[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (250)

Search Parameters:
Keywords = brown thermogenesis

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 3188 KiB  
Article
Cold Exposure Rejuvenates the Metabolic Phenotype of Panx1−/− Mice
by Filippo Molica, Avigail Ehrlich, Graziano Pelli, Olga M. Rusiecka, Christophe Montessuit, Marc Chanson and Brenda R. Kwak
Biomolecules 2024, 14(9), 1058; https://doi.org/10.3390/biom14091058 - 25 Aug 2024
Viewed by 442
Abstract
Pannexin1 (Panx1) ATP channels are important in adipocyte biology, potentially influencing energy storage and expenditure. We compared the metabolic phenotype of young (14 weeks old) and mature (20 weeks old) wild-type (WT) and Panx1−/− mice exposed or not to cold (6 °C) [...] Read more.
Pannexin1 (Panx1) ATP channels are important in adipocyte biology, potentially influencing energy storage and expenditure. We compared the metabolic phenotype of young (14 weeks old) and mature (20 weeks old) wild-type (WT) and Panx1−/− mice exposed or not to cold (6 °C) during 28 days, a condition promoting adipocyte browning. Young Panx1−/− mice weighed less and exhibited increased fat mass, improved glucose tolerance, and lower insulin sensitivity than WT mice. Their energy expenditure and respiratory exchange ratio (RER) were increased, and their fatty acid oxidation decreased. These metabolic effects were no longer observed in mature Panx1−/− mice. The exposure of mature mice to cold exacerbated their younger metabolic phenotype. The white adipose tissue (WAT) of cold-exposed Panx1−/− mice contained more small-sized adipocytes, but, in contrast to WT mice, white adipocytes did not increase their expression of Ucp1 nor of other markers of browning adipocytes. Interestingly, Glut4 expression was already enhanced in the WAT of young Panx1−/− mice kept at 22 °C as compared to WT mice. Thus, Panx1 deletion exerts overall beneficial metabolic effects in mice that are pre-adapted to chronic cold exposure. Panx1−/− mice show morphological characteristics of WAT browning, which are exacerbated upon cold exposure, an effect that appears to be associated with Ucp1-independent thermogenesis. Full article
Show Figures

Figure 1

Figure 1
<p>Experimental design and timeline of temperature changes. Fourteen-week-old mice were weighed and allocated into three experimental groups, such that, per genotype, the average body weight (BW) of the 3 groups was similar. Mice underwent echoMRI and were housed in pairs in a temperature-, light- and humidity-controlled thermostatic chamber or Phenomaster. Indirect calorimetry was performed during one week at 22 °C on group 1 (G1; top), after which the mice were killed, and the tissue/blood samples were obtained. Mice from G2 (middle) remained at 22 °C for a 6-week period, and calorimetry was performed for the last 4 weeks. Then, OGTT or ITT experiments were performed, echoMRI and body weight were measured, mice were killed, and tissue/blood samples were obtained. Mice from G3 (bottom) had one week of acclimatization at 14 °C and were subsequently exposed to a 4-week period of cold (6 °C). Then, OGTT or ITT experiments were performed, echoMRI and body weight were measured, mice were killed, and tissue/blood samples were obtained. Experiments in the Phenomaster metabolic screening platform allowed for respirometry (indirect calorimetry), monitoring of food and water consumption, and examination of locomotor activity. Each cage was considered as one pooled sample for these metabolic measurements.</p>
Full article ">Figure 2
<p>Panx1 deletion enhances glucose over fat metabolism in 14-week-old mice. (<b>A</b>) Body weights of all 14-week-old WT (black dots) and <span class="html-italic">Panx1<sup>−/−</sup></span> (red dots) mice before separating them into 3 groups. (<b>B</b>) Fat mass of WT and <span class="html-italic">Panx1<sup>−/−</sup></span> mice (G1) measured by echoMRI. Serum total cholesterol (<b>C</b>), FFA (<b>D</b>), and TG (<b>E</b>) levels in WT and <span class="html-italic">Panx1<sup>−/−</sup></span> mice. Indirect calorimetry allowed the quantification of energy expenditure (<b>F</b>), cumulative food intake (<b>G</b>), cumulative water consumption (<b>H</b>), respiratory exchange ratio (<b>K</b>), and FA oxidation (<b>L</b>) in WT and <span class="html-italic">Panx1<sup>−/−</sup></span> mice (mice were housed two per cage; each cage was considered as one pooled sample; N = 6 cages). The active period of mice is indicated in light blue. Ghrelin (<b>I</b>), leptin (<b>J</b>), and FGF21 (<b>N</b>) levels measured by ELISA in serum of WT and <span class="html-italic">Panx1<sup>−/−</sup></span> mice. (<b>M</b>) Oral glucose tolerance test and (<b>O</b>) insulin tolerance test of WT (OGTT: N = 29; ITT: N = 20) and <span class="html-italic">Panx1<sup>−/−</sup></span> (OGTT: N = 30; ITT: N = 20) mice. Data are shown as individual data points, except for time lines where the number of mice is specified, and expressed as mean ± SEM. * <span class="html-italic">p</span> ≤ 0.05, <sup>†</sup> <span class="html-italic">p</span> ≤ 0.01, and <sup>‡</sup> <span class="html-italic">p</span> ≤ 0.001.</p>
Full article ">Figure 3
<p>The metabolic phenotype of <span class="html-italic">Panx1<sup>−/−</sup></span> mice normalizes between 14 and 20 weeks of age. (<b>A</b>) Body weight of 20-week-old WT (black dots) and <span class="html-italic">Panx1<sup>−/−</sup></span> mice (red dots). (<b>B</b>) The weight gain of the mice between 14 and 20 weeks of age was also measured. (<b>C</b>) Fat mass and (<b>D</b>) difference in fat mass (as compared to their fat mass at 14 weeks) of WT and <span class="html-italic">Panx1<sup>−/−</sup></span> mice were measured by echoMRI. Serum total cholesterol (<b>E</b>), FFA (<b>F</b>), and TG (<b>G</b>) levels in WT and <span class="html-italic">Panx1<sup>−/−</sup></span> mice. Energy expenditure (<b>H</b>), cumulative food intake (<b>I</b>), cumulative water consumption (<b>J</b>), respiratory exchange ratio (<b>M</b>), and FA oxidation (<b>N</b>) in WT and <span class="html-italic">Panx1<sup>−/−</sup></span> mice were measured by indirect calorimetry (N = 3 cages). The active period of mice is indicated in light blue. Ghrelin (<b>K</b>), leptin (<b>L</b>), and FGF21 (<b>Q</b>) levels measured by ELISA in serum of WT and <span class="html-italic">Panx1<sup>−/−</sup></span> mice. (<b>O</b>) Oral glucose tolerance test and (<b>P</b>) insulin tolerance test of WT (OGTT: N = 15; ITT: N = 8) and <span class="html-italic">Panx1<sup>−/−</sup></span> (OGTT: N = 15; ITT: N = 8) mice. Data are shown as individual data points, except for time lines where the number of mice is specified, and expressed as mean ± SEM. * <span class="html-italic">p</span> ≤ 0.05 and <sup>†</sup> <span class="html-italic">p</span> ≤ 0.01.</p>
Full article ">Figure 4
<p>Cold exposure evokes different metabolic changes in <span class="html-italic">Panx1<sup>−/−</sup></span> and WT mice. (<b>A</b>) Body weights of 20-week-old WT (black dots) and <span class="html-italic">Panx1<sup>−/−</sup></span> (red dots) after 4 weeks of cold exposure at 6 °C. (<b>B</b>) The weight gain of the mice between 14 and 20 weeks of age was also measured. (<b>C</b>) Fat mass and (<b>D</b>) difference in fat mass (as compared to their fat mass at 14 weeks) of cold-exposed WT and <span class="html-italic">Panx1<sup>−/−</sup></span> mice were measured by echoMRI. Serum total cholesterol (<b>E</b>), FFA (<b>F</b>), and TG (<b>G</b>) levels in cold-exposed WT and <span class="html-italic">Panx1<sup>−/−</sup></span> mice. Energy expenditure (<b>H</b>), cumulative food intake (<b>I</b>), cumulative water consumption (<b>J</b>), respiratory exchange ratio (<b>M</b>), and FA oxidation (<b>N</b>) in WT and <span class="html-italic">Panx1<sup>−/−</sup></span> mice were measured by indirect calorimetry (N = 3 cages) during cold exposure. The active period of mice is indicated in light blue. Ghrelin (<b>K</b>), leptin (<b>L</b>), and FGF21 (<b>P</b>) levels measured by ELISA in serum of cold-exposed WT and <span class="html-italic">Panx1<sup>−/−</sup></span> mice. (<b>O</b>) Oral glucose tolerance test and (<b>Q</b>) insulin tolerance test of cold-exposed WT (OGTT: N = 15; ITT: N = 8) and <span class="html-italic">Panx1<sup>−/−</sup></span> (OGTT: N = 15; ITT: N = 8) mice. Data are shown as individual data points, except for time lines where the number of mice is specified, and expressed as mean ± SEM. * <span class="html-italic">p</span> ≤ 0.05, <sup>†</sup> <span class="html-italic">p</span> ≤ 0.01, and <sup>‡</sup> <span class="html-italic">p</span> ≤ 0.001.</p>
Full article ">Figure 5
<p>Cold-induced WAT morphological changes are amplified in <span class="html-italic">Panx1<sup>−/−</sup></span> mice. (<b>A</b>,<b>B</b>) Illustrative Western blots (Western blot original images can be found in <a href="#app1-biomolecules-14-01058" class="html-app">Supplementary Materials</a>) and (<b>C</b>,<b>D</b>) quantification of Glut4 expression in WAT (<b>C</b>) and skeletal muscle (<b>D</b>) from 14-week-old or 20-week-old WT and <span class="html-italic">Panx1<sup>−/−</sup></span> mice housed either at 22 °C or at 6 °C during 4 weeks. Glut4 expression was normalized towards beta actin. (<b>E</b>) Weight of WAT pads from 14-week-old or 20-week-old WT and <span class="html-italic">Panx1<sup>−/−</sup></span> mice at 22 °C or exposed to cold. (<b>F</b>) Hematoxylin and eosin staining on paraffin sections from WAT of WT and <span class="html-italic">Panx1<sup>−/−</sup></span> mice at 22 °C or exposed to cold. Quantification of the numbers (<b>G</b>) of adipocytes in the WAT of WT and <span class="html-italic">Panx1<sup>−/−</sup></span> mice at 22 °C or exposed to cold. Data are shown as individual data points and expressed as mean ± SEM. (<b>H</b>) Quantification of the proportion of small (≤50 μm), medium (51–69 μm), large (70–89 μm), and very large (≥90 μm) adipocytes in WAT of WT and <span class="html-italic">Panx1<sup>−/−</sup></span> mice at 22 °C or exposed to cold. * <span class="html-italic">p</span> ≤ 0.05 and <sup>‡</sup> <span class="html-italic">p</span> ≤ 0.001.</p>
Full article ">Figure 6
<p><span class="html-italic">Panx1<sup>−/−</sup></span> mice do not increase expression of mitochondrial oxidation genes upon cold exposure. (<b>A</b>) Relative mRNA expression of Panx1 in mitochondria extracted from WAT of 20-week-old WT (white bar, black dots) and <span class="html-italic">Panx1<sup>−/−</sup></span> (white bar, red dots) mice at 22 °C or exposed to cold (6 °C). BAT mRNA extracts from WT mice (grey bar) were used as positive control. Ucp1 (<b>B</b>) mRNA expression was established by qPCR in WAT from WT (black dots) and <span class="html-italic">Panx1<sup>−/−</sup></span> (red dots) mice at 22 °C or exposed to cold. (<b>C</b>) Ucp1 protein expression was determined by Western blot (Western blot original images can be found in <a href="#app1-biomolecules-14-01058" class="html-app">Supplementary Materials</a>). Ucp1 expression was normalized towards beta actin. Cidea (<b>D</b>), Cpt1a (<b>E</b>), and Cpt1b (<b>F</b>) mRNA expressions were assessed in WAT from WT (black dots) and <span class="html-italic">Panx1<sup>−/−</sup></span> (red dots) mice at 22 °C or exposed to cold. mRNA expression of all investigated genes was normalized to B2m. Data are shown as individual data points and expressed as mean ± SEM. * <span class="html-italic">p</span> ≤ 0.05 and <sup>†</sup> <span class="html-italic">p</span> ≤ 0.01.</p>
