[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

Journals

Article Types

Countries / Regions

Search Results (133)

Search Parameters:
Keywords = PRDM1

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
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 ">
14 pages, 3676 KiB  
Article
Creating Meiotic Recombination-Regulating DNA Sites by SpEDIT in Fission Yeast Reveals Inefficiencies, Target-Site Duplications, and Ectopic Insertions
by Reine U. Protacio, Seth Dixon, Mari K. Davidson and Wayne P. Wahls
Biomolecules 2024, 14(8), 1016; https://doi.org/10.3390/biom14081016 - 16 Aug 2024
Viewed by 432
Abstract
Recombination hotspot-activating DNA sites (e.g., M26, CCAAT, Oligo-C) and their binding proteins (e.g., Atf1-Pcr1 heterodimer; Php2-Php3-Php5 complex, Rst2, Prdm9) regulate the distribution of Spo11 (Rec12)-initiated meiotic recombination. We sought to create 14 different candidate regulatory DNA sites via bp substitutions [...] Read more.
Recombination hotspot-activating DNA sites (e.g., M26, CCAAT, Oligo-C) and their binding proteins (e.g., Atf1-Pcr1 heterodimer; Php2-Php3-Php5 complex, Rst2, Prdm9) regulate the distribution of Spo11 (Rec12)-initiated meiotic recombination. We sought to create 14 different candidate regulatory DNA sites via bp substitutions in the ade6 gene of Schizosaccharomyces pombe. We used a fission yeast-optimized CRISPR-Cas9 system (SpEDIT) and 196 bp-long dsDNA templates with centrally located bp substitutions designed to ablate the genomic PAM site, create specific 15 bp-long DNA sequences, and introduce a stop codon. After co-transformation with a plasmid that encoded both the guide RNA and Cas9 enzyme, about one-third of colonies had a phenotype diagnostic for DNA sequence changes at ade6. PCR diagnostics and DNA sequencing revealed a diverse collection of alterations at the target locus, including: (A) complete or (B) partial template-directed substitutions; (C) non-homologous end joinings; (D) duplications; (E) bp mutations, and (F) insertions of ectopic DNA. We concluded that SpEDIT can be used successfully to generate a diverse collection of DNA sequence elements within a reporter gene of interest. However, its utility is complicated by low efficiency, incomplete template-directed repair events, and undesired alterations to the target locus. Full article
(This article belongs to the Special Issue Two Billion Years of Sex)
Show Figures

Figure 1

Figure 1
<p><b>CRISPR-Based Approach to Generate Candidate Regulatory DNA Sites Within the <span class="html-italic">ade6</span> Reporter Gene.</b> (<b>a</b>) Diagram of the <span class="html-italic">ade6</span> ORF (<span class="html-italic">box</span>) shows the position of the recombination-initiating dsDNA break (<span class="html-italic">DSB, scissors</span>), dsDNA template molecule for recombination-mediated repair (<span class="html-italic">grey box</span>), features within that template (<span class="html-italic">colored rectangles</span>), and recombination events (×) between the genomic DNA and the template DNA. Successful genome editing (<span class="html-italic">SpEDIT</span>) transfers the features from the template to the genome. (<b>b</b>) As in panel (<b>a</b>) but zoomed in to show the relevant DNA sequences and the position of the guide RNA. Base pair substitutions (<span class="html-italic">bold lowercase</span>) are used to generate a stop codon (<span class="html-italic">red</span>), to create the HES sequence (<span class="html-italic">green</span>), and to inactivate PAM (<span class="html-italic">blue</span>). The same symbols and color codes are used in subsequent figures. (<b>c</b>) Sequences of the HES DNA elements (<span class="html-italic">green</span>) and bp substitutions needed to create those DNA sequences; the adjacent stop codon (<span class="html-italic">red</span>) provides a marker for analyses of meiotic recombination.</p>
Full article ">Figure 2
<p><b>Efficiency of Targeting at the <span class="html-italic">ade6</span> Locus.</b> (<b>a</b>) G418-resistant transformants were screened for red colony color on YEA media, which is diagnostic for mutations within the <span class="html-italic">ade6</span> target gene. The structures and DNA sequences of the <span class="html-italic">ade6</span> locus from candidate (red) colonies were analyzed. (<b>b</b>) Agarose gel shows examples of PCR products that had the expected length and unexpected lengths. (<b>c</b>) Summary of results lists each class of alteration detected within the <span class="html-italic">ade6</span> gene; data for each class are presented in subsequent figures.</p>
Full article ">Figure 3
<p><b>Complete (Successful) Templated-Directed Modifications of the <span class="html-italic">ade6</span> Target Gene.</b> (<b>a</b>) Diagram of approach and results. Correctly positioned recombination events (×) are required to transfer all of the desired elements (<span class="html-italic">color-coded rectangles</span>) into the genome. (<b>b</b>) The relevant DNA sequences are shown for the parental strain (<span class="html-italic">ade6<sup>+</sup></span>) and for one correctly edited, representative clone for each of the 14 different HES DNA site elements, plus the one control (<span class="html-italic">CON</span>).</p>
Full article ">Figure 4
<p><b>Incomplete Template-Directed Modifications of the <span class="html-italic">ade6</span> Gene.</b> (<b>a</b>) Diagram of approach and results. Some recombination events (×) flanking the DSB transfer only a subset of the desired elements (<span class="html-italic">color-coded rectangles</span>) into the genome. (<b>b</b>) Representative sequences show correct editing very close to the DSB, at the PAM site (<span class="html-italic">blue</span>), but no editing at the more distal bp substitutions (<span class="html-italic">green</span> and <span class="html-italic">red</span> in panel (<b>a</b>)).</p>
Full article ">Figure 5
<p><b>Template-Directed Duplications of the Targeted DNA Region Within <span class="html-italic">ade6</span>.</b> Diagrams show organization of elements for (<b>a</b>) normal-length, wild-type <span class="html-italic">ade6</span> and (<b>b–e</b>) a subset of clones with increased length. These clones had one or more additional copies of DNA sequence from the template molecule (<span class="html-italic">grey box</span>) at the correct position in the target. Some clones (<b>b</b>,<b>e</b>) had incomplete incorporation of features distal to the DSB (as in <a href="#biomolecules-14-01016-f004" class="html-fig">Figure 4</a>) and some clones (<b>c</b>–<b>e</b>) had additional, unplanned bp mutations (<span class="html-italic">pink</span>). The diagrams show how the concatemerization of templates prior to repair and the positions of subsequent recombination events (×) produce these types of clones. (<b>f</b>) A mechanism for concatemers. Overlapping microhomologies (underlined) near the 5′ and 3′ ends of the linear, dsDNA template molecules (<span class="html-italic">top</span>) likely contribute to the formation of the junctions (<span class="html-italic">bottom</span>) within the concatemers.</p>
Full article ">Figure 6
<p><b>Template-Independent Modifications of <span class="html-italic">ade6</span> by Non-homologous End Joining.</b> (<b>a</b>) Diagram of findings and mechanism. (<b>b</b>) DNA sequences of representative clones that had an <span class="html-italic">ade6</span> mutant phenotype and mutations (<span class="html-italic">pink</span>; ‡ in panel (<b>a</b>)) within the <span class="html-italic">ade6</span> ORF without incorporating any of the engineered bp substitutions. The short insertions or deletions (<span class="html-italic">pink</span>) cluster precisely at the site of the DSB; these indels are diagnostic for non-homologous end joining. The +1 and −2 frameshift mutations within the <span class="html-italic">ade6</span> ORF each trigger the use of an out-of-frame nonsense codon (STOP) nearby.</p>
Full article ">Figure 7
<p><b>Insertions of Ectopic DNA within <span class="html-italic">ade6</span>.</b> The diagram shows the organization of the <span class="html-italic">ade6</span> gene for a subset of clones with increased length. (<b>a</b>) Location of DSB within <span class="html-italic">ade6</span> and the non-homologous integrations of (<b>b</b>) template DNA sequences or (<b>c</b>) salmon sperm carrier DNA. The DNA sequences of the insertion junctions show that the ectopic DNAs integrated precisely at the site of the DSB by a non-homologous end-joining mechanism.</p>
Full article ">
10 pages, 1139 KiB  
Article
Adaptive Evolution and Functional Differentiation of Testis-Expressed Genes in Theria
by Yukako Katsura, Shuji Shigenobu and Yoko Satta
Animals 2024, 14(16), 2316; https://doi.org/10.3390/ani14162316 - 9 Aug 2024
Viewed by 407
Abstract
Gene expression patterns differ in different tissues, and the expression pattern of genes in the mammalian testis is known to be extremely variable in different species. To clarify how the testis transcriptomic pattern has evolved in particular species, we examined the evolution of [...] Read more.
Gene expression patterns differ in different tissues, and the expression pattern of genes in the mammalian testis is known to be extremely variable in different species. To clarify how the testis transcriptomic pattern has evolved in particular species, we examined the evolution of the adult testis transcriptome in Theria using 10 species: two marsupials (opossum and Tasmanian devil), six eutherian (placental) mammals (human, chimpanzee, bonobo, gorilla, rhesus macaque, and mouse), and two outgroup species (platypus and chicken). We show that 22 testis-expressed genes are marsupial-specific, suggesting their acquisition in the stem lineage of marsupials after the divergence from eutherians. Despite the time length of the eutherian stem lineage being similar to that of the marsupial lineage, acquisition of testis-expressed genes was not found in the stem lineage of eutherians; rather, their expression patterns differed by species, suggesting rapid gene evolution in the eutherian ancestors. Fifteen testis-expressed genes are therian-specific, and for three of these genes, the evolutionary tempo is markedly faster in eutherians than in marsupials. Our phylogenetic analysis of Rho GTPase-activating protein 28 (ARHGAP28) suggests the adaptive evolution of this gene in the eutherians, probably together with the expression pattern differentiation. Full article
(This article belongs to the Section Animal Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Acquisition/loss of testis-expressed genes in 10 species. The genes at each clade are listed in <a href="#animals-14-02316-t001" class="html-table">Table 1</a>, <a href="#animals-14-02316-t002" class="html-table">Table 2</a>, <a href="#app1-animals-14-02316" class="html-app">Tables S2 and S3</a>, and the left and right of the ‘/’ symbols mean acquisition and loss of testis-expressed genes, respectively.</p>
Full article ">Figure 2
<p>The phylogenetic trees, branch lengths, and dn/ds ratios for three therian-specific testis genes in 10 species. (<b>A</b>) <span class="html-italic">ARHGAP28</span> genes (1833 bp). (<b>B</b>) <span class="html-italic">SYNM</span> genes (2235 bp). (<b>C</b>) <span class="html-italic">PDZRN3</span> genes (2256 bp). The trees were constructed by the NJ method using the number of nucleotide differences. The average and standard deviation of branch lengths in six eutherians are shown above the black bold branch, and those in two marsupials are shown above the gray bold branch. The dn/ds ratio in six eutherian or two marsupial pairs is shown under the branch. ** and * mean <span class="html-italic">p</span> &lt; 0.01 and <span class="html-italic">p</span> &lt; 0.05, respectively, and are supported by Fisher’s exact test.</p>
Full article ">Figure 2 Cont.
<p>The phylogenetic trees, branch lengths, and dn/ds ratios for three therian-specific testis genes in 10 species. (<b>A</b>) <span class="html-italic">ARHGAP28</span> genes (1833 bp). (<b>B</b>) <span class="html-italic">SYNM</span> genes (2235 bp). (<b>C</b>) <span class="html-italic">PDZRN3</span> genes (2256 bp). The trees were constructed by the NJ method using the number of nucleotide differences. The average and standard deviation of branch lengths in six eutherians are shown above the black bold branch, and those in two marsupials are shown above the gray bold branch. The dn/ds ratio in six eutherian or two marsupial pairs is shown under the branch. ** and * mean <span class="html-italic">p</span> &lt; 0.01 and <span class="html-italic">p</span> &lt; 0.05, respectively, and are supported by Fisher’s exact test.</p>
Full article ">
17 pages, 22700 KiB  
Article
Identification of Schizophrenia Susceptibility Loci in the Urban Taiwanese Population
by Chih-Chung Huang, Yi-Guang Wang, Chun-Lun Hsu, Ta-Chuan Yeh, Wei-Chou Chang, Ajeet B. Singh, Chin-Bin Yeh, Yi-Jen Hung, Kuo-Sheng Hung and Hsin-An Chang
Medicina 2024, 60(8), 1271; https://doi.org/10.3390/medicina60081271 - 6 Aug 2024
Viewed by 623
Abstract
Background and Objectives: Genomic studies have identified several SNP loci associated with schizophrenia in East Asian populations. Environmental factors, particularly urbanization, play a significant role in schizophrenia development. This study aimed to identify schizophrenia susceptibility loci and characterize their biological functions and [...] Read more.
Background and Objectives: Genomic studies have identified several SNP loci associated with schizophrenia in East Asian populations. Environmental factors, particularly urbanization, play a significant role in schizophrenia development. This study aimed to identify schizophrenia susceptibility loci and characterize their biological functions and molecular pathways in Taiwanese urban Han individuals. Materials and Methods: Participants with schizophrenia were recruited from the Taiwan Precision Medicine Initiative at Tri-Service General Hospital. Genotype–phenotype association analysis was performed, with significant variants annotated and analyzed for functional relevance. Results: A total of 137 schizophrenia patients and 26,129 controls were enrolled. Ten significant variants (p < 1 × 10−5) and 15 expressed genes were identified, including rs1010840 (SOWAHC and RGPD6), rs11083963 (TRPM4), rs11619878 (LINC00355 and LINC01052), rs117010638 (AGBL1 and MIR548AP), rs1170702 (LINC01680 and LINC01720), rs12028521 (KAZN and PRDM2), rs12859097 (DMD), rs1556812 (ATP11A), rs78144262 (LINC00977), and rs9997349 (ENPEP). These variants and associated genes are involved in immune response, blood pressure regulation, muscle function, and the cytoskeleton. Conclusions: Identified variants and associated genes suggest a potential genetic predisposition to schizophrenia in the Taiwanese urban Han population, highlighting the importance of potential comorbidities, considering population-specific genetic and environmental interactions. Full article
(This article belongs to the Section Psychiatry)
Show Figures

