Heart J. 41:371–80
    [Google Scholar]
  79. 79. 
    Anroedh S, Hilvo M, Akkerhuis KM, Kauhanen D, Koistinen K et al. 2018. Plasma concentrations of molecular lipid species predict long-term clinical outcome in coronary artery disease patients. J. Lipid Res. 59:1729–37
    [Google Scholar]
  80. 80. 
    Cheng JM, Suoniemi M, Kardys I, Vihervaara T, de Boer SP et al. 2015. Plasma concentrations of molecular lipid species in relation to coronary plaque characteristics and cardiovascular outcome: results of the ATHEROREMO-IVUS study. Atherosclerosis 243:560–66
    [Google Scholar]
  81. 81. 
    Hilvo M, Simolin H, Metso J, Ruuth M, Oorni K et al. 2018. PCSK9 inhibition alters the lipidome of plasma and lipoprotein fractions. Atherosclerosis 269:159–65
    [Google Scholar]
  82. 82. 
    Mantovani A, Bonapace S, Lunardi G, Salgarello M, Dugo C et al. 2018. Association of plasma ceramides with myocardial perfusion in patients with coronary artery disease undergoing stress myocardial perfusion scintigraphy. Arterioscler. Thromb. Vasc. Biol. 38:2854–61
    [Google Scholar]
  83. 83. 
    Pan W, Yu J, Shi R, Yan L, Yang T et al. 2014. Elevation of ceramide and activation of secretory acid sphingomyelinase in patients with acute coronary syndromes. Coron. Artery Dis. 25:230–35
    [Google Scholar]
  84. 84. 
    Tarasov K, Ekroos K, Suoniemi M, Kauhanen D, Sylvanne T et al. 2014. Molecular lipids identify cardiovascular risk and are efficiently lowered by simvastatin and PCSK9 deficiency. J. Clin. Endocrinol. Metab. 99:E45–52
    [Google Scholar]
  85. 85. 
    Ng TW, Ooi EM, Watts GF, Chan DC, Meikle PJ, Barrett PH 2015. Association of plasma ceramides and sphingomyelin with VLDL apoB-100 fractional catabolic rate before and after rosuvastatin treatment. J. Clin. Endocrinol. Metab. 100:2497–501
    [Google Scholar]
  86. 86. 
    Mayo Clinic Laboratories 2016. Ceramides, plasma [a test in focus]. Mayo Clinic Laboratories July 28. https://news.mayocliniclabs.com/2016/07/28/ceramides-plasma-a-test-in-focus/
    [Google Scholar]
  87. 87. 
    Chokshi A, Drosatos K, Cheema FH, Ji R, Khawaja T et al. 2012. Ventricular assist device implantation corrects myocardial lipotoxicity, reverses insulin resistance, and normalizes cardiac metabolism in patients with advanced heart failure. Circulation 125:2844–53
    [Google Scholar]
  88. 88. 
    Chaurasia B, Kaddai VA, Lancaster GI, Henstridge DC, Sriram S et al. 2016. Adipocyte ceramides regulate subcutaneous adipose browning, inflammation, and metabolism. Cell Metab 24:820–34
    [Google Scholar]
  89. 89. 
    Holland WL, Bikman BT, Wang LP, Yuguang G, Sargent KM et al. 2011. Lipid-induced insulin resistance mediated by the proinflammatory receptor TLR4 requires saturated fatty acid-induced ceramide biosynthesis in mice. J. Clin. Investig. 121:1858–70
    [Google Scholar]
  90. 90. 
    Zhang QJ, Holland WL, Wilson L, Tanner JM, Kearns D et al. 2012. Ceramide mediates vascular dysfunction in diet-induced obesity by PP2A-mediated dephosphorylation of the eNOS-Akt complex. Diabetes 61:1848–59
    [Google Scholar]
  91. 91. 
    Ussher JR, Koves TR, Cadete VJ, Zhang L, Jaswal JS et al. 2010. Inhibition of de novo ceramide synthesis reverses diet-induced insulin resistance and enhances whole-body oxygen consumption. Diabetes 59:2453–64
    [Google Scholar]
  92. 92. 
    Jiang M, Li C, Liu Q, Wang A, Lei M 2019. Inhibiting ceramide synthesis attenuates hepatic steatosis and fibrosis in rats with non-alcoholic fatty liver disease. Front. Endocrinol. 10:665
    [Google Scholar]
  93. 93. 