Full article ">Figure 7
<p>Schematic illustrating the main findings of our study. Upon chronic cold exposure (6 °C) <span class="html-italic">Panx1<sup>−/−</sup></span> mice exhibit increased energy expenditure and have an increased ghrelin/leptin ratio resulting in increased food and water intake. Their serum TG and FFAs, as well as fat mass, are relatively decreased as compared to WT mice under the same conditions. WAT of cold-exposed <span class="html-italic">Panx1<sup>−/−</sup></span> mice contain smaller adipocytes than WT mice with increased levels of Glut4. While chronic cold exposure induced Glut4, Cidea, Ucp1, and Cpt1 in WAT of WT mice, chronic cold exposure of <span class="html-italic">Panx1<sup>−/−</sup></span> mice failed to induce Ucp1, Cpt1, and Glut4 in their WAT, pointing to Ucp1-independent thermogenesis in <span class="html-italic">Panx1<sup>−/−</sup></span> mice.</p>
Full article ">
16 pages, 5679 KiB  
Article
TRPV1 Activation Antagonizes High-Fat Diet-Induced Obesity at Thermoneutrality and Enhances UCP-1 Transcription via PRDM-16
by Padmamalini Baskaran, Noah Gustafson and Nicolas Chavez
Pharmaceuticals 2024, 17(8), 1098; https://doi.org/10.3390/ph17081098 - 21 Aug 2024
Viewed by 891
Abstract
Body weight is a balance between energy intake and energy expenditure. Energy expenditure is mainly governed by physical activity and adaptive thermogenesis. Adaptive dietary thermogenesis in brown and beige adipose tissue occurs through mitochondrial uncoupling protein (UCP-1). Laboratory mice, when housed at an [...] Read more.
Body weight is a balance between energy intake and energy expenditure. Energy expenditure is mainly governed by physical activity and adaptive thermogenesis. Adaptive dietary thermogenesis in brown and beige adipose tissue occurs through mitochondrial uncoupling protein (UCP-1). Laboratory mice, when housed at an ambient temperature of 22–24 °C, maintain their body temperature by dietary thermogenesis, eating more food compared to thermoneutrality. Humans remain in the thermoneutral zone (TNZ) without expending extra energy to maintain normal body temperature. TRPV1 activation by capsaicin (CAP) inhibited weight gain in mice housed at ambient temperature by activating UCP-1-dependent adaptive thermogenesis. Hence, we evaluated the effect of CAP feeding on WT and UCP-1−/− mice maintained under thermoneutral conditions. Our research presents novel findings that TRPV1 activation by CAP at thermoneutrality counters obesity in WT mice and promotes PRDM-16-dependent UCP-1 transcription. CAP fails to inhibit weight gain in UCP-1−/− mice housed at thermoneutrality and in adipose tissue-specific PRDM-16−/− mice. In vitro, capsaicin treatment increases UCP-1 transcription in PRDM-16 overexpressing cells. Our data indicate for the first time that TRPV1 activation counters obesity at thermoneutrality permissive for UCP-1 and the enhancement of PRDM-16 is not beneficial in the absence of UCP-1. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>TRPV1 expression and activity in WAT and BAT. (<b>A</b>). Western blot showing TRPV1 expression in mouse tissues. HEK<sup>TRPV1</sup> is the positive control. (<b>B</b>). TRPV1 mRNA expression normalized to the 18s RNA in these tissues (<span class="html-italic">n</span> = 5). (<b>C</b>). Representative micrograph showing the immunohistochemical detection of TRPV1 expression in the iWAT preadipocytes of WT and TRPV1<sup>−/−</sup> mice fed various diets. (<b>D</b>). Quantification of the fluorescence intensity (arbitrary units). (<b>E</b>). Representative traces of CAP-stimulated TRPV1 currents in the primary brown preadipocytes of WT and TRPV1<sup>−/−</sup> mice at −60 mV. (<b>F</b>). Average currents ± S.E.M. in these cells (<span class="html-italic">n</span> = 6 to 8). (<b>G</b>). CAP-stimulated TRPV1 currents in NCD- or HFD (±CAP)-fed primary brown preadipocytes from WT mice. (<b>H</b>). Average currents ± S.E.M. in these cells (<span class="html-italic">n</span> = 9 to 11). ** <span class="html-italic">p</span> &lt; 0.01, significantly different.</p>
Full article ">Figure 2
<p>Rectal temperatures of WT mice at ambient temperature (<b>A</b>) and in the TNZ (<b>B</b>). Rectal temperatures of WT mice fed the NCD or HFD or HFD + CAP at ambient temperature and in the TNZ measured every week from 6 weeks till 38 weeks of age using a thermometer. Average mean temperatures ± S.E.Ms. of these mice (<span class="html-italic">n</span> = 8 for each condition). ** <span class="html-italic">p</span> &lt; 0.01, significantly different.</p>
Full article ">Figure 3
<p>CAP counters HFD-induced weight gain in WT mice at ambient temperature and at thermoneutrality. Weight gain plotted against the feeding week for NCD- or HFD (±CAP, 0.01% in diet)-fed WT and UCP-1<sup>−/−</sup> mice at ambient temperature (<b>A</b>,<b>B</b>) and at thermoneutrality (<b>E</b>,<b>F</b>). Mean energy and water intake (±S.E.M.) of these mice at ambient temperature (<b>C</b>,<b>D</b>) and at thermoneutrality (<b>G</b>,<b>H</b>). (<b>I</b>). Weight gain plotted against the feeding week for NCD- or HFD (±CAP, 0.01% in diet)-fed WT mice that received the same number of calories that WT mice received at ambient temperature.</p>
Full article ">Figure 4
<p>CAP feeding increases the respiratory quotient (respiratory exchange ratio, RER= VCO2/VO2) and energy expenditure (EE) in WT mice maintained at thermoneutrality. RER (<b>A</b>,<b>C</b>), VCO2 (<b>E</b>,<b>G</b>), VO2 (<b>F</b>,<b>H</b>), EE (<b>I</b>,<b>K</b>), and locomotor activity (<b>M</b>,<b>O</b>) of NCD- or HFD (± CAP, 0.01% in diet)-fed WT and UCP-1<sup>−/−</sup> mice at thermoneutrality. Means ± S.E.Ms. for the RER (<b>B</b>,<b>D</b>), EE (<b>J</b>,<b>L</b>), and locomotor activities (<b>N</b>,<b>P</b>) of these mice. ** represents statistical significance at <span class="html-italic">p</span> &lt; 0.01 for <span class="html-italic">n</span> = 8 mice for each condition.</p>
Full article ">Figure 5
<p>Effect of CAP feeding on the mRNA levels of adipogenic and thermogenic genes in the BAT of NCD- or HFD (±CAP)-fed WT and UCP-1<sup>−/−</sup> mice at thermoneutrality. Mean mRNA levels ± S.E.Ms. for Bmp4 (<b>A</b>), Bmp8a (<b>B</b>), Bmp8b (<b>C</b>), Cidea (<b>D</b>), CoxII (<b>E</b>), Dio2 (<b>F</b>), Foxc2 (<b>G</b>), Pgc-1α (<b>H</b>), and Sirt-1 (<b>I</b>) in the BAT of these mice. For quantitative RT-PCR experiments, 18s ribosomal RNA was used as control. ** represents statistical significance at <span class="html-italic">p</span> &lt; 0.01 for <span class="html-italic">n</span> = 4 experiments.</p>
Full article ">Figure 6
<p>Effect of CAP feeding on the Ucp-1 (<b>A</b>) and Prdm-16 (<b>B</b>) mRNAs normalized to 18s RNA in NCD- or HFD (± CAP)-fed WT and UCP-1<sup>−/−</sup> mice. Mean Ucp-1 and Prdm-16 mRNA levels normalized to 18s RNA ± S.E.Ms in the BAT of these mice. ** represents statistical significance at <span class="html-italic">p</span> &lt; 0.01 for <span class="html-italic">n</span> = 4 experiments.</p>
Full article ">Figure 7
<p>CAP does not counter obesity in <sup>AD</sup>PRDM-16<sup>−/−</sup> mice. (<b>A</b>) Mean body weight gain in NCD- or HFD (± CAP, 0.01% in diet)-fed <sup>AT</sup>PRDM-16<sup>−/−</sup> mice (<span class="html-italic">n</span> = 4). (<b>B</b>) Average daily energy and water intake in these mice (<span class="html-italic">n</span> = 6). (<b>C</b>) UCP-1 mRNA levels in the sWAT and BAT of these mice (<span class="html-italic">n</span> = 6 experiments). ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 8
<p>PRDM-16 overexpression increases UCP-1 transcription and CAP treatment enhances it further. (<b>A</b>) Micrographs showing UCP-1-GFP expression in HEK TRPV1 cells from the control (basal; 1), PRDM-16 (2), PRDM-16 + CAP (1 μM; 3), PRDM-16, CPZ (10 μM; TRPV1 inhibitor) + CAP (1 μM), or CAP (1 μM) treatment groups. The magnification is 10x, and the scale bar is 100 μm. (<b>B</b>) Mean intensity of UCP-1-GFP normalized to the control (basal) group ± S.E.M. for <span class="html-italic">n</span> = 3 independent experiments. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 9
<p>Model describing the effect of TRPV1 activation by CAP on PRDM-16-dependent UCP-1 expression and therefore thermogenesis. (<b>A</b>). The HFD inhibits TRPV1 [<a href="#B10-pharmaceuticals-17-01098" class="html-bibr">10</a>], SiRT-1, and PRDM-16 [<a href="#B9-pharmaceuticals-17-01098" class="html-bibr">9</a>]. Inhibition of SiRT-1 suppresses the deacetylation of PRDM-16 [<a href="#B9-pharmaceuticals-17-01098" class="html-bibr">9</a>] and PRDM-16-dependent UCP-1 transcription. (<b>B</b>). CAP counters the effect of the HFD. (<b>C</b>). In UCP-1 KO mice, CAP activates the TRPV1-SIRT-1-PRDM-16-dependent signaling axis, but it fails to stimulate thermogenesis.</p>
Full article ">
16 pages, 2636 KiB  
Article
Adaptation of Brown Adipose Tissue in Response to Chronic Exposure to the Environmental Pollutant 1,1-Dichloro-2,2-bis(p-chlorophenyl) Ethylene (DDE) and/or a High-Fat Diet in Male Wistar Rats
by Vincenzo Migliaccio, Ilaria Di Gregorio, Serena Penna, Giuliana Panico, Assunta Lombardi and Lillà Lionetti
Nutrients 2024, 16(16), 2616; https://doi.org/10.3390/nu16162616 - 8 Aug 2024
Viewed by 660
Abstract
Brown adipose tissue (BAT) participates in thermogenesis and energy homeostasis. Studies on factors capable of influencing BAT function, such as a high-fat diet (HFD) or exposure to environmental pollutants, could be useful for finding metabolic targets for maintaining energy homeostasis. We evaluated the [...] Read more.
Brown adipose tissue (BAT) participates in thermogenesis and energy homeostasis. Studies on factors capable of influencing BAT function, such as a high-fat diet (HFD) or exposure to environmental pollutants, could be useful for finding metabolic targets for maintaining energy homeostasis. We evaluated the effect of chronic exposure to dichlorodiphenyldichloroethylene (DDE), the major metabolite of dichlorodiphenyltrichloroethane (DDT), and/or a HFD on BAT morphology, mitochondrial mass, dynamics, and oxidative stress in rats. To this end, male Wistar rats were treated for 4 weeks with a standard diet, or a HFD alone, or together with DDE. An increase in paucilocular adipocytes and the lipid droplet size were observed in HFD-treated rats, which was associated with a reduction in mitochondrial mass and in mitochondrial fragmentation, as well as with increased oxidative stress and upregulation of the superoxide dismutase-2. DDE administration mimics most of the effects induced by a HFD on BAT, and it aggravates the increase in the lipid droplet size when administered together with a HFD. Considering the known role of oxidative stress in altering BAT functionality, it could underlie the ability of both DDE and a HFD to induce similar metabolic adaptations in BAT, leading to reduced tissue thermogenesis, which can result in a predisposition to the onset of energy homeostasis disorders. Full article
(This article belongs to the Section Nutrition and Metabolism)
Show Figures

