Loss of UCP1 function augments recruitment of futile lipid cycling for thermogenesis in murine brown fat

Josef Oeckl, Petra Janovska, Katerina Adamcova, Kristina Bardova, Sarah Brunner, Sebastian Dieckmann, Josef Ecker, Tobias Fromme, Jiri Funda, Thomas Gantert, Piero Giansanti, Maria Soledad Hidrobo, Ondrej Kuda, Bernhard Kuster, Yongguo Li, Radek Pohl, Sabine Schmitt, Sabine Schweizer, Hans Zischka, Petr Zouhar, Jan Kopecky, Martin Klingenspor, Josef Oeckl, Petra Janovska, Katerina Adamcova, Kristina Bardova, Sarah Brunner, Sebastian Dieckmann, Josef Ecker, Tobias Fromme, Jiri Funda, Thomas Gantert, Piero Giansanti, Maria Soledad Hidrobo, Ondrej Kuda, Bernhard Kuster, Yongguo Li, Radek Pohl, Sabine Schmitt, Sabine Schweizer, Hans Zischka, Petr Zouhar, Jan Kopecky, Martin Klingenspor

Abstract

Objective: Classical ATP-independent non-shivering thermogenesis enabled by uncoupling protein 1 (UCP1) in brown adipose tissue (BAT) is activated, but not essential for survival, in the cold. It has long been suspected that futile ATP-consuming substrate cycles also contribute to thermogenesis and can partially compensate for the genetic ablation of UCP1 in mouse models. Futile ATP-dependent thermogenesis could thereby enable survival in the cold even when brown fat is less abundant or missing.

Methods: In this study, we explore different potential sources of UCP1-independent thermogenesis and identify a futile ATP-consuming triglyceride/fatty acid cycle as the main contributor to cellular heat production in brown adipocytes lacking UCP1. We uncover the mechanism on a molecular level and pinpoint the key enzymes involved using pharmacological and genetic interference.

Results: ATGL is the most important lipase in terms of releasing fatty acids from lipid droplets, while DGAT1 accounts for the majority of fatty acid re-esterification in UCP1-ablated brown adipocytes. Furthermore, we demonstrate that chronic cold exposure causes a pronounced remodeling of adipose tissues and leads to the recruitment of lipid cycling capacity specifically in BAT of UCP1-knockout mice, possibly fueled by fatty acids from white fat. Quantification of triglyceride/fatty acid cycling clearly shows that UCP1-ablated animals significantly increase turnover rates at room temperature and below.

Conclusion: Our results suggest an important role for futile lipid cycling in adaptive thermogenesis and total energy expenditure.

Keywords: Brown adipose tissue; Fatty acids; Futile substrate cycle; Lipolysis; Re-esterification; UCP1-independent thermogenesis.

Copyright © 2022 The Author(s). Published by Elsevier GmbH.. All rights reserved.

Figures

Figure 1
Figure 1
Proteins associated with futile calcium and lipid cycling are acutely upregulated during active lipolysis in brown UCP1-knockout adipocytes. A) Schematic depiction of sample generation, processing, and subsequent proteome analysis. Primary cultures of fully differentiated 129Sv/S1 brown wild type (WT) and UCP1-knockout (UCP1KO) adipocytes were stimulated with 500 nM isoproterenol (Iso) for 30 min. WT and UCP1KO samples were processed and analyzed separately. Pathway enrichment comparing Iso versus no treatment (Con) was performed within each genotype. The complete set of processed mass spectrometry data can be found in Supp. Table 2. B) Two-dimensional annotation enrichment analysis showing biological processes and pathways, which are significantly regulated upon adrenergic stimulation in at least one of the two genotypes. Terms that are preferentially upregulated in stimulated UCP1KO adipocytes and downregulated or not affected in WT cells are located above the x-axis and near or to the left of the y-axis. Values between 0 

Figure 2

Futile calcium cycling does not…

Figure 2

Futile calcium cycling does not contribute to cellular thermogenesis in murine brown UCP1-knockout…

Figure 2
Futile calcium cycling does not contribute to cellular thermogenesis in murine brown UCP1-knockout adipocytes. A) Seahorse XF24 extracellular flux measurement of primary cultures of fully differentiated 129Sv/S1 brown wild type (WT) and UCP1-knockout (UCP1KO) adipocytes and quantification of the oligomycin-sensitive part of stimulated oxygen consumption rate (OCR). n = 10–12 wells from two independent biological experiments. A two-tailed Welch-test was applied. Asterisk (∗) indicates a significant difference between the two groups. Isoproterenol (Iso), oligomycin (Oligo), carbonyl cyanide-4-(trifluoromethoxy)phenylhydrazone (FCCP), antimycin A (AA). B) & C) XF96 extracellular flux measurements of primary cultures of fully differentiated 129Sv/S1 brown UCP1KO adipocytes. B) Atp2a2 was knocked down in brown UCP1KO adipocytes with DsiRNA as described in “Material & Methods” and the isoproterenol-induced increase in OCR was calculated. n = 17–25 wells from three independent biological experiments. A two-tailed t-test was applied. Not statistically significant (n.s.). C) Brown UCP1KO adipocytes were acutely treated with thapsigargin (5 μM final) or BAPTA (20 μM final) for 1 h and the response to isoproterenol was monitored (left). Compounds were added as an injection via port “A”. The isoproterenol-induced increase in OCR over basal OCR (middle) and over OCR after the addition of compounds (right) was calculated. n = 18–23 wells from three independent biological experiments. One-way ANOVA followed by Tukey's HSD was applied. “a” indicates a significant difference from the control group. Not statistically significant (n.s.).

Figure 3

A futile cycle of lipolysis…

Figure 3

A futile cycle of lipolysis and re-esterification of fatty acids mediates non-shivering thermogenesis…