    Dekker MJ, Baker C, Naples M, Samsoondar J, Zhang R et al. 2013. Inhibition of sphingolipid synthesis improves dyslipidemia in the diet-induced hamster model of insulin resistance: evidence for the role of sphingosine and sphinganine in hepatic VLDL-apoB100 overproduction. Atherosclerosis 228:98–109
    [Google Scholar]
  94. 94. 
    Shimabukuro M, Zhou YT, Levi M, Unger RH 1998. Fatty acid-induced β cell apoptosis: a link between obesity and diabetes. PNAS 95:2498–502
    [Google Scholar]
  95. 95. 
    Yang Q, Graham TE, Mody N, Preitner F, Peroni OD et al. 2005. Serum retinol binding protein 4 contributes to insulin resistance in obesity and type 2 diabetes. Nature 436:356–62
    [Google Scholar]
  96. 96. 
    Bikman BT, Guan Y, Shui G, Siddique MM, Holland WL et al. 2012. Fenretinide prevents lipid-induced insulin resistance by blocking ceramide biosynthesis. J. Biol. Chem. 287:17426–37
    [Google Scholar]
  97. 97. 
    Preitner F, Mody N, Graham TE, Peroni OD, Kahn BB 2009. Long-term Fenretinide treatment prevents high-fat diet-induced obesity, insulin resistance, and hepatic steatosis. Am. J. Physiol. Endocrinol. Metab. 297:E1420–29
    [Google Scholar]
  98. 98. 
    Hammerschmidt P, Ostkotte D, Nolte H, Gerl MJ, Jais A et al. 2019. CerS6-derived sphingolipids interact with Mff and promote mitochondrial fragmentation in obesity. Cell 177:1536–52.e23
    [Google Scholar]
  99. 99. 
    Turpin-Nolan SM, Hammerschmidt P, Chen W, Jais A, Timper K et al. 2019. CerS1-derived C18:0 ceramide in skeletal muscle promotes obesity-induced insulin resistance. Cell Rep 26:1–10.e7
    [Google Scholar]
  100. 100. 
    Xia JY, Holland WL, Kusminski CM, Sun K, Sharma AX et al. 2015. Targeted induction of ceramide degradation leads to improved systemic metabolism and reduced hepatic steatosis. Cell Metab 22:266–78
    [Google Scholar]
  101. 101. 
    Chakraborty M, Lou C, Huan C, Kuo MS, Park TS et al. 2013. Myeloid cell-specific serine palmitoyltransferase subunit 2 haploinsufficiency reduces murine atherosclerosis. J. Clin. Investig. 123:1784–97
    [Google Scholar]
  102. 102. 
    Bharath LP, Ruan T, Li Y, Ravindran A, Wan X et al. 2015. Ceramide-initiated protein phosphatase 2a activation contributes to arterial dysfunction in vivo. Diabetes 64:3914–26
    [Google Scholar]
  103. 103. 
    Reforgiato MR, Milano G, Fabrias G, Casas J, Gasco P et al. 2016. Inhibition of ceramide de novo synthesis as a postischemic strategy to reduce myocardial reperfusion injury. Basic Res. Cardiol. 111:12
    [Google Scholar]
  104. 104. 
    Russo SB, Baicu CF, Van Laer A, Geng T, Kasiganesan H et al. 2012. Ceramide synthase 5 mediates lipid-induced autophagy and hypertrophy in cardiomyocytes. J. Clin. Investig. 122:3919–30
    [Google Scholar]
  105. 105. 
    Raichur S, Brunner B, Bielohuby M, Hansen G, Pfenninger A et al. 2019. The role of C16:0 ceramide in the development of obesity and type 2 diabetes: CerS6 inhibition as a novel therapeutic approach. Mol. Metab. 21:36–50
    [Google Scholar]
  106. 106. 
    Gosejacob D, Jäger PS, Vom Dorp K, Frejno M, Carstensen AC et al. 2016. Ceramide synthase 5 is essential to maintain C16:0-ceramide pools and contributes to the development of diet-induced obesity. J. Biol. Chem. 291:6989–7003
    [Google Scholar]
  107. 107. 
    Turner N, Lim XY, Toop HD, Osborne B, Brandon AE et al. 2018. A selective inhibitor of ceramide synthase 1 reveals a novel role in fat metabolism. Nat. Commun. 9:3165
    [Google Scholar]
  108. 108. 