Figure 1

Figure 1
<p>The study encompassed 137 schizophrenia patients as case groups and 25,927 participants as the control group. Genotyping was conducted using the Taiwan Precision Medicine (TPM) array chip. A dataset of 493,852 SNPs underwent filtration, with 280,153 SNPs passing through and subsequently subjected to chi−squared testing for the detection of risk factors.</p>
Full article ">Figure 2
<p>Manhattan plot of associated variants results in schizophrenia patients. Following chi−squared testing, 280,153 variants were detected among TSGH TPMI participants. Among these, ten highly significant variants were identified based on a <span class="html-italic">p</span>-value &lt; 10<sup>−5</sup> (green dots above the blue dashed line): rs78144262, rs9997349, rs1010840, rs11083963, rs11619878, rs1170702, rs117010638, rs12028521, rs12859097, and rs1556812.</p>
Full article ">Figure 3
<p>Interpretation associations results in Q-Q (quantile-quantile) plot. The distribution of low <span class="html-italic">p</span>-values (represented by high observed −log10(<span class="html-italic">p</span>) values) indicates true associations with schizophrenia. The genomic inflation factor, lambda, was determined based on the median observed and expected chi-square values.</p>
Full article ">Figure 4
<p>Top 10 significant Gene Ontology biological process (GO BP) functions relative to variants. The BP functions of genes associated with identified risk variants were characterized using the GO database. Selected GO terms exhibited with <span class="html-italic">p</span>-values &lt; 0.05.</p>
Full article ">Figure 5
<p>Top 10 significant pathways relative to variants analysis. Genes associated with identified risk variants were characterized using the Reactome database. Selected pathways exhibited with <span class="html-italic">p</span>-values &lt; 0.05.</p>
Full article ">Figure 6
<p>Gene functional and molecular pathway annotation network of AGBL1, ATP11A, DMD, ENPEP, and TPM. The functional annotation network was constructed using GO biological process (green circles) and Reactome pathway (pink circles).</p>
Full article ">
12 pages, 1839 KiB  
Article
Dynamic Changes in Histone Modifications Are Associated with Differential Chromatin Interactions
by Yumin Nie and Mengjie Wang
Genes 2024, 15(8), 988; https://doi.org/10.3390/genes15080988 - 26 Jul 2024
Viewed by 616
Abstract
Eukaryotic genomes are organized into chromatin domains through long-range chromatin interactions which are mediated by the binding of architectural proteins, such as CTCF and cohesin, and histone modifications. Based on the published Hi-C and ChIP-seq datasets in human monocyte-derived macrophages, we identified 206 [...] Read more.
Eukaryotic genomes are organized into chromatin domains through long-range chromatin interactions which are mediated by the binding of architectural proteins, such as CTCF and cohesin, and histone modifications. Based on the published Hi-C and ChIP-seq datasets in human monocyte-derived macrophages, we identified 206 and 127 differential chromatin interactions (DCIs) that were not located within transcription readthrough regions in influenza A virus- and interferon β-treated cells, respectively, and found that the binding positions of CTCF and RAD21 within more than half of the DCI sites did not change. However, five histone modifications, H3K4me3, H3K27ac, H3K36me3, H3K9me3, and H3K27me3, showed significantly more dramatic changes than CTCF and RAD21 within the DCI sites. For H3K4me3, H3K27ac, H3K36me3, and H3K27me3, significantly more dramatic changes were observed outside than within the DCI sites. We further applied a motif scanning approach to discover proteins that might correlate with changes in histone modifications and chromatin interactions and found that PRDM9, ZNF384, and STAT2 frequently bound to DNA sequences corresponding to 1 kb genomic intervals with gains or losses of a histone modification within the DCI sites. This study explores the dynamic regulation of chromatin interactions and extends the current knowledge of the relationship between histone modifications and chromatin interactions. Full article
(This article belongs to the Section Epigenomics)
Show Figures

