Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Hepatocyte ATF3 protects against atherosclerosis by regulating HDL and bile acid metabolism

Abstract

Activating transcription factor (ATF)3 is known to have an anti-inflammatory function, yet the role of hepatic ATF3 in lipoprotein metabolism or atherosclerosis remains unknown. Here we show that overexpression of human ATF3 in hepatocytes reduces the development of atherosclerosis in Western-diet-fed Ldlr−/− or Apoe−/− mice, whereas hepatocyte-specific ablation of Atf3 has the opposite effect. We further show that hepatic ATF3 expression is inhibited by hydrocortisone. Mechanistically, hepatocyte ATF3 enhances high-density lipoprotein (HDL) uptake, inhibits intestinal fat and cholesterol absorption and promotes macrophage reverse cholesterol transport by inducing scavenger receptor group B type 1 (SR-BI) and repressing cholesterol 12α-hydroxylase (CYP8B1) in the liver through its interaction with p53 and hepatocyte nuclear factor 4α, respectively. Our data demonstrate that hepatocyte ATF3 is a key regulator of HDL and bile acid metabolism and atherosclerosis.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: HC represses hepatic ATF3 expression by activating the cAMP–PKA pathway.
Fig. 2: Hepatocyte ATF3 regulates HDL uptake via SR-BI.
Fig. 3: Hepatocyte ATF3 regulates SR-BI expression and plasma HDL-C levels by interacting with p53.
Fig. 4: Hepatocyte ATF3 is indispensable for regulating intestinal fat and cholesterol absorption as well as macrophage RCT.
Fig. 5: Hepatocyte ATF3 inhibits intestinal cholesterol and fat absorption by an ATF3–HNF4α–CYP8B1 pathway.
Fig. 6: Overexpression of hATF3 in hepatocytes attenuates the development of atherosclerosis independently of ApoE or LDLR.
Fig. 7: Ablation of hepatocyte ATF3 aggravates the development of atherosclerosis.

Similar content being viewed by others

Data availability

The data that support the findings of this study are available from the corresponding author upon request. RNA-seq data were deposited in the GEO repository (accession no. GSE148301). The genes regulated by ATF3 are presented in Supplementary Data 1. Source data are provided with this paper.

References

  1. Linton, M. F., Tao, H., Linton, E. F. & Yancey, P. G. SR-BI: a multifunctional receptor in cholesterol homeostasis and atherosclerosis. Trends Endocrinol. Metab. 28, 461–472 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Cuchel, M. & Rader, D. J. Macrophage reverse cholesterol transport: key to the regression of atherosclerosis? Circulation 113, 2548–2555 (2006).

    Article  PubMed  Google Scholar 

  3. Rosenson, R. S. et al. Cholesterol efflux and atheroprotection: advancing the concept of reverse cholesterol transport. Circulation 125, 1905–1919 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  4. Zhang, Y. et al. Hepatic expression of scavenger receptor class B type I (SR-BI) is a positive regulator of macrophage reverse cholesterol transport in vivo. J. Clin. Invest. 115, 2870–2874 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Braun, A. et al. Loss of SR-BI expression leads to the early onset of occlusive atherosclerotic coronary artery disease, spontaneous myocardial infarctions, severe cardiac dysfunction, and premature death in apolipoprotein E-deficient mice. Circ. Res. 90, 270–276 (2002).

    Article  CAS  PubMed  Google Scholar 

  6. Kozarsky, K. F., Donahee, M. H., Glick, J. M., Krieger, M. & Rader, D. J. Gene transfer and hepatic overexpression of the HDL receptor SR-BI reduces atherosclerosis in the cholesterol-fed LDL receptor-deficient mouse. Arterioscler. Thromb. Vasc. Biol. 20, 721–727 (2000).

    Article  CAS  PubMed  Google Scholar 

  7. Zanoni, P. et al. Rare variant in scavenger receptor BI raises HDL cholesterol and increases risk of coronary heart disease. Science 351, 1166–1171 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Li-Hawkins, J. et al. Cholic acid mediates negative feedback regulation of bile acid synthesis in mice. J. Clin. Invest. 110, 1191–1200 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Slatis, K. et al. Abolished synthesis of cholic acid reduces atherosclerotic development in apolipoprotein E knockout mice. J. Lipid Res. 51, 3289–3298 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  10. Miyake, J. H. et al. Transgenic expression of cholesterol-7-α-hydroxylase prevents atherosclerosis in C57BL/6J mice. Arterioscler. Thromb. Vasc. Biol. 22, 121–126 (2002).

    Article  CAS  PubMed  Google Scholar 

  11. Thompson, M. R., Xu, D. & Williams, B. R. ATF3 transcription factor and its emerging roles in immunity and cancer. J. Mol. Med. 87, 1053–1060 (2009).

    Article  CAS  PubMed  Google Scholar 

  12. Hai, T., Wolford, C. C. & Chang, Y. S. ATF3, a hub of the cellular adaptive-response network, in the pathogenesis of diseases: is modulation of inflammation a unifying component? Gene Expr. 15, 1–11 (2010).

    Article  CAS  PubMed  Google Scholar 

  13. Gilchrist, M. et al. Systems biology approaches identify ATF3 as a negative regulator of Toll-like receptor 4. Nature 441, 173–178 (2006).

    Article  CAS  PubMed  Google Scholar 

  14. Whitmore, M. M. et al. Negative regulation of TLR-signaling pathways by activating transcription factor-3. J. Immunol. 179, 3622–3630 (2007).

    Article  CAS  PubMed  Google Scholar 

  15. De Nardo, D. et al. High-density lipoprotein mediates anti-inflammatory reprogramming of macrophages via the transcriptional regulator ATF3. Nat. Immunol. 15, 152–160 (2014).

    Article  PubMed  CAS  Google Scholar 

  16. Yao, B. C. et al. Chronic stress: a critical risk factor for atherosclerosis. J. Int. Med. Res. 47, 1429–1440 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  17. Black, P. H. & Garbutt, L. D. Stress, inflammation and cardiovascular disease. J. Psychosom. Res. 52, 1–23 (2002).

    Article  PubMed  Google Scholar 

  18. Vieweg, W. V. et al. Treatment of depression in patients with coronary heart disease. Am. J. Med. 119, 567–573 (2006).

    Article  CAS  PubMed  Google Scholar 

  19. Rosengren, A. et al. Association of psychosocial risk factors with risk of acute myocardial infarction in 11119 cases and 13648 controls from 52 countries (the INTERHEART study): case–control study. Lancet 364, 953–962 (2004).

