Volume 4, Issue 4, Pages (August 2013)

Slides:



Advertisements
Similar presentations
Peter L. Lee, Yuefeng Tang, Huawei Li, David A. Guertin 
Advertisements

Volume 4, Issue 12, Pages (December 2015)
Volume 10, Issue 4, Pages (October 2009)
Volume 12, Issue 1, Pages (July 2010)
Won Hee Choi, Jiyun Ahn, Chang Hwa Jung, Jung Sook Seo, Tae Youl Ha 
Volume 20, Issue 3, Pages (July 2017)
Volume 11, Issue 8, Pages (May 2015)
Volume 126, Issue 3, Pages (March 2004)
Volume 3, Issue 3, Pages (March 2006)
Won Hee Choi, Jiyun Ahn, Chang Hwa Jung, Jung Sook Seo, Tae Youl Ha 
Beneficial Effects of Subcutaneous Fat Transplantation on Metabolism
Volume 3, Issue 5, Pages (May 2006)
Volume 16, Issue 10, Pages (September 2016)
Volume 20, Issue 4, Pages (October 2014)
Irs1 Serine 307 Promotes Insulin Sensitivity in Mice
Grzegorz Sumara, Olga Sumara, Jason K. Kim, Gerard Karsenty 
Volume 142, Issue 5, Pages e1 (May 2012)
Beneficial Effects of Subcutaneous Fat Transplantation on Metabolism
Volume 6, Issue 3, Pages (September 2007)
Volume 18, Issue 2, Pages (August 2013)
Volume 22, Issue 3, Pages (March 2015)
Volume 21, Issue 11, Pages (December 2017)
Antidiabetic Effects of IGFBP2, a Leptin-Regulated Gene
Metabolic Stress Signaling Mediated by Mixed-Lineage Kinases
Volume 17, Issue 5, Pages (May 2013)
Volume 7, Issue 4, Pages (May 2014)
Volume 20, Issue 1, Pages (July 2014)
Volume 20, Issue 3, Pages (September 2014)
Volume 16, Issue 7, Pages (August 2016)
Protection against High-Fat-Diet-Induced Obesity in MDM2C305F Mice Due to Reduced p53 Activity and Enhanced Energy Expenditure  Shijie Liu, Tae-Hyung.
Antidiabetic Effects of IGFBP2, a Leptin-Regulated Gene
Volume 17, Issue 10, Pages (December 2016)
Volume 14, Issue 10, Pages (March 2016)
Regulation of Hepatic Energy Metabolism and Gluconeogenesis by BAD
Heat Shock Transcription Factor 1 Is a Key Determinant of HCC Development by Regulating Hepatic Steatosis and Metabolic Syndrome  Xiongjie Jin, Demetrius.
Cold-Inducible SIRT6 Regulates Thermogenesis of Brown and Beige Fat
Volume 23, Issue 6, Pages (June 2016)
Volume 13, Issue 8, Pages (November 2015)
Induction of Hepatitis by JNK-Mediated Expression of TNF-α
Volume 1, Issue 1, Pages (January 2005)
Volume 16, Issue 4, Pages (October 2012)
Volume 3, Issue 2, Pages (February 2006)
Volume 17, Issue 8, Pages (November 2016)
Volume 10, Issue 1, Pages (July 2009)
Volume 8, Issue 4, Pages (October 2008)
Volume 20, Issue 12, Pages (September 2017)
Volume 9, Issue 4, Pages (November 2014)
Volume 5, Issue 5, Pages (May 2007)
Volume 1, Issue 4, Pages (April 2005)
Volume 24, Issue 8, Pages e7 (August 2018)
Volume 22, Issue 2, Pages (August 2015)
Volume 6, Issue 1, Pages (July 2007)
Volume 6, Issue 3, Pages (September 2007)
Volume 13, Issue 12, Pages (December 2015)
Prevention of Steatosis by Hepatic JNK1
Volume 8, Issue 5, Pages (November 2008)
Volume 20, Issue 3, Pages (September 2014)
Volume 159, Issue 2, Pages (October 2014)
Identification of SH2-B as a key regulator of leptin sensitivity, energy balance, and body weight in mice  Decheng Ren, Minghua Li, Chaojun Duan, Liangyou.
Volume 5, Issue 6, Pages (June 2007)
Volume 7, Issue 3, Pages (March 2008)
Volume 22, Issue 2, Pages (August 2015)
Mice with AS160/TBC1D4-Thr649Ala Knockin Mutation Are Glucose Intolerant with Reduced Insulin Sensitivity and Altered GLUT4 Trafficking  Shuai Chen, David.
Volume 20, Issue 4, Pages (October 2014)
Volume 6, Issue 4, Pages (October 2007)
Volume 4, Issue 5, Pages (November 2006)
Adipose Fatty Acid Oxidation Is Required for Thermogenesis and Potentiates Oxidative Stress-Induced Inflammation  Jieun Lee, Jessica M. Ellis, Michael J.
Volume 3, Issue 3, Pages (March 2006)
Volume 26, Issue 1, Pages 1-10.e7 (January 2019)
Volume 16, Issue 3, Pages (September 2012)
Presentation transcript:

Volume 4, Issue 4, Pages 681-688 (August 2013) Role of the Mixed-Lineage Protein Kinase Pathway in the Metabolic Stress Response to Obesity  Shashi Kant, Tamera Barrett, Anastassiia Vertii, Yun Hee Noh, Dae Young Jung, Jason K. Kim, Roger J. Davis  Cell Reports  Volume 4, Issue 4, Pages 681-688 (August 2013) DOI: 10.1016/j.celrep.2013.07.019 Copyright © 2013 The Authors Terms and Conditions

Cell Reports 2013 4, 681-688DOI: (10.1016/j.celrep.2013.07.019) Copyright © 2013 The Authors Terms and Conditions

Figure 1 MLK Deficiency Protects Mice against the HFD-Induced Obesity and Insulin Resistance (A) WT and MLK-deficient male mice (8–10 weeks old) were fed either a chow diet (ND) or a HFD (16 weeks). The weight of the mice was measured (mean ± SE; n = 10). (B) Insulin tolerance tests (ITT) on MLK-deficient and WT mice fed either a chow diet (ND) or a HFD (16 weeks) were performed by intraperitoneal injection of insulin (0.75 U/kg body mass). The concentration of blood glucose was measured (mean ± SE; n = 10). (C) Glucose tolerance tests (GTT) on chow-fed (ND) and HFD-fed WT and MLK-deficient mice were performed by measurement of blood glucose concentration in animals following intraperitoneal injection of glucose (2 g/kg). The data presented represent the mean ± SE (n = 10). (D) Glucose-induced insulin release. The effect of administration of glucose (2 g/kg body mass) by intraperitoneal injection on blood insulin concentration was examined (mean ± SE; n = 10). (E) Insulin sensitivity was measured using a hyperinsulinemic-euglycemic clamp in conscious HFD-fed (3 weeks) WT and MLK-deficient mice. The steady state glucose infusion rate (GIR) during the clamp, whole body glucose turnover, hepatic glucose production (HGP) during the clamp, the hepatic insulin action (expressed as insulin-mediated percent suppression of basal HGP), and whole body glycogen plus lipid synthesis are presented (mean ± SE; n = 7 ∼ 8). (F) Sections prepared from epididymal adipose tissue and liver of chow-fed and HFD-fed WT and MLK-deficient mice were stained with hematoxylin and eosin. The panels are representative of five mice per group. Statistically significant differences between WT and MLK-deficient mice are indicated (∗p < 0.05; ∗∗p < 0.01; ∗∗∗p < 0.001; ∗∗∗∗p < 0.0001). See also Figure S2. Cell Reports 2013 4, 681-688DOI: (10.1016/j.celrep.2013.07.019) Copyright © 2013 The Authors Terms and Conditions