Figure 1

Figure 1
<p>Hi-C contact maps of mock- and IAV-treated MDMs and two DCIs on chromosome 11. The blue and green square represented a strengthened and weakened DCI, respectively.</p>
Full article ">Figure 2
<p>Genomic distances between DCI sites and patterns of CTCF and RAD21 binding within DCI sites after IAV infection. (<b>A</b>) Proportions of DCI sites with different distances after IAV and IFN-β treatment of MDMs. (<b>B</b>) Gains and losses of CTCF and RAD21 binding within DCI sites after IAV infection. (<b>C</b>) Changes in the patterns of CTCF and RAD21 binding within DCI sites after IAV infection. For two genomic regions where a DCI occurred, 0, 1, and 2 represent binding peaks of CTCF or RAD21 overlapped with neither, either, and both of the genomic regions, respectively.</p>
Full article ">Figure 3
<p>Changes in the binding of CTCF and RAD21 and histone modifications within DCI sites after IAV infection. (<b>A</b>) Changes within all DCI sites. (<b>B</b>) Changes within weakened and strengthened loops. (<b>C</b>,<b>D</b>) Gains and losses of CTCF and RAD21 binding and histone modifications within weakened and strengthened loops. Here, 0 and 1 represent a transcription factor or histone modification being absent or present within an interval, respectively. *, **, and *** represent a <span class="html-italic">p</span>-value of less than 0.05, 0.005, and 0.001, respectively.</p>
Full article ">Figure 4
<p>Changes in the binding of CTCF and RAD21 and histone modifications around DCI sites after IAV infection (<b>A</b>) and IFN-β treatment (<b>B</b>). * <span class="html-italic">p</span>-value of less than 0.05 produced by a Wilcoxon rank-sum test.</p>
Full article ">Figure 5
<p>Transcription factors most frequently bound within four types of intervals within DCI sites after IAV infection. Nodes represent transcription factors and four types of intervals. For each type of interval, W and S represent weakened and strengthened loops, respectively, while 0→1 and 1→0 represent gain and loss of a histone modification, respectively. The width of the lines between transcription factors and types of intervals is proportional to the occurrence number of transcription factors within the intervals.</p>
Full article ">Figure 6
<p>Colocalization of CTCF, RAD21, and five histone modifications in mock-, IAV-, and IFN-β-treated MDMs. The lower and upper triangular heat maps show pairwise similarities in mock and IAV (or IFN-β) conditions, respectively.</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 ">
23 pages, 6503 KiB  
Article
Identification of miRNAs and Their Target Genes Associated with Sunitinib Resistance in Clear Cell Renal Cell Carcinoma Patients
by María Armesto, Stéphane Nemours, María Arestín, Iraide Bernal, Jon Danel Solano-Iturri, Manuel Manrique, Laura Basterretxea, Gorka Larrinaga, Javier C. Angulo, David Lecumberri, Ane Miren Iturregui, José I. López and Charles H. Lawrie
Int. J. Mol. Sci. 2024, 25(13), 6881; https://doi.org/10.3390/ijms25136881 - 22 Jun 2024
Viewed by 1046
Abstract
Sunitinib has greatly improved the survival of clear cell renal cell carcinoma (ccRCC) patients in recent years. However, 20–30% of treated patients do not respond. To identify miRNAs and genes associated with a response, comparisons were made between biopsies from responder and non-responder [...] Read more.
Sunitinib has greatly improved the survival of clear cell renal cell carcinoma (ccRCC) patients in recent years. However, 20–30% of treated patients do not respond. To identify miRNAs and genes associated with a response, comparisons were made between biopsies from responder and non-responder ccRCC patients. Using integrated transcriptomic analyses, we identified 37 miRNAs and 60 respective target genes, which were significantly associated with the NF-kappa B, PI3K-Akt and MAPK pathways. We validated expression of the miRNAs (miR-223, miR-155, miR-200b, miR-130b) and target genes (FLT1, PRDM1 and SAV1) in 35 ccRCC patients. High levels of miR-223 and low levels of FLT1, SAV1 and PRDM1 were associated with worse overall survival (OS), and combined miR-223 + SAV1 levels distinguished responders from non-responders (AUC = 0.92). Using immunohistochemical staining of 170 ccRCC patients, VEGFR1 (FLT1) expression was associated with treatment response, histological grade and RECIST (Response Evaluation Criteria in Solid Tumors) score, whereas SAV1 and BLIMP1 (PRDM1) were associated with metachronous metastatic disease. Using in situ hybridisation (ISH) to detect miR-155 we observed higher tumoural cell expression in non-responders, and non-tumoural cell expression with increased histological grade. In summary, our preliminary analysis using integrated miRNA-target gene analyses identified several novel biomarkers in ccRCC patients that surely warrant further investigation. Full article
(This article belongs to the Section Molecular Oncology)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the workflow used in this study.</p>
Full article ">Figure 2
<p>Heatmap of unsupervised cluster analyses depicting expression of (<b>A</b>) mature miRNAs, (<b>B</b>) pre-miRNAs, (<b>C</b>) snoRNAs and scaRNAs, (<b>D</b>) lncRNA and (<b>E</b>) coding genes in ccRCC cases. The dendrogram at the side shows the distribution of the RNAs, and at the top the relationship between patient samples (blue responder and red non-responder) is shown.</p>
Full article ">Figure 3
<p>String visualisation network of miRNA–target gene interactions associated with sunitinib resistance in ccRCC patients.</p>
Full article ">Figure 4
<p>Gene ontology and pathway mapping of miRNA targeted genes. Terms are functionally grouped based on shared genes (kappa score) and are shown in different colours. The node size represents the degree of significance.</p>
Full article ">Figure 5
<p>Box and whisker plots of levels of differentially expressed miRNAs measured by qRT-PCR in NR and R ccRCC cases. (<b>A</b>) <span class="html-italic">miR-17-3p</span>; (<b>B</b>) <span class="html-italic">miR-99a-5p</span>; (<b>C</b>) <span class="html-italic">miR-223-3p</span>; (<b>D</b>) <span class="html-italic">miR-155</span>; (<b>E</b>) <span class="html-italic">miR-484</span>; (<b>F</b>) <span class="html-italic">miR-200b-3p;</span> (<b>G</b>) <span class="html-italic">miR-200c-3p</span>; (<b>H</b>) <span class="html-italic">miR-150-5p</span>; (<b>I</b>) <span class="html-italic">miR-130b-3p</span>. Significant differences (<span class="html-italic">p</span> &lt; 0.05) are denoted by asterisks (*).</p>
Full article ">Figure 6
<p>Box and whisker plots of levels of differentially expressed genes measured by qRT-PCR in NR and R ccRCC cases. (<b>A</b>) <span class="html-italic">CD274</span>; (<b>B</b>) <span class="html-italic">EPAS1</span>; (<b>C</b>) <span class="html-italic">VEGFA</span>; (<b>D</b>) <span class="html-italic">FLT1</span>; (<b>E</b>) <span class="html-italic">ZEB1</span>; (<b>F</b>) <span class="html-italic">LRP6;</span> (<b>G</b>) <span class="html-italic">PTBP2</span>; (<b>H</b>) <span class="html-italic">PRDM1</span>; (<b>I</b>) <span class="html-italic">SAV1</span>. Significant differences (<span class="html-italic">p</span> &lt; 0.05) are denoted by asterisks (*).</p>
Full article ">Figure 7
<p>Kaplan–Meier survival curves in univariate analysis of expression levels of (<b>A</b>) <span class="html-italic">miR-223-3p</span>, (<b>B</b>) <span class="html-italic">PRDM1</span>, (<b>C</b>) <span class="html-italic">FLT1</span> and (<b>D</b>) <span class="html-italic">SAV1</span> as a function of overall survival (OS) in months.</p>
Full article ">Figure 8
<p>Examples of <span class="html-italic">miR-155</span> expression detection by ISH in ccRCC cases demonstrating (<b>A</b>) positive expression in tumour cells, (<b>B</b>) positive expression in non-tumour cells and (<b>C</b>) negative expression.</p>
Full article ">Figure 9
<p>Schematic summary of main findings in this study.</p>
Full article ">
10 pages, 1325 KiB  
Review
An Eye into the Aorta: The Role of Extracellular Matrix Regulatory Genes ZNF469 and PRDM5, from Their Previous Association with Brittle Cornea Syndrome to Their Novel Association with Aortic and Arterial Aneurysmal Diseases
by Peyton Moore, Adam Wolf and Mohanakrishnan Sathyamoorthy
Int. J. Mol. Sci. 2024, 25(11), 5848; https://doi.org/10.3390/ijms25115848 - 28 May 2024
Viewed by 855
Abstract
The extracellular matrix is a complex network of proteins and other molecules that are essential for the support, integrity, and structure of cells and tissues within the human body. The genes ZNF469 and PRDM5 each produce extracellular-matrix-related proteins that, when mutated, have been [...] Read more.
The extracellular matrix is a complex network of proteins and other molecules that are essential for the support, integrity, and structure of cells and tissues within the human body. The genes ZNF469 and PRDM5 each produce extracellular-matrix-related proteins that, when mutated, have been shown to result in the development of brittle cornea syndrome. This dysfunction results from aberrant protein function resulting in extracellular matrix disruption. Our group recently identified and published the first known associations between variants in these genes and aortic/arterial aneurysms and dissection diseases. This paper delineates the proposed effects of mutated ZNF469 and PRDM5 on various essential extracellular matrix components, including various collagens, TGF-B, clusterin, thrombospondin, and HAPLN-1, and reviews our recent reports associating single-nucleotide variants to these genes’ development of aneurysmal and dissection diseases. Full article
(This article belongs to the Special Issue The Role of Extracellular Matrix Proteins in Pathogenesis)
Show Figures