Figure 1

Figure 1
<p>Effects of DDE and/or HFD on BAT morphology, lipid droplet size, and perilipin 1 level. (<b>A</b>) BAT sections stained with hematoxylin and eosin. ND and ND+DDE groups showed a multilocular distribution of fatty depots in brown adipocytes. Both lipid droplets (LD) and nuclei (arrowhead) are evidenced in histological images. In HFD and HFD+DDE groups, a change in lipid droplet size was detected (the symbol * indicates paucilocular adipocytes). Magnification used: 20×. Scale bar applied (in ND): 50 μm. (<b>B</b>) Total lipid droplet diameter, and (<b>C</b>) the percentage of small (diameter &lt; 10 μm) medium (diameter between 10 and 30 μm), and large (diameter &gt; 30 μm) lipid droplets. (<b>D</b>) Representative image of PLIN-1 Western blot is shown, and Tubulin has been used as the loading control. (<b>E</b>) Histogram represents the quantification of Western blot data. Data were normalized to the values obtained for the ND group, set as 1, and were graphically represented as mean ± SEM of 6 animals per group. Two-way ANOVA analysis, followed by Tukey’s post-test were applied. Bars labeled with different letters are significantly different (<span class="html-italic">p</span> &lt; 0.05). (<b>F</b>) PLIN1 immunohistochemistry detected in BAT sections. Immunostaining was evidenced around droplets and in cell cytosol (black arrows). Magnification used: 20×. Scale bar applied (in ND): 50 μm. Two-way ANOVA analysis results: lipid droplet mean size; effect of the diet (<span class="html-italic">p</span> &lt; 0.0001); effect of DDE (<span class="html-italic">p</span> &lt; 0.0001); interaction (<span class="html-italic">p</span> &lt; 0.0001). Small Lipid droplet percentage: effect of diet (<span class="html-italic">p</span> &lt; 0.0001); effect of DDE (<span class="html-italic">p</span> &lt; 0.0001); interaction significant (<span class="html-italic">p</span> &lt; 0.0001). Medium lipid droplet percentage: effect of diet (<span class="html-italic">p</span> &lt; 0.0001); effect of DDE (<span class="html-italic">p</span> &lt; 0.0001); interaction (<span class="html-italic">p</span> &lt; 0.0001). Large lipid droplets: effect of diet (<span class="html-italic">p</span> &lt; 0.0001); effect of DDE (<span class="html-italic">p</span> = 0.0009); interaction (<span class="html-italic">p</span> = 0.0007). PLIN1 levels: effect of diet (<span class="html-italic">p</span> = 0.0002); effect of DDE (<span class="html-italic">p</span> &lt; 0.0001); interaction (<span class="html-italic">p</span> = 0.8152).</p>
Full article ">Figure 2
<p>Effect of DDE and/or HFD treatments on UCP1 levels and COX activity. (<b>A</b>) COX activity detected both in total homogenates and in isolated mitochondria. The symbol * indicates statistical differences between HFD and ND groups tested by using <span class="html-italic">t</span>-test analyses. (<b>B</b>) Representative histological section of UCP1 immunostaining. Immunostaining was evidenced by black arrows. (<b>C</b>) Western blot of UCP1, detected in BAT lysate and in isolated mitochondria. Tubulin has been used as a loading control. (<b>D</b>) Histograms represent the quantification of Western blot data. Data were normalized to the values obtained for the ND group, set as 1, and were graphically represented as mean ± SEM of 6 animals per group. Two-way ANOVA analysis, followed by Tukey’s post-test were applied. Bars labelled with different letters are significantly different (<span class="html-italic">p</span> &lt; 0.05). Statistical narrative results evidenced the following effects: UCP1 levels detected in BAT lysate: effect of diet n.s. (<span class="html-italic">p</span> = 0.3889); effect of DDE (<span class="html-italic">p</span> = 0.0044); interaction (<span class="html-italic">p</span> = 0.0082). UCP1 levels detected in isolated mitochondria: effect of diet (<span class="html-italic">p</span> = 0.0001); effect of DDE (<span class="html-italic">p</span> = 0.0020); the interaction n.s. (<span class="html-italic">p</span> = 0.1920). UCP1 homogenate/mitochondria ratio: effect of diet (<span class="html-italic">p</span> &lt; 0.0001); effect of DDE (<span class="html-italic">p</span> &lt; 0.0001); interaction (<span class="html-italic">p</span> &lt; 0.0001). COX activity in BAT homogenates: effect of diet (<span class="html-italic">p</span> = 0.0271); effect of DDE (<span class="html-italic">p</span> = 0.0014); interaction (<span class="html-italic">p</span> &lt; 0.0001). COX activity in BAT mitochondria: effect of diet (<span class="html-italic">p</span> &lt; 0.0001); effect of DDE (<span class="html-italic">p</span> &lt; 0.0001); interaction (<span class="html-italic">p</span> = 0.0004).</p>
Full article ">Figure 3
<p>Effect of DDE and/or HFD treatments on VDAC and PGC1α levels. (<b>A</b>) Representative Western blot of VDAC detected in BAT lysate and in isolated mitochondria. (<b>B</b>) Histograms represent the quantification of Western blot data. (<b>C</b>) Representative Western blot of PGC1-α detected in BAT lysate. (<b>D</b>) Histograms represent the quantification of Western blot data. Tubulin has been used as a loading control. Data were normalized to the values obtained for the ND group, set as 1, and were graphically represented as mean ± SEM of 6 animals per group. Two-way ANOVA analysis, followed by Tukey’s post-test were applied. Bars labelled with different letters are significantly different (<span class="html-italic">p</span> &lt; 0.05). Statistical narrative results evidenced the following effects: VDAC levels detected in BAT lysate: effect of diet (<span class="html-italic">p</span> &lt; 0.0001); effect of DDE (<span class="html-italic">p</span> &lt; 0.0001); interaction (<span class="html-italic">p</span> &lt; 0.0001). VDAC levels detected in isolated mitochondria: effect of diet not significant (n.s.) (<span class="html-italic">p</span> = 0.6777); effect of DDE n.s. (<span class="html-italic">p</span> = 0.8630); interaction n.s. (<span class="html-italic">p</span> = 0.2387). VDAC homogenate/mitochondria ratio: effect of diet (<span class="html-italic">p</span> &lt; 0.0001); effect of DDE (<span class="html-italic">p</span> = 0.0002); interaction (<span class="html-italic">p</span> = 0.0004). PGC1 alfa levels: effect of diet (<span class="html-italic">p</span> = 0.0008); effect of DDE (<span class="html-italic">p</span> = 0.0099); interaction (<span class="html-italic">p</span> = 0.0073).</p>
Full article ">Figure 4
<p>Effect of DDE and/or HFD on mitochondrial dynamics. (<b>A</b>) Representative Western Blot of MFN2 and DRP1 and their ratio detected in BAT lysate. (<b>B</b>) Histograms represent the quantification of Western blot data. (<b>C</b>) Representative Western Blot of MFN2 and DRP1 and their ratio detected in BAT isolated mitochondria. (<b>D</b>) Histograms represent the quantification of Western blot data. Tubulin has been used as a loading control. Data were normalized to the values obtained for the ND group, set as 1, and were graphically represented as mean ± SEM of 6 animals per group. Two-way ANOVA analysis, followed by Tukey’s post-test were applied. Bars labelled with different letters are significantly different (<span class="html-italic">p</span> &lt; 0.05). (<b>E</b>) MFN2 and (<b>F</b>) DRP1 immunolocalization. Black arrows indicate immunoreactive evidence of MFN2 and DRP1 localization in the cytosol of brown adipocytes. Magnification used: 20×. Scale bar applied (in ND): 50 μm. Two-way ANOVA analysis: BAT lysates MFN2 levels: effect of diet not significant (n.s.) <span class="html-italic">p</span> = 0.5726; effect of DDE n.s. <span class="html-italic">p</span> = 0.3592; interaction n.s. <span class="html-italic">p</span> = 0.2725. BAT lysates DRP1 levels: effect of diet n.s. <span class="html-italic">p</span> = 0.9103; effect of DDE n.s. <span class="html-italic">p</span> = 0.5189; interaction n.s. <span class="html-italic">p</span> = 0.2211. MFN2/DRP1 ratio detected in BAT lysate: effect of diet n.s. <span class="html-italic">p</span> = 0.9488; effect of DDE: <span class="html-italic">p</span> = 0.2093; interaction n.s. <span class="html-italic">p</span> = 0.6270. Isolated mitochondria MFN2 levels: effect of diet n.s. <span class="html-italic">p</span> = 0.1725; effect of DDE n.s. <span class="html-italic">p</span> = 0.3849; interaction n.s. <span class="html-italic">p</span> = 0.5635. Isolated mitochondria DRP1 levels: effect of diet <span class="html-italic">p</span> = 0.0404; effect of DDE <span class="html-italic">p</span> = 0.0186; interaction <span class="html-italic">p</span> = 0.0010. MFN2/DRP1 ratio detected in isolated mitochondria: effect of diet n.s. <span class="html-italic">p</span> = 0.1451; effect of DDE <span class="html-italic">p</span> = 0.0134; interaction n.s. <span class="html-italic">p</span> = 0.0235.</p>
Full article ">Figure 5
<p>Effect of DDE and/or HFD treatments on oxidative stress parameters detected in BAT. (<b>A</b>) Total ROS and (<b>B</b>) TBARS levels in total BAT homogenates. (<b>C</b>) Representative Western blot of mitochondrial SOD2. (<b>D</b>) Histograms represent the quantification of Western blot data. Tubulin has been used as a loading control. For statistical analysis, two-way ANOVA analysis, followed by Tukey’s post-test were applied. Bars labelled with different letters are significantly different (<span class="html-italic">p</span> &lt; 0.05). Statistical narrative results evidenced the following effects: ROS levels detected in BAT lysate: effect of diet (<span class="html-italic">p</span> &lt; 0.0001); effect of DDE (<span class="html-italic">p</span> &lt; 0.0001); interaction (<span class="html-italic">p</span> &lt; 0.0001). TBARS levels in BAT lysate: effect of diet (<span class="html-italic">p</span> = 0.0008); effect of DDE (<span class="html-italic">p</span> &lt; 0.0001); interaction (<span class="html-italic">p</span> = 0.0326). SOD2 levels detected in isolated mitochondria: effect of diet (<span class="html-italic">p</span> = 0.0020); effect of DDE (<span class="html-italic">p</span> = 0.0001); the interaction n.s. (<span class="html-italic">p</span> = 0.6733).</p>
Full article ">
17 pages, 5065 KiB  
Article
A Mixture of Lactobacillus HY7601 and KY1032 Regulates Energy Metabolism in Adipose Tissue and Improves Cholesterol Disposal in High-Fat-Diet-Fed Mice
by Kippeum Lee, Hyeon-Ji Kim, Joo-Yun Kim, Jae-Jung Shim and Jae-Hwan Lee
Nutrients 2024, 16(15), 2570; https://doi.org/10.3390/nu16152570 - 5 Aug 2024
Viewed by 966
Abstract
We aimed to characterize the anti-obesity and anti-atherosclerosis effects of Lactobacillus curvatus HY7601 and Lactobacillus plantarum KY1032 using high-fat diet (HFD)-fed obese C57BL/6 mice. We divided the mice into control (CON), HFD, HFD with 108 CFU/kg/day probiotics (HFD + KL, HY7301:KY1032 = [...] Read more.
We aimed to characterize the anti-obesity and anti-atherosclerosis effects of Lactobacillus curvatus HY7601 and Lactobacillus plantarum KY1032 using high-fat diet (HFD)-fed obese C57BL/6 mice. We divided the mice into control (CON), HFD, HFD with 108 CFU/kg/day probiotics (HFD + KL, HY7301:KY1032 = 1:1), and HFD with 109 CFU/kg/day probiotics (HFD + KH, HY7301:KY1032 = 1:1) groups and fed/treated them during 7 weeks. The body mass, brown adipose tissue (BAT), inguinal white adipose tissue (iWAT), and epididymal white adipose tissue (eWAT) masses and the total cholesterol and triglyceride concentrations were remarkably lower in probiotic-treated groups than in the HFD group in a dose-dependent manner. In addition, the expression of uncoupling protein 1 in the BAT, iWAT, and eWAT was significantly higher in probiotic-treated HFD mice than in the HFD mice, as demonstrated by immunofluorescence staining and Western blotting. We also measured the expression of cholesterol transport genes in the liver and jejunum and found that the expression of those encoding liver-X-receptor α, ATP-binding cassette transporters G5 and G8, and cholesterol 7α-hydroxylase were significantly higher in the HFD + KH mice than in the HFD mice. Thus, a Lactobacillus HY7601 and KY1032 mixture with 109 CFU/kg/day concentration can assist with body weight regulation through the management of lipid metabolism and thermogenesis. Full article
(This article belongs to the Special Issue Nutritional and Metabolic Changes Affecting Adipose Tissue Biology)
Show Figures