Figure 2 MLK Deficiency Reduces HFD-Induced JNK Activation and Suppression of AKT (A and B) Chow-fed (ND) and HFD -fed WT and MLK-deficient mice (16 weeks) were fasted overnight and then treated by intraperitoneal injection of saline (control) or 1.5 U/kg insulin (15 min). Extracts prepared from epididymal fat pads were examined by (A) immunoprecipitation with an antibody to IRS-1 and immunoblot analysis with antibodies to IRS-1/phosphotyrosine and (B) immunoblot analysis using antibodies to pThr308-AKT, pSer473-AKT, and AKT (upper panel) and by multiplexed ELISA (lower panel). Relative AKT phosphorylation (calculated as the ratio pSer473-AKT/AKT) is presented (mean ± SE; n = 6; ∗p < 0.05). (C–H) WT and MLK-deficient mice were fed a chow diet (ND) or a HFD (16 weeks) and starved overnight. The blood concentrations of insulin (E), glucose (F), leptin (G), resistin (H), IL1β (I), and IL6 (J) are presented (mean ± SE; n = 10). Statistically significant differences between MLK-deficient and WT mice are indicated (∗p < 0.05; ∗∗p < 0.01). (I) MAPK activation in epididymal adipose tissue of WT and MLK-deficient mice was measured by immunoblot analysis by probing with antibodies to MAPK and phospho-MAPK. (J) WT and MLK-deficient primary embryo fibroblasts were treated with 0.5% (w.v.) BSA or 0.75 mM palmitate/0.5% BSA (16 hr). Cell lysates were examined by immunoblot analysis with antibodies to pJNK and JNK. Cell Reports 2013 4, 681-688DOI: (10.1016/j.celrep.2013.07.019) Copyright © 2013 The Authors Terms and Conditions

Figure 3 MLK Deficiency Causes Increased Energy Expenditure Food and water consumption, gas exchange (VO2 and VCO2), energy expenditure, and physical activity were measured using metabolic cages (mean ± SE; n = 6). Statistically significant differences between WT and MLK-deficient mice are indicated (∗p < 0.05; ∗∗p < 0.01; ∗∗∗p < 0.001). Cell Reports 2013 4, 681-688DOI: (10.1016/j.celrep.2013.07.019) Copyright © 2013 The Authors Terms and Conditions

Figure 4 The Sympathoadrenal System Causes Increased Thermogenesis in Mlk2−/− Mlk3−/− Mice (A and B) WT and MLK-deficient mice were fed a chow diet (ND) or a HFD (16 weeks). The blood concentration of epinephrine and norepinephrine was measured (mean ± SE; n = 8). (C) The body temperature of the mice was measured (mean ± SE; n = 8). (D–H) The expression of Dio2 (D), Ucp1 (E), Ucp3 (F), Ppargc1a (G), and Lpl (H) messenger RNA (mRNA) expression in interscapular brown fat was measured by quantitative RT-PCR and was normalized to the amount of Gapdh mRNA in each sample (mean ± SE; n = 8). (I) Sections prepared from interscapular brown fat of chow-fed (ND) and HFD-fed WT and MLK-deficient mice were stained with hematoxylin and eosin. The panels presented are representative of five mice per group. (J) The body temperature of WT and MLK-deficient mice was measured (mean ± SE; n = 10). The effect of treatment of the mice with solvent (control) or the β3 adrenergic receptor antagonist SR59230A was examined. (K–M) The expression of Ppargc1a (K), Ucp1 (L), and Ucp3 (M) mRNA expression in brown fat was measured by quantitative RT-PCR and was normalized to the amount of Gapdh mRNA in each sample (mean ± SE; n = 10). Statistically significant differences between WT and MLK-deficient mice are indicated (∗p < 0.05; ∗∗p < 0.01). Cell Reports 2013 4, 681-688DOI: (10.1016/j.celrep.2013.07.019) Copyright © 2013 The Authors Terms and Conditions

Figure S1 Expression of MLK Isoforms, Related to Results (A) Total RNA isolated from liver and epididymal white adipose tissue of WT mice fed a chow diet (ND) or a high fat diet (HFD) for 16 wks. was employed to measure the expression of Map3k9, Map3k10, Map3k11, and Bc021891 mRNA by quantitative Taqman© RT-PCR analysis. The data are normalized for the amount of Gapdh mRNA detected in each sample (mean ± SE; n = 4). (B) The expression of Map3k9, Map3k10, Map3k11, and Bc021891 mRNA by gastrocnemius muscle, brown adipose tissue, whole brain, and hypothalamus of chow-fed WT mice is presented. Cell Reports 2013 4, 681-688DOI: (10.1016/j.celrep.2013.07.019) Copyright © 2013 The Authors Terms and Conditions

Figure S2 Comparison of WT and MLK-Deficient Mouse Tissue Mass, Related to Figure 1 Map3k10−/− Map3k11−/− and WT mice were fed a chow diet (ND) or a HFD (16 weeks). The weight of epididymal white fat, interscapular brown fat, quadriceps muscle, liver, and heart were measured (mean ± SE; n = 10). Statistically significant differences between Map3k10−/− Map3k11−/− and WT are indicated (∗p < 0.05; ∗∗p < 0.01; ∗∗∗p < 0.001). White epididymal fat pad mass in MLK-deficient mice was decreased compared with WT mice. However, the total number of adipocytes per epididymal fat pad was greater in MLK-deficient mice (1,390,000 ± 140,000 adipocytes) than WT mice (720,000 ± 120,000 adipocytes) (mean ± SEM; n = 5; p < 0.05). Cell Reports 2013 4, 681-688DOI: (10.1016/j.celrep.2013.07.019) Copyright © 2013 The Authors Terms and Conditions

Figure S3 Pituitary Hormones and Hypothalamic Gene Expression, Related to Results (A) Blood concentration of pituitary peptide hormones. WT and Map3k10−/− Map3k11−/− mice were fed a chow diet (ND) or a HFD (16 wks.) and then fasted overnight. The concentration of adrenocorticotropic hormone (ACTH), thyroid stimulating hormone (TSH), and growth hormone (GH) in the blood was measured by multiplexed ELISA (mean ± SE, n = 10). Statistically significant differences between WT and Map3k10−/− Map3k11−/− mice are indicated (∗p < 0.05). (B) Hypothalamic gene expression. WT and Map3k10−/− Map3k11−/− mice were fed a chow diet (ND) or a HFD (16 wk) and then fasted overnight. The expression of thyrotropin releasing hormone (Trh), leptin receptor (Lepr), pro-opiomelanocortin (Pomc), Agouti-related peptide (Agrp), somatostatin (Sst) and pro-melanin-concentrating hormone (Pmch) mRNA in the hypothalamus was measured by quantitative RT–PCR. The data were normalized to the expression of Gapdh mRNA in each sample and are presented as the mean ± SE (n = 6). Statistically significant differences between WT and Map3k10−/− Map3k11−/− mice are indicated (∗p < 0.05; ∗∗p < 0.01). Cell Reports 2013 4, 681-688DOI: (10.1016/j.celrep.2013.07.019) Copyright © 2013 The Authors Terms and Conditions

Figure S4 Thyroid Hormone-Stimulated Gene Expression in WT and Map3k10−/− Map3k11−/− Mice, Related to Results (A) The expression of uncoupling protein 1 (Ucp1), phosphoenolpyruvate carboxykinase (Pck1), Spot14 (Thrsp), Lactate dehydrogenase β (Ldhb), and glucose transporter 4 (Slc2A4) in brown fat of WT and Map3k10−/− Map3k11−/− mice was measured by quantitative RT–PCR. The data were normalized to the expression of Gapdh mRNA in each sample and are presented as the mean ± SE (n = 6). Statistically significant differences between WT and Map3k10−/− Map3k11−/− mice are indicated (∗,p < 0.05; ∗∗,p < 0.01). (B–G) WT and Map3k10−/− Map3k11−/− mice were fed a HFD (16 weeks) together with PTU in the drinking water. No significant differences (p > 0.05) between the WT and MLK-deficient mice were detected in body temperature (B), body weight (C), insulin tolerance (D), blood glucose concentration of fed (E) and over-night starved (F) mice or fasted insulin concentration (G). The data are presented as the mean ± SE (n = 10). Cell Reports 2013 4, 681-688DOI: (10.1016/j.celrep.2013.07.019) Copyright © 2013 The Authors Terms and Conditions