Figure 1

Figure 1
<p>Simplified illustration of the ZNF469 protein product. The protein includes proline-rich and arginine-rich domains, and 7 C2H2 zinc fingers that are responsible for the proteins’ role in regulation of gene expression. Various mutations have been identified in ECM-related disease, with a few depicted above [<a href="#B6-ijms-25-05848" class="html-bibr">6</a>,<a href="#B17-ijms-25-05848" class="html-bibr">17</a>].</p>
Full article ">Figure 2
<p>Simplified protein structure of the PRDM5 protein. The protein contains an intact PR domain at the NH-terminus, with 16 zinc-finger domains near the C terminus. The PR- and zinc-finger-domains are responsible for the proteins’ role in transcriptional regulation. Two variants in the <span class="html-italic">PRDM5</span> gene that have been reported in prior ECM-related diseases are shown [<a href="#B17-ijms-25-05848" class="html-bibr">17</a>].</p>
Full article ">Figure 3
<p>A representation of disruptions in ECM genes such as <span class="html-italic">ZNF469</span> and <span class="html-italic">PRDM5</span> affecting ECM components such as collagens, clusterin, elastin, or other proteins, resulting in phenotypes such as BCS, keratoconus, osteogenesis imperfecta, Ehlers–Danlos syndrome, and TAAD [<a href="#B17-ijms-25-05848" class="html-bibr">17</a>].</p>
Full article ">Figure 4
<p>Summary figure demonstrating that variants in the <span class="html-italic">ZNF469</span> and <span class="html-italic">PRDM5</span> gene may cause disruption in ECM components. These disruptions have been correlated with the development of the ECM-related disease BCS in the past, and newer associations are being reported in thoracic aortic aneurysm and dissection formation. Variants in these genes are linking corneal and aortic diseases for the first time [<a href="#B17-ijms-25-05848" class="html-bibr">17</a>].</p>
Full article ">
10 pages, 3182 KiB  
Article
Anti-Obesity Activities of the Compounds from Perilla frutescens var. acuta and Chemical Profiling of the Extract
by Isoo Youn, Donglan Piao, Jisu Park, Seung A Ock, Sujin Han, Ah-Reum Han, Sunhye Shin and Eun Kyoung Seo
Molecules 2024, 29(11), 2465; https://doi.org/10.3390/molecules29112465 - 23 May 2024
Cited by 1 | Viewed by 1114
Abstract
Perilla frutescens var. acuta (Lamiaceae) is widely used not only as an oil or a spice, but also as a traditional medicine to treat colds, coughs, fever, and indigestion. As an ongoing effort, luteolin-7-O-diglucuronide (1), apigenin-7-O-diglucuronide ( [...] Read more.
Perilla frutescens var. acuta (Lamiaceae) is widely used not only as an oil or a spice, but also as a traditional medicine to treat colds, coughs, fever, and indigestion. As an ongoing effort, luteolin-7-O-diglucuronide (1), apigenin-7-O-diglucuronide (2), and rosmarinic acid (3) isolated from P. frutescens var. acuta were investigated for their anti-adipogenic and thermogenic activities in 3T3-L1 cells. Compound 1 exhibited a strong inhibition against adipocyte differentiation by suppressing the expression of Pparg and Cebpa over 52.0% and 45.0%, respectively. Moreover, 2 inhibited the expression of those genes in a dose-dependent manner [Pparg: 41.7% (5 µM), 62.0% (10 µM), and 81.6% (50 µM); Cebpa: 13.8% (5 µM), 18.4% (10 µM), and 37.2% (50 µM)]. On the other hand, the P. frutescens var. acuta water extract showed moderate thermogenic activities. Compounds 1 and 3 also induced thermogenesis in a dose-dependent manner by stimulating the mRNA expressions of Ucp1, Pgc1a, and Prdm16. Moreover, an LC-MS/MS chromatogram of the extract was acquired using UHPLC-MS2 and it was analyzed by feature-based molecular networking (FBMN) and the Progenesis QI software (version 3.0). The chemical profiling of the extract demonstrated that flavonoids and their glycoside derivatives, including those isolated earlier as well as rosmarinic acid, are present in P. frutescens var. acuta. Full article
(This article belongs to the Collection Bioactive Natural Molecules from Functional Foods)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Structures of luteolin-7-<span class="html-italic">O</span>-diglucuronide (<b>1</b>), apigenin-7-<span class="html-italic">O</span>-diglucuronide (<b>2</b>), and rosmarinic acid (<b>3</b>). (<b>b</b>) Cell viability of the <span class="html-italic">P. frutescens</span> var. <span class="html-italic">acuta</span> water extract (PFW) and <b>1</b>–<b>3</b> against a 3T3-L1 cell line. (<b>c</b>) Inhibitory effect against the preadipocyte differentiation of PFW (0, 10, 50, and 100 μg/mL). (<b>d</b>–<b>f</b>) Inhibition of preadipocyte differentiation by <b>1</b>–<b>3</b> (0, 5, 10, and 50 μM). 3T3-L1 preadipocytes were treated with each component dissolved in DMSO during differentiation. The mRNA levels were determined by quantitative real-time polymerase chain reaction (RT-PCR) with normalization relative to <span class="html-italic">18s</span> rRNA. Data are presented as means ± standard error of mean (<span class="html-italic">n</span> = 3). Different letters (a and b) indicate significant differences at <span class="html-italic">p</span> &lt; 0.05 by Duncan’s multiple comparison test. <span class="html-italic">Pparg</span>, peroxisome proliferator-activated receptor gamma; <span class="html-italic">Cebpa</span>, CCAAT/enhancer-binding protein α; <span class="html-italic">P</span>, <span class="html-italic">p</span>-value.</p>
Full article ">Figure 2
<p>Thermogenetic induction of (<b>a</b>) the PFW extract (0, 10, 50, and 100 μg/mL) and (<b>b</b>–<b>d</b>) <b>1</b>–<b>3</b> (0, 5, 10, and 50 μM) on 3T3-L1 adipocytes. The compounds were identified as luteolin-7-<span class="html-italic">O</span>-diglucuronide (<b>1</b>), apigenin-7-<span class="html-italic">O</span>-diglucuronide (<b>2</b>), and rosmarinic acid (<b>3</b>). Fully differentiated 3T3-L1 preadipocytes were treated with each component dissolved in DMSO for 24 h with a 4 h treatment of CL316,243 (CL; 10 μM). The mRNA levels were determined by quantitative RT-PCR with normalization relative to 18s rRNA. Data are presented as means ± standard error of mean (<span class="html-italic">n</span> = 3). A two-way ANOVA was conducted to determine the effects of the concentration of the PFW extract and <b>1</b>–<b>3</b> (Conc), CL treatment (CL), and the interaction of Conc and CL (Conc*CL). Different letters (a, b and c) indicate significant differences at <span class="html-italic">p</span> &lt; 0.05 by Duncan’s multiple comparison test. <span class="html-italic">Ucp1</span>, uncoupling protein 1; <span class="html-italic">Pgc1a</span>, peroxisome proliferator-activated receptor gamma coactivator 1-α; <span class="html-italic">Prdm16</span>, PR domain-containing 16; <span class="html-italic">P</span>, <span class="html-italic">p</span>-value.</p>
Full article ">Figure 3
<p>Feature-based molecular networking (FBMN) of the PFW extract. A total of 34 single nodes and 10 clusters (nodes ≥ 2) were found in the network, and a total of seven compounds (<b>3</b> and <b>P1</b>–<b>P6</b>) were identified using the FBMN analysis. Compounds in the orange boxes (<b>3</b>, <b>P2</b>, <b>P5</b>, and <b>P6</b>) indicate the components that were simultaneously identified by the FBMN and Progenesis QI (ProQI) analyses. Although the node “637.104” in <b>Cluster 1</b> was not identified in the FBMN analysis, it was assigned to compound <b>1</b> (luteolin-7-<span class="html-italic">O</span>-diglucuronide/luteolin-7-<span class="html-italic">O</span>-[<span class="html-italic">β</span>-D-glucuronosyl-(1→2)-<span class="html-italic">β</span>-D-glucuronide]) by the ProQI software ver. 3.0.</p>
Full article ">Figure 4
<p>Example of the component prediction of the extract using ProQI. Compound <b>1</b> was identified as luteolin-7-<span class="html-italic">O</span>-diglucuronide by analyzing the fragmentation patterns of the peak “<span class="html-italic">m</span>/<span class="html-italic">z</span> 637.1031”. This peak was divided into two substructures: a flavonoid (<span class="html-italic">m</span>/<span class="html-italic">z</span> 285.0392) and a diglucuronide (<span class="html-italic">m</span>/<span class="html-italic">z</span> 351.0555).</p>
Full article ">
14 pages, 2144 KiB  
Article
Exploration of Genome-Wide Recombination Rate Variation Patterns at Different Scales in Pigs
by Zuoquan Chen, Meng Zhou, Yingchun Sun, Xi Tang, Zhiyan Zhang and Lusheng Huang
Animals 2024, 14(9), 1345; https://doi.org/10.3390/ani14091345 - 29 Apr 2024
Viewed by 901
Abstract
Meiotic recombination is a prevalent process in eukaryotic sexual reproduction organisms that plays key roles in genetic diversity, breed selection, and species evolution. However, the recombination events differ across breeds and even within breeds. In this study, we initially computed large-scale population recombination [...] Read more.
Meiotic recombination is a prevalent process in eukaryotic sexual reproduction organisms that plays key roles in genetic diversity, breed selection, and species evolution. However, the recombination events differ across breeds and even within breeds. In this study, we initially computed large-scale population recombination rates of both sexes using approximately 52 K SNP genotypes in a total of 3279 pigs from four different Chinese and Western breeds. We then constructed a high-resolution historical recombination map using approximately 16 million SNPs from a sample of unrelated individuals. Comparative analysis of porcine recombination events from different breeds and at different resolutions revealed the following observations: Firstly, the 1Mb-scale pig recombination maps of the same sex are moderately conserved among different breeds, with the similarity of recombination events between Western pigs and Chinese indigenous pigs being lower than within their respective groups. Secondly, we identified 3861 recombination hotspots in the genome and observed medium- to high-level correlation between historical recombination rates (0.542~0.683) and estimates of meiotic recombination rates. Third, we observed that recombination hotspots are significantly far from the transcription start sites of pig genes, and the silico–predicted PRDM9 zinc finger domain DNA recognition motif is significantly enriched in the regions of recombination hotspots compared to recombination coldspots, highlighting the potential role of PRDM9 in regulating recombination hotspots in pigs. Our study analyzed the variation patterns of the pig recombination map at broad and fine scales, providing a valuable reference for genomic selection breeding and laying a crucial foundation for further understanding the molecular mechanisms of pig genome recombination. Full article
Show Figures