    Article  PubMed  Google Scholar 

  20. Brindley, D. N., McCann, B. S., Niaura, R., Stoney, C. M. & Suarez, E. C. Stress and lipoprotein metabolism: modulators and mechanisms. Metabolism 42, 3–15 (1993).

    Article  CAS  PubMed  Google Scholar 

  21. Catalina-Romero, C. et al. The relationship between job stress and dyslipidemia. Scand. J. Public Health 41, 142–149 (2013).

    Article  CAS  PubMed  Google Scholar 

  22. Djindjic, N., Jovanovic, J., Djindjic, B., Jovanovic, M. & Jovanovic, J. J. Associations between the occupational stress index and hypertension, type 2 diabetes mellitus, and lipid disorders in middle-aged men and women. Ann. Occup. Hyg. 56, 1051–1062 (2012).

    CAS  PubMed  Google Scholar 

  23. Assadi, S. N. What are the effects of psychological stress and physical work on blood lipid profiles? Medicine 96, e6816 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Heidt, T. et al. Chronic variable stress activates hematopoietic stem cells. Nat. Med. 20, 754–758 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Russell, G. & Lightman, S. The human stress response. Nat. Rev. Endocrinol. 15, 525–534 (2019).

    Article  PubMed  Google Scholar 

  26. Troxler, R. G., Sprague, E. A., Albanese, R. A., Fuchs, R. & Thompson, A. J. The association of elevated plasma cortisol and early atherosclerosis as demonstrated by coronary angiography. Atherosclerosis 26, 151–162 (1977).

    Article  CAS  PubMed  Google Scholar 

  27. Dekker, M. J. et al. Salivary cortisol is related to atherosclerosis of carotid arteries. J. Clin. Endocrinol. Metab. 93, 3741–3747 (2008).

    Article  CAS  PubMed  Google Scholar 

  28. Alevizaki, M., Cimponeriu, A., Lekakis, J., Papamichael, C. & Chrousos, G. P. High anticipatory stress plasma cortisol levels and sensitivity to glucocorticoids predict severity of coronary artery disease in subjects undergoing coronary angiography. Metabolism 56, 222–226 (2007).

    Article  CAS  PubMed  Google Scholar 

  29. Neary, N. M. et al. Hypercortisolism is associated with increased coronary arterial atherosclerosis: analysis of noninvasive coronary angiography using multidetector computerized tomography. J. Clin. Endocrinol. Metab. 98, 2045–2052 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Hermanowski-Vosatka, A. et al. 11β-HSD1 inhibition ameliorates metabolic syndrome and prevents progression of atherosclerosis in mice. J. Exp. Med. 202, 517–527 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Cholongitas, E. et al. Relative adrenal insufficiency is associated with the clinical outcome in patients with stable decompensated cirrhosis. Ann. Hepatol. 16, 584–590 (2017).

    Article  CAS  PubMed  Google Scholar 

  32. Spadaro, L. et al. Apolipoprotein AI and HDL are reduced in stable cirrhotic patients with adrenal insufficiency: a possible role in glucocorticoid deficiency. Scand. J. Gastroenterol. 50, 347–354 (2015).

    Article  CAS  PubMed  Google Scholar 

  33. Hayashi, R. Glucocorticoid replacement affects serum adiponectin levels and HDL-C in patients with secondary adrenal insufficiency. J. Clin. Endocrinol. Metab 104, 5814–5822 (2019).

    Article  PubMed  Google Scholar 

  34. Werumeus Buning, J. et al. Downregulation of cholesteryl ester transfer protein by glucocorticoids: a randomised study on HDL. Eur. J. Clin. Invest. 47, 494–503 (2017).

    Article  CAS  PubMed  Google Scholar 

  35. Nilsson, A. G. et al. Long-term safety of once-daily, dual-release hydrocortisone in patients with adrenal insufficiency: a phase 3b, open-label, extension study. Eur. J. Endocrinol. 176, 715–725 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Guarnotta, V., Ciresi, A., Pillitteri, G. & Giordano, C. Improved insulin sensitivity and secretion in prediabetic patients with adrenal insufficiency on dual-release hydrocortisone treatment: a 36-month retrospective analysis. Clin. Endocrinol. 88, 665–672 (2018).

    Article  CAS  Google Scholar 

  37. Yan, C., Lu, D., Hai, T. & Boyd, D. D. Activating transcription factor 3, a stress sensor, activates p53 by blocking its ubiquitination. EMBO J. 24, 2425–2435 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Reynier, M. O. et al. Comparative effects of cholic, chenodeoxycholic, and ursodeoxycholic acids on micellar solubilization and intestinal absorption of cholesterol. J. Lipid Res. 22, 467–473 (1981).

    Article  CAS  PubMed  Google Scholar 

  39. Tint, G. S., Salen, G. & Shefer, S. Effect of ursodeoxycholic acid and chenodeoxycholic acid on cholesterol and bile acid metabolism. Gastroenterology 91, 1007–1018 (1986).

    Article  CAS  PubMed  Google Scholar 

  40. Lanzini, A. & Northfield, T. C. Effect of ursodeoxycholic acid on biliary lipid coupling and on cholesterol absorption during fasting and eating in subjects with cholesterol gallstones. Gastroenterology 95, 408–416 (1988).

    Article  CAS  PubMed  Google Scholar 

  41. Hardison, W. G. & Grundy, S. M. Effect of ursodeoxycholate and its taurine conjugate on bile acid synthesis and cholesterol absorption. Gastroenterology 87, 130–135 (1984).

    Article  CAS  PubMed  Google Scholar 

  42. Salvioli, G., Lugli, R. & Pradelli, J. M. Cholesterol absorption and sterol balance in normal subjects receiving dietary fiber or ursodeoxycholic acid. Dig. Dis. Sci. 30, 301–307 (1985).

    Article  CAS  PubMed  Google Scholar 

  43. Chevre, R. et al. Therapeutic modulation of the bile acid pool by Cyp8b1 knockdown protects against nonalcoholic fatty liver disease in mice. FASEB J. 32, 3792–3802 (2018).