Figure 1

Figure 1
<p>Effects of <span class="html-italic">Lactobacillus</span> HY7601 and KY1032 on the body and tissue masses of the mice. (<b>A</b>) Body masses of the mice after 7 weeks. (<b>B</b>) Images of brown adipose tissue (BAT), inguinal white adipose tissue (inguinal WAT), and epididymal adipose tissue (epididymal WAT) samples. Masses of the (<b>C</b>) epididymal WAT, (<b>D</b>) inguinal WAT, (<b>E</b>) BAT, (<b>F</b>) livers, and (<b>G</b>) spleens of the mice. (<b>H</b>) Food intake and (<b>I</b>) water intake per unit of body weight. Data are expressed as mean ± SD (n = 12). Groups accompanied by different letters were significantly different: <span class="html-italic">p</span> &lt; 0.05 (a &gt; ab &gt; b &gt; bc &gt; c &gt; d). CON, control; HFD, HFD-fed obese mice; HFD-KL, 10<sup>8</sup> CFU/kg/day <span class="html-italic">Lactobacillus</span> HY7601 and KY103 plus HFD; HFD-KH, 10<sup>9</sup> CFU/kg/day <span class="html-italic">Lactobacillus</span> HY7601 and KY103 plus HFD.</p>
Full article ">Figure 2
<p>Histology of adipose tissue depots in each of the mouse groups. (<b>A</b>) Brown adipose tissue (BAT, top), inguinal white adipose tissue (WAT, middle), and epididymal white adipose tissue (bottom), stained with hematoxylin and eosin (scale bar = 50 μm). (<b>B</b>) Number of adipocytes in epididymal WAT and inguinal WAT, determined using ImageJ (version 1.53t). N = 6–8 mice/group. Data are expressed as mean ± SD (n = 12). Groups accompanied by different letters were significantly different: <span class="html-italic">p</span> &lt; 0.05 (a &gt; b &gt; c).</p>
Full article ">Figure 3
<p>Effects of <span class="html-italic">Lactobacillus</span> HY7601 and KY1032 on serum lipid and cholesterol-related parameters in HFD-fed mice. Serum concentrations of (<b>A</b>) adiponectin, (<b>B</b>) triglyceride (TG), (<b>C</b>) total cholesterol (T-Chol), (<b>D</b>) low-density lipoprotein cholesterol (LDL-Chol), (<b>E</b>) high-density lipoprotein cholesterol (HDL-Chol), and (<b>F</b>) blood urea nitrogen (BUN). Data are mean ± SD (n = 12). Groups accompanied by different letters were significantly different: <span class="html-italic">p</span> &lt; 0.05 (a &gt; ab &gt; b &gt; c).</p>
Full article ">Figure 4
<p>Effects of <span class="html-italic">Lactobacillus</span> HY7601 and KY1032 on the rectal temperature and serum parameters related to glucose metabolism in HFD-fed mice. (<b>A</b>) Core temperature, measured using a thermometer. Serum concentrations of (<b>B</b>) glucose, (<b>C</b>) glycated or glycosylated hemoglobin A1c (HbA1C), (<b>D</b>) creatine kinase (CK), and (<b>E</b>) lactate. Data are expressed as mean ± SD (n = 12). Groups accompanied by different letters were significantly different: <span class="html-italic">p</span> &lt; 0.05 (a &gt; ab &gt; b &gt; c).</p>
Full article ">Figure 5
<p>Adipose sections immunostained for uncoupling protein 1 (UCP1) in the various mouse groups. Representative images of (<b>A</b>) BAT and (<b>B</b>) inguinal WAT (bright field, UCP1 (red), DAPI (blue), and merged UCP1 and DAPI).</p>
Full article ">Figure 6
<p>Effects of <span class="html-italic">Lactobacillus</span> HY7601 and KY1032 on the expression of key thermogenic proteins in (<b>A</b>) BAT, (<b>B</b>) inguinal WAT, and (<b>C</b>) epididymal WAT. Western blot data for sirtuin 1 (SirT1), UCP1, phosphorylated-AMP-activated protein kinase (<span class="html-italic">p</span>-AMPK), AMPK, peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC1α), and glyceraldehyde 3-phosphate dehydrogenase (GAPDH). Groups accompanied by different letters were significantly different: <span class="html-italic">p</span> &lt; 0.05 (a &gt; ab &gt; b &gt; c).</p>
Full article ">Figure 7
<p>Effects of <span class="html-italic">Lactobacillus</span> HY7601 and KY1032 on the circulating concentrations of liver-related enzymes and the liver mRNA expression of the mouse groups. Serum activities of (<b>A</b>) alanine transferase (ALT), (<b>B</b>) aspartate transaminase (AST), and (<b>C</b>) lactate dehydrogenase (LDH). Expression of the genes encoding (<b>D</b>) 3-hydroxy-3-methylglutaryl-coenzyme A reductase (<span class="html-italic">Hmgcr</span>), (<b>E</b>) sterol regulatory element-binding protein 2 (<span class="html-italic">Srdbp2</span>), (<b>F</b>) peroxisome proliferator-activated receptor alpha (<span class="html-italic">PPARa</span>), (<b>G</b>) ATP-binding cassette (ABC) transporter G5 (<span class="html-italic">Abcg5</span>), (<b>H</b>) <span class="html-italic">Agcg8</span>, (<b>I</b>) liver X receptor alpha (<span class="html-italic">LXRb</span>), (<b>J</b>) <span class="html-italic">LXRβ</span>, and (<b>K</b>) cholesterol 7 alpha-hydroxylase (<span class="html-italic">Cyp7a1</span>), normalized to that of <span class="html-italic">Gapdh</span>. Groups accompanied by different letters were significantly different: <span class="html-italic">p</span> &lt; 0.05 (a &gt; ab &gt; b &gt; bc &gt; c).</p>
Full article ">Figure 8
<p>Effects of the <span class="html-italic">Lactobacillus</span> HY7601 and KY1032 mixture on cholesterol-metabolism-related parameters in the jejuna of the mouse groups. Expression of genes encoding (<b>A</b>) ATP-binding cassette (ABC) transporter G5 (<span class="html-italic">Abcg5</span>), (<b>B</b>) <span class="html-italic">Agcg8</span>, (<b>C</b>) liver X receptor alpha (<span class="html-italic">LXRα</span>), and (<b>D</b>) NPC1-like intracellular cholesterol transporter 1 (<span class="html-italic">Npcl1</span>), normalized to that of <span class="html-italic">Gapdh</span>. Fecal concentrations of (<b>E</b>) bile acids and (<b>F</b>) total cholesterol. Groups accompanied by different letters were significantly different: <span class="html-italic">p</span> &lt; 0.05 (a &gt; ab &gt; b &gt; c).</p>
Full article ">
13 pages, 1507 KiB  
Article
TLCD4 as Potential Transcriptomic Biomarker of Cold Exposure
by Bàrbara Reynés, Estefanía García-Ruiz, Evert M. van Schothorst, Jaap Keijer, Paula Oliver and Andreu Palou
Biomolecules 2024, 14(8), 935; https://doi.org/10.3390/biom14080935 - 1 Aug 2024
Viewed by 571
Abstract
(1) Background: Cold exposure induces metabolic adaptations that can promote health benefits, including increased energy disposal due to lipid mobilization in adipose tissue (AT). This study aims to identify easily measurable biomarkers mirroring the effect of cold exposure on AT. (2) Methods: Transcriptomic [...] Read more.
(1) Background: Cold exposure induces metabolic adaptations that can promote health benefits, including increased energy disposal due to lipid mobilization in adipose tissue (AT). This study aims to identify easily measurable biomarkers mirroring the effect of cold exposure on AT. (2) Methods: Transcriptomic analysis was performed in peripheral blood mononuclear cells (PBMCs) and distinct AT depots of two animal models (ferrets and rats) exposed to cold, and in PBMCs of cold-exposed humans. (3) Results: One week of cold exposure (at 4 °C) affected different metabolic pathways and gene expression in the AT of ferrets, an animal model with an AT more similar to humans than that of rodents. However, only one gene, Tlcd4, was affected in the same way (overexpressed) in aortic perivascular and inguinal AT depots and in PBMCs, making it a potential biomarker of interest. Subsequent targeted analysis in rats showed that 1 week at 4 °C also induced Tlcd4 expression in brown AT and PBMCs, while 1 h at 4 °C resulted in reduced Tlcd4 mRNA levels in retroperitoneal white AT. In humans, no clear effects were observed. Nevertheless, decreased PBMC TLCD4 expression was observed after acute cold exposure in women with normal weight, although this effect could be attributed to short-term fasting during the procedure. No effect was evident in women with overweight or in normal-weight men. (4) Conclusions: Our results obtained for different species point toward TLCD4 gene expression as a potential biomarker of cold exposure/fat mobilization that could tentatively be used to address the effectiveness of cold exposure-mimicking therapies. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Schematic overview of regulated genes in aPVAT, IAT, and PBMCs in control (22 °C) vs. one-week cold-exposed (4 °C) male ferrets. The overlapping genes between aPVAT, IAT, and PBMCs are those equally regulated. Student’s <span class="html-italic">t</span>-test, <span class="html-italic">p</span>-value &lt; 0.05 (<b>A</b>) or <span class="html-italic">p</span>-value &lt; 0.01 (<b>C</b>). The detailed manual classification of the 36 genes equally regulated in the three types of samples (Student’s <span class="html-italic">t</span>-test, <span class="html-italic">p</span>-value &lt; 0.05) (<b>B</b>). <span class="html-italic">TLCD4</span> gene expression in aPVAT, IAT, and PBMCs of the same set of ferrets (<b>D</b>). The mRNA expression was measured by real-time RT-qPCR. The results represent the means ± SEM (<span class="html-italic">n</span> = 4–7) of the ratios of specific mRNA levels relative to <span class="html-italic">Mettl2b</span>, expressed as a percentage, where the control group was set as 100%. Statistics: the * symbol shows the significance of the cold-exposed groups vs. the control group (U Mann–Whitney, <span class="html-italic">p</span> &lt; 0.05, or indicated when different).</p>
Full article ">Figure 2
<p><span class="html-italic">Tlcd4</span> gene expression in the inguinal and retroperitoneal WAT, in BAT and PBMCs of female rats of different ages (from 1 to 6 months) housed at different room temperatures: 22 °C (Control) or 4 °C for one week (Cold). mRNA expression was measured by real-time RT-qPCR. Results represent means ± SEM (<span class="html-italic">n</span> = 4–6) of ratios of specific mRNA levels relative to <span class="html-italic">Lrp10</span>, expressed as a percentage of the value of 1-month-old control animals that was set to 100%. Statistics: the * symbol shows the significance of the cold-exposed groups vs. the control group (U Mann–Whitney, <span class="html-italic">p</span> &lt; 0.05, or indicated when different).</p>
Full article ">Figure 3
<p><span class="html-italic">Tlcd4</span> gene expression in BAT, retroperitoneal (RWAT), inguinal WAT (IAT), and PBMCs of female rats housed at different room temperatures: 22 °C (Control) or 4 °C for 1 (Cold 1 h) or 2 h (Cold 2 h). mRNA expression was measured by real-time RT-qPCR. Results represent means ± SEM (<span class="html-italic">n</span> = 6–8) of ratios of specific mRNA levels relative to <span class="html-italic">Lrp10</span> or <span class="html-italic">Gdi</span>, expressed as a percentage, where the control group was set to 100%. Statistics: the * symbol shows the significance of the cold-exposed groups vs. the control group (U Mann–Whitney, <span class="html-italic">p</span> &lt; 0.05, or indicated when different). Bars not sharing common letters (a, b) are significantly different (one-way ANOVA, <span class="html-italic">p</span> &lt; 0.05). DMS post hoc was used after one-way ANOVA analysis.</p>
Full article ">Figure 4
<p>Effects of cold exposure on <span class="html-italic">TLCD4</span> gene expression in PBMCs of women and men with normal weight, and women with overweight or obesity. <span class="html-italic">TLCD4</span> mRNA expression was measured by real-time RT-qPCR in PBMC samples collected 4 h (T1) and 6 h (T2) after feeding, with (<b>A</b>) or without (<b>B</b>) cold exposure in this timelapse. Results represent means ± SEM (<span class="html-italic">n</span> = 4–8) of ratios of specific mRNA levels relative to <span class="html-italic">RPLP0</span>, expressed as a percentage, where T1 was set as 100%. Statistics: the * symbol shows the significance of the cold-exposed groups vs. the control group (Wilcoxon test, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
16 pages, 1751 KiB  
Review
Adipose Tissue Dysfunction Related to Climate Change and Air Pollution: Understanding the Metabolic Consequences
by Radoslav Stojchevski, Preethi Chandrasekaran, Nikola Hadzi-Petrushev, Mitko Mladenov and Dimiter Avtanski
Int. J. Mol. Sci. 2024, 25(14), 7849; https://doi.org/10.3390/ijms25147849 - 18 Jul 2024
Viewed by 1361
Abstract
Obesity, a global pandemic, poses a major threat to healthcare systems worldwide. Adipose tissue, the energy-storing organ during excessive energy intake, functions as a thermoregulator, interacting with other tissues to regulate systemic metabolism. Specifically, brown adipose tissue (BAT) is positively associated with an [...] Read more.
Obesity, a global pandemic, poses a major threat to healthcare systems worldwide. Adipose tissue, the energy-storing organ during excessive energy intake, functions as a thermoregulator, interacting with other tissues to regulate systemic metabolism. Specifically, brown adipose tissue (BAT) is positively associated with an increased resistance to obesity, due to its thermogenic function in the presence of uncoupled protein 1 (UCP1). Recently, studies on climate change and the influence of environmental pollutants on energy homeostasis and obesity have drawn increasing attention. The reciprocal relationship between increasing adiposity and increasing temperatures results in reduced adaptive thermogenesis, decreased physical activity, and increased carbon footprint production. In addition, the impact of climate change makes obese individuals more prone to developing type 2 diabetes mellitus (T2DM). An impaired response to heat stress, compromised vasodilation, and sweating increase the risk of diabetes-related comorbidities. This comprehensive review provides information about the effects of climate change on obesity and adipose tissue, the risk of T2DM development, and insights into the environmental pollutants causing adipose tissue dysfunction and obesity. The effects of altered dietary patterns on adiposity and adaptation strategies to mitigate the detrimental effects of climate change are also discussed. Full article
(This article belongs to the Special Issue Lipidomics and Lipid Metabolism in Health and Disease)
Show Figures

Figure 1

Figure 1
<p>Adipose tissue responses to thermal challenges. (<b>a</b>) In BAT, polarized macrophages, due to cold exposure, directly activate beta-adrenergic signaling, thereby increasing heat production. (<b>b</b>) Energy expenditure is triggered by the body’s cooling mechanisms when the temperature exceeds the thermoneutral zone. (<b>c</b>) The elevated thermogenic activity of brown and beige adipocytes due to cold exposure is a result of increased glucose and free fatty acid uptake. (<b>d</b>) The browning of WAT is an important mechanism in which cold exposure triggers an increase in the oxidative metabolic rates of brown and beige adipocytes. This is essential for maintaining core body temperature during prolonged cold exposure. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 2
<p>Addressing global challenges related to obesity and climate. (<b>a</b>) Obesity is triggered by high levels of greenhouse gas emissions due to increased food intake. (<b>b</b>) An important factor contributing to obesity is the increased time spent in the thermoneutral zone and decreased thermogenesis as a consequence. (<b>c</b>) A protein-rich diet is recommended due to it having the highest effect on diet-induced thermogenesis. (<b>d</b>) The increase in the consumption of processed foods is attributed to increased temperatures negatively impacting crop yields and agriculture. (<b>e</b>) Physical inactivity, as a result of extreme temperatures, contributes to weight gain. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 3
<p>Effects of heat exposure and air pollution on insulin resistance and diabetes complications. (<b>a</b>) Exposure to high temperatures increases sweat secretion and peripheral vasodilation. This, in turn, dissipates heat and maintains optimal body temperature. (<b>b</b>) Impairment of insulin signaling stimulates insulin resistance in various tissues due to dehydration caused by heat exposure. (<b>c</b>) Glucose intolerance is a consequence of disruption of thermoregulation by impairment of orthostatic response. (<b>d</b>) Air pollutants, specifically PM2.5, increase risk of glucose intolerance and type 2 DM-associated cardiovascular diseases. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">
19 pages, 6813 KiB  
Article
Dietary Pectin from Premna microphylla Turcz Leaves Prevents Obesity by Regulating Gut Microbiota and Lipid Metabolism in Mice Fed High-Fat Diet
by Jiaobei Gao, Mengxue Zhang, Li Zhang, Nan Wang, Yan Zhao, Daoyuan Ren and Xingbin Yang
Foods 2024, 13(14), 2248; https://doi.org/10.3390/foods13142248 - 17 Jul 2024
Viewed by 700
Abstract
The present study was designed to investigate the protective effects of pectin extracted from Premna microphylla Turcz leaves (PTP) against high-fat-diet (HFD)-induced lipid metabolism disorders and gut microbiota dysbiosis in obese mice. PTP was made using the acid extraction method, and it was [...] Read more.
The present study was designed to investigate the protective effects of pectin extracted from Premna microphylla Turcz leaves (PTP) against high-fat-diet (HFD)-induced lipid metabolism disorders and gut microbiota dysbiosis in obese mice. PTP was made using the acid extraction method, and it was found to be an acidic pectin that had relative mole percentages of 32.1%, 29.2%, and 26.2% for galacturonic acid, arabinose, and galactose, respectively. The administration of PTP in C57BL/6J mice inhibited the HFD-induced abnormal weight gain, visceral obesity, and dyslipidemia, and also improved insulin sensitivity, as revealed by the improved insulin tolerance and the decreased glucose levels during an insulin sensitivity test. These effects were linked to increased energy expenditure, as demonstrated by the upregulation of thermogenesis-related protein UCP1 expression in the brown adipose tissue (BAT) of PTP-treated mice. 16S rRNA gene sequencing revealed that PTP dramatically improved the HFD-induced gut dysbiosis by lowering the ratio of Firmicutes to Bacteroidetes and the quantity of potentially harmful bacteria. These findings may provide a theoretical basis for us to understand the functions and usages of PTP in alleviating obesity. Full article
(This article belongs to the Section Nutraceuticals, Functional Foods, and Novel Foods)
Show Figures