Figure 1

Figure 1
<p>Comparison of scatter plots of the ACM among different populations by gender. (<b>A</b>) Scatter-overlay boxplots of the average number of meiotic crossover events per division among different populations of males. The X-axis represents different populations; the Y-axis represents the ACM. “*” indicates a <span class="html-italic">p</span>-value &lt; 0.05; “**” indicates a <span class="html-italic">p</span>-value &lt; 0.01; “***” indicates a <span class="html-italic">p</span>-value &lt; 0.0001; NS indicates no significant difference. (<b>B</b>) Scatter-overlay boxplots of the average number of crossovers per meiotic crossover event among different populations of females.</p>
Full article ">Figure 2
<p>Autosomal recombination landscapes constructed by different software and parameters. From (<b>A</b>–<b>C</b>), the 1 Mb window recombination rates of pryho, FastEPRR 5 kb window, and FastEPRR 10 KB window, respectively.</p>
Full article ">Figure 3
<p>Chromosome distribution of recombination coldspots and hotspots. Comparison of GC content and Pi in coldspot and hotspot intervals. (<b>A</b>) Fine recombination map of chromosome 18. The red bar line represents the recombination rate in the possible recombination hotspots, the cyan bar line represents the recombination rate in the non-recombination hotspots, and the blue line below the X-axis represents the location of the possible recombination coldspots. (<b>B</b>) Line plots of GC content in different intervals. (<b>C</b>) Line plots of Pi in different intervals.</p>
Full article ">Figure 4
<p>(<b>A</b>) Comparison of the number of recombination hotspots and 1000 random points that have intersections near the TSS. The pink line represents the distribution density of the number of random points with intersections near the TSS 1000 times, and the blue line represents the number of recombination hotspots with intersections with the TSS. The <span class="html-italic">p</span>-value is calculated by <span class="html-italic">t</span>-test. (<b>B</b>) Chromosome 18 recombination hotspots and the distribution of H3K4me3 peaks visualized in IGV, with partial regions magnified. Among them, red represents the position interval of the recombination hotspot, green represents the position interval of the H3K4me3 peak, and blue represents the annotation gene of the pig reference genome.</p>
Full article ">Figure 5
<p>Predicted <span class="html-italic">PRDM9</span> gene zinc finger domain 15-mer DNA recognition sequence. Four different colors and letters represent four different bases, respectively.</p>
Full article ">
0 pages, 2142 KiB  
Article
Genomic Inbreeding and Runs of Homozygosity Analysis of Cashmere Goat
by Qian Zhao, Chang Huang, Qian Chen, Yingxiao Su, Yanjun Zhang, Ruijun Wang, Rui Su, Huijuan Xu, Shucai Liu, Yuehui Ma, Qianjun Zhao and Shaohui Ye
Animals 2024, 14(8), 1246; https://doi.org/10.3390/ani14081246 - 22 Apr 2024
Viewed by 1254
Abstract
Cashmere goats are valuable genetic resources which are famous worldwide for their high-quality fiber. Runs of homozygosity (ROHs) have been identified as an efficient tool to assess inbreeding level and identify related genes under selection. However, there is limited research on ROHs in [...] Read more.
Cashmere goats are valuable genetic resources which are famous worldwide for their high-quality fiber. Runs of homozygosity (ROHs) have been identified as an efficient tool to assess inbreeding level and identify related genes under selection. However, there is limited research on ROHs in cashmere goats. Therefore, we investigated the ROH pattern, assessed genomic inbreeding levels and examined the candidate genes associated with the cashmere trait using whole-genome resequencing data from 123 goats. Herein, the Inner Mongolia cashmere goat presented the lowest inbreeding coefficient of 0.0263. In total, we identified 57,224 ROHs. Seventy-four ROH islands containing 50 genes were detected. Certain identified genes were related to meat, fiber and milk production (FGF1, PTPRM, RERE, GRID2, RARA); fertility (BIRC6, ECE2, CDH23, PAK1); disease or cold resistance and adaptability (PDCD1LG2, SVIL, PRDM16, RFX4, SH3BP2); and body size and growth (TMEM63C, SYN3, SDC1, STRBP, SMG6). 135 consensus ROHs were identified, and we found candidate genes (FGF5, DVL3, NRAS, KIT) were associated with fiber length or color. These findings enhance our comprehension of inbreeding levels in cashmere goats and the genetic foundations of traits influenced by selective breeding. This research contributes significantly to the future breeding, reservation and use of cashmere goats and other goat breeds. Full article
(This article belongs to the Section Animal Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Population genetics structures analyses and linkage disequilibrium. (<b>A</b>) Principle component analysis (PCA); each point represents a single individual. The colors in B and D represent the same groups as indication in the legend of A. (<b>B</b>) Phylogenetic tree constructed by the neighbor-joining (NJ) method. (<b>C</b>) Population structure plot of seven goat populations at K = 2–5, (<b>D</b>) LD decay map measured by r² over distance between SNPs in seven populations.</p>
Full article ">Figure 2
<p>The distribution of ROHs identified in different populations across autosomes. (<b>A</b>) This graph is a bar chart showing the number of ROHs on different chromosomes. ROHs are divided into three length classes: small (0–0.3 Mb), medium (0.3–1.5 Mb) and large (&gt;1.5 Mb). (<b>B</b>) Relationship between the number of runs of homozygosity (ROH) per individual and the total length of the genome covered by them. The x-axis shows the sum total length of ROHs (Mb), and the y-axis shows the total number of ROHs. Every circle represents a different individual within a population, with the groups labeled as ALG, HSC, IMC and so on. (<b>C</b>) Violin plots and boxplots of sum total length of ROHs (SROHs). The form of each violin reflects the distribution density of cohort, with wider sections of the violin plot indicating higher frequency of observation at that y -value. Inside each violin, there is a boxplot that shows the median, the interquartile range (the length of the box) and the outliers (the points outside the whiskers). (<b>D</b>) The descriptive statistics for the ROHs, categorized by ROH length class across different populations, are presented as mean counts of ROHs (Y-axis) by class of ROH length.</p>
Full article ">Figure 3
<p>Manhattan plot of SNPs in ROHs for HSC (<b>A</b>) and BOE (<b>B</b>). The x-axis exhibits positions along each chromosome.</p>
Full article ">Figure 4
<p>(<b>A</b>) A Gene Ontology (GO) functional enrichment analysis. (<b>B</b>) KEGG enrichment pathways of gene annotation in the consensus ROHs of IMC and HSC breeds.</p>
Full article ">
14 pages, 3153 KiB  
Article
Ancistrocladus tectorius Extract Inhibits Obesity by Promoting Thermogenesis and Mitochondrial Dynamics in High-Fat Diet-Fed Mice
by Minju Kim, Jin Hyub Paik, Hwa Lee, Min Ji Kim, Sang Mi Eum, Soo Yong Kim, Sangho Choi, Ho-Yong Park, Hye Gwang Jeong and Tae-Sook Jeong
Int. J. Mol. Sci. 2024, 25(7), 3743; https://doi.org/10.3390/ijms25073743 - 27 Mar 2024
Viewed by 1158
Abstract
Root extracts of Ancistrocladus tectorius (AT), a shrub native to China, have been shown to have antiviral and antitumor activities, but the anti-obesity effects of AT aerial parts, mainly the leaves and stems, have not been investigated. This study is the first to [...] Read more.
Root extracts of Ancistrocladus tectorius (AT), a shrub native to China, have been shown to have antiviral and antitumor activities, but the anti-obesity effects of AT aerial parts, mainly the leaves and stems, have not been investigated. This study is the first to investigate the anti-obesity effects and molecular mechanism of AT 70% ethanol extract in 3T3-L1 adipocytes and high-fat diet (HFD)-fed C57BL/6J mice. Treatment with AT extract inhibited lipid accumulation in 3T3-L1 cells and decreased the expression of adipogenesis-related genes. AT extract also upregulated the mRNA expression of genes related to mitochondrial dynamics in 3T3-L1 adipocytes. AT administration for 12 weeks reduced body weight and organ weights, including liver, pancreas, and white and brown adipose tissue, and improved plasma profiles such as glucose, insulin, homeostasis model assessment of insulin resistance, triglyceride (TG), and total cholesterol in HFD-fed mice. AT extract reduced HFD-induced hepatic steatosis with levels of liver TG and lipogenesis-related genes. AT extract upregulated thermogenesis-related genes such as Cidea, Pgc1α, Ucp1, Prdm16, Adrb1, and Adrb3 and mitochondrial dynamics-related genes such as Mff, Opa1, and Mfn2 in brown adipose tissue (BAT). Therefore, AT extract effectively reduced obesity by promoting thermogenesis and the mitochondrial dynamics of BAT in HFD-fed mice. Full article
(This article belongs to the Special Issue Insulin Resistance and Metabolic Syndrome)
Show Figures