    Article  CAS  PubMed  Google Scholar 

  44. Bertaggia, E. et al. Cyp8b1 ablation prevents Western diet-induced weight gain and hepatic steatosis because of impaired fat absorption. Am. J. Physiol. Endocrinol. Metab. 313, E121–E133 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  45. Watanabe, M. et al. Bile acids lower triglyceride levels via a pathway involving FXR, SHP, and SREBP-1c. J. Clin. Invest. 113, 1408–1418 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Inoue, Y. et al. Regulation of bile acid biosynthesis by hepatocyte nuclear factor 4α. J. Lipid Res. 47, 215–227 (2006).

    Article  CAS  PubMed  Google Scholar 

  47. Jahan, A. & Chiang, J. Y. Cytokine regulation of human sterol 12α-hydroxylase (CYP8B1) gene. Am. J. Physiol. Gastrointest. Liver Physiol. 288, G685–G695 (2005).

    Article  CAS  PubMed  Google Scholar 

  48. Kliewer, S. A. & Mangelsdorf, D. J. Bile acids as hormones: the FXR–FGF15/19 pathway. Dig. Dis. 33, 327–331 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  49. Davis, R. A., Miyake, J. H., Hui, T. Y. & Spann, N. J. Regulation of cholesterol-7α-hydroxylase: BAREly missing a SHP. J. Lipid Res. 43, 533–543 (2002).

    Article  CAS  PubMed  Google Scholar 

  50. Wang, J. et al. Studies on LXR- and FXR-mediated effects on cholesterol homeostasis in normal and cholic acid-depleted mice. J. Lipid Res. 47, 421–430 (2006).

    Article  CAS  PubMed  Google Scholar 

  51. Peet, D. J. et al. Cholesterol and bile acid metabolism are impaired in mice lacking the nuclear oxysterol receptor LXRα. Cell 93, 693–704 (1998).

    Article  CAS  PubMed  Google Scholar 

  52. Hong, C. et al. The LXR–Idol axis differentially regulates plasma LDL levels in primates and mice. Cell Metab. 20, 910–918 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Wang, N., Arai, T., Ji, Y., Rinninger, F. & Tall, A. R. Liver-specific overexpression of scavenger receptor BI decreases levels of very low density lipoprotein ApoB, low density lipoprotein ApoB, and high density lipoprotein in transgenic mice. J. Biol. Chem. 273, 32920–32926 (1998).

    Article  CAS  PubMed  Google Scholar 

  54. Bjorklund, M. M. et al. Induction of atherosclerosis in mice and hamsters without germline genetic engineering. Circ. Res. 114, 1684–1689 (2014).

    Article  CAS  PubMed  Google Scholar 

  55. Joseph, J. J. & Golden, S. H. Cortisol dysregulation: the bidirectional link between stress, depression, and type 2 diabetes mellitus. Ann. N. Y. Acad. Sci. 1391, 20–34 (2017).

    Article  PubMed  Google Scholar 

  56. Huszar, D. et al. Increased LDL cholesterol and atherosclerosis in LDL receptor-deficient mice with attenuated expression of scavenger receptor B1. Arterioscler. Thromb. Vasc. Biol. 20, 1068–1073 (2000).

    Article  CAS  PubMed  Google Scholar 

  57. Arai, T., Wang, N., Bezouevski, M., Welch, C. & Tall, A. R. Decreased atherosclerosis in heterozygous low density lipoprotein receptor-deficient mice expressing the scavenger receptor BI transgene. J. Biol. Chem. 274, 2366–2371 (1999).

    Article  CAS  PubMed  Google Scholar 

  58. Huby, T. et al. Knockdown expression and hepatic deficiency reveal an atheroprotective role for SR-BI in liver and peripheral tissues. J. Clin. Invest. 116, 2767–2776 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Ji, Y. et al. Hepatic scavenger receptor BI promotes rapid clearance of high density lipoprotein free cholesterol and its transport into bile. J. Biol. Chem. 274, 33398–33402 (1999).

    Article  CAS  PubMed  Google Scholar 

  60. Houten, S. M., Watanabe, M. & Auwerx, J. Endocrine functions of bile acids. EMBO J. 25, 1419–1425 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Wang, D. Q., Tazuma, S., Cohen, D. E. & Carey, M. C. Feeding natural hydrophilic bile acids inhibits intestinal cholesterol absorption: studies in the gallstone-susceptible mouse. Am. J. Physiol. Gastrointest. Liver Physiol. 285, G494–G502 (2003).

    Article  CAS  PubMed  Google Scholar 

  62. Wang, J. et al. Relative roles of ABCG5/ABCG8 in liver and intestine. J. Lipid Res. 56, 319–330 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Huang, Y. et al. Lactobacillus acidophilus ATCC 4356 prevents atherosclerosis via inhibition of intestinal cholesterol absorption in apolipoprotein E-knockout mice. Appl. Environ. Microbiol. 80, 7496–7504 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  64. Davidson, M. H. et al. Inhibition of intestinal cholesterol absorption with ezetimibe increases components of reverse cholesterol transport in humans. Atherosclerosis 230, 322–329 (2013).

    Article  CAS  PubMed  Google Scholar 

  65. Lim, J. H., Lee, H. J., Pak, Y. K., Kim, W. H. & Song, J. Organelle stress-induced activating transcription factor-3 downregulates low-density lipoprotein receptor expression in Sk-Hep1 human liver cells. Biol. Chem. 392, 377–385 (2011).

    Article  CAS  PubMed  Google Scholar 

  66. Xu, Y. et al. A metabolic stress-inducible miR-34a-HNF4α pathway regulates lipid and lipoprotein metabolism. Nat. Commun. 6, 7466 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Kozarsky, K. F. et al. Overexpression of the HDL receptor SR-BI alters plasma HDL and bile cholesterol levels. Nature 387, 414–417 (1997).

    Article  CAS  PubMed  Google Scholar 

  68. Xu, Y. et al. Lipocalin-2 protects against diet-induced nonalcoholic fatty liver disease by targeting hepatocytes. Hepatol. Commun. 3, 763–775 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Zhang, Y., Castellani, L. W., Sinal, C. J., Gonzalez, F. J. & Edwards, P. A. Peroxisome proliferator-activated receptor-γ coactivator 1α (PGC-1α) regulates triglyceride metabolism by activation of the nuclear receptor FXR. Genes Dev. 18, 157–169 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Ge, X. et al. Aldo-keto reductase 1B7 is a target gene of FXR and regulates lipid and glucose homeostasis. J. Lipid Res. 52, 1561–1568 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Li, Y. et al. Hepatic forkhead box protein A3 regulates ApoA-I (apolipoprotein A-I) expression, cholesterol efflux, and atherogenesis. Arterioscler. Thromb. Vasc. Biol. 39, 1574–1587 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Voshol, P. J. et al. Postprandial chylomicron formation and fat absorption in multidrug resistance gene 2 P-glycoprotein-deficient mice. Gastroenterology 118, 173–182 (2000).