Figure 1

Figure 1
<p>Chemical analysis of pectin from dietary <span class="html-italic">Premna microphylla</span> Turcz leaves (PTP). (<b>A</b>) FT-IR spectra. (<b>B</b>) Scanning electron microscope images at different magnifications. The HPLC chromatograms of PMP (1-pheny-3-methyl-5-pyrazolone) derivatives of eight standard monosaccharides (<b>C</b>) and component monosaccharides released by hydrolyzing PTP as pectin (<b>D</b>). Peaks: (1) mannose, (2) ribose, (3) rhamnose, (4) glucuronic acid, (5) galacturonic acid, (6) glucose, (7) galactose, (8) arabinose.</p>
Full article ">Figure 2
<p>Effects of PTP on body weight, water intake, food consumption, and basic parameters for 12 consecutive weeks of high-fat diet (HFD) feeding in mice (<b>A</b>–<b>F</b>). Hepatosomatic index (HI) = liver weight (g)/body weight (g); fat index (FI) = total fat weight (g)/body weight (g); total fat weight (g) = epididymal white adipose tissue (eWAT) weight (g) + inguinal white adipose tissue (iWAT) weight (g) + mesentery adipose tissue (MAT) weight (g). One-way ANOVA followed by Tukey’s multiple comparisons test was performed for all groups. Values are expressed as means ± SD (<span class="html-italic">n</span> = 8). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 and ns indicates no significant difference.</p>
Full article ">Figure 3
<p>Effects of PTP on histopathological changes in hepatocytes in the liver and three different adipocytes stained with H&amp;E, and original magnification 200×. The diameters of adipocytes were determined by ImageJ software (version 1.54) (<span class="html-italic">n</span> = 8 per group). (<b>A</b>) Histopathological alterations of the livers and the adipose tissues stained by H&amp;E (original magnification of 400×). (<b>B</b>–<b>D</b>) The diameters of epididymal white adipose tissue (eWAT) and inguinal white adipose tissue (iWAT) as well as brown adipose tissue (BAT), respectively. # <span class="html-italic">p</span> &lt; 0.05 and ### <span class="html-italic">p</span> &lt; 0.001, versus the ND mice. ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001, versus the HFD mice.</p>
Full article ">Figure 4
<p>Effects of PTP on lipid metabolism and its related factors, adiponectin (ADPN) and lipopolysaccharide (LPS), in HFD-fed mice. (<b>A</b>–<b>F</b>) Serum TC, TG, LDL-C, HDL-C, ADPN, and LPS levels. (<b>G</b>–<b>J</b>) Hepatic TC, TG, LDL-C, and HDL-C levels. Data are expressed as mean ± SD (<span class="html-italic">n</span> = 8 per group). One-way ANOVA followed by Tukey’s multiple comparisons test were performed for all groups. # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01, and ### <span class="html-italic">p</span> &lt; 0.001, versus the ND mice. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 and ns indicates no significant difference, versus the HFD mice.</p>
Full article ">Figure 5
<p>Effects of PTP on glucose tolerance (OGTT) and insulin sensibility test (IST) in HFD-fed mice. (<b>A</b>) OGTT. (<b>B</b>) IST. (<b>C</b>,<b>D</b>) Areas under the curve (AUCs) for OGTT and IST. Data are expressed as mean ± SD (<span class="html-italic">n</span> = 8 per group). One-way ANOVA followed by Tukey’s multiple comparisons test for were performed for all groups. ### <span class="html-italic">p</span> &lt; 0.001, versus the ND mice. * <span class="html-italic">p</span> &lt; 0.05, versus the HFD mice using Tukey’s multiple comparisons test for all groups. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Expression of thermogenic protein UCP1 in inguinal white adipose tissue (iWAT) and brown adipose tissue (BAT), and original magnification 400×. (<b>A</b>) iWAT. (<b>B</b>) BAT. (<b>C</b>) Score of immunohistochemical analysis (<span class="html-italic">n</span> = 8 per group). ### <span class="html-italic">p</span> &lt; 0.001, versus the ND mice. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001, versus the HFD mice.</p>
Full article ">Figure 7
<p>Effects of PTP on intestinal microflora structure of mice fed with HFD. (<b>A</b>) Venn diagram of bacteria detected at the ASV level. (<b>B</b>,<b>C</b>) Beta diversity analysis of intestinal microbiota using the non-metric multidimensional scaling (NMDS) and principal co-ordinates analysis (PCoA). (<b>D</b>) LEfSe analysis of microbiota. (<b>E</b>) Bacterial community at the phylum level.</p>
Full article ">Figure 8
<p>Effect of PTP on the abundance of intestinal flora in HFD-fed mice. (<b>A</b>) Relative abundance of <span class="html-italic">Faecalibaculum</span> at the genus level. (<b>B</b>) Relative abundance of <span class="html-italic">Romboutsia</span> at the genus level. (<b>C</b>) Relative abundance of <span class="html-italic">Enterorhabdus</span> at the genus level. (<b>D</b>) Relative abundance of <span class="html-italic">norank-f-Muribaculaceae</span> at the genus level. (<b>E</b>) Relative abundance of <span class="html-italic">Firmicutes</span> at the phylum level. (<b>F</b>) Relative abundance of <span class="html-italic">Bacteroidota</span> at the phylum level. (<b>G</b>) Heatmap comparison and hierarchical clustering dendrogram based on the relative abundance at the genus level. One-way ANOVA followed. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001, versus the HFD mice.</p>
Full article ">
17 pages, 3019 KiB  
Article
Role of Brown Adipose Tissue in Metabolic Health and Efficacy of Drug Treatment for Obesity
by Natalia O. Markina, Georgy A. Matveev, German G. Zasypkin, Tatiana I. Golikova, Daria V. Ryzhkova, Yulia A. Kononova, Sergey D. Danilov and Alina Yu. Babenko
J. Clin. Med. 2024, 13(14), 4151; https://doi.org/10.3390/jcm13144151 - 16 Jul 2024
Viewed by 939
Abstract
(1) Background: Brown adipose tissue (BAT) is responsible for non-shivering thermogenesis, and its activation has become a new object as both a determinant of metabolic health and a target for therapy. This study aimed to identify the relationships between the presence of [...] Read more.
(1) Background: Brown adipose tissue (BAT) is responsible for non-shivering thermogenesis, and its activation has become a new object as both a determinant of metabolic health and a target for therapy. This study aimed to identify the relationships between the presence of BAT, parameters that characterize metabolic health (glucose, lipids, blood pressure (BP)), and the dynamics of body mass index (BMI) during weight-reducing therapy. (2) Methods: The study included 72 patients with obesity. We investigated metabolic parameters, anthropometric parameters, and BP. Dual-energy X-ray absorptiometry (DXA) and positron emission tomography and computed tomography (PET/CT) imaging with 18F-fluorodeoxyglucose (18F-FDG) were performed. (3) Results: Before weight-reducing therapy, BAT was revealed only in 19% patients with obesity. The presence of BAT was associated with a lower risk of metabolic deviations that characterize metabolic syndrome: shorter waist circumference (WC) (p = 0.02) and lower levels of glucose (p = 0.03) and triglycerides (p = 0.03). Thereafter, patients were divided into four groups according to the type of therapy (only lifestyle modification or with Liraglutide or Reduxin or Reduxin Forte). We did not find a relationship between the presence of BAT and response to therapy: percent weight reduction was 10.4% in patients with BAT and 8.5% in patients without BAT (p = 0.78) during six months of therapy. But we noted a significant positive correlation between the volume of BAT and the effectiveness of weight loss at 3 months (r = 0.52, p = 0.016). The dynamic analysis of BAT after 6 months of therapy showed a significant increase in the volume of cold-induced metabolically active BAT, as determined by PET/CT with 18F-FDG in the Liraglutide group (p = 0.04) and an increase in the activity of BAT standardized uptake value (SUV mean and SUV max) in the Reduxin (p = 0.02; p = 0.01, respectively) and Liraglutide groups (p = 0.02 in both settings). (4) Conclusions: The presence of brown adipose tissue is associated with a lower risk of metabolic abnormalities. In general, our study demonstrated that well-established drugs in the treatment of obesity (Liraglutide and Reduxin) have one more mechanism for implementing their effects. These drugs have the ability to increase the activity of BAT. A significant positive relationship between the total volume of BAT and the percentage of weight loss may further determine the priority mechanism of the weight-reducing effect of these medicaments. Full article
(This article belongs to the Topic Metabolic Syndrome, Biomarkers and Lifestyles)
Show Figures