Figure 1

Figure 1
<p>Effect of AT extract and its BuOH fraction on cytotoxicity and adipogenesis in 3T3-L1 adipocytes. Cytotoxicity of AT extract (<b>A</b>) and its BuOH fraction (<b>D</b>) in 3T3-L1 cells. (<b>B</b>) Quantification of lipid accumulation of AT extract (<b>B</b>) and its BuOH fraction (<b>E</b>) in differentiated adipocytes assessed by Oil Red O staining. The stained lipid was extracted from the cells with isopropanol and its absorbance was monitored spectrophotometrically at 490 nm. (<b>C</b>,<b>F</b>) Relative mRNA levels of adipogenesis-related genes were determined 4 days after sample treatment. AT 70E: 70% EtOH extract of AT; AT BuOH Fr.: BuOH fraction of AT 70% EtOH extract. Values are presented as means ± SD. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.01 versus pre-adipocytes, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 versus adipocytes.</p>
Full article ">Figure 2
<p>Effect of the AT extract and its BuOH fraction on genes related to mitochondrial dynamics in mature 3T3-L1 adipocytes. Two days after confluence, 3T3-L1 pre-adipocytes (day 0) were treated with samples at 10 or 20 μg/mL every other day for 4 days. The mRNA expression was measured on fully differentiated adipocytes (Day 8). AT 70E: 70% EtOH extract of AT; AT BuOH Fr.: BuOH fraction of AT 70% EtOH extract. Values are expressed as mean ± SD. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.01 versus pre-adipocytes, ** <span class="html-italic">p</span> &lt; 0.01 versus adipocytes.</p>
Full article ">Figure 3
<p>Effect of AT extract on oral glucose tolerance test (OGTT) in HFD-fed mice. (<b>A</b>) Blood glucose levels were measured in tail vein blood at 0, 15, 30, 45, 60, 90, 120, and 150 min. (<b>B</b>) Area under the curve (AUC) of changed glucose levels during OGTT. Data are expressed as mean ± SE (<span class="html-italic">n</span> = 10). Different letters (a, b, c) within a variable are significantly different at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Effect of AT extract on hepatic lipid metabolism in HFD-fed mice. (<b>A</b>) Histology of livers stained with hematoxylin and eosin (H&amp;E) (×200 magnification); (<b>B</b>) hepatic TG and (<b>C</b>) TC levels; (<b>D</b>) Lipogenesis-related mRNA expression levels were measured by real-time qRT-PCR. Data are expressed as mean ± SE. Different letters (a, b) within a variable are significantly different at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>Effects of AT extract on adipocyte hypertrophy in HFD-fed mice. (<b>A</b>) Histology of WAT and BAT stained with H&amp;E (200× magnification). (<b>B</b>) Quantitative measurement of adipocyte size. Data are expressed as mean ± SE. Different letters (a, b, c) within a variable are significantly different at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Effect of AT extract on the thermogenesis in BAT of HFD-fed mice. (<b>A</b>–<b>C</b>) The mRNA expression levels were measured by real-time qRT-PCR in the BAT. Data are expressed as mean ± SE. Different letters (a, b, c) within a variable are significantly different at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 7
<p>Effect of AT extract on the mitochondrial dynamics in BAT of HFD-fed mice. The mRNA expression levels were measured by real-time qRT-PCR in the BAT. Data are expressed as mean ± SE. Different letters (a, b, c) within a variable are significantly different at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 8
<p>Molecular mechanisms of the anti-obesity effects of AT extract. AT extract suppresses obesity by enhancing thermogenesis and mitochondrial dynamics in BAT.</p>
Full article ">
19 pages, 10099 KiB  
Article
Reproductive Ability Disparity in the Pacific Whiteleg Shrimp (Penaeus vannamei): Insights from Ovarian Cellular and Molecular Levels
by Jianchun Zhang, Jie Kong, Jiawang Cao, Ping Dai, Baolong Chen, Jian Tan, Xianhong Meng, Kun Luo, Qiang Fu, Peiming Wei, Sheng Luan and Juan Sui
Biology 2024, 13(4), 218; https://doi.org/10.3390/biology13040218 - 27 Mar 2024
Viewed by 1259
Abstract
The Pacific whiteleg shrimp (Penaeus vannamei) is a highly significant species in shrimp aquaculture. In the production of shrimp larvae, noticeable variations in the reproductive capacity among female individuals have been observed. Some females experience slow gonadal development, resulting in the [...] Read more.
The Pacific whiteleg shrimp (Penaeus vannamei) is a highly significant species in shrimp aquaculture. In the production of shrimp larvae, noticeable variations in the reproductive capacity among female individuals have been observed. Some females experience slow gonadal development, resulting in the inability to spawn, while others undergo multiple maturations and contribute to the majority of larval supply. Despite numerous studies that have been conducted on the regulatory mechanisms of ovarian development in shrimp, the factors contributing to the differences in reproductive capacity among females remain unclear. To elucidate the underlying mechanisms, this study examined the differences in the ovarian characteristics between high and low reproductive bulks at different maturity stages, focusing on the cellular and molecular levels. Transmission electron microscopy analysis revealed that the abundance of the endoplasmic reticulum, ribosomes, mitochondria, and mitochondrial cristae in oocytes of high reproductive bulk was significantly higher than that of the low reproductive bulk in the early stages of ovarian maturation (stages I and II). As the ovaries progressed to late-stage maturation (stages III and IV), differences in the internal structures of oocytes between females with different reproductive capacities gradually diminished. Transcriptome analysis identified differentially expressed genes (DEGs) related to the mitochondria between two groups, suggesting that energy production processes might play a crucial role in the observed variations in ovary development. The expression levels of the ETS homology factor (EHF) and PRDI-BF1 and RIZ homology domain containing 9 (PRDM9), which were significantly different between the two groups, were compared using qRT-PCR in individuals at different stages of ovarian maturation. The results showed a significantly higher expression of the EHF gene in the ovaries of high reproductive bulk at the II and IV maturity stages compared to the low reproductive bulk, while almost no expression was detected in the eyestalk tissue of the high reproductive bulk. The PRDM9 gene was exclusively expressed in ovarian tissue, with significantly higher expression in the ovaries of the high reproductive bulk at the four maturity stages compared to the low reproductive bulk. Fluorescence in situ hybridization further compared the expression patterns of EHF and PRDM9 in the ovaries of individuals with different fertility levels, with both genes showing stronger positive signals in the high reproductive bulk at the four ovarian stages. These findings not only contribute to our understanding of the regulatory mechanisms involved in shrimp ovarian development, but also provide valuable insights for the cultivation of new varieties aimed at improving shrimp fecundity. Full article
Show Figures