    Article  CAS  PubMed  Google Scholar 

  73. Goudriaan, J. R. et al. CD36 deficiency in mice impairs lipoprotein lipase-mediated triglyceride clearance. J. Lipid Res. 46, 2175–2181 (2005).

    Article  CAS  PubMed  Google Scholar 

  74. Khalifeh-Soltani, A. et al. Mfge8 promotes obesity by mediating the uptake of dietary fats and serum fatty acids. Nat. Med. 20, 175–183 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Turley, S. D., Herndon, M. W. & Dietschy, J. M. Reevaluation and application of the dual-isotope plasma ratio method for the measurement of intestinal cholesterol absorption in the hamster. J. Lipid Res. 35, 328–339 (1994).

    Article  CAS  PubMed  Google Scholar 

  76. Xu, Y. Macrophage miR-34a is a key regulator of cholesterol efflux and atherosclerosis. Mol. Ther. 202–216 (2019).

  77. Rossi, S. S., Converse, J. L. & Hofmann, A. F. High pressure liquid chromatographic analysis of conjugated bile acids in human bile: simultaneous resolution of sulfated and unsulfated lithocholyl amidates and the common conjugated bile acids. J. Lipid Res. 28, 589–595 (1987).

    Article  CAS  PubMed  Google Scholar 

  78. Wang, D. Q., Lammert, F., Paigen, B. & Carey, M. C. Phenotypic characterization of Lith genes that determine susceptibility to cholesterol cholelithiasis in inbred mice: pathophysiology of biliary lipid secretion. J. Lipid Res. 40, 2066–2079 (1999).

    Article  CAS  PubMed  Google Scholar 

  79. Naik, S. U. et al. Pharmacological activation of liver X receptors promotes reverse cholesterol transport in vivo. Circulation 113, 90–97 (2006).

    Article  CAS  PubMed  Google Scholar 

  80. Xu, Y. et al. FXR activation increases reverse cholesterol transport by modulating bile acid composition and cholesterol absorption. Hepatology 64, 1072–1085 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Roberts, D. C. et al. An alternative procedure for incorporating radiolabelled cholesteryl ester into human plasma lipoproteins in vitro. Biochem. J. 226, 319–322 (1985).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank L. Yang’s group for providing human data. The non-clinical work was supported in part by the National Institutes of Health, grants R01DK102619 (Y. Zhang), R01HL103227 (Y. Zhang and L. Yin), R01HL142086 (Y. Zhang), R01DK118941 (Y. Zhang) and R01DK118805 (Y. Zhang).

Author information

Authors and Affiliations

Authors

Contributions

Y.X., Y.L., K.J. and Y. Zhang conceived and designed the study and guided the interpretation of the results. Y.X. performed 80% of the studies and data analysis. K.J. generated floxed Atf3 mice. Y.L. and K.J. performed a number of in vivo studies. Y. Zhang supervised the project. Y. Zhang and Y.X. prepared the manuscript. X.P., Y. Zhu, S.H., S.C. and L. Yin performed various studies. The human data were provided by L.C., Y.T. and L. Yang; H.H.W. and D.Q.-H.W. performed the BA composition assays. All authors discussed the results and approved the final version of the manuscript.

Corresponding author

Correspondence to Yanqiao Zhang.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Metabolism thanks Alexander Bartelt and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Primary Handling Editor: Christoph Schmitt.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 Hepatic ATF3 is repressed by stress signaling.

a. Mouse or human primary hepatocytes were treated with vehicle (Veh), 500 nM angiotensin II (A-II), 500 nM dexamethasone (DEX), 250 μM forskolin (FSK), 100 nM glucagon (GCG) or 250 nM hydrocortisone (HC) (n = 4). After 24 h, protein levels were determined. **P < 1E-6 for Veh versus A-II, DEX, FSK, GCG or HC, respectively for mouse or human primary hepatocytes. b. HepG2 cells were treated with vehicle, angiotensin II (A-II), dexamethasone (DEX), forskolin (FSK), glucagon (GCG) or hydrocortisone (HC) as described in (a) (n = 4). mRNA levels were determined after 6 h (top panel) and protein levels were determined after 24 h (bottom panel). **P = 0.000019, 3E-6, 1E-6, 0.0031 or 2E-6 for Veh versus A-II, DEX, FSK, GSK or HC, respectively. c–f. Mouse primary hepatocytes (c; n = 4), human primary hepatocytes (d; n = 3 for Veh group and n = 4 for cAMP group) or HepG2 cells (e and f; n = 4) were treated with vehicle (Veh) or 500 μM db-cAMP (cAMP). mRNA levels were analyzed after 6 h (c–e) and protein levels in HepG2 cells were analyzed after 24 h (f). In (c), **P = 0.0034 versus Veh. In (d), **P = 0.000051 versus Veh. In (e), *P = 0.01 versus Veh. g and h. Mouse (g) or human (h) primary hepatocytes were pretreated with 10 μM H89 or 1 μM PKI 14–22 (PKI) for 2 h, followed by treatment with hydrocortisone (HC) for 6 h. mRNA levels were determined (n = 4). In (g), **P < 1E-6 for HC versus Veh, HC + H89 or HC + PKI, respectively. In (h), *P = 0.04 for HC versus Veh, and **P = 0.0061 or 0.0016 for HC versus HC + H89 or HC + PKI, respectively. i. C57BL/6J mice were i.p. injected with either vehicle or hydrocortisone (HC; 2 mg/kg) once a day for 7 days (n = 8). Plasma cholesterol lipoprotein profile was analyzed by FPLC. j–l. Atf3fl/fl mice and hepatocyte-specific Atf3−/− (L-Atf3−/−) mice were i.p. injected with either vehicle or hydrocortisone (HC; 2 mg/kg) once a day for 7 days. Hepatic mRNA levels were determined (j; n = 8 for the L-Atf3−/−+HC group, and n = 7 for 3 other groups). Plasma cholesterol lipoprotein profile was analyzed (k). Plasma ALT (left panel) and ASL (right panel) (l) levels were determined (n = 6 for the L-Atf3−/−+Vehicle group, and n = 7 for 3 other groups). In (j), for Atf3 expression, **P = 0.002 or 0.000072 for Atf3fl/fl + HC versus Atf3fl/fl + Veh or L-Atf3−/−+HC, respectively. For Scarb1 expression, **P = 2E-6 or <1E-6 for Atf3fl/fl + HC versus Atf3fl/fl + Veh or L-Atf3−/−+HC, respectively. For Cyp7a1 expression, **P = 1E-6 or <1E-6 for Atf3fl/fl + HC versus Atf3fl/fl + Veh or L-Atf3−/−+HC, respectively. For Cyp8b1 expression, **P = 0.0014 or 2E-6 for Atf3fl/fl + HC versus Atf3fl/fl + Veh or L-Atf3−/−+HC, respectively. In (l), for ALT, P = 0.94 or 0.89 for vehicle versus HC for Atf3fl/fl or L-Atf3−/− mice, respectively. For AST, P = 0.98 or 0.99 for vehicle versus HC for Atf3fl/fl or L-Atf3−/− mice, respectively. m. C57BL/6J mice were i.p. injected with either vehicle or hydrocortisone (HC; 2 mg/kg) once a day for 7 days (n = 8). The correlation between hepatic Atf3 mRNA and plasma HDL-C levels was determined. All the data are expressed as mean ± SEM. All the data points are biological replicates. A two-tailed Student’s t-test (c-e), one-way (a, b, g, h) or two-way (j, l) ANOVA with Turkey’s post hoc test for multiple comparisons, or a two-tailed Pearson correlation analysis (m) was used for statistical analysis. NS, not significant.