    Li Z, Zhang H, Liu J, Liang CP, Li Y et al. 2011. Reducing plasma membrane sphingomyelin increases insulin sensitivity. Mol. Cell. Biol. 31:4205–18
    [Google Scholar]
  109. 109. 
    Park M, Kaddai V, Ching J, Fridianto KT, Sieli RJ et al. 2016. A role for ceramides, but not sphingomyelins, as antagonists of insulin signaling and mitochondrial metabolism in C2C12 myotubes. J. Biol. Chem. 291:23978–88
    [Google Scholar]
  110. 110. 
    Bijl N, Sokolovic M, Vrins C, Langeveld M, Moerland PD et al. 2009. Modulation of glycosphingolipid metabolism significantly improves hepatic insulin sensitivity and reverses hepatic steatosis in mice. Hepatology 50:1431–41
    [Google Scholar]
  111. 111. 
    Zhao H, Przybylska M, Wu IH, Zhang J, Maniatis P et al. 2009. Inhibiting glycosphingolipid synthesis ameliorates hepatic steatosis in obese mice. Hepatology 50:85–93
    [Google Scholar]
  112. 112. 
    Zhao H, Przybylska M, Wu IH, Zhang J, Siegel C et al. 2007. Inhibiting glycosphingolipid synthesis improves glycemic control and insulin sensitivity in animal models of type 2 diabetes. Diabetes 56:1210–18
    [Google Scholar]
  113. 113. 
    Chatterjee S, Bedja D, Mishra S, Amuzie C, Avolio A et al. 2014. Inhibition of glycosphingolipid synthesis ameliorates atherosclerosis and arterial stiffness in apolipoprotein E−/− mice and rabbits fed a high-fat and -cholesterol diet. Circulation 129:2403–13
    [Google Scholar]
  114. 114. 
    Bietrix F, Lombardo E, van Roomen CP, Ottenhoff R, Vos M et al. 2010. Inhibition of glycosphingolipid synthesis induces a profound reduction of plasma cholesterol and inhibits atherosclerosis development in APOE*3 Leiden and low-density lipoprotein receptor−/− mice. Arterioscler. Thromb. Vasc. Biol. 30:931–37
    [Google Scholar]
  115. 115. 
    Glaros EN, Kim WS, Rye KA, Shayman JA, Garner B 2008. Reduction of plasma glycosphingolipid levels has no impact on atherosclerosis in apolipoprotein E-null mice. J. Lipid Res. 49:1677–81
    [Google Scholar]
  116. 116. 
    Kabayama K, Sato T, Saito K, Loberto N, Prinetti A et al. 2007. Dissociation of the insulin receptor and caveolin-1 complex by ganglioside GM3 in the state of insulin resistance. PNAS 104:13678–83
    [Google Scholar]
  117. 117. 
    Inokuchi J. 2007. Insulin resistance as a membrane microdomain disorder. Yakugaku Zasshi 127:579–86
    [Google Scholar]
  118. 118. 
    Yamashita T, Hashiramoto A, Haluzik M, Mizukami H, Beck S et al. 2003. Enhanced insulin sensitivity in mice lacking ganglioside GM3. PNAS 100:3445–49
    [Google Scholar]
  119. 119. 
    Chavez JA, Siddique MM, Wang ST, Ching J, Shayman JA, Summers SA 2014. Ceramides and glucosylceramides are independent antagonists of insulin signaling. J. Biol. Chem. 289:723–34
    [Google Scholar]
  120. 120. 
    Gudz TI, Tserng KY, Hoppel CL 1997. Direct inhibition of mitochondrial respiratory chain complex III by cell-permeable ceramide. J. Biol. Chem. 272:24154–48
    [Google Scholar]
  121. 121. 
    Di Paola M, Cocco T, Lorusso M 2000. Ceramide interaction with the respiratory chain of heart mitochondria. Biochemistry 39:6660–68
    [Google Scholar]
  122. 122. 
    Zigdon H, Kogot-Levin A, Park JW, Goldschmidt R, Kelly S et al. 2013. Ablation of ceramide synthase 2 causes chronic oxidative stress due to disruption of the mitochondrial respiratory chain. J. Biol. Chem. 288:4947–56
    [Google Scholar]
  123. 123. 
    Obeid LM, Linardic CM, Karolak LA, Hannun YA 1993. Programmed cell death induced by ceramide. Science 259:1769–71
    [Google Scholar]
  124. 124. 
    Ganesan V, Perera MN, Colombini D, Datskovskiy D, Chadha K, Colombini M 2010. Ceramide and activated Bax act synergistically to permeabilize the mitochondrial outer membrane. Apoptosis 15:553–62
    [Google Scholar]
  125. 125. 
    Dadsena S, Bockelmann S, Mina JGM, Hassan DG, Korneev S et al. 2019. Ceramides bind VDAC2 to trigger mitochondrial apoptosis. Nat. Commun. 10:1832
    [Google Scholar]
  126. 126. 
    Siskind LJ, Kolesnick RN, Colombini M 2006. Ceramide forms channels in mitochondrial outer membranes at physiologically relevant concentrations. Mitochondrion 6:118–25
    [Google Scholar]
  127. 127. 
    Siskind LJ, Davoody A, Lewin N, Marshall S, Colombini M 2003. Enlargement and contracture of C2-ceramide channels. Biophys. J. 85:1560–75
    [Google Scholar]
  128. 128. 
    Siskind LJ, Kolesnick RN, Colombini M 2002. Ceramide channels increase the permeability of the mitochondrial outer membrane to small proteins. J. Biol. Chem. 277:26796–803
    [Google Scholar]
  129. 129. 
    Siskind LJ, Colombini M. 2000. The lipids C2- and C16-ceramide form large stable channels. Implications for apoptosis. J. Biol. Chem. 275:38640–44
    [Google Scholar]
  130. 130. 
    Salinas M, Lopez-Valdaliso R, Martin D, Alvarez A, Cuadrado A 2000. Inhibition of PKB/Akt1 by C2-ceramide involves activation of ceramide-activated protein phosphatase in PC12 cells. Mol. Cell. Neurosci. 15:156–69
    [Google Scholar]
  131. 131. 
    Teruel T, Hernandez R, Lorenzo M 2001. Ceramide mediates insulin resistance by tumor necrosis factor-α in brown adipocytes by maintaining Akt in an inactive dephosphorylated state. Diabetes 50:2563–71
    [Google Scholar]
  132. 132. 
    Zinda MJ, Vlahos CJ, Lai MT 2001. Ceramide induces the dephosphorylation and inhibition of constitutively activated Akt in PTEN negative U87mg cells. Biochem. Biophys. Res. Commun. 280:1107–15
    [Google Scholar]
  133. 133. 
    Dobrowsky RT, Kamibayashi C, Mumby MC, Hannun YA 1993. Ceramide activates heterotrimeric protein phosphatase 2A. J. Biol. Chem. 268:15523–30
    [Google Scholar]
  134. 134. 
    Powell DJ, Hajduch E, Kular G, Hundal HS 2003. Ceramide disables 3-phosphoinositide binding to the pleckstrin homology domain of protein kinase B (PKB)/Akt by a PKCζ-dependent mechanism. Mol. Cell. Biol. 23:7794–808
    [Google Scholar]
  135. 135. 
    Powell DJ, Turban S, Gray A, Hajduch E, Hundal HS 2004. Intracellular ceramide synthesis and protein kinase Cζ activation play an essential role in palmitate-induced insulin resistance in rat L6 skeletal muscle cells. Biochem. J. 382:619–29
    [Google Scholar]
  136. 136. 
    Blouin CM, Prado C, Takane KK, Lasnier F, Garcia-Ocana A et al. 2010. Plasma membrane subdomain compartmentalization contributes to distinct mechanisms of ceramide action on insulin signaling. Diabetes 59:600–10
    [Google Scholar]
  137. 137. 
    Stratford S, Hoehn KL, Liu F, Summers SA 2004. Regulation of insulin action by ceramide: dual mechanisms linking ceramide accumulation to the inhibition of Akt/protein kinase B. J. Biol. Chem. 279:36608–15
    [Google Scholar]
  138. 138. 
    Jiang C, Xie C, Li F, Zhang L, Nichols RG et al. 2015. Intestinal farnesoid X receptor signaling promotes nonalcoholic fatty liver disease. J. Clin. Investig. 125:386–402
    [Google Scholar]
  139. 139. 
    Alexaki A, Clarke BA, Gavrilova O, Ma Y, Zhu H et al. 2017. De novo sphingolipid biosynthesis is required for adipocyte survival and metabolic homeostasis. J. Biol. Chem. 292:3929–39
    [Google Scholar]
  140. 140. 
    Lee SY, Lee HY, Song JH, Kim GT, Jeon S et al. 2017. Adipocyte-specific deficiency of de novo sphingolipid biosynthesis leads to lipodystrophy and insulin resistance. Diabetes 66:2596–609
    [Google Scholar]
  141. 141. 
    Eguchi J, Wang X, Yu S, Kershaw EE, Chiu PC et al. 2011. Transcriptional control of adipose lipid handling by IRF4. Cell Metab 13:249–59
    [Google Scholar]
  142. 142. 
    Wang ZV, Deng Y, Wang QA, Sun K, Scherer PE 2010. Identification and characterization of a promoter cassette conferring adipocyte-specific gene expression. Endocrinology 151:2933–39
    [Google Scholar]
  143. 143. 
    Brown MS, Goldstein JL. 2008. Selective versus total insulin resistance: a pathogenic paradox. Cell Metab 7:95–96
    [Google Scholar]
  144. 144. 
    Petersen MC, Jurczak MJ. 2016. CrossTalk opposing view: intramyocellular ceramide accumulation does not modulate insulin resistance. J. Physiol. 594:3171–74
    [Google Scholar]
  145. 145. 
    Summers SA, Goodpaster BH. 2016. CrossTalk proposal: intramyocellular ceramide accumulation does modulate insulin resistance. J. Physiol. 594:3167–70
    [Google Scholar]
  146. 146. 
    Hu W, Ross J, Geng T, Brice SE, Cowart LA 2011. Differential regulation of dihydroceramide desaturase by palmitate versus monounsaturated fatty acids: implications for insulin resistance. J. Biol. Chem. 286:16596–605
    [Google Scholar]
  147. 147. 
    Watson ML, Coghlan M, Hundal HS 2009. Modulating serine palmitoyl transferase (SPT) expression and activity unveils a crucial role in lipid-induced insulin resistance in rat skeletal muscle cells. Biochem. J. 417:791–801
    [Google Scholar]
  148. 148. 
    Boon J, Hoy AJ, Stark R, Brown RD, Meex RC et al. 2013. Ceramides contained in LDL are elevated in type 2 diabetes and promote inflammation and skeletal muscle insulin resistance. Diabetes 62:401–10
    [Google Scholar]
  149. 149. 
    Shimabukuro M, Higa M, Zhou YT, Wang MY, Newgard CB, Unger RH 1998. Lipoapoptosis in beta-cells of obese prediabetic fa/fa rats. Role of serine palmitoyltransferase overexpression. J. Biol. Chem. 273:32487–90
    [Google Scholar]
  150. 150. 
    Boslem E, Meikle PJ, Biden TJ 2012. Roles of ceramide and sphingolipids in pancreatic β-cell function and dysfunction. Islets 4:177–87
    [Google Scholar]
  151. 151. 
    Kelpe CL, Moore PC, Parazzoli SD, Wicksteed B, Rhodes CJ, Poitout V 2003. Palmitate inhibition of insulin gene expression is mediated at the transcriptional level via ceramide synthesis. J. Biol. Chem. 278:30015–21
    [Google Scholar]
  152. 152. 
    Sjöholm A. 1995. Ceramide inhibits pancreatic β-cell insulin production and mitogenesis and mimics the actions of interleukin-1β. FEBS Lett 367:283–86
    [Google Scholar]
  153. 153. 
    Veret J, Bellini L, Giussani P, Ng C, Magnan C, Le Stunff H 2014. Roles of sphingolipid metabolism in pancreatic β cell dysfunction induced by lipotoxicity. J. Clin. Med. 3:646–62
    [Google Scholar]
  154. 154. 
    Zhu Q, Shan X, Miao H, Lu Y, Xu J et al. 2009. Acute activation of acid ceramidase affects cytokine-induced cytotoxicity in rat islet β-cells. FEBS Lett 583:2136–41
    [Google Scholar]
  155. 155. 
    El-Assaad W, Buteau J, Peyot ML, Nolan C, Roduit R et al. 2003. Saturated fatty acids synergize with elevated glucose to cause pancreatic beta-cell death. Endocrinology 144:4154–63
    [Google Scholar]
  156. 156. 
    Oh YS, Bae GD, Baek DJ, Park EY, Jun HS 2018. Fatty acid-induced lipotoxicity in pancreatic beta-cells during development of type 2 diabetes. Front. Endocrinol. 9:384
    [Google Scholar]
  157. 157. 
    Tian G, Sol ER, Xu Y, Shuai H, Tengholm A 2015. Impaired cAMP generation contributes to defective glucose-stimulated insulin secretion after long-term exposure to palmitate. Diabetes 64:904–15
    [Google Scholar]
  158. 158. 
    Lucchetta M, Rudilosso S, Costa S, Bruttomesso D, Ruggero S et al. 2010. Anti-ganglioside autoantibodies in type 1 diabetes. Muscle Nerve 41:50–53
    [Google Scholar]
  159. 159. 
    Dotta F, Falorni A, Tiberti C, Dionisi S, Anastasi E et al. 1997. Autoantibodies to the GM2–1 islet ganglioside and to GAD-65 at type 1 diabetes onset. J. Autoimmun. 10:585–88
    [Google Scholar]
  160. 160. 
    Misasi R, Dionisi S, Farilla L, Carabba B, Lenti L et al. 1997. Gangliosides and autoimmune diabetes. Diabetes Metab. Rev. 13:163–79
    [Google Scholar]
  161. 161. 
    Dionisi S, Dotta F, Diaz-Horta O, Carabba B, Viglietta V, Di Mario U 1997. Target antigens in autoimmune diabetes: pancreatic gangliosides. Ann. Inst. Super. Sanita 33:433–35
    [Google Scholar]
  162. 162. 
    Park SK, La Salle DT, Cerbie J, Cho JM, Bledsoe A et al. 2019. Elevated arterial shear rate increases indexes of endothelial cell autophagy and nitric oxide synthase activation in humans. Am. J. Physiol. Heart Circ. Physiol. 316:H106–12
    [Google Scholar]
  163. 163. 
    Li W, Yang X, Xing S, Bian F, Yao W et al. 2014. Endogenous ceramide contributes to the transcytosis of oxLDL across endothelial cells and promotes its subendothelial retention in vascular wall. Oxid. Med. Cell. Longev. 2014:823071
    [Google Scholar]
  164. 164. 
    Gao D, Pararasa C, Dunston CR, Bailey CJ, Griffiths HR 2012. Palmitate promotes monocyte atherogenicity via de novo ceramide synthesis. Free Radic. Biol. Med. 53:796–806
    [Google Scholar]
  165. 165. 
    Contreras C, Gonzalez-Garcia I, Martinez-Sanchez N, Seoane-Collazo P, Jacas J et al. 2014. Central ceramide-induced hypothalamic lipotoxicity and ER stress regulate energy balance. Cell Rep 9:366–77
    [Google Scholar]
  166. 166. 
    Campana M, Bellini L, Rouch C, Rachdi L, Coant N et al. 2018. Inhibition of central de novo ceramide synthesis restores insulin signaling in hypothalamus and enhances β-cell function of obese Zucker rats. Mol. Metab. 8:23–36
    [Google Scholar]
  167. 167. 
    Nordstrom V, Willershäuser M, Herzer S, Rozman J, von Bohlen und Halbach O et al. 2013. Neuronal expression of glucosylceramide synthase in central nervous system regulates body weight and energy homeostasis. PLOS Biol 11:e1001506
    [Google Scholar]
  168. 168. 
    Herzer S, Meldner S, Grone HJ, Nordström V 2015. Fasting-induced lipolysis and hypothalamic insulin signaling are regulated by neuronal glucosylceramide synthase. Diabetes 64:3363–76
    [Google Scholar]
  169. 169. 
    Chaurasia B, Summers SA. 2015. Ceramides—lipotoxic inducers of metabolic disorders. Trends Endocrinol. Metab. 26:538–50
    [Google Scholar]
  170. 170. 
    Stockli J, Fisher-Wellman KH, Chaudhuri R, Zeng XY, Fazakerley DJ et al. 2017. Metabolomic analysis of insulin resistance across different mouse strains and diets. J. Biol. Chem. 292:19135–45
    [Google Scholar]
  171. 171. 
    Hotamisligil GS. 2006. Inflammation and metabolic disorders. Nature 444:860–67
    [Google Scholar]
  172. 172. 
    de Mello VD, Lankinen M, Schwab U, Kolehmainen M, Lehto S et al. 2009. Link between plasma ceramides, inflammation and insulin resistance: association with serum IL-6 concentration in patients with coronary heart disease. Diabetologia 52:2612–15
    [Google Scholar]
  173. 173. 
    Schilling JD, Machkovech HM, He L, Sidhu R, Fujiwara H et al. 2013. Palmitate and lipopolysaccharide trigger synergistic ceramide production in primary macrophages. J. Biol. Chem. 288:2923–32
    [Google Scholar]
  174. 174. 
    Majumdar I, Mastrandrea LD. 2012. Serum sphingolipids and inflammatory mediators in adolescents at risk for metabolic syndrome. Endocrine 41:442–49
    [Google Scholar]
  175. 175. 
    Shi H, Kokoeva MV, Inouye K, Tzameli I, Yin H, Flier JS 2006. TLR4 links innate immunity and fatty acid-induced insulin resistance. J. Clin. Investig. 116:3015–25
    [Google Scholar]
  176. 176. 
    Lancaster GI, Langley KG, Berglund NA, Kammoun HL, Reibe S et al. 2018. Evidence that TLR4 is not a receptor for saturated fatty acids but mediates lipid-induced inflammation by reprogramming macrophage metabolism. Cell Metab 27:1096–110.e5
    [Google Scholar]
  177. 177. 
    Sims K, Haynes CA, Kelly S, Allegood JC, Wang E et al. 2010. Kdo2-lipid A, a TLR4-specific agonist, induces de novo sphingolipid biosynthesis in RAW264.7 macrophages, which is essential for induction of autophagy. J. Biol. Chem. 285:38568–79
    [Google Scholar]
  178. 178. 
    Kolesnick R. 1992. Ceramide: a novel second messenger. Trends Cell Biol 2:232–36
    [Google Scholar]
  179. 179. 
    Vandanmagsar B, Youm YH, Ravussin A, Galgani JE, Stadler K et al. 2011. The NLRP3 inflammasome instigates obesity-induced inflammation and insulin resistance. Nat. Med. 17:179–88
    [Google Scholar]
  180. 180. 
    Guzman M, Galve-Roperh I, Sanchez C 2001. Ceramide: a new second messenger of cannabinoid action. Trends Pharmacol. Sci. 22:19–22
    [Google Scholar]
  181. 181. 
    Velasco G, Galve-Roperh I, Sanchez C, Blazquez C, Haro A, Guzman M 2005. Cannabinoids and ceramide: two lipids acting hand-by-hand. Life Sci 77:1723–31
    [Google Scholar]
  182. 182. 
    Gatta-Cherifi B, Cota D. 2016. New insights on the role of the endocannabinoid system in the regulation of energy balance. Int. J. Obes. 40:210–19
    [Google Scholar]
  183. 183. 
    Cinar R, Godlewski G, Liu J, Tam J, Jourdan T et al. 2014. Hepatic cannabinoid-1 receptors mediate diet-induced insulin resistance by increasing de novo synthesis of long-chain ceramides. Hepatology 59:143–53
    [Google Scholar]
  184. 184. 
    Ikeda K, Maretich P, Kajimura S 2018. The common and distinct features of brown and beige adipocytes. Trends Endocrinol. Metab. 29:191–200
    [Google Scholar]
  185. 185. 
    Jiang C, Xie C, Lv Y, Li J, Krausz KW et al. 2015. Intestine-selective farnesoid X receptor inhibition improves obesity-related metabolic dysfunction. Nat. Commun. 6:10166
    [Google Scholar]
  186. 186. 
    Yamauchi T, Kamon J, Minokoshi Y, Ito Y, Waki H et al. 2002. Adiponectin stimulates glucose utilization and fatty-acid oxidation by activating AMP-activated protein kinase. Nat. Med. 8:1288–95
    [Google Scholar]
  187. 187. 
    Villa NY, Kupchak BR, Garitaonandia I, Smith JL, Alonso E et al. 2009. Sphingolipids function as downstream effectors of a fungal PAQR. Mol. Pharmacol. 75:866–75
    [Google Scholar]
  188. 188. 
    Kupchak BR, Garitaonandia I, Villa NY, Smith JL, Lyons TJ 2009. Antagonism of human adiponectin receptors and their membrane progesterone receptor paralogs by TNFα and a ceramidase inhibitor. Biochemistry 48:5504–6
    [Google Scholar]
  189. 189. 
    Holland WL, Xia JY, Johnson JA, Sun K, Pearson MJ et al. 2017. Inducible overexpression of adiponectin receptors highlight the roles of adiponectin-induced ceramidase signaling in lipid and glucose homeostasis. Mol. Metab. 6:267–75
    [Google Scholar]
  190. 190. 
    Tanabe H, Fujii Y, Okada-Iwabu M, Iwabu M, Nakamura Y et al. 2015. Crystal structures of the human adiponectin receptors. Nature 520:312–16
    [Google Scholar]
  191. 191. 
    Xi Y, Li H. 2020. Role of farnesoid X receptor in hepatic steatosis in nonalcoholic fatty liver disease. Biomed. Pharmacother. 121:109609
    [Google Scholar]
  192. 192. 
    Xie C, Jiang C, Shi J, Gao X, Sun D et al. 2017. An intestinal farnesoid X receptor-ceramide signaling axis modulates hepatic gluconeogenesis in mice. Diabetes 66:613–26
    [Google Scholar]
  193. 193. 
    Wang CN, O'Brien L, Brindley DN 1998. Effects of cell-permeable ceramides and tumor necrosis factor-α on insulin signaling and glucose uptake in 3T3-L1 adipocytes. Diabetes 47:24–31
    [Google Scholar]
  194. 194. 
    Hyde R, Hajduch E, Powell DJ, Taylor PM, Hundal HS 2005. Ceramide down-regulates System A amino acid transport and protein synthesis in rat skeletal muscle cells. FASEB J 19:461–63
    [Google Scholar]
  195. 195. 
    Guenther GG, Peralta ER, Rosales KR, Wong SY, Siskind LJ, Edinger AL 2008. Ceramide starves cells to death by downregulating nutrient transporter proteins. PNAS 105:17402–7
    [Google Scholar]
  196. 196. 
    Kogot-Levin A, Saada A. 2014. Ceramide and the mitochondrial respiratory chain. Biochimie 100:88–94
    [Google Scholar]
  197. 197. 
    Chaurasia B, Holland WL, Summers SA 2018. Does this schlank make me look fat. Trends Endocrinol. Metab. 29:597–99
    [Google Scholar]
  198. 198. 
    Shiffman D, Pare G, Oberbauer R, Louie JZ, Rowland CM et al. 2014. A gene variant in CERS2 is associated with rate of increase in albuminuria in patients with diabetes from ONTARGET and TRANSCEND. PLOS ONE 9:e106631
    [Google Scholar]
  199. 199. 
    Köttgen A, Pattaro C, Böger CA, Fuchsberger C, Olden M et al. 2010. New loci associated with kidney function and chronic kidney disease. Nat. Genet. 42:376–84
    [Google Scholar]
  200. 200. 
    Price PM, Hirschhorn K, Safirstein RL 2010. Chronic kidney disease and GWAS: “the proper study of mankind is man.”. Cell Metab 11:451–52
    [Google Scholar]
  201. 201. 
    Hojjati MR, Li Z, Jiang XC 2005. Serine palmitoyl-CoA transferase (SPT) deficiency and sphingolipid levels in mice. Biochim. Biophys. Acta 1737:44–51
    [Google Scholar]
  202. 202. 
    Ohta E, Ohira T, Matsue K, Ikeda Y, Fujii K et al. 2009. Analysis of development of lesions in mice with serine palmitoyltransferase (SPT) deficiency—Sptlc2 conditional knockout mice. Exp. Anim. 58:515–24
    [Google Scholar]
  203. 203. 
    Li Z, Kabir I, Tietelman G, Huan C, Fan J et al. 2018. Sphingolipid de novo biosynthesis is essential for intestine cell survival and barrier function. Cell Death Dis 9:173
    [Google Scholar]
  204. 204. 
    Jennemann R, Rabionet M, Gorgas K, Epstein S, Dalpke A et al. 2012. Loss of ceramide synthase 3 causes lethal skin barrier disruption. Hum. Mol. Genet. 21:586–608
    [Google Scholar]
  205. 205. 
    Ebel P, Imgrund S, Vom Dorp K, Hofmann K, Maier H et al. 2014. Ceramide synthase 4 deficiency in mice causes lipid alterations in sebum and results in alopecia. Biochem. J. 461:147–58
    [Google Scholar]
  206. 206. 
    Jennemann R, Kaden S, Sandhoff R, Nordstrom V, Wang S et al. 2012. Glycosphingolipids are essential for intestinal endocytic function. J. Biol. Chem. 287:32598–616
    [Google Scholar]
/content/journals/10.1146/annurev-physiol-031620-093815
Loading
/content/journals/10.1146/annurev-physiol-031620-093815
Loading

Data & Media loading...

  • Article Type: Review Article
This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error