Figure 1

Figure 1
<p>Maturation frequency of 632 shrimp females in a production cycle (45 d) of the <span class="html-italic">P. vannamei</span>.</p>
Full article ">Figure 2
<p>Transmission electron microscopy of ovaries with high and low reproductive bulks in stages I–IV. (<b>a</b>) Long tubular rough endoplasmic reticulum in oocytes of high reproductive bulk in stage I; (<b>b</b>) long tubular rough endoplasmic reticulum in oocytes of low reproductive bulk in stage I; (<b>c</b>) organelles within the cytoplasm of high reproductive bulk in stage II; (<b>d</b>) mitochondria of high reproductive bulk in stage II; (<b>e</b>) organelles within the cytoplasm of low reproductive bulk in stage II; (<b>f</b>) mitochondria of low reproductive bulk in stage II; (<b>g</b>) mitochondria of low reproductive bulk in stage III; (<b>h</b>) long tubular rough endoplasmic reticulum in oocytes of low reproductive bulk in stage III; (<b>i</b>,<b>j</b>) intracytoplasmic yolk granules and cortical rods of high and low reproductive bulks in stage IV. RER: Long tubular rough endoplasmic reticulum; M: Mitochondria; R: Ribosome; G: Golgi apparatus; Y: yolk granules; CR: Cortical rod.</p>
Full article ">Figure 3
<p>Differential gene volcano map of high vs. low reproductive bulk. The abscissa is the log2 value of the fold difference; the ordinate is the −log10 value of the <span class="html-italic">p</span>−value or FDR value; the red scatter points represent upregulated differentially expressed genes; the yellow scatter points represent downregulated differentially expressed genes; the blue scatter points represent genes that do not meet the threshold screening.</p>
Full article ">Figure 4
<p>GO annotation classification statistics of high vs. low reproductive bulk. The abscissa is the secondary GO term, and the ordinate is the number of genes in the term. Orange indicates upregulation, and blue indicates downregulation. <span class="html-italic">p</span> &lt; 0.05 and |log<sub>2</sub>FC| &gt; 1.</p>
Full article ">Figure 5
<p>GO enrichment of differentially expressed genes in the Top 20. The abscissa is Gene Ratio, and the ordinate is GO Term. The size of the bubbles indicates the number of differences in enrichment into the GO term. The color of the bubbles indicates significant enrichment in the GO term.</p>
Full article ">Figure 6
<p>KEGG signaling pathway enrichment analysis of differentially expressed genes in ovarian tissue of high vs. low reproductive bulk. The abscissa is the number of genes, the ordinate is the pathway name, each column represents a pathway, and the column height represents the number of genes contained in the pathway. Different colors represent different first level classifications.</p>
Full article ">Figure 7
<p>KEGG enrichment of differentially expressed genes in the Top 20. The abscissa is gene ratio, and the ordinate is the pathway name. The size of the bubble indicates the number of differentially expressed genes enriched in the pathway. The color of the bubble indicates the significance of enrichment in the pathway.</p>
Full article ">Figure 8
<p>The expression pattern of the EHF gene in different tissues of high and low reproductive bulks at different ovary stages. * represents a significant difference between the high and low reproductive bulks at the same stage (<span class="html-italic">p</span> &lt; 0.05), ** represents a highly significant difference between the high and low reproductive bulks at the same stage (<span class="html-italic">p</span> &lt; 0.01). I–IV represents different stages of ovarian development.</p>
Full article ">Figure 9
<p>Fluorescence in situ hybridization results of the EHF gene in the I–IV ovaries of the high and low reproductive bulks. (<b>a</b>–<b>d</b>) Fluorescence in situ hybridization results of I–IV stage ovaries in the high reproductive bulk; (<b>e</b>–<b>h</b>) Fluorescence in situ hybridization results of I–IV stage ovaries in the low reproductive bulk.</p>
Full article ">Figure 10
<p>The expression pattern of PRDM9 gene in ovary of high and low reproductive bulks at different ovary stages. * represents significant difference between the high and low reproductive bulks at the same stage (<span class="html-italic">p</span> &lt; 0.05), ** represents highly significant difference between the high and low reproductive bulks at the same stage (<span class="html-italic">p</span> &lt; 0.01). I–IV represents different stages of ovarian development.</p>
Full article ">Figure 11
<p>Fluorescence in situ hybridization results of the PRDM9 gene in the I–IV ovaries of the high and low reproductive bulks. (<b>a</b>–<b>d</b>) Fluorescence in situ hybridization results of I–IV stage ovaries in the high reproductive bulk; (<b>e</b>–<b>h</b>) Fluorescence in situ hybridization results of I–IV stage ovaries in the low reproductive bulk.</p>
Full article ">
30 pages, 6292 KiB  
Review
Progesterone-Related Diabetes Mellitus in the Bitch: Current Knowledge, the Role of Pyometra, and Relevance in Practice
by Álan Gomes Pöppl, José Lucas Xavier Lopes, Taís Bock Nogueira, Denise Iparraguirre da Silva and Bruna dos Santos Machado
Animals 2024, 14(6), 890; https://doi.org/10.3390/ani14060890 - 14 Mar 2024
Cited by 1 | Viewed by 2286
Abstract
Progesterone-related diabetes mellitus (PRDM) in dogs is known for its particular potential for diabetes remission. This narrative review aims to provide relevant detailed information on (1) the canine estrus cycle and its impact on canine diabetes mellitus (CDM) etiology and management, (2) the [...] Read more.
Progesterone-related diabetes mellitus (PRDM) in dogs is known for its particular potential for diabetes remission. This narrative review aims to provide relevant detailed information on (1) the canine estrus cycle and its impact on canine diabetes mellitus (CDM) etiology and management, (2) the role of pyometra as a further cause of insulin resistance, and (3) useful individual therapeutic and preventive strategies. PRDM is recognized due to diestrus, exogenous progestogen exposure, pregnancy, and P4-production ovarian dysfunction. Pyometra represents additional inflammatory and septic negative influence on insulin sensitivity, and its diagnosis associated with CDM is therapeutically challenging. The estrus cycle’s hormone fluctuations seem to modulate peripheric insulin sensibility by influencing insulin receptor (IR) affinity and its binding capacity, as well as modulating tyrosine kinase activity. Pyometra was shown to negatively influence IR compensatory mechanisms to insulin resistance causing glucose intolerance. Spaying and pregnancy termination may cause diabetes remission in PRDM cases in a median time of 10 days (1–51). Pharmacological annulment of progesterone effects may benefit patients unable to undergo surgery; however, remission chances are virtually null. The ALIVE (Agreeing Language in Veterinary Endocrinology) project proposed new criteria for CDM diagnoses and subclinical diabetes recognition. These new concepts may increase the frequency of a PRDM diagnosis and, even more, its relevance. Spaying represents a preventive measure against pyometra and PRDM that should be individually assessed in light of its recognized benefits and harms. Full article
(This article belongs to the Section Animal Physiology)
Show Figures

Figure 1

Figure 1
<p>Hormonal fluctuations during the canine estrus cycle. The scheme represents pituitary gonadotrophin (FSH, LH, and prolactin) patterns, as well as resultant ovarian E2 and P4 responses. The P4 influence on basal GH concentration and pulsatility pattern are also represented. The FSH line (blue) represents fluctuations between 15 and 40 and 200 and 400 ng/mL during the estrus cycle, while the LH line (yellow) represents variations from 0.4 to 1.5 to 5 to 40 ng/mL, and the prolactin line (green) shows fluctuations ranging from 0.5 to 2 to 5 to 30 ng/mL. E2 (pink) and P4 (red) lines range from 5 to 10 pg/mL and 0.2 to 0.5 ng/mL when in basal levels to 45 to 120 pg/mL and 15 to 90 ng/mL, respectively [<a href="#B37-animals-14-00890" class="html-bibr">37</a>]. Basal GH concentration during anestrus varies around 1.4 ± 0.2 µg<sup>−1</sup> and can reach values around 2.3 ± 0.2 µg<sup>−1</sup> 18–20 days after ovulation. During this early diestrus, the GH pulsatile secretion pattern is marked as reduced from ~5 to ~2 peaks/12 h as well as the pulse duration from ~41 to ~11 min compared with anestrus. All these GH pulse and concentration abnormalities were slowly mitigated during the subsequent weeks following P4 throwback to basal levels [<a href="#B47-animals-14-00890" class="html-bibr">47</a>].</p>
Full article ">Figure 2
<p>Conditions under progesterone influence that are involved with the progesterone-related diabetes mellitus (PRDM) subtype in dogs. Pyometra is a progesterone-dependent disease secondary to diestrus, exogenous progestogen exposure, or ovarian dysfunctions adding inflammatory and septic negative influence on insulin sensitivity and PRDM risk. To date, progesterone-secreting adrenal tumors were not associated with diabetes in dogs, despite being considered a common complication in cats. Pink circles represent general genetic and environmental factors that may influence PRDM occurrence and that are further detailed later in the text.</p>
Full article ">Figure 3
<p>Hypothetical causal diagram for P4-related diabetes mellitus in dogs [<a href="#B2-animals-14-00890" class="html-bibr">2</a>,<a href="#B5-animals-14-00890" class="html-bibr">5</a>,<a href="#B64-animals-14-00890" class="html-bibr">64</a>,<a href="#B123-animals-14-00890" class="html-bibr">123</a>]. The arrows represent evidence obtained from studies with dogs, but also from studies with humans, animal models, and cell cultures. The main determinants of glycemia (purple) are beta-cell function and peripheric insulin sensitivity (red), both influenced by intrinsic factors such as age and genetics (blue). Factors shown in yellow can potentially be modified by proper patient management. Glycemia can also affect beta-cell function by glucotoxicity (heavy double arrow). Serial unknown factors may exert positive or negative effects on beta-cell function (“black box”) and influence the patient’s ability to increase insulin secretion to avoid hyperglycemia in an insulin-resistant environment. Those factors are potentially also determinants of diabetic remission after resolution of P4-related conditions.</p>
Full article ">Figure 4
<p>Ten-year-old mixed-breed dog with diabetes diagnosis three weeks after estrus. (<b>a</b>) The general patient’s appearance shows weight loss and increased thick folding of the skin over the withers (<b>b</b>) and neck (<b>d</b>). (<b>c</b>) Discrete widening of the interdental space and gingival hyperplasia. (<b>e</b>) A palpable flat mammary mass measuring nearly 3 cm in its highest diameter was identified during the physical exam (red circle). Insulin therapy was started (NPH, 0.4 U/kg, q12h) associated with a diabetic commercial food (Royal Canin Diabetic). On day 5, the insulin dose was adjusted to 0.55 U/kg, q12h, due to initial poor clinical and glycemic responses. Three days later, a hypoglycemic episode was documented. The insulin dose was reduced back to the initial dose, but new hypoglycemic episodes were documented in the following days. Insulin was interrupted on day 11 and diabetic remission was associated with natural P4 reduction in the second diestrus half. The owner refuses to proceed with spaying surgery, and the dog relapsed into diabetes five months later after the subsequent estrus. At this time, the bitch was spayed and underwent a mastectomy without achieving diabetic remission a second time. Her insulin need was dramatically reduced after surgery (0.2 U/kg, q12h). The dog was followed until death four years later and never achieved remission again despite maintaining a relatively “low” insulin requirement over time.</p>
Full article ">Figure 5
<p>Nine-year-old Frech Bulldog presented with poorly controlled diabetes diagnosed after the last estrus five weeks before. (<b>a</b>) Thick skin folds in the neck and (<b>b</b>) marked widened interdental space with gingival hyperplasia and discrete enlarged tongue. The bitch also had a mammary mass and was submitted to ovariohysterectomy and mastectomy. Diabetic remission was achieved four weeks later. At fourteen years, the dog relapsed due to chronic kidney disease. (<b>c</b>) Note the clear reduction in soft tissue overgrowth and skin folding; (<b>d</b>) however, bony changes documented in the mouth persisted.</p>
Full article ">Figure 6
<p>Five-year-old American Stratford Shire bitch presented with diabetes diagnosed weeks after the last heat. Acromegalic features such as (<b>a</b>) increased thick skin folds, (<b>b</b>) widened interdental space, and (<b>c</b>) discrete macroglossia and prognathism were observed. Panting and snoring were also reported by the owners. The bitch was submitted to ovariohysterectomy, and (<b>d</b>) an enlarged uterus filled with a serosanguinous content diagnosed as hematometra and (<b>e</b>) bilateral polycystic ovaries were found. (<b>f</b>) Weight gain and muscle mass were restored after adequate diabetic control, as well as skin folds becoming regressed; however, despite that diabetic remission was not achieved, glycemic control was consistently improved after surgery with a ~40% smaller insulin dose.</p>
Full article ">Figure 7
<p>Evolution of acromegalic features over time in a mixed-breed intact dog. (<b>a</b>) Facial and neck skin aspect by the age of five years. By the age of eight years, there was (<b>b</b>,<b>c</b>) skin folding on the ventral neck and over the wither, and (<b>d</b>) increased interdental space was starting to become evident. (<b>e</b>) Mammary glands were well demarked despite being in anestrus and that mating never occurred. Ovariohysterectomy was performed as a preventive measure against PRDM when she was nine years old. (<b>f</b>) Increased skin folding in the ventral neck by the time surgery was performed. The bitch never developed diabetes and died two years later due to a liver tumor.</p>
Full article ">
19 pages, 3426 KiB  
Article
The Differences in the Developmental Stages of the Cardiomyocytes and Endothelial Cells in Human and Mouse Embryos at the Single-Cell Level
by Chuyu Liu and Ning-Yi Shao
Int. J. Mol. Sci. 2024, 25(6), 3240; https://doi.org/10.3390/ijms25063240 - 13 Mar 2024
Cited by 1 | Viewed by 1409
Abstract
Our research focuses on expression patterns in human and mouse embryonic cardiomyocytes and endothelial cells at the single-cell level. We analyzed single-cell datasets containing different species, cardiac chambers, and cell types. We identified developmentally dynamic genes associated with different cellular lineages in the [...] Read more.
Our research focuses on expression patterns in human and mouse embryonic cardiomyocytes and endothelial cells at the single-cell level. We analyzed single-cell datasets containing different species, cardiac chambers, and cell types. We identified developmentally dynamic genes associated with different cellular lineages in the heart and explored their expression and possible roles during cardiac development. We used dynamic time warping, a method that aligns temporal sequences, to compare these developmental stages across two species. Our results indicated that atrial cardiomyocytes from E9.5 to E13.5 in mice corresponded to a human embryo age of approximately 5–6 weeks, whereas in ventricular cardiomyocytes, they corresponded to a human embryo age of 13–15 weeks. The endothelial cells in mouse hearts corresponded to 6–7-week-old human embryos. Next, we focused on expression changes in cardiac transcription factors over time in different species and chambers, and found that Prdm16 might be related to interspecies cardiomyocyte differences. Moreover, we compared the developmental trajectories of cardiomyocytes differentiated from human pluripotent stem cells and embryonic cells. This analysis explored the relationship between their respective developments and provided compelling evidence supporting the relevance of our dynamic time-warping results. These significant findings contribute to a deeper understanding of cardiac development across different species. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Cell types in the human embryonic heart, identified through an scRNA-seq analysis. The uniform manifold approximation and projection (UMAP) shows all filtered cells. (<b>A</b>) Unbiased clusters are indicated by different colors, identified using Louvain clustering. (<b>B</b>) The colors indicate the developmental stages from 5 weeks (5W) to 17W of gestation. (<b>C</b>) The colors represent the locations from which the cells were sourced. (<b>D</b>) Cell-type marker genes in the clusters were identified in (<b>A</b>). Bar: CMs-A, dark red; CMs-V, green; ECs, pink. The dot sizes represent the fraction of cells within each cluster. The color shades of the dots represent the mean expression level of the genes in the cluster. Same as below. (<b>E</b>) In the UMAP visualization, cardiomyocyte chamber-specific and endothelial cell-specific markers are highlighted. (<b>F</b>) The heatmap displays the Z-score-scaled average expression levels of DEGs for each cluster identified in (<b>A</b>) within the human embryonic heart.</p>
Full article ">Figure 2
<p>Cell types identified in the cardiac muscle and endothelial trajectory of mouse embryos through an scRNA-seq analysis. The UMAP visualization represents all cells within the cardiac muscle lineage. (<b>A</b>) The different colors represent distinct clusters. (<b>B</b>) The shades of color correspond to developmental stages, ranging from day 9.5 to day 13.5 of gestation. (<b>C</b>) The different colors show various trajectories. (<b>D</b>) The cell-type marker genes in the clusters identified in (<b>A</b>). Bar: CMs-A, blue; CMs-V, purple. (<b>E</b>) The UMAP visualization highlights the markers specific to cardiomyocyte chambers. (<b>F</b>) The heatmap indicates the Z-score-scaled average expression levels of DEGs for each cluster, as identified in (<b>A</b>), in the cardiac muscle. The color of the dot corresponds to those in (<b>A</b>). (<b>G</b>–<b>I</b>) All cells within the endothelial lineage are presented using UMAP visualization. (<b>G</b>) The different colors correspond to distinct clusters. (<b>H</b>) The shades of color represent developmental stages from day 9.5 to day 13.5 of gestation. (<b>I</b>) The various colors denote the trajectories of different cell types. (<b>J</b>) The cell-type marker genes in the clusters identified in (<b>G</b>). (<b>K</b>) The heatmap exhibits the Z-score-scaled average expression levels of DEGs within each cluster identified in G in the endothelial lineage. The color of the dot corresponds to those in (<b>G</b>).</p>
Full article ">Figure 3
<p>Identification of heart-related developmentally dynamic genes (DDGs). (<b>A</b>) A PCA (principal component analysis) conducted based on a 1:1 orthologue across both species delineated CMs-A and CMs-V according to the pseudo-bulk transcriptomes. The shades of color represent the developmental stages of the samples, spanning from 5 to 17 weeks (5W–17W) of gestation in humans and from embryonic day 9.5 to 13.5 in mice. (<b>B</b>) As in (<b>A</b>), the PCA delineated ECs in both species. (<b>C</b>) Expression patterns of DDGs in cardiomyocytes and endothelial cells. From left to right: CMs-A, CMs-V, and ECs. In the graph, each point represents an individual gene, and the lines show cubic spline curves. The different colors denote the species: green for humans and orange for mice. The <span class="html-italic">x</span>-axis represents developmental stages on the log2 scale (up: human; down: mouse), while the shaded grey area represents a 95% confidence interval. The <span class="html-italic">y</span>-axis represents standardized expression levels of all the genes with a mean of 0 and a standard deviation (SD) of 1. The titles on top of each panel show the gene number in each group. Same as below.</p>
Full article ">Figure 4
<p>The alignment of dynamic time warping (DTW) between humans and mice to compare developmental stages. (<b>A</b>) Comparison of the stage correspondence based on DTW alignment in CMs-A between humans and mice. (<b>B</b>) DTW alignment in CMs-V. (<b>C</b>) DTW alignment in ECs.</p>
Full article ">Figure 5
<p>Transcription factors in DEGs and the combined analysis with the <span class="html-italic">Prdm16<sup>cKO</sup></span> cardiomyocytes in mice. The volcano plots present the DEGs in cardiomyocytes. The <span class="html-italic">x</span>-axis corresponds to fold changes measured on a log2 scale. The <span class="html-italic">y</span>-axis corresponds to the <span class="html-italic">p</span>-value measured on a −log10 scale. (<b>A</b>) The impacts of the species in cardiomyocyte development. (<b>B</b>) The impacts of the species in cardiomyocyte development when controlling for the effect of chambers.</p>
Full article ">Figure 6
<p>The combined analysis with cardiomyocytes from different hPSC-derived chambers and dynamic time warping (DTW). (<b>A</b>) A PCA based on a 1:1 orthologue within the datasets of embryonic heart cardiomyocytes and hPSC-derived cardiomyocytes. The origin of cardiomyocytes is represented by different shapes, and the developmental stages are indicated by the sizes of the dots. (<b>B</b>) The 3D PCA, showing PCs 3, 4, and 5, with the same characteristics as in (<b>A</b>). (<b>C</b>) The stage correspondence was compared based on the DTW alignment in atrial cardiomyocytes (CMs-A) between the two species. hPSC-derived cardiomyocytes were added preceding the embryonic stage. The same was performed in (<b>D</b>). (<b>D</b>) DTW alignment in CMs-V.</p>
Full article ">
Back to TopTop