Source data

Extended Data Fig. 2 Over-expression of hepatic ATF3 reduces plasma HDL-C levels and bile acid hydrophobicity indices in C57BL/6J mice.

C57BL/6J mice were i.v. injected with AAV8-ALB-Null or AAV8-ALB-hATF3 (n = 8). After 2 months, mice were euthanized. a–d. Plasma levels of triglyceride (TG), total cholesterol (TC) (a), HDL-C, non-HDL-C (b), ALT and AST (c) were quantified. Plasma cholesterol lipoprotein profile was analyzed by FPLC (d). Chol, cholesterol. In (a), **P = 0.00024 versus AAV-Null for TC. In (b), **P = 0.00009 versus AAV-Null for HDL. In (c), **P = 0.0024 versus AAV-Null for ALT, and P = 0.079 for AAV-Null versus AAV-ATF3 for AST. e. RNA sequencing was performed using liver samples (n = 4 per group). The volcano plot shows many hepatic genes were differentially regulated by ATF3. f–i. Hepatic mRNA levels were quantified by qRT-PCR (f; n = 8). Hepatic proteins were analyzed by Western blot assays (g, h) and protein levels were quantified (i; n = 6). In (f), **P = 0.000013, 0.00014, 0.00053, 0.00015 or 0.0021 versus AAV-Null for Cyp7a1, Cyp8b1, Scarb1, Ldlr and Apoe expression, respectively. In (i), *P = 0.022 versus AAV-Null for LDLR, and **P = 0.00026, 0.0012, 0.0046, 0.004 or 0.0032 versus AAV-Null for CYP7A1, CYP8B1, ATF3 and ApoE expression, respectively. j,k. Plasma total bile acid (BA) levels (j; n = 8) and biliary BA composition (k; n = 8 for the Null group, and n = 7 for the ATF3 group) were determined. TUDC, tauroursodeoxycholic acid. TCA, taurocholic acid. TCDCA, taurochenodeoxycholic acid. In (j), *P = 0.025 versus Null. In (k), *P = 0.012 and **P = 0.0027 or 0.0006 versus AAV-Null for TCDCA, TUDC or TCA, respectively. l. HepG2 cells were transfected with the pGL3-Cyp7a1 or pGL3-Cyp8b1 luciferase promoters together with pCMV-ATF3 (n = 8) or pCMV-Empty (n = 7). After 36 h, relative luciferase units (RLU) were determined. **P = 0.0063 or 0.00018 versus CMV-Empty for pGL3-Cyp7a1 or pGL3-Cyp8b1, respectively. All the data are expressed as mean ± SEM. All the data points are biological replicates. A two-tailed Student’s t-test (a, c, f, i–l) or two-way ANOVA with Turkey’s post hoc test for multiple comparisons (b) was used for statistical analysis.

Source data

Extended Data Fig. 3 Over-expression of hepatic ATF3 reduces plasma HDL-C and LDL-C levels and increases plasma bile acid levels in db/db mice.

db/db mice were i.v. injected with AAV8-ALB-Null or AAV8-ALB-hATF3 (n = 8). After 2 months, mice were euthanized. a–d. Plasma total cholesterol (Chol) (a), HDL-C, non-HDL-C (b), ALT and AST (c) levels were determined. Plasma cholesterol lipoprotein profile was analyzed by FPLC (d). In (a), *P = 0.01 versus Null. In (b), **P = 0.0012 or 0.000021 versus AAV-Null for HDL or non-HDL, respectively. In (c), *P = 0.02 or 0.027 versus AAV-Null for ALT or AST, respectively. e. Plasma bile acid (BA) levels. *P = 0.042 versus Null. f and g. Hepatic proteins were analyzed by Western blot assays (f) and protein levels were quantified (g). **P = 3E-6, 8E-6, 0.00074, 7E-6, 0.0039 or 0.0065 versus AAV-Null for CYP7A1, CYP8B1, ATF3, SR-BI, LDLR or ApoE expression, respectively. All the data are expressed as mean ± SEM. All the data points are biological replicates. A two-tailed Student’s t-test (a, c, e, g) or two-way ANOVA with Turkey’s post hoc test for multiple comparisons (b) was used for statistical analysis.

Source data

Extended Data Fig. 4 Loss of hepatocyte ATF3 increases plasma HDL-C levels and reduces plasma bile acid levels.