Figure 1

Figure 1
<p>Dynamics of weight loss percentage in patients with and without BAT (<b>A</b>). Trend in weight loss in kilograms in patients with BAT versus without BAT (<b>B</b>).</p>
Full article ">Figure 2
<p>Changes in triglyceride lowering after 6 months of therapy in patients with BAT versus without BAT. Abbreviations: BAT—brown adipose tissue, TG—triglycerides.</p>
Full article ">Figure 3
<p>Percentage weight loss in patients with BAT versus without BAT in lifestyle modification group (<b>A</b>). Weight loss in kilograms in patients with and without BAT in lifestyle modification group (<b>B</b>).</p>
Full article ">Figure 4
<p>Percentage weight loss in patients with BAT versus without BAT who received Reduxin (<b>A</b>). Weight loss in kilograms in patients with and without BAT who received Reduxin (<b>B</b>).</p>
Full article ">
15 pages, 3274 KiB  
Article
Carnosic Acid (CA) Induces a Brown Fat-like Phenotype, Increases Mitochondrial Biogenesis, and Activates AMPK in 3T3-L1 Adipocytes
by Filip Vlavcheski, Rebecca E. K. MacPherson, Val Fajardo, Newman Sze and Evangelia Tsiani
Biomedicines 2024, 12(7), 1569; https://doi.org/10.3390/biomedicines12071569 - 15 Jul 2024
Viewed by 718
Abstract
Adipose tissue plays a crucial role in regulating metabolic homeostasis, and its dysfunction in obesity leads to insulin resistance and type 2 diabetes (T2D). White adipose tissue (WAT) primarily stores energy as lipids, while brown adipose tissue (BAT) regulates thermogenesis by dissipating energy [...] Read more.
Adipose tissue plays a crucial role in regulating metabolic homeostasis, and its dysfunction in obesity leads to insulin resistance and type 2 diabetes (T2D). White adipose tissue (WAT) primarily stores energy as lipids, while brown adipose tissue (BAT) regulates thermogenesis by dissipating energy as heat. The process of browning involves the transdifferentiation of WAT into brown-like or beige adipocytes, which exhibit a similar phenotype as BAT. The browning of WAT is an attractive approach against obesity and T2D, and the activation of the energy sensor AMP-activated protein kinase (AMPK) has been shown to play a role in browning. Carnosic acid (CA), a polyphenolic diterpene, found in many plants including rosemary, is reported to possess potent antioxidant, anti-inflammatory, and anti-hyperglycemic properties. The limited evidence available indicates that CA activates AMPK and may have anti-obesity and antidiabetic potential; however, the effects in adipocyte browning remain largely unexplored. This study aimed to examine the effects of CA on the markers of adipocyte browning. The treatment of 3T3L1 adipocytes with CA activated AMPK, reduced lipid accumulation, and increased the expression of browning protein markers (UCP-1, PGC-1α, PRDM16, and TFAM) and mitochondrial biogenesis. The use of compound C, an AMPK inhibitor, significantly attenuated the effects of CA, indicating AMPK involvement. These studies demonstrate that CA can activate AMPK and stimulate the browning of white adipocytes. Future animal and human studies are required to examine the effects of CA in vivo. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of CA on 3T3-L1 adipocyte lipid content. Fully differentiated 3T3-L1 adipocytes were pretreated for 1 h without (C) or with CC (25 mM), followed by a treatment without (C) or with CA (10 μM) or MET (5 mM) for 24 h in serum-deprived media. After treatment: (<b>A</b>) The cells were stained with Oil Red O (ORO) and microscopic images were taken using the color field filter on a Cytation Gen5 multimode imaging microscope (×20). (<b>B</b>) Oil Red O was extracted from the cells and the intensity of the supernatant was measured at 490 nm using an ELISA plate reader. (<b>C</b>) Fully differentiated 3T3-L1 adipocytes were treated with CA (10 μM), β<sub>3</sub>-adrenergic agonist (CL 316 243) (1 μM), or the PPARγ activator rosiglitazone (ROSI) (10 μM) for 24 h followed by ORO stain and absorbance measurements. The results are the mean ± standard error (SE) of four to six independent experiments, expressed as a percent of the control: ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. control and ### <span class="html-italic">p</span> &lt; 0.001, as indicated.</p>
Full article ">Figure 2
<p>Effects of CA on mitochondrial density: Fully differentiated 3T3-L1 adipocytes were treated without (C) or with CA (10 μM) in the presence or absence of CC for 24 h followed by exposure to 250 nM MitoTracker reagent and 2.5 mg/mL Hoechst blue for 30 min. The cells were then fixed and visualized with Cytation5 using TexasRed (abs/em 644/665 nm) and DAPI/Hoechst filter. Pictures of the plate were taken automatically at the same time using the Cytation5 recommended protocol using the Hoechst/DAPI filter to detect the nuclei (<b>A</b>). The intensity of the red florescence was expressed in arbitrary units (<b>B</b>). Hoechst blue images were merged with the MitoTracker Red and pictures were created. The data are the mean ± SE of five to six separate experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. control, # <span class="html-italic">p</span> &lt; 0.05, and ### <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>CA and MET increase AMPK and ACC phosphorylation in 3T3-L1 adipocytes. Fully differentiated 3T3-L1 adipocytes were incubated without (C) or with carnosic acid (CA) (10 μM) or metformin (MET) (5 mM) for 24 h in the absence or presence of compound C (CC) (25 μM). After treatment, the cells were lysed and SDS-PAGE was performed, followed by immunoblotting using specific antibodies to recognize the total and phosphorylated (Thr172) levels of AMPK and ACC. Representative blots are shown (<b>A</b>,<b>B</b>). The densitometry of the bands was measured and expressed in arbitrary units as a percent of the control (<b>C</b>,<b>D</b>). The data are the mean ± SE of seven to eight separate experiments., *** <span class="html-italic">p</span> &lt; 0.001 vs. control, ## <span class="html-italic">p</span> &lt; 0.01, ### <span class="html-italic">p</span> &lt; 0.001, as indicated.</p>
Full article ">Figure 4
<p>Effects of CA on UCP-1 levels. Fully differentiated 3T3-L1 adipocytes were pretreated for 1 h without (C) or with CC (25 mM), followed by treatment without (C) or with CA (10 μM) or MET (5 mM) for 24 h in serum-deprived media. After treatment, the cells are lysed and SDS-PAGE was performed followed by immunoblotting using specific antibodies to recognize the total levels of UCP-1 or β-actin and immunostaining using an anti-UCP-1 primary antibody and AlexaFluor488 secondary antibody. Hoechst blue stain was used to label the nuclei. Representative blots are shown (<b>A</b>). The densitometry of the bands was measured and expressed in arbitrary units as a percent of the control (<b>B</b>). Images were taken with Cytation5, a florescence microscope using Green Florescent Protein (GFP) and DAPI/Hoechst filter (<b>C</b>). The intensity of the green florescence was measured using ImageJ and is expressed in arbitrary units (<b>D</b>). The data are the mean ± SE of six to eleven separate experiments. ** <span class="html-italic">p</span> &lt; 0.01 vs. control, # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01, as indicated.</p>
Full article ">Figure 5
<p>Effects of CA on browning markers: Fully differentiated 3T3-L1 adipocytes were incubated without (C) or with CA (10 μM) or MET (5 mM) for 24 h in the absence and presence of CC (25 μM). After treatment, the cells are lysed and SDS-PAGE was performed, which was followed by immunoblotting using specific antibodies to recognize the total levels of PPARγ, PRDM18, PGC-1α, TFAM, or β-actin. Representative blots are shown (<b>A</b>). The densitometry of the bands was measured and is expressed in arbitrary units as a percent of the control (<b>B</b>–<b>E</b>). The data are the mean ± SE of seven to nine separate experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. control, # <span class="html-italic">p</span> &lt; 0.05, and ## <span class="html-italic">p</span> &lt; 0.01, as indicated.</p>
Full article ">Figure 6
<p>Effects of CA and CC on 3T3-L1 adipocyte viability. (<b>A</b>) Fully differentiated adipocytes were treated without (C) or with a range of CA concentrations (5 to 100 μM) or with their corresponding vehicle (DMSO) concentrations for 24 h followed by incubation with MTT. The formazan dye was then solubilized, and absorbance was measured at 570 nm. Cell viability is expressed as a percent of the control (C) untreated cells (<b>B</b>) Fully differentiated adipocytes were treated without (C) or with CA in the absence or presence of CC for 24 h followed by MTT assay. The dye was solubilized and read at 570 nm. The values were expressed as a percent of the control and are the mean ± SEM of three independent experiments.</p>
Full article ">Figure 7
<p>Effects of CA and MET on GSK3β. Fully differentiated 3T3-L1 adipocytes were incubated without (C) or with CA (10 μM) or MET (5 mM) for 24 h in the absence or the presence of CC (25 μM). After the treatment, the cells are lysed and SDS-PAGE was performed, which was followed by immunoblotting using specific antibodies to recognize the total and phosphorylated (Ser9) levels of GSK3β. Representative blots are shown (<b>A</b>). The densitometry of the bands was measured and expressed as a percent of control (<b>B</b>). The data are the mean ± SE of two separate experiments. * <span class="html-italic">p</span> &lt; 0.05, and ## <span class="html-italic">p</span> &lt; 0.01, as indicated.</p>
Full article ">Figure 8
<p>CA activated AMPK, inhibited GSK3β, and increased the expression of browning (UCP-1, PRDM16, and PPARγ) and mitochondrial biogenesis protein markers (PGC-1α and TFAM). Use of compound C, an AMPK inhibitor, significantly attenuated the effects of CA. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">
20 pages, 1198 KiB  
Review
Metabolic Effects of Ketogenic Diets: Exploring Whole-Body Metabolism in Connection with Adipose Tissue and Other Metabolic Organs
by Yusra Ahmad, Dong Soo Seo and Younghoon Jang
Int. J. Mol. Sci. 2024, 25(13), 7076; https://doi.org/10.3390/ijms25137076 - 27 Jun 2024
Cited by 2 | Viewed by 2902
Abstract
The ketogenic diet (KD) is characterized by minimal carbohydrate, moderate protein, and high fat intake, leading to ketosis. It is recognized for its efficiency in weight loss, metabolic health improvement, and various therapeutic interventions. The KD enhances glucose and lipid metabolism, reducing triglycerides [...] Read more.
The ketogenic diet (KD) is characterized by minimal carbohydrate, moderate protein, and high fat intake, leading to ketosis. It is recognized for its efficiency in weight loss, metabolic health improvement, and various therapeutic interventions. The KD enhances glucose and lipid metabolism, reducing triglycerides and total cholesterol while increasing high-density lipoprotein levels and alleviating dyslipidemia. It significantly influences adipose tissue hormones, key contributors to systemic metabolism. Brown adipose tissue, essential for thermogenesis and lipid combustion, encounters modified UCP1 levels due to dietary factors, including the KD. UCP1 generates heat by uncoupling electron transport during ATP synthesis. Browning of the white adipose tissue elevates UCP1 levels in both white and brown adipose tissues, a phenomenon encouraged by the KD. Ketone oxidation depletes intermediates in the Krebs cycle, requiring anaplerotic substances, including glucose, glycogen, or amino acids, for metabolic efficiency. Methylation is essential in adipogenesis and the body’s dietary responses, with DNA methylation of several genes linked to weight loss and ketosis. The KD stimulates FGF21, influencing metabolic stability via the UCP1 pathways. The KD induces a reduction in muscle mass, potentially involving anti-lipolytic effects and attenuating proteolysis in skeletal muscles. Additionally, the KD contributes to neuroprotection, possesses anti-inflammatory properties, and alters epigenetics. This review encapsulates the metabolic effects and signaling induced by the KD in adipose tissue and major metabolic organs. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Highlights of the molecular aspects of each organ, including adipose tissue, on the impact of the ketogenic diet on body metabolism (red arrow indicates the effect of KD on molecular aspects; dark arrows indicate metabolic effects or molecular aspects).</p>
Full article ">
35 pages, 1938 KiB  
Review
White-to-Beige and Back: Adipocyte Conversion and Transcriptional Reprogramming
by Stanislav Boychenko, Vera S. Egorova, Andrew Brovin and Alexander D. Egorov
Pharmaceuticals 2024, 17(6), 790; https://doi.org/10.3390/ph17060790 - 16 Jun 2024
Viewed by 1797
Abstract
Obesity has become a pandemic, as currently more than half a billion people worldwide are obese. The etiology of obesity is multifactorial, and combines a contribution of hereditary and behavioral factors, such as nutritional inadequacy, along with the influences of environment and reduced [...] Read more.
Obesity has become a pandemic, as currently more than half a billion people worldwide are obese. The etiology of obesity is multifactorial, and combines a contribution of hereditary and behavioral factors, such as nutritional inadequacy, along with the influences of environment and reduced physical activity. Two types of adipose tissue widely known are white and brown. While white adipose tissue functions predominantly as a key energy storage, brown adipose tissue has a greater mass of mitochondria and expresses the uncoupling protein 1 (UCP1) gene, which allows thermogenesis and rapid catabolism. Even though white and brown adipocytes are of different origin, activation of the brown adipocyte differentiation program in white adipose tissue cells forces them to transdifferentiate into “beige” adipocytes, characterized by thermogenesis and intensive lipolysis. Nowadays, researchers in the field of small molecule medicinal chemistry and gene therapy are making efforts to develop new drugs that effectively overcome insulin resistance and counteract obesity. Here, we discuss various aspects of white-to-beige conversion, adipose tissue catabolic re-activation, and non-shivering thermogenesis. Full article
(This article belongs to the Special Issue Anti-obesity and Anti-aging Natural Products)
Show Figures

Figure 1

Figure 1
<p>The scheme of adipocyte differentiation.</p>
Full article ">Figure 2
<p>Chemical substances that affect white-to-beige adipocyte conversion.</p>
Full article ">Figure 3
<p>AAV vectors in white-to-beige conversion.</p>
Full article ">
19 pages, 3023 KiB  
Review
Molecular Regulation of Thermogenic Mechanisms in Beige Adipocytes
by Siqi Yang, Yingke Liu, Xiaoxu Wu, Rongru Zhu, Yuanlu Sun, Shuoya Zou, Dongjie Zhang and Xiuqin Yang
Int. J. Mol. Sci. 2024, 25(12), 6303; https://doi.org/10.3390/ijms25126303 - 7 Jun 2024
Viewed by 861
Abstract
Adipose tissue is conventionally recognized as a metabolic organ responsible for storing energy. However, a proportion of adipose tissue also functions as a thermogenic organ, contributing to the inhibition of weight gain and prevention of metabolic diseases. In recent years, there has been [...] Read more.
Adipose tissue is conventionally recognized as a metabolic organ responsible for storing energy. However, a proportion of adipose tissue also functions as a thermogenic organ, contributing to the inhibition of weight gain and prevention of metabolic diseases. In recent years, there has been significant progress in the study of thermogenic fats, particularly brown adipose tissue (BAT). Despite this progress, the mechanism underlying thermogenesis in beige adipose tissue remains highly controversial. It is widely acknowledged that beige adipose tissue has three additional thermogenic mechanisms in addition to the conventional UCP1-dependent thermogenesis: Ca2+ cycling thermogenesis, creatine substrate cycling thermogenesis, and triacylglycerol/fatty acid cycling thermogenesis. This paper delves into these three mechanisms and reviews the latest advancements in the molecular regulation of thermogenesis from the molecular genetic perspective. The objective of this review is to provide readers with a foundation of knowledge regarding the beige fats and a foundation for future research into the mechanisms of this process, which may lead to the development of new strategies for maintaining human health. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Brown and beige adipocytes’ growth and development. Mesenchymal stem cells (MSCs) differentiate into two types of precursor cells: MYF5<sup>+</sup> stem cells and MYF5<sup>−</sup> stem cells. MYF5<sup>+</sup> stem cells can differentiate into brown precursor and myoblast; MYF5<sup>−</sup> stem cells can differentiate into white precursor. Interestingly, when exposed to cold stimulation, beige adipocytes produce both MYF5<sup>+</sup> and MYF5<sup>−</sup>. Beige adipocytes can be transdifferentiated from mature white adipocytes.</p>
Full article ">Figure 2
<p>Molecular regulation of heat production in beige adipocytes. In response to cold stimulation, neurons exocytose catecholamines (dark brown circles), which bind to β-adrenoreceptors (β-AR) coupled to the Gs alpha subunit, activating adenylyl cyclases to increase cAMP (pink circles), thereby activating protein kinase A (PKA) and exchange proteins directly activated by cAMP 1 (EPAC1). EPAC1 promotes thermogenic adipocyte proliferation while inhibiting white preadipocyte proliferation. PKA promotes the phosphorylation of a number of genes, resulting in the promotion of transcription of <span class="html-italic">UCP1</span> and other thermogenic genes. BMP9 is a protein that activates thermogenesis, or heat production, by modulating brown and beige adipocyte activation. By PKA and AMP-activated protein kinase (AMPK), the phosphorylation of hormone-sensitive lipase (pHSL) can promote lipolysis, which produces free fatty acids (FFAs). SIRT1 can induce the browning of WAT via deacetylation of PPARγ. CIDEA is located in lipids and regulates lipid fusion and enhances PPARγ binding to the Ucp1 enhancer, thereby promoting <span class="html-italic">UCP1</span> transcription. SOX4 forms a complex with PRDM16 and PPARγ, which promotes their binding and positively regulates the transcription of the gene UCP1.</p>
Full article ">Figure 3
<p>The three novel UCP1-independent mechanisms: Ca<sup>2+</sup> cycling thermogenesis, creatine substrate cycling thermogenesis, and triacylglycerol/fatty acid cycling thermogenesis. ATP, adenosine triphosphate; ADP, adenosine diphosphate; Pi, phosphate; RYR2, ryanodine receptor 2; IP3R, inositol trisphosphate receptor; SERCA2B, sarco/endoplasmic reticulum Ca <sup>2+</sup>-ATPase 2B; TNAP, tissue-non-specific alkaline phosphatase; CKB, creatine kinase B; TAG, triacylglycerol; FAs, fatty acids; G, glycerol; GyK, glycerol kinase; G3P, glycerol 3-phosphate; mGDP, mitochondrial G3P dehydrogenase; DHAP, dihydroxyacetone-3-phosphate; NAD+, nicotinamide adenine dinucleotide; NADH, nicotinamide adenine dinucleotide.</p>
Full article ">
22 pages, 11727 KiB  
Article
Polyphenol Compound 18a Modulates UCP1-Dependent Thermogenesis to Counteract Obesity
by Xueping Wen, Yufei Song, Mei Zhang, Yiping Kang, Dandan Chen, Hui Ma, Fajun Nan, Yanan Duan and Jingya Li
Biomolecules 2024, 14(6), 618; https://doi.org/10.3390/biom14060618 - 23 May 2024
Viewed by 911
Abstract
Recent studies increasingly suggest that targeting brown/beige adipose tissues to enhance energy expenditure offers a novel therapeutic approach for treating metabolic diseases. Brown/beige adipocytes exhibit elevated expression of uncoupling protein 1 (UCP1), which is a thermogenic protein that efficiently converts energy into heat, [...] Read more.
Recent studies increasingly suggest that targeting brown/beige adipose tissues to enhance energy expenditure offers a novel therapeutic approach for treating metabolic diseases. Brown/beige adipocytes exhibit elevated expression of uncoupling protein 1 (UCP1), which is a thermogenic protein that efficiently converts energy into heat, particularly in response to cold stimulation. Polyphenols possess potential anti-obesity properties, but their pharmacological effects are limited by their bioavailability and distribution within tissue. This study discovered 18a, a polyphenol compound with a favorable distribution within adipose tissues, which transcriptionally activates UCP1, thereby promoting thermogenesis and enhancing mitochondrial respiration in brown adipocytes. Furthermore, in vivo studies demonstrated that 18a prevents high-fat-diet-induced weight gain and improves insulin sensitivity. Our research provides strong mechanistic evidence that UCP1 is a complex mediator of 18a-induced thermogenesis, which is a critical process in obesity mitigation. Brown adipose thermogenesis is triggered by 18a via the AMPK-PGC-1α pathway. As a result, our research highlights a thermogenic controlled polyphenol compound 18a and clarifies its underlying mechanisms, thus offering a potential strategy for the thermogenic targeting of adipose tissue to reduce the incidence of obesity and its related metabolic problems. Full article
(This article belongs to the Topic Natural Products and Drug Discovery)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p><b>18a</b> exhibits favorable distribution in adipose tissues and activates UCP1 expression in C3H10-T1/2. (<b>A</b>) The plasma concentration of compound <b>18a</b> in C57BL/6J mice at 30 min following oral administration of 20 mg/kg; (<b>B</b>) the tissue distribution of <b>18a</b> in C57BL/6J mice; (<b>C</b>) structure of <b>18a</b>; (<b>D</b>) effect of <b>18a</b> on the transcriptional levels of thermogenic genes; (<b>E</b>) expressions of indicated protein by <b>18a</b>; (<b>F</b>,<b>G</b>) protein level statistics of PGC-1α (<b>F</b>), UCP1 (<b>G</b>) compared to α-tubulin. <span class="html-italic">n</span> = 4 per group. Data presented as the means ± SEM. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 versus the DMSO group. Original images can be found in <a href="#app1-biomolecules-14-00618" class="html-app">Figure S1F</a>.</p>
Full article ">Figure 2
<p><b>18a</b> enhances the thermogenic program and increases mitochondrial respiration on primary brown and beige adipocytes. Effect of <b>18a</b> on (<b>A</b>) mRNA levels of thermogenic genes in primary brown adipocytes; (<b>B</b>) protein levels of PGC-1α and UCP1 expression in primary brown adipocytes; (<b>C</b>,<b>D</b>) comparison of relative protein levels of PGC-1α (<b>C</b>) and UCP1 (<b>D</b>) compared to α-tubulin; (<b>E</b>) detection of OCR; (<b>F</b>) basal respiration and uncoupled respiration statistics; (<b>G</b>) mRNA levels of thermogenic genes in primary beige adipocytes; (<b>H</b>) protein levels in primary beige adipocytes; (<b>I</b>,<b>J</b>) comparison of relative protein levels of PGC-1α (<b>I</b>) and UCP1 (<b>J</b>) compared to α-tubulin; (<b>K</b>) detection of OCR; (<b>L</b>) basal respiration and uncoupled respiration statistics; <span class="html-italic">n</span> = 4 per group. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 versus the DMSO group. Original images can be found in <a href="#app1-biomolecules-14-00618" class="html-app">Figure S2A</a>.</p>
Full article ">Figure 3
<p><b>18a</b> enhances the activation of the thermogenic program and reduces adipose tissue weight in mice. (<b>A</b>) UCP1 reporter gene expression after <b>18a</b> administration; (<b>B</b>,<b>C</b>) fluorescence statistics; (<b>D</b>) rectal temperature changes in mice after 4 °C exposure; (<b>E</b>) infrared thermal images of mice after 6 h cold stimuli; (<b>F</b>) quantitative statistics of interscapular skin surface temperature; (<b>G</b>) weights of iBAT, iWAT, and eWAT in mice; (<b>H</b>) H&amp;E staining of adipose tissues; (<b>I</b>) cell area counts of iWAT; (<b>J</b>) Cell area counts of eWAT; (<b>K</b>) immunohistochemical staining of UCP1 in iBAT and iWAT; (<b>L</b>) detection of thermogenic mRNA levels in iBAT; (<b>M</b>) detection of thermogenic mRNA levels in iWAT; (<b>N</b>,<b>O</b>) Western blot analysis of indicated proteins in iBAT (<b>N</b>) and iWAT (<b>O</b>); (<b>P</b>,<b>Q</b>) comparison of PGC-1α (<b>P</b>) and UCP1 (<b>Q</b>) protein levels compared to α-tubulin in iBAT (<b>R</b>,<b>S</b>); comparison of PGC-1α (<b>R</b>) and UCP1 (<b>S</b>) protein levels compared to α-tubulin in iWAT. <span class="html-italic">n</span> = 4–6 per group, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 versus the Vehicle group. Original images can be found in <a href="#app1-biomolecules-14-00618" class="html-app">Figure S2C</a>.</p>
Full article ">Figure 4
<p>The thermogenic effects of <b>18a</b> are nullified under thermoneutral conditions. (<b>A</b>) Schematic diagram of the experimental procedure; (<b>B</b>) rectal temperature of 22 °C and 30 °C mice subjected to cold stimuli; (<b>C</b>) representative pictures of infrared thermal images; (<b>D</b>) statistics of the interscapular skin temperatures of the mice; (<b>E</b>) weights of iBAT, iWAT, and eWAT; (<b>F</b>) percentage of body composition as a percentage of body weight; <span class="html-italic">n</span> = 6, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, ns = non-significant. <b>18a</b> group versus the Vehicle group; # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01, and ### <span class="html-italic">p</span> &lt; 0.001, 30 °C group versus the 22 °C group.</p>
Full article ">Figure 5
<p><b>18a</b> resists weight gain and enhances glucose tolerance in obese mice. (<b>A</b>) Body weight changes per week; (<b>B</b>) body mass index statistics; (<b>C</b>) oral glucose tolerance test (GTT); (<b>D</b>) area under the curve (AUC) of GTT; (<b>E</b>) changes in rectal temperature after 4 °C exposure; (<b>F</b>) infrared thermal images in the HFD-Vehicle or HFD-<b>18a</b> group after exposure to 4 °C condition for 6 h; (<b>G</b>) quantitative statistics of interscapular skin surface temperature; (<b>H</b>) weights of adipose tissues; (<b>I</b>) representative hematoxylin and eosin staining from iBAT, iWAT, and eWAT sections; (<b>J</b>,<b>K</b>) quantification of adipocyte area of iWAT (<b>J</b>) and eWAT (<b>K</b>). <span class="html-italic">n</span> = 6, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 versus HFD-Vehicle group.</p>
Full article ">Figure 6
<p>UCP1 deficiency reverses the metabolic benefits of <b>18a</b>. (<b>A</b>) Body weight of high-fat diet WT and HO mice after <b>18a</b> treatment; (<b>B</b>) body mass index statistics; (<b>C</b>,<b>D</b>) GTT (<b>C</b>) and AUC statistics of WT and HO mice (<b>D</b>); (<b>E</b>,<b>F</b>) ITT (<b>E</b>) and AUC statistics of WT and HO mice; (<b>F</b>); (<b>G</b>) rectal temperature changes of WT and HO mice upon exposure to cold stimuli; (<b>H</b>) representative images of infrared thermography of mice exposed to 4 °C condition for 6 h; (<b>I</b>) iBAT skin temperature statistics; (<b>J</b>) adipose tissue weight of WT mice and HO mice; (<b>K</b>) H&amp;E staining of adipose tissue; (<b>L</b>) iWAT cell area statistics; (<b>M</b>) eWAT cell area statistics; <span class="html-italic">n</span> = 5–6, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 and ns = non-significant, <b>18a</b> group versus the Vehicle group. # <span class="html-italic">p</span> &lt; 0.05 and ## <span class="html-italic">p</span> &lt; 0.01, WT-Vehicle group versus the HO-Vehicle group.</p>
Full article ">Figure 7
<p><b>18a</b> exerts thermogenesis-promoting effects through AMPK-PGC-1α. (<b>A</b>) Western blot of p-AMPKα, AMPKα, p-ACC, and ACC in <b>18a</b>-treated C3H10-T1/2; (<b>B</b>) target proteins in C3H10-T1/2 treated with dorsomorphin followed <b>18a</b>; (<b>C</b>) mRNA levels of Pgc-1α and Ucp1 after AMPKα knockdown; (<b>D</b>) expression of PGC-1α and UCP1 proteins in C3H10-T1/2 after AMPKα knockdown; (<b>E</b>,<b>F</b>) comparison of PGC-1α (<b>E</b>) and UCP1 (<b>F</b>) protein levels to α-Tubulin; (<b>G</b>) mRNA levels of Pgc-1α and Ucp1 after PGC-1α knockdown; (<b>H</b>) protein levels of PGC-1α and UCP1 after PGC-1α knockdown; (<b>I</b>) graphical abstract. <span class="html-italic">n</span> = 3 per group. Data are presented as the means ± SEM. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, <b>18a</b> group versus the DMSO group; # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01, and ### <span class="html-italic">p</span> &lt; 0.001, siRNA group versus the Scramble group. Original images can be found in <a href="#app1-biomolecules-14-00618" class="html-app">Figure S9</a>.</p>
Full article ">
17 pages, 2112 KiB  
Review
The Role of Exerkines in Obesity-Induced Disruption of Mitochondrial Homeostasis in Thermogenic Fat
by Hui Shao, Huijie Zhang and Dandan Jia
Metabolites 2024, 14(5), 287; https://doi.org/10.3390/metabo14050287 - 17 May 2024
Viewed by 1332
Abstract
There is a notable correlation between mitochondrial homeostasis and metabolic disruption. In this review, we report that obesity-induced disruption of mitochondrial homeostasis adversely affects lipid metabolism, adipocyte differentiation, oxidative capacity, inflammation, insulin sensitivity, and thermogenesis in thermogenic fat. Elevating mitochondrial homeostasis in thermogenic [...] Read more.
There is a notable correlation between mitochondrial homeostasis and metabolic disruption. In this review, we report that obesity-induced disruption of mitochondrial homeostasis adversely affects lipid metabolism, adipocyte differentiation, oxidative capacity, inflammation, insulin sensitivity, and thermogenesis in thermogenic fat. Elevating mitochondrial homeostasis in thermogenic fat emerges as a promising avenue for developing treatments for metabolic diseases, including enhanced mitochondrial function, mitophagy, mitochondrial uncoupling, and mitochondrial biogenesis. The exerkines (e.g., myokines, adipokines, batokines) released during exercise have the potential to ameliorate mitochondrial homeostasis, improve glucose and lipid metabolism, and stimulate fat browning and thermogenesis as a defense against obesity-associated metabolic diseases. This comprehensive review focuses on the manifold benefits of exercise-induced exerkines, particularly emphasizing their influence on mitochondrial homeostasis and fat thermogenesis in the context of metabolic disorders associated with obesity. Full article
Show Figures

Figure 1

Figure 1
<p>Mitochondrial homeostasis in thermogenesis fat. Obesity results in the disruption of mitochondrial homeostasis in thermogenesis fat, affecting mitochondrial oxidative phosphorylation, dynamics, biogenesis, and mitophagy. This is evidenced by the increased mitochondrial reactive oxygen species (mtROS), mitochondrial fragmentation, mitochondrial DNA (mtDNA) damage, and the accumulation of dysfunctional mitochondria. Mfn1/2, mitofusin 1 and 2; Opa1, optic atrophy 1; Drp-1, dynamin-related protein 1; Fis-1, mitochondrial fission protein 1; TFAM, mitochondrial transcription factor A; TFB2M, mitochondrial transcription factors B2; mtRNAP, mitochondrial RNA polymerase; TEFM, transcription elongation factor of mitochondria; MTERF1, mitochondrial transcription termination factor 1; MAPP1LC3, microtubule-associated protein 1 light chain 3; BCL2L13, Bcl-2-like protein 13; FUNDC1, FUN14 domain-containing protein 1; BNIP3, BCL2 interacting protein 3; FKBP8, FK506 binding protein 8. Red arrows indicate an increase (up arrow).</p>
Full article ">Figure 2
<p>The effect of myokines on white adipose tissue browning. Exercise training promotes the browning of WAT by triggering the secretion of various myokines from skeletal muscle. This is substantiated by reductions in adipocyte size, heightened lipid and glucose metabolism, increased FFA oxidation and uptake, enhanced mitochondrial biogenesis, improved insulin sensitivity, brown adipogenesis, and thermogenesis. Myokines are secreted from skeletal muscle in response to exercise. Red arrows indicate an increase (up arrow) or a decrease (down arrow).</p>
Full article ">Figure 3
<p>Exercise-induced thermogenesis in adipocytes. Exercise not only activates UCP1-dependent thermogenesis but also induces UCP1-independent thermogenesis by increasing various futile cycles, such as creatine futile cycling, Ca<sup>2+</sup> futile cycling, and leptin-induced TAG-fatty acid cycling. The mechanisms of UCP1-independent thermogenesis involve respiratory rate, fatty acid oxidation, mitochondrial biogenesis, and function. Red arrows indicate an increase (up arrow).</p>
Full article ">Figure 4
<p>The impacts of regular exercise-induced adipokines and batokines on obesity. Adipokines and batokines induced by regular exercise play crucial roles in influencing glucose tolerance, fatty acid metabolism, insulin sensitivity, inflammation, mitochondrial homeostasis, the sympathetic neural network, as well as thermogenesis in both BAT and WAT. T3, thyroid hormone; IGFBP2, insulin-like growth factor binding protein 2; IGF-1, insulin-like growth factor 1. Red arrows indicate an increase (up arrow).</p>
Full article ">
31 pages, 3677 KiB  
Review
Unveiling the Potential of Natural Compounds: A Comprehensive Review on Adipose Thermogenesis Modulation
by Jaeeun Shin, Yeonho Lee, Seong Hun Ju, Young Jae Jung, Daehyeon Sim and Sung-Joon Lee
Int. J. Mol. Sci. 2024, 25(9), 4915; https://doi.org/10.3390/ijms25094915 - 30 Apr 2024
Cited by 1 | Viewed by 1008
Abstract
The process of adipocyte browning has recently emerged as a novel therapeutic target for combating obesity and obesity-related diseases. Non-shivering thermogenesis is the process of biological heat production in mammals and is primarily mediated via brown adipose tissue (BAT). The recruitment and activation [...] Read more.
The process of adipocyte browning has recently emerged as a novel therapeutic target for combating obesity and obesity-related diseases. Non-shivering thermogenesis is the process of biological heat production in mammals and is primarily mediated via brown adipose tissue (BAT). The recruitment and activation of BAT can be induced through chemical drugs and nutrients, with subsequent beneficial health effects through the utilization of carbohydrates and fats to generate heat to maintain body temperature. However, since potent drugs may show adverse side effects, nutritional or natural substances could be safe and effective as potential adipocyte browning agents. This review aims to provide an extensive overview of the natural food compounds that have been shown to activate brown adipocytes in humans, animals, and in cultured cells. In addition, some key genetic and molecular targets and the mechanisms of action of these natural compounds reported to have therapeutic potential to combat obesity are discussed. Full article
(This article belongs to the Section Bioactives and Nutraceuticals)
Show Figures

Figure 1

Figure 1
<p>The mechanism of action of curcumin on adipose browning. Curcumin induces adipose browning through activation of UCP1 via β3-AR, TGR5, and AMPK phosphorylation. Curcumin up-regulated PPARγ, PRDM16, and phosphorylated AMPK, which in turn activated Ppargc1a and UCP1 [<a href="#B35-ijms-25-04915" class="html-bibr">35</a>,<a href="#B37-ijms-25-04915" class="html-bibr">37</a>,<a href="#B39-ijms-25-04915" class="html-bibr">39</a>]. Additionally, curcumin exerted beneficial effects by altering gut-microbiome symbiosis and bile acids. BAs are known to activate TGR5, which triggers the production of cAMP. Activation of TGR5 through curcumin supplementation induced the phosphorylation of CREB, up-regulating UCP1-dependent thermogenesis [<a href="#B49-ijms-25-04915" class="html-bibr">49</a>]. Curcumin was also shown to increase the level of norepinephrine in serum upon supplementation [<a href="#B40-ijms-25-04915" class="html-bibr">40</a>,<a href="#B42-ijms-25-04915" class="html-bibr">42</a>]. Norepinepherine, up-regulated by curcumin, activates β3-AR, leading to the induction of DIO2 expression. Increased DIO2 enhances the conversion of T4 to T3 within adipocytes. Subsequently, T3 activates the expression of genes involved in thermogenesis, PRDM16, PPARγ, PPARGC1A, and UCP1 [<a href="#B40-ijms-25-04915" class="html-bibr">40</a>,<a href="#B42-ijms-25-04915" class="html-bibr">42</a>]. Overall, curcumin’s mechanism of action involves the activation of key regulators such as UCP1 through β3-AR, TGR5, and AMPK phosphorylation, alongside modulation of PPARγ, PRDM16, and AMPK phosphorylation. Ppargc1a activates mitochondrial biogenesis while UCP1 increases thermogenesis. The diagram includes key components such as norepinephrine (NE), bile acids (BAs), adenylyl cyclase (AC), 5′ AMP-activated protein kinase (AMPK), beta-3 adrenergic receptor (β3-AR), cAMP-response element-binding protein (CREB), iodothyronine deiodinase 2 (Dio2), phosphorylation (P), peroxisome proliferator-activated receptor γ coactivator-1α (PPARGC1A), protein kinase A (PKA), PR domain zinc finger protein 16 (PRDM16), peroxisome proliferator-activated receptor gamma (PPARG), thyroxine (T4), triiodothyronine (T3), Takeda G protein-coupled receptor 5 (TGR5), and uncoupling protein-1 (UCP1).</p>
Full article ">Figure 2
<p>The mechanism of action of berberine on adipose browning. Berberine induces thermogenesis and adipocyte browning via stimuli of FGF21 and DIO and activating the AMPK/SIRT1/PPARGC1A pathway. Berberine phosphorylated AMPK, which in turn activated SIRT1. Phosphorylated SIRT1 promotes the deacetylation and activation of PPARGC1A and the induction of PPARγ, leading to the induction of UCP1 expression and the subsequent enhancement of thermogenesis in brown adipose tissue [<a href="#B54-ijms-25-04915" class="html-bibr">54</a>,<a href="#B55-ijms-25-04915" class="html-bibr">55</a>]. In addition, berberine was also shown to regulate UCP1 expression by augmenting FGF21 levels through AMPK phosphorylation [<a href="#B56-ijms-25-04915" class="html-bibr">56</a>]. On the other hand, berberine supplementation in DIO mice was shown to increase UCP1 alongside DIO2 expression levels; however, further study is needed to elucidate the exact mechanism of action [<a href="#B62-ijms-25-04915" class="html-bibr">62</a>]. This signaling cascade resulted in an increase in thermogenic markers, along with an increase in UCP1. The diagram includes key components such as 5′ AMP-activated protein kinase (AMPK), peroxisome proliferator-activated receptor gamma (PPARG), iodothyronine deiodinase 2 (Dio2), phosphorylation (P), fibroblast growth factor 21 (FGF21), iodothyronine deiodinase 2 (DIO2), peroxisome proliferator-activated receptor γ coactivator-1α (PPARGC1A), sirtuin-1 (SIRT1), and uncoupling protein-1 (UCP1).</p>
Full article ">Figure 3
<p>The mechanism of action of resveratrol on adipose browning. Resveratrol induces adipose browning via the activation of the mTORC1, AMPK-SIRT1, and TGR5—PPARGC1A pathways. Upon the activation of resveratrol, AMPK phosphorylates and activates SIRT1, a nicotinamide adenine dinucleotide (NAD+)-dependent deacetylase. Phosphorylated SIRT1, in turn, exerts its enzymatic activity to deacetylate and activate PPARGC1A, a master regulator of mitochondrial biogenesis and UCP1 expression. Subsequently, activated PPARGC1A promotes the transcription of UCP1, leading to the induction of thermogenesis in adipose tissue [<a href="#B74-ijms-25-04915" class="html-bibr">74</a>,<a href="#B77-ijms-25-04915" class="html-bibr">77</a>]. Additionally, resveratrol increases the concentration of bile acid and lithocholic acid, which in turn activates TGR5 receptor. The activation of the TGR5 receptor stimulates AC, resulting in elevated levels of cAMP. It then activates PKA, which phosphorylates and activates DIO2. Activated DIO2 enhances the expression of PPARGC1A, promoting the transcription of UCP1, facilitating thermogenesis and energy dissipation in brown adipose tissue [<a href="#B85-ijms-25-04915" class="html-bibr">85</a>]. The diagram includes key components such as bile acids (BAs), 5′ AMP-activated protein kinase (AMPK), cyclic adenosine monophosphate (cAMP), type 2 iodothyronine deiodinase (DIO2), adenyl cyclase (AC), the cAMP-response element-binding protein (CREB), phosphorylation (P), peroxisome proliferator-activated receptor γ coactivator-1α (PPARGC1A), sirtuin-1 (SIRT1), Takeda G protein-coupled receptor 5 (TGR5), and uncoupling protein-1 (UCP1).</p>
Full article ">Figure 4
<p>A diagram illustrating the potential mechanisms of corylin, formononetin, ginsenoside, phytol, luteolin, menthol, nobiletin, and rutin-induced adipose browning. The signaling pathways responsible for the up-regulation of UCP1 and its transcriptional regulators are depicted. Corylin induces adipose browning by activating SIRT1 and β3-AR; however, the underlying mechanisms are unknown [<a href="#B101-ijms-25-04915" class="html-bibr">101</a>]. Formononetin [<a href="#B102-ijms-25-04915" class="html-bibr">102</a>], ginsenoside [<a href="#B95-ijms-25-04915" class="html-bibr">95</a>,<a href="#B96-ijms-25-04915" class="html-bibr">96</a>,<a href="#B99-ijms-25-04915" class="html-bibr">99</a>,<a href="#B100-ijms-25-04915" class="html-bibr">100</a>], phytol [<a href="#B103-ijms-25-04915" class="html-bibr">103</a>,<a href="#B104-ijms-25-04915" class="html-bibr">104</a>], and luteolin [<a href="#B105-ijms-25-04915" class="html-bibr">105</a>,<a href="#B106-ijms-25-04915" class="html-bibr">106</a>] induce adipose browning by activating the AMPK-PPARGC1A pathway. Menthol stimulates adipose browning through the AMPK-PPARGC1A pathway, as well as by stimulating TRPM8 [<a href="#B48-ijms-25-04915" class="html-bibr">48</a>,<a href="#B107-ijms-25-04915" class="html-bibr">107</a>,<a href="#B108-ijms-25-04915" class="html-bibr">108</a>,<a href="#B109-ijms-25-04915" class="html-bibr">109</a>]. Nobiletin [<a href="#B110-ijms-25-04915" class="html-bibr">110</a>,<a href="#B111-ijms-25-04915" class="html-bibr">111</a>] and rutin [<a href="#B112-ijms-25-04915" class="html-bibr">112</a>,<a href="#B113-ijms-25-04915" class="html-bibr">113</a>,<a href="#B114-ijms-25-04915" class="html-bibr">114</a>] stimulate adipose browning by up-regulating PPARGC1a and UCP1; however, the signaling components are yet to be elucidated. The diagram includes key components such as adenylyl cyclase (AC), 5′ AMP-activated protein kinase (AMPK), activating transcription factor 2 (ATF2), beta-3 adrenergic receptor (β3-AR), cAMP-response element-binding protein (CREB), iodothyronine deiodinase 2 (Dio2), phosphorylation (P), peroxisome proliferator-activated receptor γ coactivator-1α (PPARGC1A), peroxisome proliferator-activated receptor α (PPARα), calcium/calmodulin-dependent protein kinase 2 (CAMKK2), sirtuin-1 (SIRT1), transient receptor potential cation channel subfamily M member 8 (TRPM8), and uncoupling protein-1 (UCP1).</p>
Full article ">Figure 5
<p>A schematic diagram illustrating the browning effects of natural compounds on adipose thermogenesis regulation. The consumption of natural compounds improves and enhances adipose browning through signaling pathways such as SIRT1, PPARA/G, AMPK, and PPARGC1A. The up-regulation of these markers results in increased levels of UCP1, in other words, thermogenic capacity, which could be an effective strategy in regulating metabolic health.</p>
Full article ">
Back to TopTop