Figure 3
A futile cycle of lipolysis and re-esterification of fatty acids mediates non-shivering thermogenesis in brown UCP1-knockout adipocytes. A) – F) XF96 extracellular flux measurements of primary cultures of fully differentiated 129Sv/S1 brown UCP1-knockout (UCP1KO) adipocytes. A) Cells were pre-treated with an HSL inhibitor (HSLi, 20 μM final), an ATGL inhibitor (ATGLi, 40 μM final), or both inhibitors (ATGLi + HSLi) for 1 h prior to the measurement and inhibitors were present in the assay medium throughout the measurement. Isoproterenol-induced increase in oxygen consumption rate (OCR) was calculated. n = 18–24 wells from three independent biological experiments. One-way ANOVA followed by Tukey's HSD was applied. “a” indicates a significant difference from the control group, “b” from the HSLi group, and “c” from the ATGLi group. Not statistically significant (n.s.). B) Oligomycin delivered via the second injection (port “B”) was replaced with a combination of the ATGL and the HSL inhibitor. Portion of stimulated OCR sensitive to oligomycin or lipase-inhibitors was calculated. n = 21 wells from three independent biological experiments. A two-tailed t-test was applied. Not statistically significant (n.s.). C) Cells were pre-treated with a long-chain acyl-CoA synthetase inhibitor (Triacsin C, 5 μM final) for 1 h prior to the measurement and triacsin C was present in the assay medium throughout the measurement. Isoproterenol-induced increase in oxygen consumption rate (OCR) was calculated. n = 15–22 wells from three independent biological experiments. A two-tailed t-test was applied. Asterisk (∗) indicates a significant difference between the two groups. D) Cells were pre-treated with a DGAT1 inhibitor (DGAT1i, 40 μM final), a DGAT2 inhibitor (DGAT2i, 40 μM final), or both inhibitors (DGAT1i + DGAT2i) for 16 h prior to the measurement and inhibitors were present in the assay medium throughout the measurement. Isoproterenol-induced increase in oxygen consumption rate (OCR) was calculated. n = 20–21 wells from three independent biological experiments. One-way ANOVA followed by Tukey's HSD was applied. “a” indicates a significant difference from the control group, “b” from the DGAT2i group, and “c” from the DGAT1i group. Not statistically significant (n.s.). E) Dgat1 was knocked down in brown UCP1KO adipocytes with DsiRNA as described in “Material & Methods” and the isoproterenol-induced increase in OCR was calculated. n = 27–30 wells from three independent biological experiments. A two-tailed t-test was applied. Asterisk (∗) indicates a significant difference between the two groups. F) Palmitate conjugated to bovine serum albumin (Palmitate:BSA, 160 μM Palmitate:28 μM BSA final) or just BSA (28 μM final) was added to the BSA-free assay medium immediately prior to starting the measurement and ATP-linked OCR was quantified. n = 37–43 wells from three independent biological experiments. A two-tailed t-test was applied. Asterisk (∗) indicates a significant difference between the two groups. G) 129Sv/S1 brown wild type (WT) and UCP1KO cells were pre-treated as described in “Material & Methods”, glycerol and free fatty acids (FFAs) in the medium were quantified, and the amount of re-esterified FFA was calculated. 2-deoxyglucose (2DG, 50 mM final). n = two independent biological experiments (within each experiment, the content of 8 wells of each treatment level was pooled). Kruskal–Wallis two-way ANOVA followed by multiple comparisons with a Bonferroni-corrected Mann–Whitney U test was applied. Asterisk (∗) indicates a significant difference from the control group of the respective genotype. Number sign (#) indicates a significant difference between the two genotypes within one treatment level.

Figure 4

Glycolysis fuels UCP1-independent thermogenesis in…

Figure 4

Glycolysis fuels UCP1-independent thermogenesis in brown adipocytes lacking UCP1. A) – C) XF96…

Figure 4
Glycolysis fuels UCP1-independent thermogenesis in brown adipocytes lacking UCP1. A) – C) XF96 extracellular flux measurements of primary cultures of fully differentiated 129Sv/S1 brown UCP1-knockout (UCP1KO) adipocytes. A) Cells were assayed in glucose-free medium. Glucose (25 mM final) or buffer was delivered via the second injection (port “B”) and the isoproterenol-induced increase in OCR was calculated. n = 10–16 wells from two independent biological experiments. A two-tailed t-test was applied. Asterisk (∗) indicates a significant difference between the two groups. B) Cells were pre-treated with 2-deoxyglucose (2DG, 50 mM final) or a combination of 2DG and pyruvate (2DG + Pyruvate; 2DG 50 mM final, pyruvate 5 mM final). Isoproterenol-induced increase in OCR was calculated. n = 16–28 wells from three independent biological experiments. One-way ANOVA followed by Tukey's HSD was applied. Asterisk (∗) indicates a significant difference from the control group. C) Gpd1 was knocked down in brown UCP1KO adipocytes with DsiRNA as described in “Material & Methods” and the isoproterenol-induced increase in OCR was calculated. n = 28–31 wells from three independent biological experiments. A two-tailed t-test was applied. Asterisk (∗) indicates a significant difference between the two groups.

Figure 5

Brown UCP1-knockout adipocytes have a…

Figure 5

Brown UCP1-knockout adipocytes have a higher number of peridroplet mitochondria and mitochondria are…

Figure 5
Brown UCP1-knockout adipocytes have a higher number of peridroplet mitochondria and mitochondria are tightly associated with lipid droplets. A) Representative electron micrograph of a fully differentiated 129Sv/S1 brown UCP1-knockout (UCP1KO) cell including lipid droplets, mitochondria, and peridroplet mitochondria (PDM). Scale bar represents 1 μm “LD” indicates a lipid droplet, “M” a mitochondrion, and “PDM” a peridroplet mitochondrion. B) (Left) Relative proportion of PDM in relation to the total number of mitochondria in 129Sv/S1 brown wild type (WT) and UCP1KO adipocytes. (Right) Relative proportion of lipid droplet perimeter occupied by mitochondria. n = 35–46 pictures from two independent biological experiments. A two-tailed t-test was applied. Asterisk (∗) indicates a significant difference between the two groups.

Figure 6

Futile lipid cycling is recruited…

Figure 6

Futile lipid cycling is recruited for non-shivering thermogenesis in adipose tissues of UCP1-knockout…

Figure 6
Futile lipid cycling is recruited for non-shivering thermogenesis in adipose tissues of UCP1-knockout mice. A) Graphical summary of futile triglyceride/fatty acid (TG/FA) cycling in adipocytes. In response to a cold sensation, norepinephrine, here mimicked by isoproterenol, binds to adrenergic receptors and triggers the downstream cAMP-PKA-signalling cascade, which leads to the activation of lipolysis. ATGL and HSL-mediated hydrolysis of TGs and diglycerides (DGs) causes an immediate increase in cellular FA levels. The majority of newly released FA in brown UCP1KO adipocytes is re-esterified onto DGs and possibly to a lesser extent onto monoglycerides (MGs) as well as G3P. In this setting, more glucose is taken up to replenish the TCA cycle, maintain protonmotive force, and most importantly to provide G3P backbones serving as FA acceptor molecules. Primarily DGAT1 but potentially also other acyltransferases, such as glycerol-3-phosphate acyltransferases (GPATs) and 1-acylglycerol-3-phosphate O-acyltransferase (AGPATs), catalyze the esterification of FA derived from intracellular TG stores or taken up from the circulation, which is preceded by the ATP-dependent activation. The breakdown of TGs is simultaneously counterbalanced by re-synthesis, and thus lipid flux in both directions accelerates causing ATP turnover to increase without altering metabolite levels. Thereby, ATP consumption, its anaerobic glycolytic provision, and generation of ATP via oxidative phosphorylation directly linked to the flux through the mitochondrial electron transport chain (ETC) can be adjusted by enhancing or reducing lipid flux, which represents the theoretical foundation of TG/FA cycling. This figure was created using Servier Medical Art templates, which are licensed under a Creative Commons Attribution 3.0 Unported License; https://smart.servier.com. B) Futile TG/FA cycling activity and de novo lipogenesis in epididymal white adipose tissue (eWAT) of C57BL/6J wild type (WT) and UCP1-knockout (UCP1KO) mice acclimated to different ambient temperatures: 30 °C “warm-acclimated” (WA), 20 °C “mild cold-acclimated” (MCA), or 6 °C “cold-acclimated” (CA), for at least three weeks. n = 9–10 animals from two independent experiments. Relative enrichment is shown in top panels; enrichment per depot considering total tissue weight (Table 3) is shown in bottom panels. A two-way ANOVA followed by Tukey's HSD was applied. “a” indicates a significant difference from the WA group of the respective genotype. “b” indicates a significant difference from the MCA group of the respective genotype. “c” indicates a significant difference between the two genotypes within one treatment level. C) Futile TG/FA cycling activity and de novo lipogenesis in interscapular brown adipose tissue (iBAT) of WA, MCA, and CA C57BL/6J WT and UCP1KO mice. n = 9–10 animals from two independent experiments. Relative enrichment is shown in top panels; enrichment per depot considering total tissue weight (Table 3) is shown in bottom panels. A two-way ANOVA followed by Tukey's HSD was applied. “a” indicates a significant difference from the WA group of the respective genotype. “b” indicates a significant difference from the MCA group of the respective genotype. “c” indicates a significant difference between the two genotypes within one treatment level. D) Regulation of proteins potentially involved in TG/FA cycling in iBAT of C57BL/6N WT and UCP1KO mice acclimated to 23 °C or 5 °C. The complete set of processed mass spectrometry data can be found in Supp. Table 4. n = 4 animals. Column labels represent individual mice: WT (WT23) and UCP1KO (UCP1KO23) mice acclimated to 23 °C, WT (WT5) and UCP1KO (UCP1KO5) mice acclimated to 5 °C. Row labels represent gene symbols of proteins. Normalized protein intensities were scaled by calculating z-scores for each protein. Cell color indicates z-score. Abbreviations: Long-chain acyl-CoA synthetase (Acsl), acylglycerol kinase (Agk), 1-acylglycerol-3-phosphate O-acyltransferase (Agpat), diacylglycerol O-acyltransferase (Dgat), diacylglycerol kinase epsilon (Dgke), glycerol kinase (Gk), mitochondrial glycerol-3-phosphate acyltransferase (Gpam), glycerol-3-phosphate acyltransferase (Gpat), glycerol-3-phosphate dehydrogenase (Gpd), hormone-sensitive lipase (Lipe), monoglyceride lipase (Mgll), adipose tissue triglyceride lipase (Pnpla2).

Figure 7

Expression of selected genes in…

Figure 7

Expression of selected genes in iBAT of C57BL/6J wild type (WT) and UCP1-knockout…

Figure 7
Expression of selected genes in iBAT of C57BL/6J wild type (WT) and UCP1-knockout (UCP1KO) mice acclimated to different ambient temperatures: 30 °C “warm-acclimated” (WA), 20 °C “mild cold-acclimated” (MCA), or 6 °C “cold-acclimated” (CA), for at least three weeks. n = 4–5 animals (data from one experiment; confirmed in two independent experiments). “a” indicates a significant difference from the WA group of the respective genotype. “b” indicates a significant difference from the MCA group of the respective genotype. “c” indicates a significant difference between the two genotypes within one treatment level. Abbreviations of genes: CD36 molecule (Cd36), diacylglycerol O-acyltransferase 1 (Dgat1), diacylglycerol O-acyltransferase 2 (Dgat2), glycerol kinase (Gk), glycerol-3-phosphate dehydrogenase 1 (Gpd1), lipoprotein lipase (Lpl).
All figures (7)
Figure 2
Figure 2
Futile calcium cycling does not contribute to cellular thermogenesis in murine brown UCP1-knockout adipocytes. A) Seahorse XF24 extracellular flux measurement of primary cultures of fully differentiated 129Sv/S1 brown wild type (WT) and UCP1-knockout (UCP1KO) adipocytes and quantification of the oligomycin-sensitive part of stimulated oxygen consumption rate (OCR). n = 10–12 wells from two independent biological experiments. A two-tailed Welch-test was applied. Asterisk (∗) indicates a significant difference between the two groups. Isoproterenol (Iso), oligomycin (Oligo), carbonyl cyanide-4-(trifluoromethoxy)phenylhydrazone (FCCP), antimycin A (AA). B) & C) XF96 extracellular flux measurements of primary cultures of fully differentiated 129Sv/S1 brown UCP1KO adipocytes. B) Atp2a2 was knocked down in brown UCP1KO adipocytes with DsiRNA as described in “Material & Methods” and the isoproterenol-induced increase in OCR was calculated. n = 17–25 wells from three independent biological experiments. A two-tailed t-test was applied. Not statistically significant (n.s.). C) Brown UCP1KO adipocytes were acutely treated with thapsigargin (5 μM final) or BAPTA (20 μM final) for 1 h and the response to isoproterenol was monitored (left). Compounds were added as an injection via port “A”. The isoproterenol-induced increase in OCR over basal OCR (middle) and over OCR after the addition of compounds (right) was calculated. n = 18–23 wells from three independent biological experiments. One-way ANOVA followed by Tukey's HSD was applied. “a” indicates a significant difference from the control group. Not statistically significant (n.s.).
Figure 3
Figure 3
A futile cycle of lipolysis and re-esterification of fatty acids mediates non-shivering thermogenesis in brown UCP1-knockout adipocytes. A) – F) XF96 extracellular flux measurements of primary cultures of fully differentiated 129Sv/S1 brown UCP1-knockout (UCP1KO) adipocytes. A) Cells were pre-treated with an HSL inhibitor (HSLi, 20 μM final), an ATGL inhibitor (ATGLi, 40 μM final), or both inhibitors (ATGLi + HSLi) for 1 h prior to the measurement and inhibitors were present in the assay medium throughout the measurement. Isoproterenol-induced increase in oxygen consumption rate (OCR) was calculated. n = 18–24 wells from three independent biological experiments. One-way ANOVA followed by Tukey's HSD was applied. “a” indicates a significant difference from the control group, “b” from the HSLi group, and “c” from the ATGLi group. Not statistically significant (n.s.). B) Oligomycin delivered via the second injection (port “B”) was replaced with a combination of the ATGL and the HSL inhibitor. Portion of stimulated OCR sensitive to oligomycin or lipase-inhibitors was calculated. n = 21 wells from three independent biological experiments. A two-tailed t-test was applied. Not statistically significant (n.s.). C) Cells were pre-treated with a long-chain acyl-CoA synthetase inhibitor (Triacsin C, 5 μM final) for 1 h prior to the measurement and triacsin C was present in the assay medium throughout the measurement. Isoproterenol-induced increase in oxygen consumption rate (OCR) was calculated. n = 15–22 wells from three independent biological experiments. A two-tailed t-test was applied. Asterisk (∗) indicates a significant difference between the two groups. D) Cells were pre-treated with a DGAT1 inhibitor (DGAT1i, 40 μM final), a DGAT2 inhibitor (DGAT2i, 40 μM final), or both inhibitors (DGAT1i + DGAT2i) for 16 h prior to the measurement and inhibitors were present in the assay medium throughout the measurement. Isoproterenol-induced increase in oxygen consumption rate (OCR) was calculated. n = 20–21 wells from three independent biological experiments. One-way ANOVA followed by Tukey's HSD was applied. “a” indicates a significant difference from the control group, “b” from the DGAT2i group, and “c” from the DGAT1i group. Not statistically significant (n.s.). E) Dgat1 was knocked down in brown UCP1KO adipocytes with DsiRNA as described in “Material & Methods” and the isoproterenol-induced increase in OCR was calculated. n = 27–30 wells from three independent biological experiments. A two-tailed t-test was applied. Asterisk (∗) indicates a significant difference between the two groups. F) Palmitate conjugated to bovine serum albumin (Palmitate:BSA, 160 μM Palmitate:28 μM BSA final) or just BSA (28 μM final) was added to the BSA-free assay medium immediately prior to starting the measurement and ATP-linked OCR was quantified. n = 37–43 wells from three independent biological experiments. A two-tailed t-test was applied. Asterisk (∗) indicates a significant difference between the two groups. G) 129Sv/S1 brown wild type (WT) and UCP1KO cells were pre-treated as described in “Material & Methods”, glycerol and free fatty acids (FFAs) in the medium were quantified, and the amount of re-esterified FFA was calculated. 2-deoxyglucose (2DG, 50 mM final). n = two independent biological experiments (within each experiment, the content of 8 wells of each treatment level was pooled). Kruskal–Wallis two-way ANOVA followed by multiple comparisons with a Bonferroni-corrected Mann–Whitney U test was applied. Asterisk (∗) indicates a significant difference from the control group of the respective genotype. Number sign (#) indicates a significant difference between the two genotypes within one treatment level.
Figure 4
Figure 4
Glycolysis fuels UCP1-independent thermogenesis in brown adipocytes lacking UCP1. A) – C) XF96 extracellular flux measurements of primary cultures of fully differentiated 129Sv/S1 brown UCP1-knockout (UCP1KO) adipocytes. A) Cells were assayed in glucose-free medium. Glucose (25 mM final) or buffer was delivered via the second injection (port “B”) and the isoproterenol-induced increase in OCR was calculated. n = 10–16 wells from two independent biological experiments. A two-tailed t-test was applied. Asterisk (∗) indicates a significant difference between the two groups. B) Cells were pre-treated with 2-deoxyglucose (2DG, 50 mM final) or a combination of 2DG and pyruvate (2DG + Pyruvate; 2DG 50 mM final, pyruvate 5 mM final). Isoproterenol-induced increase in OCR was calculated. n = 16–28 wells from three independent biological experiments. One-way ANOVA followed by Tukey's HSD was applied. Asterisk (∗) indicates a significant difference from the control group. C) Gpd1 was knocked down in brown UCP1KO adipocytes with DsiRNA as described in “Material & Methods” and the isoproterenol-induced increase in OCR was calculated. n = 28–31 wells from three independent biological experiments. A two-tailed t-test was applied. Asterisk (∗) indicates a significant difference between the two groups.
Figure 5
Figure 5
Brown UCP1-knockout adipocytes have a higher number of peridroplet mitochondria and mitochondria are tightly associated with lipid droplets. A) Representative electron micrograph of a fully differentiated 129Sv/S1 brown UCP1-knockout (UCP1KO) cell including lipid droplets, mitochondria, and peridroplet mitochondria (PDM). Scale bar represents 1 μm “LD” indicates a lipid droplet, “M” a mitochondrion, and “PDM” a peridroplet mitochondrion. B) (Left) Relative proportion of PDM in relation to the total number of mitochondria in 129Sv/S1 brown wild type (WT) and UCP1KO adipocytes. (Right) Relative proportion of lipid droplet perimeter occupied by mitochondria. n = 35–46 pictures from two independent biological experiments. A two-tailed t-test was applied. Asterisk (∗) indicates a significant difference between the two groups.
Figure 6
Figure 6
Futile lipid cycling is recruited for non-shivering thermogenesis in adipose tissues of UCP1-knockout mice. A) Graphical summary of futile triglyceride/fatty acid (TG/FA) cycling in adipocytes. In response to a cold sensation, norepinephrine, here mimicked by isoproterenol, binds to adrenergic receptors and triggers the downstream cAMP-PKA-signalling cascade, which leads to the activation of lipolysis. ATGL and HSL-mediated hydrolysis of TGs and diglycerides (DGs) causes an immediate increase in cellular FA levels. The majority of newly released FA in brown UCP1KO adipocytes is re-esterified onto DGs and possibly to a lesser extent onto monoglycerides (MGs) as well as G3P. In this setting, more glucose is taken up to replenish the TCA cycle, maintain protonmotive force, and most importantly to provide G3P backbones serving as FA acceptor molecules. Primarily DGAT1 but potentially also other acyltransferases, such as glycerol-3-phosphate acyltransferases (GPATs) and 1-acylglycerol-3-phosphate O-acyltransferase (AGPATs), catalyze the esterification of FA derived from intracellular TG stores or taken up from the circulation, which is preceded by the ATP-dependent activation. The breakdown of TGs is simultaneously counterbalanced by re-synthesis, and thus lipid flux in both directions accelerates causing ATP turnover to increase without altering metabolite levels. Thereby, ATP consumption, its anaerobic glycolytic provision, and generation of ATP via oxidative phosphorylation directly linked to the flux through the mitochondrial electron transport chain (ETC) can be adjusted by enhancing or reducing lipid flux, which represents the theoretical foundation of TG/FA cycling. This figure was created using Servier Medical Art templates, which are licensed under a Creative Commons Attribution 3.0 Unported License; https://smart.servier.com. B) Futile TG/FA cycling activity and de novo lipogenesis in epididymal white adipose tissue (eWAT) of C57BL/6J wild type (WT) and UCP1-knockout (UCP1KO) mice acclimated to different ambient temperatures: 30 °C “warm-acclimated” (WA), 20 °C “mild cold-acclimated” (MCA), or 6 °C “cold-acclimated” (CA), for at least three weeks. n = 9–10 animals from two independent experiments. Relative enrichment is shown in top panels; enrichment per depot considering total tissue weight (Table 3) is shown in bottom panels. A two-way ANOVA followed by Tukey's HSD was applied. “a” indicates a significant difference from the WA group of the respective genotype. “b” indicates a significant difference from the MCA group of the respective genotype. “c” indicates a significant difference between the two genotypes within one treatment level. C) Futile TG/FA cycling activity and de novo lipogenesis in interscapular brown adipose tissue (iBAT) of WA, MCA, and CA C57BL/6J WT and UCP1KO mice. n = 9–10 animals from two independent experiments. Relative enrichment is shown in top panels; enrichment per depot considering total tissue weight (Table 3) is shown in bottom panels. A two-way ANOVA followed by Tukey's HSD was applied. “a” indicates a significant difference from the WA group of the respective genotype. “b” indicates a significant difference from the MCA group of the respective genotype. “c” indicates a significant difference between the two genotypes within one treatment level. D) Regulation of proteins potentially involved in TG/FA cycling in iBAT of C57BL/6N WT and UCP1KO mice acclimated to 23 °C or 5 °C. The complete set of processed mass spectrometry data can be found in Supp. Table 4. n = 4 animals. Column labels represent individual mice: WT (WT23) and UCP1KO (UCP1KO23) mice acclimated to 23 °C, WT (WT5) and UCP1KO (UCP1KO5) mice acclimated to 5 °C. Row labels represent gene symbols of proteins. Normalized protein intensities were scaled by calculating z-scores for each protein. Cell color indicates z-score. Abbreviations: Long-chain acyl-CoA synthetase (Acsl), acylglycerol kinase (Agk), 1-acylglycerol-3-phosphate O-acyltransferase (Agpat), diacylglycerol O-acyltransferase (Dgat), diacylglycerol kinase epsilon (Dgke), glycerol kinase (Gk), mitochondrial glycerol-3-phosphate acyltransferase (Gpam), glycerol-3-phosphate acyltransferase (Gpat), glycerol-3-phosphate dehydrogenase (Gpd), hormone-sensitive lipase (Lipe), monoglyceride lipase (Mgll), adipose tissue triglyceride lipase (Pnpla2).
Figure 7
Figure 7
Expression of selected genes in iBAT of C57BL/6J wild type (WT) and UCP1-knockout (UCP1KO) mice acclimated to different ambient temperatures: 30 °C “warm-acclimated” (WA), 20 °C “mild cold-acclimated” (MCA), or 6 °C “cold-acclimated” (CA), for at least three weeks. n = 4–5 animals (data from one experiment; confirmed in two independent experiments). “a” indicates a significant difference from the WA group of the respective genotype. “b” indicates a significant difference from the MCA group of the respective genotype. “c” indicates a significant difference between the two genotypes within one treatment level. Abbreviations of genes: CD36 molecule (Cd36), diacylglycerol O-acyltransferase 1 (Dgat1), diacylglycerol O-acyltransferase 2 (Dgat2), glycerol kinase (Gk), glycerol-3-phosphate dehydrogenase 1 (Gpd1), lipoprotein lipase (Lpl).

References

    1. Newsholme E.A., Crabtree B. Substrate cycles in metabolic regulation and in heat generation. Biochem Soc Symp. 1976;(41):61–109.
    1. Newsholme E.A. Substrate cycles: their metabolic, energetic and thermic consequences in man. Biochem Soc Symp. 1978;(43):183–205.
    1. Newsholme E.A., Arch J.R., Brooks B., Surholt B. The role of substrate cycles in metabolic regulation. Biochem Soc Trans. 1983;11(1):52–56.
    1. Newsholme E.A., Board Sounding. A possible metabolic basis for the control of body weight. N Engl J Med. 1980;302(7):400–405.
    1. Newsholme E.A., Challiss R.A.J., Crabtree B. Substrate cycles: their role in improving sensitivity in metabolic control. Trends in Biochemical Sciences. 1984;9(6):277–280.
    1. Newsholme E.A., Crabtree B., Higgins S.J., Thornton S.D., Start C. The activities of fructose diphosphatase in flight muscles from the bumble-bee and the role of this enzyme in heat generation. Biochem J. 1972;128(1):89–97.
    1. Surholt B., Newsholme E.A. The rate of substrate cycling between glucose and glucose 6-phosphate in muscle and fat-body of the hawk moth (Acherontia atropos) at rest and during flight. Biochem J. 1983;210(1):49–54.
    1. Surholt B., Greive H., Baal T., Bertsch A. Warm-up and substrate cycling in flight muscles of male bumblebees, Bombus terrestris. Comparative Biochemistry and Physiology Part A: Physiology. 1991;98(2):299–303.
    1. Morrissette J.M., Franck J.P.G., Block B.A. Characterization of ryanodine receptor and Ca2+-ATPase isoforms in the thermogenic heater organ of blue marlin (Makaira nigricans) Journal of Experimental Biology. 2003;206(5):805–812.
    1. Brooks B., Arch J.R., Newsholme E.A. Effects of hormones on the rate of the triacylglycerol/fatty acid substrate cycle in adipocytes and epididymal fat pads. FEBS Lett. 1982;146(2):327–330.
    1. Guan H.P., Li Y., Jensen M.V., Newgard C.B., Steppan C.M., Lazar M.A. A futile metabolic cycle activated in adipocytes by antidiabetic agents. Nat Med. 2002;8(10):1122–1128.
    1. Wolfe R.R., Herndon D.N., Jahoor F., Miyoshi H., Wolfe M. Effect of severe burn injury on substrate cycling by glucose and fatty acids. N Engl J Med. 1987;317(7):403–408.
    1. Reidy S.P., Weber J.-M. Accelerated substrate cycling: a new energy-wasting role for leptin in vivo. American Journal of Physiology-Endocrinology and Metabolism. 2002;282(2):E312–E317.
    1. Elia M., Zed C., Neale G., Livesey G. The energy cost of triglyceride-fatty acid recycling in nonobese subjects after an overnight fast and four days of starvation. Metabolism. 1987;36(3):251–255.
    1. Patel D., Kalhan S. Glycerol metabolism and triglyceride-fatty acid cycling in the human newborn: effect of maternal diabetes and intrauterine growth retardation. Pediatric Research. 1992;31(1):52–58.
    1. Blondin D.P., Nielsen S., Kuipers E.N., Severinsen M.C., Jensen V.H., Miard S., et al. Human Brown adipocyte thermogenesis is driven by β2-AR stimulation. Cell Metab. 2020;32(2):287–300. e7.
    1. Flachs P., Adamcova K., Zouhar P., Marques C., Janovska P., Viegas I., et al. Induction of lipogenesis in white fat during cold exposure in mice: link to lean phenotype. Int J Obes (Lond) 2017;41(3):372–380.
    1. Bardova K., Funda J., Pohl R., Cajka T., Hensler M., Kuda O., et al. Additive effects of omega-3 fatty acids and thiazolidinediones in mice fed a high-fat diet: triacylglycerol/fatty acid cycling in adipose tissue. Nutrients. 2020;12(12)
    1. Hui S., Cowan A.J., Zeng X., Yang L., TeSlaa T., Li X., et al. Quantitative fluxomics of circulating metabolites. Cell Metabolism. 2020;32(4):676–688. e4.
    1. Kalderon B., Mayorek N., Berry E., Zevit N., Bar-Tana J. Fatty acid cycling in the fasting rat. Am J Physiol Endocrinol Metab. 2000;279(1):E221–E227.
    1. Solinas G., Summermatter S., Mainieri D., Gubler M., Pirola L., Wymann M.P., et al. The direct effect of leptin on skeletal muscle thermogenesis is mediated by substrate cycling between de novo lipogenesis and lipid oxidation. FEBS Letters. 2004;577(3):539–544.
    1. Bal N.C., Periasamy M. Uncoupling of sarcoendoplasmic reticulum calcium ATPase pump activity by sarcolipin as the basis for muscle non-shivering thermogenesis. Philosophical Transactions of the Royal Society B: Biological Sciences. 2020;375(1793):20190135.
    1. Veliova M., Ferreira C.M., Benador I.Y., Jones A.E., Mahdaviani K., Brownstein A.J., et al. Blocking mitochondrial pyruvate import in brown adipocytes induces energy wasting via lipid cycling. EMBO Rep. 2020;21(12):e49634.
    1. Ikeda K., Kang Q., Yoneshiro T., Camporez J.P., Maki H., Homma M., et al. UCP1-independent signaling involving SERCA2b-mediated calcium cycling regulates beige fat thermogenesis and systemic glucose homeostasis. Nat Med. 2017;23(12):1454–1465.
    1. Kazak L., Chouchani E.T., Jedrychowski M.P., Erickson B.K., Shinoda K., Cohen P., et al. A creatine-driven substrate cycle enhances energy expenditure and thermogenesis in beige fat. Cell. 2015;163(3):643–655.
    1. Enerbäck S., Jacobsson A., Simpson E.M., Guerra C., Yamashita H., Harper M.-E., et al. Mice lacking mitochondrial uncoupling protein are cold-sensitive but not obese. Nature. 1997;387(6628):90–94.
    1. Keipert S., Kutschke M., Lamp D., Brachthäuser L., Neff F., Meyer C.W., et al. Genetic disruption of uncoupling protein 1 in mice renders brown adipose tissue a significant source of FGF21 secretion. Molecular Metabolism. 2015;4(7):537–542.
    1. Meyer C.W., Willershäuser M., Jastroch M., Rourke B.C., Fromme T., Oelkrug R., et al. Adaptive thermogenesis and thermal conductance in wild-type and UCP1-KO mice. Am J Physiol Regul Integr Comp Physiol. 2010;299(5):R1396–R1406.
    1. Keipert S., Kutschke M., Ost M., Schwarzmayr T., van Schothorst E.M., Lamp D., et al. Long-term cold adaptation does not require FGF21 or UCP1. Cell Metab. 2017;26(2):437–446. e5.
    1. Schweizer S., Oeckl J., Klingenspor M., Fromme T. Substrate fluxes in brown adipocytes upon adrenergic stimulation and uncoupling protein 1 ablation. Life Sci Alliance. 2018;1(6):e201800136.
    1. Ukropec J., Anunciado R.P., Ravussin Y., Hulver M.W., Kozak L.P. UCP1-independent thermogenesis in white adipose tissue of cold-acclimated Ucp1-/- mice. Journal of Biological Chemistry. 2006;281(42):31894–31908.
    1. Grimpo K., Völker M.N., Heppe E.N., Braun S., Heverhagen J.T., Heldmaier G. Brown adipose tissue dynamics in wild-type and UCP1-knockout mice: in vivo insights with magnetic resonance. Journal of Lipid Research. 2014;55(3):398–409.
    1. Antonacci M.A., McHugh C., Kelley M., McCallister A., Degan S., Branca R.T. Direct detection of brown adipose tissue thermogenesis in UCP1-/- mice by hyperpolarized (129)Xe MR thermometry. Sci Rep. 2019;9(1):14865.
    1. Chondronikola M., Volpi E., Børsheim E., Porter C., Saraf M.K., Annamalai P., et al. Brown adipose tissue activation is linked to distinct systemic effects on lipid metabolism in humans. Cell Metab. 2016;23(6):1200–1206.
    1. Dieckmann S., Strohmeyer A., Willershäuser M., Maurer S.F., Wurst W., Marschall S., et al. Susceptibility to diet-induced obesity at thermoneutral conditions is independent of UCP1. American Journal of Physiology-Endocrinology and Metabolism. 2022;322(2):E85–E100.
    1. Kroupova P., van Schothorst E.M., Keijer J., Bunschoten A., Vodicka M., Irodenko I., et al. Omega-3 phospholipids from krill oil enhance intestinal fatty acid oxidation more effectively than omega-3 triacylglycerols in high-fat diet-fed obese mice. Nutrients. 2020;12(7):2037.
    1. Kuda O., Brezinova M., Rombaldova M., Slavikova B., Posta M., Beier P., et al. Docosahexaenoic acid-derived fatty acid esters of hydroxy fatty acids (FAHFAs) with anti-inflammatory properties. Diabetes. 2016;65(9):2580–2590.
    1. Flachs P., Rossmeisl M., Kuda O., Kopecky J. Stimulation of mitochondrial oxidative capacity in white fat independent of UCP1: a key to lean phenotype. Biochim Biophys Acta. 2013;1831(5):986–1003.
    1. Bederman I.R., Foy S., Chandramouli V., Alexander J.C., Previs S.F. Triglyceride synthesis in epididymal adipose tissue: contribution of glucose and non-glucose carbon sources. J Biol Chem. 2009;284(10):6101–6108.
    1. Oeckl J., Bast-Habersbrunner A., Fromme T., Klingenspor M., Li Y. Isolation, culture, and functional analysis of murine thermogenic adipocytes. STAR Protocols. 2020;1(3):100118.
    1. Schweiger M., Schreiber R., Haemmerle G., Lass A., Fledelius C., Jacobsen P., et al. Adipose triglyceride lipase and hormone-sensitive lipase are the major enzymes in adipose tissue triacylglycerol catabolism. J Biol Chem. 2006;281(52):40236–40241.
    1. Adamcova K., Horakova O., Bardova K., Janovska P., Brezinova M., Kuda O., et al. Reduced number of adipose lineage and endothelial cells in epididymal fat in response to omega-3 PUFA in mice fed high-fat diet. Mar Drugs. 2018;16(12)
    1. Livak K.J., Schmittgen T.D. Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods. 2001;25(4):402–408.
    1. Zouhar P., Janovska P., Stanic S., Bardova K., Funda J., Haberlova B., et al. A pyrexic effect of FGF21 independent of energy expenditure and UCP1. Molecular Metabolism. 2021;53:101324.
    1. Ruprecht B., Wang D., Chiozzi R.Z., Li L.H., Hahne H., Kuster B. Hydrophilic strong anion exchange (hSAX) chromatography enables deep fractionation of tissue proteomes. Methods Mol Biol. 2017;1550:69–82.
    1. Yu P., Petzoldt S., Wilhelm M., Zolg D.P., Zheng R., Sun X., et al. Trimodal mixed mode chromatography that enables efficient offline two-dimensional peptide fractionation for proteome analysis. Analytical Chemistry. 2017;89(17):8884–8891.
    1. Cox J., Mann M. MaxQuant enables high peptide identification rates, individualized p.p.b.-range mass accuracies and proteome-wide protein quantification. Nature Biotechnology. 2008;26(12):1367–1372.
    1. Käll L., Canterbury J.D., Weston J., Noble W.S., MacCoss M.J. Semi-supervised learning for peptide identification from shotgun proteomics datasets. Nature Methods. 2007;4(11):923–925.
    1. Vizcaíno J.A., Côté R.G., Csordas A., Dianes J.A., Fabregat A., Foster J.M., et al. The PRoteomics IDEntifications (PRIDE) database and associated tools: status in 2013. Nucleic Acids Research. 2012;41(D1):D1063–D1069.
    1. Tyanova S., Temu T., Sinitcyn P., Carlson A., Hein M.Y., Geiger T., et al. The Perseus computational platform for comprehensive analysis of (prote)omics data. Nature Methods. 2016;13(9):731–740.
    1. Team R.C. 2013. R: a language and environment for statistical computing.
    1. Cox J., Mann M. 1D and 2D annotation enrichment: a statistical method integrating quantitative proteomics with complementary high-throughput data. BMC Bioinformatics. 2012;13(16):1–11.
    1. Schindelin J., Arganda-Carreras I., Frise E., Kaynig V., Longair M., Pietzsch T., et al. Fiji: an open-source platform for biological-image analysis. Nature Methods. 2012;9(7):676–682.
    1. Ball E.G., Jungas R.L. On the action of hormones which accelerate the rate of oxygen consumption and fatty acid release in rat adipose tissue in vitro. Proc Natl Acad Sci U S A. 1961;47(7):932–941.
    1. Li Y., Fromme T., Schweizer S., Schöttl T., Klingenspor M. Taking control over intracellular fatty acid levels is essential for the analysis of thermogenic function in cultured primary brown and brite/beige adipocytes. EMBO Rep. 2014;15(10):1069–1076.
    1. Braun K., Oeckl J., Westermeier J., Li Y., Klingenspor M. Non-adrenergic control of lipolysis and thermogenesis in adipose tissues. The Journal of Experimental Biology. 2018;221(Suppl 1):jeb165381.
    1. Edens N.K., Leibel R.L., Hirsch J. Mechanism of free fatty acid re-esterification in human adipocytes in vitro. J Lipid Res. 1990;31(8):1423–1431.
    1. Yu X.X., Lewin D.A., Forrest W., Adams S.H. Cold elicits the simultaneous induction of fatty acid synthesis and β-oxidation in murine brown adipose tissue: prediction from differential gene expression and confirmation in vivo. The FASEB Journal. 2002;16(2):155–168.
    1. Chitraju C., Walther T.C., Farese R.V., Jr. The triglyceride synthesis enzymes DGAT1 and DGAT2 have distinct and overlapping functions in adipocytes. J Lipid Res. 2019;60(6):1112–1120.
    1. Yen C.-L.E., Stone S.J., Koliwad S., Harris C., Farese R.V., Jr. Thematic review series: glycerolipids. DGAT enzymes and triacylglycerol biosynthesis. Journal of Lipid Research. 2008;49(11):2283–2301.
    1. Chitraju C., Mejhert N., Haas J.T., Diaz-Ramirez L.G., Grueter C.A., Imbriglio J.E., et al. Triglyceride synthesis by DGAT1 protects adipocytes from lipid-induced ER stress during lipolysis. Cell Metabolism. 2017;26(2):407–418. e3.
    1. Li Y., Li Z., Ngandiri D.A., Llerins Perez M., Wolf A., Wang Y. The molecular brakes of adipose tissue lipolysis. Frontiers in Physiology. 2022;13
    1. Maurer S.F., Fromme T., Mocek S., Zimmermann A., Klingenspor M. Uncoupling protein 1 and the capacity for nonshivering thermogenesis are components of the glucose homeostatic system. American Journal of Physiology-Endocrinology and Metabolism. 2020;318(2):E198–E215.
    1. Olsen J.M., Csikasz R.I., Dehvari N., Lu L., Sandström A., Öberg A.I., et al. β(3)-Adrenergically induced glucose uptake in brown adipose tissue is independent of UCP1 presence or activity: mediation through the mTOR pathway. Mol Metab. 2017;6(6):611–619.
    1. Hankir M.K., Klingenspor M. Brown adipocyte glucose metabolism: a heated subject. EMBO Reports. 2018;19(9):e46404.
    1. Houstĕk J., Cannon B., Lindberg O. Gylcerol-3-phosphate shuttle and its function in intermediary metabolism of hamster brown-adipose tissue. Eur J Biochem. 1975;54(1):11–18.
    1. Benador I.Y., Veliova M., Mahdaviani K., Petcherski A., Wikstrom J.D., Assali E.A., et al. Mitochondria bound to lipid droplets have unique bioenergetics, composition, and dynamics that support lipid droplet expansion. Cell Metab. 2018;27(4):869–885. e6.
    1. Schreiber R., Diwoky C., Schoiswohl G., Feiler U., Wongsiriroj N., Abdellatif M., et al. Cold-induced thermogenesis depends on ATGL-mediated lipolysis in cardiac muscle, but not Brown adipose tissue. Cell Metabolism. 2017;26(5):753–763. e7.
    1. Shin H., Ma Y., Chanturiya T., Cao Q., Wang Y., Kadegowda A.K.G., et al. Lipolysis in Brown adipocytes is not essential for cold-induced thermogenesis in mice. Cell Metabolism. 2017;26(5):764–777. e5.
    1. Rosell M., Kaforou M., Frontini A., Okolo A., Chan Y.-W., Nikolopoulou E., et al. Brown and white adipose tissues: intrinsic differences in gene expression and response to cold exposure in mice. American Journal of Physiology-Endocrinology and Metabolism. 2014;306(8):E945–E964.
    1. Kazak L., Chouchani E.T., Stavrovskaya I.G., Lu G.Z., Jedrychowski M.P., Egan D.F., et al. UCP1 deficiency causes brown fat respiratory chain depletion and sensitizes mitochondria to calcium overload-induced dysfunction. Proc Natl Acad Sci U S A. 2017;114(30):7981–7986.
    1. Yehuda-Shnaidman E., Buehrer B., Pi J., Kumar N., Collins S. Acute stimulation of white adipocyte respiration by PKA-induced lipolysis. Diabetes. 2010;59(10):2474–2483.
    1. Keipert S., Jastroch M. Brite/beige fat and UCP1 - is it thermogenesis? Biochim Biophys Acta. 2014;1837(7):1075–1082.
    1. Wojtczak L., Lehninger A.L. Formation and disappearance of an endogenous uncoupling factor during swelling and contraction of mitochondria. Biochim Biophys Acta. 1961;51:442–456.
    1. Wojtczak L., Schönfeld P. Effect of fatty acids on energy coupling processes in mitochondria. Biochim Biophys Acta. 1993;1183(1):41–57.
    1. Penzo D., Tagliapietra C., Colonna R., Petronilli V., Bernardi P. Effects of fatty acids on mitochondria: implications for cell death. Biochimica et Biophysica Acta (BBA) - Bioenergetics. 2002;1555(1):160–165.
    1. Di Paola M., Lorusso M. Interaction of free fatty acids with mitochondria: coupling, uncoupling and permeability transition. Biochim Biophys Acta. 2006;1757(9–10):1330–1337.
    1. Sorger D., Athenstaedt K., Hrastnik C., Daum G. A yeast strain lacking lipid particles bears a defect in ergosterol formation∗. Journal of Biological Chemistry. 2004;279(30):31190–31196.
    1. Dubey R., Stivala C.E., Nguyen H.Q., Goo Y.-H., Paul A., Carette J.E., et al. Lipid droplets can promote drug accumulation and activation. Nature Chemical Biology. 2020;16(2):206–213.
    1. Gülden M., Mörchel S., Tahan S., Seibert H. Impact of protein binding on the availability and cytotoxic potency of organochlorine pesticides and chlorophenols in vitro. Toxicology. 2002;175(1–3):201–213.
    1. Lewin T.M., Kim J.-H., Granger D.A., Vance J.E., Coleman R.A. Acyl-CoA synthetase isoforms 1, 4, and 5 are present in different subcellular membranes in rat liver and can Be inhibited independently∗. Journal of Biological Chemistry. 2001;276(27):24674–24679.
    1. Golozoubova V., Cannon B., Nedergaard J. UCP1 is essential for adaptive adrenergic nonshivering thermogenesis. American Journal of Physiology-Endocrinology and Metabolism. 2006;291(2):E350–E357.
    1. Golozoubova V., Hohtola E., Matthias A., Jacobsson A., Cannon B., Nedergaard J. Only UCP1 can mediate adaptive nonshivering thermogenesis in the cold. The FASEB Journal. 2001;15(11):2048–2050.
    1. Matthias A., Ohlson K.B.E., Fredriksson J.M., Jacobsson A., Nedergaard J., Cannon B. Thermogenic responses in Brown fat cells are fully UCP1-dependent. Journal of Biological Chemistry. 2000;275(33):25073–25081.
    1. Essen G.v., Lindsund E., Cannon B., Nedergaard J. Adaptive facultative diet-induced thermogenesis in wild-type but not in UCP1-ablated mice. American Journal of Physiology-Endocrinology and Metabolism. 2017;313(5):E515–E527.
    1. Festuccia W.T., Blanchard P.G., Turcotte V., Laplante M., Sariahmetoglu M., Brindley D.N., et al. The PPARgamma agonist rosiglitazone enhances rat brown adipose tissue lipogenesis from glucose without altering glucose uptake. Am J Physiol Regul Integr Comp Physiol. 2009;296(5):R1327–R1335.
    1. Lasar D., Rosenwald M., Kiehlmann E., Balaz M., Tall B., Opitz L., et al. Peroxisome proliferator activated receptor gamma controls mature Brown adipocyte inducibility through glycerol kinase. Cell Rep. 2018;22(3):760–773.
    1. Festuccia W.T.L., Guerra-Sá R., Kawashita N.H., Garófalo M.A.R., Evangelista E.A., Rodrigues V., et al. Expression of glycerokinase in brown adipose tissue is stimulated by the sympathetic nervous system. American Journal of Physiology-Regulatory, Integrative and Comparative Physiology. 2003;284(6):R1536–R1541.
    1. Palmisano B.T., Zhu L., Eckel R.H., Stafford J.M. Sex differences in lipid and lipoprotein metabolism. Molecular Metabolism. 2018;15:45–55.
    1. Wang H., Willershäuser M., Karlas A., Gorpas D., Reber J., Ntziachristos V., et al. A dual Ucp1 reporter mouse model for imaging and quantitation of brown and brite fat recruitment. Mol Metab. 2019;20:14–27.
    1. Wang H., Willershäuser M., Li Y., Fromme T., Schnabl K., Bast-Habersbrunner A., et al. Uncoupling protein-1 expression does not protect mice from diet-induced obesity. Am J Physiol Endocrinol Metab. 2021;320(2):E333–E345.
    1. Rohm M., Schäfer M., Laurent V., Üstünel B.E., Niopek K., Algire C., et al. An AMP-activated protein kinase-stabilizing peptide ameliorates adipose tissue wasting in cancer cachexia in mice. Nat Med. 2016;22(10):1120–1130.
    1. Granneman J.G., Burnazi M., Zhu Z., Schwamb L.A. White adipose tissue contributes to UCP1-independent thermogenesis. American Journal of Physiology-Endocrinology and Metabolism. 2003;285(6):E1230–E1236.

Source: PubMed

3
Abonnere