Atf3fl/fl mice and hepatocyte-specific Atf3−/− (L-Atf3−/−) mice were fed a chow diet (n = 6). a–d. Plasma total cholesterol (Chol) (a), HDL, non-HDL-C (b), ALT and AST (c) levels were determined. Plasma cholesterol lipoprotein profile was analyzed by FPLC (d). In (a), **P = 0.0036 versus Atf3fl/fl. In (b), **P = 0.00075 for Atf3fl/fl versus L-Atf3−/− for HDL. In (c), P = 0.49 or 0.28 for Atf3fl/fl versus L-Atf3−/− for ALT or AST, respectively. e. Plasma bile acid (BA) levels. *P = 0.019 versus Atf3fl/fl. f and g. Hepatic proteins were analyzed by Western blot assays (f) and protein levels were quantified (g). **P = 2E-6, 0.000099, 0.000035 or 0.000376 versus Atf3fl/fl for ATF3, CYP7A1, CYP8B1 or SR-BI expression, respectively. h. Hepatic mRNA levels were determined by qRT-PCR. P = 0.12, 0.25 or 0.1 for Atf3fl/fl versus L-Atf3−/− for Srebp2, Hmgcr, or Hmgcs expression, respectively. All the data are expressed as mean ± SEM. All the data points are biological replicates. A two-tailed Student’s t-test (a, c, e, g, h) or two-way ANOVA with Turkey’s post hoc test for multiple comparisons (b) was used for statistical analysis.

Source data

Extended Data Fig. 5 ATF3 regulates HDL-C and LDL-C uptake by hepatocytes.

a,b. Scarb1fl/fl (Srb1fl/fl) or hepatocyte-specific Scarb1−/− (L-Srb1−/−) mice were i.v. injected with AAV8-ALB-Null or AAV8-ALB-hATF3 (a). In a separate study, Atf3+/+ or Atf3−/− mice were i.v. injected with 0.5 × 109 pfu Ad-Empty or Ad-SR-BI (b). After 10 days, primary hepatocytes were isolated from these mice. HDL uptake was carried out after hepatocytes were treated for 4 h with HDL labeled with [14C]cholesteryl Oleate (14C-HDL) (n = 5) (a, b). In (a), *P = 4E-6 or <1E-6 for Srb1fl/fl + AAV-Null versus Srb1fl/fl + AAV-ATF3 or L-Srb1−/− +AAV-Null, respectively. In (b), **P < 1E-6 for Atf3−/−+Ad-Empty versus Atf3+/++Ad-Empty, Atf3+/++Ad-SR-BI or Atf3−/−+Ad-SR-BI. c,d. Ldlr+/+ mice and Ldlr−/− mice were i.v. injected with AAV8-ALB-Null or AAV8-ALB-hATF3. After 10 days, primary hepatocytes were isolated. LDL uptake was carried out after hepatocytes were treated for 2 h (c, left panel) or 4 h (c, right panel) with LDL labeled with [14C]cholesteryl Oleate (14C-LDL) (n = 5). Western blot assays were performed (d, left panel) and LDLR protein levels were quantified (d, right panel). In (c, left panel), **P = 0.00002 or <1E-6 for Ldlr+/++AAV-Null versus Ldlr+/++AAV-ATF3 or Ldlr−/− +AAV-Null, respectively. In (c, right panel), *P = 0.000011 or <1E-6 for Ldlr+/++AAV-Null versus Ldlr+/++AAV-ATF3 or Ldlr−/− +AAV-Null, respectively. In (d), *P = 0.0038 versus Null. e. Mouse primary hepatocytes were isolated from Atf3fl/fl or L-Atf3−/− mice. LDL uptake was carried out after hepatocytes were treated for 2 or 4 h with LDL labeled with [14C]cholesteryl Oleate (14C-LDL) (n = 5). P = 0.36 or 0.58 for Atf3+/+versus Atf3−/− at 2 h or 4 h, respectively. f,g. HepG2 cells were infected with Ad-Empty or Ad-ATF3 for 24 h. Western blot assays (f) and LDL uptake (g) were performed as described above (n = 5). In (f), *P = 0.04 versus Empty. In (g), **P = 0.0016 or 0.000013 for Ad-Empty versus Ad-ATF3 at 2 h or 4 h, respectively. h,i. HepG2 cells were infected with Ad-shLacZ or Ad-shATF3 for 24 h. Protein levels were determined by Western blot assays (h, left and right panels). LDL uptake was performed as described above (n = 5) (i). In (h), **P = 0.00003 or 0.0015 versus shLacZ for ATF3 or LDLR expression, respectively. In (i), *P = 0.011 and **P = 0.00003 for shLacZ versus shAtf3 at 2 h or 4 h, respectively. All the data are expressed as mean ± SEM. All the data points are biological replicates. A two-tailed Student’s t-test (d, f, h) or two-way ANOVA with Turkey’s post hoc test for multiple comparisons (a–c, e, g, i) was used for statistical analysis. NS, not significant.

Source data

Extended Data Fig. 6 Regulation of SR-BI expression by ATF3 or p53.

a. C57BL/6J mice were i.v. injected with Ad-Empty or Ad-hP53 (n = 4). After 7 days, hepatic proteins were analyzed by Western blot assays (top panel) and protein levels were quantified (bottom panel). **P = 0.0015 versus Empty. b–d. p53fl/fl mice and hepatocyte-specific p53−/− (L-p53−/−) mice were i.v. injected with AAV8-ALB-Null or AAV8-ALB-hATF3 (n = 7 per group). After 30 days, plasma cholesterol lipoprotein profile was analyzed (b). Plasma non-HDL-C (c), ALT and AST (d) levels were quantified. In (c), P = 0.232 or 0.687 for AAV-Null versus AAV-ATF3 for p53fl/fl or L-p53−/− mice, respectively. In (d), for ALT, *P = 0.048 and **P = 0.00025 for AAV-Null versus AAV-ATF3 for p53fl/fl or L-p53−/− mice, respectively; for AST, P = 0.547 or 0.97 for AAV-Null versus AAV-ATF3 for p53fl/fl or L-p53−/− mice, respectively. e,f. p53fl/fl mice and L-p53−/− mice were i.p. injected with either vehicle or hydrocortisone (HC; 2 mg/kg) once a day for 7 days (n = 7 per group). Plasma cholesterol lipoprotein profile (e) and plasma ALT and AST levels (f) were determined. For ALT, P = 0.825 or 0.998 for Vehicle versus HC for p53fl/fl or L-p53−/− mice, respectively. For AST, P = 0.936 or 0.941 for Vehicle versus HC for p53fl/fl or L-p53−/− mice, respectively. g. HepG2 cells were transfected with pGL3-SR-BI-Luc constructs together pCMV-ATF3 or pCMV-Empty (n = 8). After 36 h, relative luciferase units (RLU) were determined. **P = 2E-6, 0.000017, 1.5E-7, 6.9E-8, 4E-7 or 1E-7 for CMV-Empty versus CMV-ATF3 for −0.29, −1.0, −0.8, −0.6, −0.4 or −0.2 kb SR-BI-Luc, respectively. h. EMSA was performed to determine the binding site for ATF3 in the proximal SR-BI promoter. Lane 1: free probe. Lane 2: binding assay for testing if ATF3 protein bound to the SR-BI DNA oligonucleotides. Lanes 3-10: competition assays using SR-BI DNA oligonucleotides containing a wild-type or mutant p53 binding site. Lanes 11 and 12: supershift assays in the presence of IgG or an ATF3 antibody. This experiment was repeated once with similar results. All the data are expressed as mean ± SEM. All the data points are biological replicates. A two-tailed Student’s t-test (a, g) or two-way ANOVA with Turkey’s post hoc test for multiple comparisons (c, d, f) was used for statistical analysis.

Source data

Extended Data Fig. 7 Hepatic ATF3 regulates fat absorption and VLDL secretion.

a. C57BL/6J mice were i.v. injected with AAV8-ALB-Null (n = 8) or AAV8-ALB-hATF3 (n = 7 per group). Fat absorption was performed by gavaging mice with [3H]triolein. Plasma radioactivity at indicated time points was measured. **P = 0.0019 for AAV-Null versus AAV-ATF3. b. Fat absorption was performed in Atf3fl/fl mice and L-Atf3−/− mice as described in (a) (n = 8). **P = 0.0011 for Atf3fl/fl versus L-Atf3−/− mice. c. C57BL/6J mice were i.v. injected with AAV8-ALB-Null or AAV8-ALB-hATF3 (n = 7). VLDL secretion was performed. **P = 0.0062 for AAV-Null versus AAV-ATF3. d. VLDL secretion was performed in Atf3fl/fl mice and L-Atf3−/− mice (n = 8). **P = 0.000056 for Atf3fl/fl versus L-Atf3−/− mice. All the data are expressed as mean ± SD. All the data points are biological replicates. Two-way ANOVA with Turkey’s post hoc test for multiple comparisons was used for statistical analysis (a-d).

Source data

Extended Data Fig. 8 ATF3 induces hepatic LXRα expression and regulates cholesterol or fat absorption independent of FXR signaling.

a. C57BL/6J mice were i.v. injected with AAV8-ALB-Null, AAV8-ALB-ATF3, and/or AAV8-ALB-CYP8B1 (n = 8). After two months, plasma ALT and AST levels were determined. For ALT, *P = 0.042 and **P = 0.001 for AAV-Null versus AAV-ATF3 for CYP8B1 and Null groups, respectively. For AST, P = 0.35 or 0.39 AAV-Null versus AAV-ATF3 for Null or CYP8B1 groups, respectively. b,c. Hnf4αfl/fl mice and L-Hnf4αfl/fl mice were i.v. injected with AAV8-ALB-Null or AAV8-ALB-hATF3 (n = 6). After two months, plasma ALT and AST levels (b) and plasma cholesterol lipoprotein profile (c) were determined. In (b), for ALT, *P = 0.011 and **P = 0.0097 for AAV-Null versus AAV-ATF3 for L-Hnf4α−/− and Hnf4αfl/fl mice, respectively. For AST, P = 0.93 or 0.16 for AAV-Null versus AAV-ATF3 for Hnf4αfl/fl or L-Hnf4α−/− mice, respectively. d. HepG2 cells were transfected with pGL3-Cyp8b1-Luc constructs together pCMV-ATF3 or pCMV-Empty (n = 6 per group). After 36 h, relative luciferase units (RLU) were determined. **P < 1E-6 for CMV-Empty versus CMV-ATF3 for −1.1, −0.9, −0.85, −0.75 or −0.66 kb Cyp8b1-Luc, respectively. e. C57BL/6J mice were i.v. injected with AAV8-ALB-Null or AAV8-ALB-hATF3 (n = 6 per group). After 2 months, hepatic or intestinal mRNA levels were determined. **P = 0.0096 or 0.0019 for AAV-Null versus AAV-ATF3 for Cyp7a1 or Fgf15, respectively. f. C57BL/6J mice were i.v. injected with AAV8-ALB-Null, AAV8-ALB-hATF3 and/or AAV8-ALB-CYP8B1 (n = 7 for the Null or ATF3 groups, n = 6 for the CYP8B1 group, and n = 8 for the ATF3 + CYP8B1 group). After two months, hepatic mRNA levels were determined. For Lxrα, **P = 1E-6 or 0.00011 for AAV-ATF3 versus AAV-Null or AAV-ATF3 + AAV-CYP8B1, respectively. g,h. Fxr+/+ mice and Fxr−/− mice were i.v. injected with AAV8-ALB-Null or AAV8-ALB-hATF3 (n = 6). After 2 months, intestinal cholesterol (g) or fat (h) absorption was determined. In (g), *P = 0.033 for Fxr+/++AAV-Null versus Fxr−/−+AAV-Null, and **P = 0.00007 or <1E-6 for AAV-Null versus AAV-ATF3 for Fxr+/+ or Fxr−/− mice, respectively. In (h), **P < 1E-6 for AAV-Null versus AAV-ATF3 for Fxr+/+ or Fxr−/− mice, respectively. The data are expressed as mean ± SEM (a, b, d-g) or mean ± SD (h). All the data points are biological replicates. A two-tailed Student’s t-test (e) and one-way (f) or two-way (a, b, d, g, h) ANOVA with Turkey’s post hoc test for multiple comparisons were used for statistical analysis. NS, not significant.

Source data

Extended Data Fig. 9 Over-expression of hepatocyte ATF3 lowers plasma lipid levels in Apoe−/−, Ldlr−/− or Apoe/−Ldlr−/− mice.

a–h. Three-months-old Apoe−/− mice were i.v. injected with AAV8-ALB-Null or AAV8-ALB-hATF3, and then fed a Western diet for 3 months. Food intake (n = 15) and body weight (n = 10) were measured (a). Plasma total cholesterol (n = 9 for the Null group, n = 13 for the ATF3 group), LDL-C (n = 8) (b), ApoA-I, ApoB (c; n = 8), cholesterol lipoprotein profile (d), triglyceride (TG) lipoprotein profile (e), and plasma ALT or AST levels (f; n = 7) were determined. Aortic roots were stained using an MOMA-2 antibody (g) and MOMA-2-positive areas were quantified (h; n = 6). In (d), the inset shows HDL fraction. In (a), P = 0.62 or 0.41versus Null for food intake or body weight, respectively. In (b), **P = 0.0002 or 0.0008 versus Null for plasma cholesterol or LDL-C, respectively. In (c), *P = 0.023 or 0.048 versus Null for plasma ApoA-I or ApoB, respectively. In (f), *P = 0.023 versus Null for AST, and **P = 0.0057 versus Null for ALT. In (h), **P = 0.0011 versus Null. i–n. Ldlr−/− mice were i.v. injected with AAV8-ALB-Null or AAV8-ALB-hATF3, and then fed a Western diet for 3 months (n = 8). Food intake and body weight were measured. Plasma total cholesterol, LDL-C (j), ApoA-I, ApoB (k), cholesterol lipoprotein profile (l), triglyceride lipoprotein profile (m), and ALT or AST levels (n) were determined. In (i), P = 0.59 or 0.13 versus Null for food intake or body weight, respectively. In (j), *P = 0.035 versus Null for cholesterol, and **P = 0.0035 versus Null for LDL-C. In (k), *P = 0.04 or 0.046 versus Null for plasma ApoA-I or ApoB, respectively. In (n), *P = 0.011 versus Null for AST, and **P = 0.0045 versus Null for ALT. o–t. Nine-weeks-old Apoe−/− mice or Apoe−/−Ldlr−/− mice were i.v. injected with AAV8-ALB-Null or AAV8-ALB-hATF3, and then fed a Western diet for 3 months. Food intake (n = 10 for Apoe−/−+AAV-Null or ATF3/Apoe−/−+AAV-ATF3, n = 15 for Apoe−/−Ldlr−/−+AAV-Null or Apoe−/−Ldlr−/−+AAV-ATF3) and body weight (n = 8 for Apoe−/−+AAV-Null or Apoe−/−+AAV-ATF3, n = 11 for Apoe−/−Ldlr−/−+AAV-Null, and n = 12 for Apoe−/−Ldlr−/−+AAV-ATF3) were measured (o). Plasma total cholesterol (p, left panel; n = 8 for Apoe−/−+AAV-Null or Apoe−/−+AAV-ATF3, n = 11 for Apoe−/−Ldlr−/−+AAV-Null, and n = 12 for Apoe−/−Ldlr−/−+AAV-ATF3), LDL-C (p, right panel; n = 10), ApoA-I (q, left panel; n = 10), ApoB (q, right panel; n = 8), cholesterol lipoprotein profile (r), triglyceride lipoprotein profile (s), and ALT or AST levels (t; n = 8 for Apoe−/−+AAV-Null or Apoe−/−+AAV-ATF3, n = 11 for Apoe−/−Ldlr−/−+AAV-Null, and n = 12 for Apoe−/−Ldlr−/−+AAV-ATF3) were determined. In (o), P = 0.288 or 0.93 for AAV-Null versus AAV-ATF3 for food intake or body weight, respectively. In (p), for plasma cholesterol, *P = 0.013 or **P = 0.000066 for AAV-Null versus AAV-ATF3 for Apoe−/− or Apoe−/−Ldlr−/− mice, respectively; for plasma LDL-C, **P = 0.0011 versus Null. In (q), *P = 0.031 or 0.022 versus Null for plasma ApoA-I or ApoB, respectively. In (t), for ALT, *P = 0.025 and **P = 0.0021 for AAV-Null versus AAV-ATF3 for Apoe−/− and Apoe−/−Ldlr−/− mice, respectively. For AST, *P = 0.019 or 0.015 for AAV-Null versus AAV-ATF3 for Apoe−/− or Apoe−/−Ldlr−/− mice, respectively. All the data are expressed as mean ± SEM. All the data points are biological replicates. A two-tailed Student’s t-test (a–c, f, h–k, n, p (LDL-C), q) or two-way ANOVA with Turkey’s post hoc test for multiple comparisons (o, p (total cholesterol), t) was used for statistical analysis.

Source data

Extended Data Fig. 10 Loss of hepatic ATF3 increases plasma lipid levels.

a. C57BL/6J (WT) mice, Ldlr−/− mice or Apoe−/− mice (on a C57BL/6J background) were fed a standard chow diet (CD) or Western diet (WD) for 3 months (n = 8). Hepatic proteins were analyzed by Western blot assays (left panel) and protein levels were quantified (right panel). **P = 0.0048, 0.000089 or 0.000062 for WT + CD versus WT + WD, Ldlr−/−+WD or Apoe−/−+WD, respectively. b–e. Six-weeks-old Atf3fl/flApoe−/− mice and L-Atf3−/−Apoe−/− mice were fed a Western diet for 3 months (n = 8). Plasma cholesterol levels were analyzed (b). Plasma cholesterol (c) or TG (d) lipoprotein profiles were determined. Plasma ALT or AST levels were quantified (e). In (b), **P = 0.0016 versus Atf3fl/flApoe−/−. In (e), *P = 0.032 and **P = 0.0001 for Atf3fl/flApoe−/− versus L-Atf3−/−Apoe−/− for AST and ALT, respectively. All the data are expressed as mean ± SEM. All the data points are biological replicates. A two-tailed Student’s t-test (b, e) or one-way ANOVA with Turkey’s post hoc test for multiple comparisons (a) was used for statistical analysis.

Source data

Supplementary information

Supplementary Information

Supplementary Data 1

Reporting Summary

Supplementary Table 1

Characteristics of participants. Plasma parameters in 24 participants are presented. A two-tailed Student’s t-test was used for statistical analysis.

Supplementary Table 2

Oligonucleotide sequences for RT–qPCR.

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 1

Unprocessed western blots.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 2

Unprocessed western blots.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 3

Unprocessed western blots.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 5

Unprocessed western blots.

Source Data Fig. 6

Statistical source data.

Source Data Fig. 6

Unprocessed western blots.

Source Data Fig. 7

Statistical source data.

Source Data Fig. 7

Unprocessed western blots.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 1

Unprocessed western blots.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 2

Unprocessed western blots.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 3

Unprocessed western blots.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 4

Unprocessed western blots.

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 5

Unprocessed western blots.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 6

Unprocessed western blots.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 9

Statistical source data.

Source Data Extended Data Fig. 10

Statistical source data.

Source Data Extended Data Fig. 10

Unprocessed western blots.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Xu, Y., Li, Y., Jadhav, K. et al. Hepatocyte ATF3 protects against atherosclerosis by regulating HDL and bile acid metabolism. Nat Metab 3, 59–74 (2021). https://doi.org/10.1038/s42255-020-00331-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s42255-020-00331-1

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing