[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

Journals

Article Types

Countries / Regions

Search Results (252)

Search Parameters:
Keywords = calcitriol

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
9 pages, 229 KiB  
Article
Eosinophilia as Monitoring Parameter for Chronic Graft-versus-Host Disease and Vitamin D Metabolism as Monitoring Parameter for Increased Infection Rates in Very Long-Term Survivors of Allogeneic Stem Cell Transplantation—A Prospective Clinical Study
by Thomas Neumann, Nadette Peters, Laila Schneidewind and William Krüger
BioMed 2024, 4(3), 293-301; https://doi.org/10.3390/biomed4030023 - 27 Aug 2024
Viewed by 488
Abstract
Background: Our aim is to investigate cardiovascular risk factors, chronic graft-versus-host disease (CGvHD), and vitamin D metabolism in very long-term survivors of adult allogeneic stem cell transplantation (alloSCT). Methods: This study is a prospective unicentric, non-interventional trial. The detailed study protocol is available [...] Read more.
Background: Our aim is to investigate cardiovascular risk factors, chronic graft-versus-host disease (CGvHD), and vitamin D metabolism in very long-term survivors of adult allogeneic stem cell transplantation (alloSCT). Methods: This study is a prospective unicentric, non-interventional trial. The detailed study protocol is available via the WHO Clinical Trial Registry. Results: We were able to include 33 patients with a mean age of 60.5 years (SD 11.1). Acute myeloid leukemia (AML) was the most frequent underlying disease (n = 12; 36.4%). The median survival time was 9.0 years (IQR 8.5–13.0). Relevant cardiovascular risk factors in the study population are the body mass index, cholesterol, LDL cholesterol, and lipoprotein(a). Cardiovascular risk factors have no significant impact on HRQoL. CGvHD of the skin as a limited disease was present in six patients (18.2%), and it has no impact on HRQoL. CGvHD was significantly associated with eosinophilia in peripheral blood (p = 0.003). Three patients (9.1%) had a shortage of calcitriol, and one patient (3.0%) took calcium substitution. The shortage is significantly associated with increased infection rates (p = 0.038). Conclusions: Cardiovascular risk factors and CGvHD need to be closely monitored. Eosinophilia might be a good and convenient monitoring parameter for CGvHD. Full article
49 pages, 3691 KiB  
Review
Anti-Inflammatory Benefits of Vitamin D and Its Analogues against Glomerulosclerosis and Kidney Diseases
by Theodora Adamantidi, George Maris, Petroula Altantsidou and Alexandros Tsoupras
Sclerosis 2024, 2(3), 217-265; https://doi.org/10.3390/sclerosis2030015 - 26 Aug 2024
Viewed by 1105
Abstract
Apart from the significant progress the scientific community has made during the last few decades, inflammation-mediated kidney-related diseases like chronic and diabetic kidney diseases (CKD and DKD) and glomerulosclerosis still continue to raise mortality rates. Recently, conventional therapeutic interventions have been put aside, [...] Read more.
Apart from the significant progress the scientific community has made during the last few decades, inflammation-mediated kidney-related diseases like chronic and diabetic kidney diseases (CKD and DKD) and glomerulosclerosis still continue to raise mortality rates. Recently, conventional therapeutic interventions have been put aside, since natural vitamin D-derived treatment has gained attention and offered several promising outcomes. Within this article, the utilization of vitamin D and its analogues as potential treatment toward kidney-related diseases, due to their anti-inflammatory, antioxidant and anti-fibrotic activity, is outlined. Vitamin D analogues including calcitriol, paricalcitol and 22-oxacalcitriol have been previously explored for such applications, but their hidden potential has yet to be further elucidated. Several clinical trials have demonstrated that vitamin D analogues’ supplementation is correlated with inflammatory signaling and oxidative stress regulation, immunity/metabolism augmentation and subsequently, kidney diseases and healthcare-related infections’ prevention, and the results of these trials are thoroughly evaluated. The highlighted research outcomes urge further study on a plethora of vitamin D analogues with a view to fully clarify their potential as substantial anti-inflammatory constituents of renal diseases-related treatment and their health-promoting properties in many kidney-associated healthcare complications and infections. Full article
Show Figures

Figure 1

Figure 1
<p>Vitamin D and its analogues in glomerulosclerosis and kidney disease therapy.</p>
Full article ">Figure 2
<p>The role of important inflammation biomarkers in CKD and other kidney-related diseases progression as well as several complications’ enhancement.</p>
Full article ">Figure 3
<p>Vitamin D’s two equal forms: derivatives. Vitamin D<sub>2</sub> has one more double bond (c<sub>22–23</sub>) in its molecule in opposition to vitamin D<sub>3</sub>.</p>
Full article ">Figure 4
<p>The production of vitamin D<sub>3</sub> from 7-DHC in the skin epidermis. 7-DHC is converted to pre-vitamin D<sub>3</sub> via the exposure to UV (UV-B) and then pre-vitamin D<sub>3</sub> is thermally rearranged to form D<sub>3</sub> reversibly. Lumisterol and tachysterol are formed through continuous exposure to UV-B irradiation, and they are reverted back to pre-vitamin D<sub>3</sub> in dark conditions.</p>
Full article ">Figure 5
<p>Vitamin D<sub>3</sub>’s 25-hydroxylation metabolism. 25-hydroxylation occurs in the liver, where vitamin D<sub>3</sub> is converted to 25-hydroxyvitamin D<sub>3</sub> (25OHD<sub>3</sub>), which is then converted in the kidney via the cytochrome CYP27B1 enzyme in two different derivatives. Via the hydroxylation in the 1α (A ring) position, 25OHD<sub>3</sub> is converted to 1α,25-dihydroxyvitamin D<sub>3</sub> (P and Ca are decreased while PTH and FGF-23 are increased) and via the hydroxylation in the 24th position, 25OHD<sub>3</sub> is converted to 24,25-dihydroxyvitamin D<sub>3</sub> (P and Ca are increased).</p>
Full article ">Figure 6
<p>Vitamin D analogues utilized in kidney-related diseases treatment.</p>
Full article ">Figure 7
<p>The roles of Vitamin D<sub>3</sub> and its active forms produced in the kidney, calcitriol, on immune cells’ modulation towards activation of anti-inflammatory cell responses and suppression of pro-inflammatory signaling.</p>
Full article ">
11 pages, 610 KiB  
Article
Exploring Vitamin D Deficiency and IGF Axis Dynamics in Colorectal Adenomas
by George Ciulei, Olga Hilda Orășan, Angela Cozma, Vasile Negrean, Teodora Gabriela Alexescu, Simina Țărmure, Florin Eugen Casoinic, Roxana Liana Lucaciu, Adriana Corina Hangan and Lucia Maria Procopciuc
Biomedicines 2024, 12(8), 1922; https://doi.org/10.3390/biomedicines12081922 - 22 Aug 2024
Viewed by 464
Abstract
(1) Colorectal cancer is a major cause of cancer-related death, with colorectal adenomas (CRAs) serving as precursors. Identifying risk factors such as vitamin D deficiency and the insulin-like growth factor (IGF) axis is crucial for prevention. (2) This case–control study included 85 participants [...] Read more.
(1) Colorectal cancer is a major cause of cancer-related death, with colorectal adenomas (CRAs) serving as precursors. Identifying risk factors such as vitamin D deficiency and the insulin-like growth factor (IGF) axis is crucial for prevention. (2) This case–control study included 85 participants (53 CRA patients and 32 controls) who underwent colonoscopy. We measured serum vitamin D3 (cholecalciferol), calcidiol (vitamin D metabolite), calcitriol (active vitamin D metabolite), insulin-like growth factor-1 (IGF-1), and insulin-like growth factor binding protein-3 (IGFBP-3) to explore their associations with CRA risk. (3) Results: We found that lower cholecalciferol levels were a significant risk factor for CRA (OR = 4.63, p = 0.004). Although no significant differences in calcidiol and calcitriol levels were observed between CRA patients and controls, calcidiol deficiency was common in the study population. IGF-1 levels inversely correlated with age, calcitriol, and IGFBP-3 in CRA patients. (4) This study highlights the potential of lower cholecalciferol levels to detect patients at risk of CRA when calcidiol values cannot, suggesting the importance of evaluating different vitamin D metabolites in cancer prevention research. Our findings underscore the need to further investigate the interactions between calcitriol, the active form of vitamin D, and the IGF axis in colorectal cancer development. Full article
Show Figures

Figure 1

Figure 1
<p>Correlation between IGF-1 and IGFBP-3 serum levels in the CRA<sup>3</sup> group, Spearman’s r coefficient, <span class="html-italic">p</span> value, and a simple linear regression line. <sup>1</sup> IGFBP-3—insulin-like growth factor binding protein-3; <sup>2</sup> IGF-1—insulin-like growth factor-1; <sup>3</sup> CRA—colorectal adenoma.</p>
Full article ">Figure 2
<p>Correlation between IGF-1 and IGFBP-3 serum levels in the control group, Spearman’s r coefficient, <span class="html-italic">p</span> value, and a simple linear regression line. <sup>1</sup> IGFBP-3—insulin-like growth factor binding protein-3; <sup>2</sup> IGF-1—insulin-like growth factor-1.</p>
Full article ">
20 pages, 8205 KiB  
Article
Dietary Supplementation of Crossbred Pigs with Glycerol, Vitamin C, and Niacinamide Alters the Composition of Gut Flora and Gut Flora-Derived Metabolites
by Panting Wei, Wenchen Sun, Shaobin Hao, Linglan Deng, Wanjie Zou, Huadong Wu, Wei Lu and Yuyong He
Animals 2024, 14(15), 2198; https://doi.org/10.3390/ani14152198 - 28 Jul 2024
Viewed by 1022
Abstract
The addition of glycerin, vitamin C, and niacinamide to pig diets increased the redness of longissimus dorsi; however, it remains unclear how these supplements affect gut microbiota and metabolites. A total of 84 piglets (20.35 ± 2.14 kg) were randomly allotted to [...] Read more.
The addition of glycerin, vitamin C, and niacinamide to pig diets increased the redness of longissimus dorsi; however, it remains unclear how these supplements affect gut microbiota and metabolites. A total of 84 piglets (20.35 ± 2.14 kg) were randomly allotted to groups A (control), B (glycerin-supplemented), C (vitamin C and niacinamide-supplemented), and D (glycerin, vitamin C and niacinamide-supplemented) during a feeding experiment. Metagenomic and metabolomic technologies were used to analyze the fecal compositions of bile acids, metabolites, and microbiota. The results showed that compared to pigs in group A, pigs in group D had lower virulence factor expressions of lipopolysaccharide (p < 0.05), fatty acid resistance system (p < 0.05), and capsule (p < 0.01); higher fecal levels of ferric ion (p < 0.05), allolithocholic acid (p < 0.01), deoxycholic acid (p < 0.05), tauroursodeoxycholic acid dihydrate (p < 0.01), glycodeoxycholic acid (p < 0.05), L-proline (p < 0.01) and calcitriol (p < 0.01); and higher (p < 0.05) abundances of iron-acquiring microbiota (Methanobrevibacter, Clostridium, Clostridiaceae, Clostridium_sp_CAG_1000, Faecalibacterium_sp_CAG_74_58_120, Eubacteriales_Family_XIII_Incertae_Sedis, Alistipes_sp_CAG_435, Alistipes_sp_CAG_514 and Methanobrevibacter_sp_YE315). Supplementation with glycerin, vitamin C, and niacinamide to pigs significantly promoted the growth of iron-acquiring microbiota in feces, reduced the expression of some virulence factor genes of fecal pathogens, and increased the fecal levels of ferric ion, L-proline, and some secondary bile acids. The administration of glycerol, vitamin C, and niacinamide to pigs may serve as an effective measure for muscle redness improvement by altering the compositions of fecal microbiota and metabolites. Full article
Show Figures

Figure 1

Figure 1
<p>The box plot presented the difference of gene number between groups. <span class="html-italic">x</span>-axis: groups; <span class="html-italic">y</span>-axis: non-redundant gene number; the upper and lower end of the box: The upper and lower interquartile range (IQR); the median line: median; upper and lower edges: maximum and minimum inner bounding value (1.5-times IQR); points outside the upper and lower edges: outliers. The number on the line between the columns is the <span class="html-italic">p</span>-value.</p>
Full article ">Figure 2
<p>The comparisons of differential KO pathways. KO pathways with <span class="html-italic">p</span>-value smaller than 0.05 in Kruskal–Wallis rank-sum test were shown in the bar plots. (<b>A</b>): Bar plots show the relative abundance of the top 15 remarkably different KO pathways (<span class="html-italic">p</span> &lt; 0.05). The abscissa shows the abundance of KO pathways, and the ordinate is the names of differential KO pathways. (<b>B</b>): Comparison of significant differences of differential KO pathways among groups A–D. The abscissa shows the names of treatment groups, and the ordinate is the relative abundance of KO pathways. Marks on the line between the columns indicate a statistical significance: * indicates <span class="html-italic">p</span> &lt; 0.05; ** indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>The comparisons of differential CAZymes family genes. Functional genes with p-value smaller than 0.05 in Kruskal–Wallis rank-sum test were shown in the bar plots. (<b>A</b>): Bar plots show the relative abundance of the top 15 remarkably different CAZy family genes (<span class="html-italic">p</span> &lt; 0.05). The abscissa shows gene abundance, and the ordinate is the names of differential CAZymes genes. (<b>B</b>): Comparison of significant differences of differential CAZymes family genes among groups A–D. The abscissa shows the names of treatment groups, and the ordinate is the relative abundance of CAZymes family gene. Marks on the line between the columns indicate a statistical significance: * indicates <span class="html-italic">p</span> &lt; 0.05; ** indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>The comparisons of virulence factors. Virulence factors with <span class="html-italic">p</span>-value smaller than 0.05 in Kruskal–Wallis rank-sum test were shown in the bar plots. (<b>A</b>): Bar plots show the relative abundance of the top 15 remarkably different virulence factors (<span class="html-italic">p</span> &lt; 0.05). The abscissa shows the abundance of virulence factors, and the ordinate is the names of differential virulence factors. (<b>B</b>): Comparison of significant differences of differential virulence factors among groups A–D. The abscissa shows the names of treatment groups, and the ordinate is the relative abundance of virulence factors. Marks on the line between the columns indicate a statistical significance: * indicates <span class="html-italic">p</span> &lt; 0.05; ** indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>Microbial community composition displaying the top abundant taxa across all samples. The taxonomic composition of the fecal microbiome is shown at various taxonomic levels: Kingdom, Phylum, Class, Order, Family, Genus, and Species. Each bar represents the relative abundance of different microbial taxa in fecal samples from four treatment groups: A (control), B (glycerin-supplemented), C (vitamin C and niacinamide-supplemented), and D (glycerin, vitamin C, and niacinamide-supplemented). The color-coded bar plot shows the average bacterial distribution in each group.</p>
Full article ">Figure 6
<p>PCoA plot based on Bray–Curtis distance matrix of fecal microbial community structures of the four treatment groups. Treatment groups are differentiated with different colored circles: A (control), B (glycerin-supplemented), C (vitamin C and niacinamide-supplemented), and D (glycerin, vitamin C, and niacinamide-supplemented). The two axes of the plot explained 35.3% of the variance, with X and Y axes explaining 20.88% and 14.42% of the total variation, respectively.</p>
Full article ">Figure 7
<p>The differential microbiota between groups based on Linear Discriminant Analysis (LDA) Effect Size (LEfSe). This figure compares the relative abundance of specific microbial taxa among the treatment groups: (A) vs. (B), (A) vs. (C), (A) vs. (D), (B) vs. (C), (B) vs. (D), and (C) vs. (D). Each plot shows taxa with significant differences in abundance between groups, with the LDA score (log 10) indicating the effect size. Orange bars represent higher abundance in the first group, while blue bars represent higher abundance in the second group. Significant taxa include various bacterial species at different taxonomic levels, with notable differences highlighted for each comparison.</p>
Full article ">Figure 8
<p>The OPLS-DA score plot of primary metabolites obtained from fecal samples analyzed by untargeted liquid chromatography/mass spectrometry. It compares the OPLS-DA score of A vs. B, A vs. C, A vs. D, B vs. C, B vs. D, and C vs. D. The <span class="html-italic">x</span>-axis (t1) represents the predicted component (inter group difference component), and the <span class="html-italic">y</span>-axis (to1) represents the orthogonal component (intra group difference component). R2Y represents the percentage of all sample variables explained by the model, Q2Y represents the percentage of all sample variables predicted by the model.</p>
Full article ">Figure 9
<p>The Volcano plots of differential metabolites between groups. Each dot in the volcano plots represents a metabolite, and the blue ones represent the significantly down-regulated metabolites, the red dots represent the significantly up-regulated metabolites, the gray dots represent the metabolites that have no significant difference between the two groups. The ordinate shows the −log10 (<span class="html-italic">p</span>-value), and the abscissa is the log2 (fold change) value. The ions with VIP &gt;  1.0 and <span class="html-italic">p</span>  &lt;  0.05 were considered to be important differential metabolites.</p>
Full article ">Figure 10
<p>The top 20 terms KO pathway enrichment bar plot of differential metabolites. The abscissa represents enrichment radio, and the ordinate represents enrichment pathways. The different colored entries represent the hierarchical classification annotations of the KEGG pathway, corresponding to KO pathway level 3 and KEGG pathway names. The length of the column represents the number of differential metabolites annotated by the pathway.</p>
Full article ">
30 pages, 1407 KiB  
Review
Enhancement of Orthodontic Tooth Movement by Local Administration of Biofunctional Molecules: A Comprehensive Systematic Review
by Cristina Dora Ciobotaru, Dana Feștilă, Elena Dinte, Alexandrina Muntean, Bianca Adina Boșca, Anca Ionel and Aranka Ilea
Pharmaceutics 2024, 16(8), 984; https://doi.org/10.3390/pharmaceutics16080984 - 25 Jul 2024
Cited by 1 | Viewed by 634
Abstract
Enhancement of orthodontic tooth movement (OTM) through local administration of biofunctional molecules has become increasingly significant, particularly for adult patients seeking esthetic and functional improvements. This comprehensive systematic review analyzes the efficacy of various biofunctional molecules in modulating OTM, focusing on the method [...] Read more.
Enhancement of orthodontic tooth movement (OTM) through local administration of biofunctional molecules has become increasingly significant, particularly for adult patients seeking esthetic and functional improvements. This comprehensive systematic review analyzes the efficacy of various biofunctional molecules in modulating OTM, focusing on the method of administration and its feasibility, especially considering the potential for topical application. A search across multiple databases yielded 36 original articles of experimental human and animal OTM models, which examined biofunctional molecules capable of interfering with the biochemical reactions that cause tooth movement during orthodontic therapy, accelerating the OTM rate through their influence on bone metabolism (Calcitriol, Prostaglandins, Recombinant human Relaxin, RANKL and RANKL expression plasmid, growth factors, PTH, osteocalcin, vitamin C and E, biocompatible reduced graphene oxide, exogenous thyroxine, sclerostin protein, a specific EP4 agonist (ONO-AE1-329), carrageenan, and herbal extracts). The results indicated a variable efficacy in accelerating OTM, with Calcitriol, Prostaglandins (PGE1 and PGE2), RANKL, growth factors, and PTH, among others, showing promising outcomes. PGE1, PGE2, and Calcitriol experiments had statistically significant outcomes in both human and animal studies and, while other molecules underwent only animal testing, they could be validated in the future for human use. Notably, only one of the animal studies explored topical administration, which also suggests a future research direction. This review concluded that while certain biofunctional molecules demonstrated potential for OTM enhancement, the evidence is not definitive. The development of suitable topical formulations for human use could offer a patient-friendly alternative to injections, emphasizing comfort and cost-effectiveness. Future research should focus on overcoming current methodological limitations and advancing translational research to confirm these biomolecules’ efficacy and safety in clinical orthodontic practice. Full article
(This article belongs to the Special Issue Biomedical Applications: Advances in Bioengineering and Drug Delivery)
Show Figures

Figure 1

Figure 1
<p>PRISMA 2020 flowchart.</p>
Full article ">Figure 2
<p>Risk of Bias assessment [<a href="#B17-pharmaceutics-16-00984" class="html-bibr">17</a>,<a href="#B18-pharmaceutics-16-00984" class="html-bibr">18</a>,<a href="#B19-pharmaceutics-16-00984" class="html-bibr">19</a>,<a href="#B20-pharmaceutics-16-00984" class="html-bibr">20</a>,<a href="#B21-pharmaceutics-16-00984" class="html-bibr">21</a>,<a href="#B22-pharmaceutics-16-00984" class="html-bibr">22</a>,<a href="#B23-pharmaceutics-16-00984" class="html-bibr">23</a>,<a href="#B24-pharmaceutics-16-00984" class="html-bibr">24</a>].</p>
Full article ">Figure 3
<p>Risk of Bias assessment [<a href="#B28-pharmaceutics-16-00984" class="html-bibr">28</a>,<a href="#B32-pharmaceutics-16-00984" class="html-bibr">32</a>,<a href="#B33-pharmaceutics-16-00984" class="html-bibr">33</a>,<a href="#B34-pharmaceutics-16-00984" class="html-bibr">34</a>,<a href="#B35-pharmaceutics-16-00984" class="html-bibr">35</a>,<a href="#B36-pharmaceutics-16-00984" class="html-bibr">36</a>,<a href="#B37-pharmaceutics-16-00984" class="html-bibr">37</a>,<a href="#B38-pharmaceutics-16-00984" class="html-bibr">38</a>,<a href="#B39-pharmaceutics-16-00984" class="html-bibr">39</a>,<a href="#B40-pharmaceutics-16-00984" class="html-bibr">40</a>,<a href="#B41-pharmaceutics-16-00984" class="html-bibr">41</a>,<a href="#B42-pharmaceutics-16-00984" class="html-bibr">42</a>,<a href="#B43-pharmaceutics-16-00984" class="html-bibr">43</a>,<a href="#B44-pharmaceutics-16-00984" class="html-bibr">44</a>,<a href="#B45-pharmaceutics-16-00984" class="html-bibr">45</a>,<a href="#B46-pharmaceutics-16-00984" class="html-bibr">46</a>,<a href="#B47-pharmaceutics-16-00984" class="html-bibr">47</a>,<a href="#B48-pharmaceutics-16-00984" class="html-bibr">48</a>,<a href="#B49-pharmaceutics-16-00984" class="html-bibr">49</a>,<a href="#B50-pharmaceutics-16-00984" class="html-bibr">50</a>,<a href="#B51-pharmaceutics-16-00984" class="html-bibr">51</a>,<a href="#B52-pharmaceutics-16-00984" class="html-bibr">52</a>,<a href="#B53-pharmaceutics-16-00984" class="html-bibr">53</a>,<a href="#B54-pharmaceutics-16-00984" class="html-bibr">54</a>,<a href="#B55-pharmaceutics-16-00984" class="html-bibr">55</a>,<a href="#B56-pharmaceutics-16-00984" class="html-bibr">56</a>,<a href="#B57-pharmaceutics-16-00984" class="html-bibr">57</a>,<a href="#B58-pharmaceutics-16-00984" class="html-bibr">58</a>].</p>
Full article ">
16 pages, 3219 KiB  
Article
Vitamin D Receptor Regulates the Expression of the Grainyhead-Like 1 Gene
by Agnieszka Taracha-Wisniewska, Emma G. C. Parks, Michal Miller, Lidia Lipinska-Zubrycka, Sebastian Dworkin and Tomasz Wilanowski
Int. J. Mol. Sci. 2024, 25(14), 7913; https://doi.org/10.3390/ijms25147913 - 19 Jul 2024
Viewed by 577
Abstract
Vitamin D plays an important pleiotropic role in maintaining global homeostasis of the human body. Its functions go far beyond skeletal health, playing a crucial role in a plethora of cellular functions, as well as in extraskeletal health, ensuring the proper functioning of [...] Read more.
Vitamin D plays an important pleiotropic role in maintaining global homeostasis of the human body. Its functions go far beyond skeletal health, playing a crucial role in a plethora of cellular functions, as well as in extraskeletal health, ensuring the proper functioning of multiple human organs, including the skin. Genes from the Grainyhead-like (GRHL) family code for transcription factors necessary for the development and maintenance of various epithelia. Even though they are involved in many processes regulated by vitamin D, a direct link between vitamin D-mediated cellular pathways and GRHL genes has never been described. We employed various bioinformatic methods, quantitative real-time PCR, chromatin immunoprecipitation, reporter gene assays, and calcitriol treatments to investigate this issue. We report that the vitamin D receptor (VDR) binds to a regulatory region of the Grainyhead-like 1 (GRHL1) gene and regulates its expression. Ectopic expression of VDR and treatment with calcitriol alters the expression of the GRHL1 gene. The evidence presented here indicates a role of VDR in the regulation of expression of GRHL1 and correspondingly a role of GRHL1 in mediating the actions of vitamin D. Full article
Show Figures

Figure 1

Figure 1
<p>Bioinformatic predictions of transcription factor binding site (TFBS) motifs for VDR in the human <span class="html-italic">GRHL1</span> gene. (<b>A</b>) Genomic location of regulatory regions of the <span class="html-italic">GRHL1</span> gene according to the Ensembl database. (<b>B</b>) VDR binding sites identified in the <span class="html-italic">GRHL1</span> gene using GTRD modules: meta-clusters, motifs, and GEM–PICS–MACS2–SISSRs clusters. (<b>C</b>) Genomic location of CpG Islands, markers of open chromatin (H3K4Me1, H3K4Me3, H3K27Ac), and transcription factor binding sites (ChIP-seq). (<b>D</b>) Multiple sequence alignment of a TFBS example across different species using T-Coffee (Pro-Coffee mode). Asterisks indicate conserved nucleotides.</p>
Full article ">Figure 2
<p>Baseline expression of <span class="html-italic">VDR</span> and <span class="html-italic">GRHL1–3</span> genes in various tissues. Based on <a href="https://www.proteinatlas.org/" target="_blank">https://www.proteinatlas.org/</a>, last accessed 11 June 2024. TPM, transcripts per million. Database version: Genotype-Tissue Expression (GTEx) RNA-seq data v8, available from the above portal.</p>
Full article ">Figure 3
<p>Overexpression of VDR alters mRNA level of <span class="html-italic">GRHL1</span> gene. (<b>A</b>) Genomic coordinates of the binding site for VDR in the promoter region of the <span class="html-italic">GRHL1</span> gene, obtained from the MotEvo database. (<b>B</b>) The mRNA expression levels of the <span class="html-italic">GRHL1–3</span> genes in HaCaT cells (1) transiently overexpressing VDR treated with ethanol, (2) transiently overexpressing VDR treated with 100 nM calcitriol, or (3) transfected with an empty vector treated with 100 nM calcitriol. The results represent relative expression of the respective target gene vs. <span class="html-italic">HPRT</span> genes. Data are shown as means ± SEM of experiments independently performed in triplicate, * significantly different at <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 4
<p>(<b>A</b>) Quantitative ChIP-PCR analysis of VDR occupancy of the <span class="html-italic">GRHL1</span> regulatory region was performed in HaCaT cells transfected with pcDNA3.1-K-DYK-VDR. Chromatin was immunoprecipitated with anti-DYK (FLAG) antibody or nonspecific antibody. The amount of DNA amplified from immunoprecipitated DNA was normalized to that amplified from input DNA. Data are shown as means ± SEM from experiments independently performed in triplicate, * significantly different at <span class="html-italic">p</span>≤ 0.05. (<b>B1</b>,<b>B2</b>) HaCaT cells were transfected with (<b>B1</b>) pcDNA3.1-K-DYK-VDR or pcDNA3.1-empty plasmid, 500 ng of the firefly luciferase vector with VDR binding site derived from the regulatory region of the <span class="html-italic">GRHL1</span> gene, and 25 ng pRL-CMV Renilla luciferase control reporter vector and treated with 100 nM calcitriol or (B2) pcDNA3.1-K-DYK-VDR, 500 ng of the firefly luciferase vector with VDR binding site derived from the regulatory region of the <span class="html-italic">GRHL1</span> gene, and 25 ng pRL-CMV Renilla luciferase control reporter vector and treated with 100 nM calcitriol or ethanol vehicle. Data are shown as means ± SEM of experiments independently performed in triplicate, * significantly different at <span class="html-italic">p</span>≤ 0.05.</p>
Full article ">
21 pages, 17481 KiB  
Review
Vitamin D in Central Nervous System: Implications for Neurological Disorders
by Bayan Sailike, Zhadyra Onzhanova, Burkitkan Akbay, Tursonjan Tokay and Ferdinand Molnár
Int. J. Mol. Sci. 2024, 25(14), 7809; https://doi.org/10.3390/ijms25147809 - 17 Jul 2024
Cited by 1 | Viewed by 984
Abstract
Vitamin D, obtained from diet or synthesized internally as cholecalciferol and ergocalciferol, influences bodily functions through its most active metabolite and the vitamin D receptor. Recent research has uncovered multiple roles for vitamin D in the central nervous system, impacting neural development and [...] Read more.
Vitamin D, obtained from diet or synthesized internally as cholecalciferol and ergocalciferol, influences bodily functions through its most active metabolite and the vitamin D receptor. Recent research has uncovered multiple roles for vitamin D in the central nervous system, impacting neural development and maturation, regulating the dopaminergic system, and controlling the synthesis of neural growth factors. This review thoroughly examines these connections and investigates the consequences of vitamin D deficiency in neurological disorders, particularly neurodegenerative diseases. The potential benefits of vitamin D supplementation in alleviating symptoms of these diseases are evaluated alongside a discussion of the controversial findings from previous intervention studies. The importance of interpreting these results cautiously is emphasised. Furthermore, the article proposes that additional randomised and well-designed trials are essential for gaining a deeper understanding of the potential therapeutic advantages of vitamin D supplementation for neurological disorders. Ultimately, this review highlights the critical role of vitamin D in neurological well-being and highlights the need for further research to enhance our understanding of its function in the brain. Full article
Show Figures

Figure 1

Figure 1
<p>Sources of vitamin D<sub>3</sub> and pathways for calcitriol biosynthesis in the body. (<b>A</b>) The classical pathway involves the synthesis of VD<sub>3</sub> in the skin under the UVB radiation from the sun. This pathway starts with 7-dehydrocholesterol (7-HDC), a derivative of cholesterol, reacting to UV radiation to form cholecalciferol (VD<sub>3</sub>) through an intermediate previtamin D<sub>3</sub> (Pre-VD<sub>3</sub>). (<b>B</b>) The alternative pathway involves dietary intake of VD<sub>3</sub> or ergocalciferol (VD<sub>2</sub>)-rich food such as fatty fish, eggs, or fortified milk products for VD<sub>3</sub> or plants and fungi for VD<sub>2</sub>. Under normal circumstances, higher quantities are synthesised via classical pathways. VD<sub>3</sub> from both sources enters the circulation, binds to vitamin D binding protein (VDBP), and is transported to metabolic tissues. (<b>C</b>) The traditional synthetic pathway of 1<math display="inline"><semantics> <mi>α</mi> </semantics></math>,25-dihydroxyvitamin D<sub>3</sub> (calcitriol/1,25-VD<sub>3</sub>), the most active metabolite of vitamin D, starts in the liver when VD<sub>3</sub> is hydroxylated at C25 by CYP2R1 (and in minority also CYP27A1) to yield 25-dihydroxyvitamin D<sub>3</sub> (calcidiol/25-VD<sub>3</sub>), which is the major circulating storage form of vitamin D. (<b>D</b>) This 25-VD<sub>3</sub> is transported by VDBP to the kidney, where it is hydroxylated at position C1 by CYP27B1 to form 1<math display="inline"><semantics> <mi>α</mi> </semantics></math>,25-VD<sub>3</sub>. The 1<math display="inline"><semantics> <mi>α</mi> </semantics></math>,25-VD<sub>3</sub> binds to its molecular target vitamin D receptor (VDR), which regulates the transcription of various target genes. (<b>E</b>) Vitamin D can also be transported directly to the brain as 25-VD<sub>3</sub> or 1<math display="inline"><semantics> <mi>α</mi> </semantics></math>,25-VD<sub>3</sub>, where both can cross the blood–brain barrier. Additionally, 25-VD<sub>3</sub> can be converted to 1<math display="inline"><semantics> <mi>α</mi> </semantics></math>,25-VD<sub>3</sub> in the brain, since the enzymes involved in its synthesis are expressed in pericytes, glial cells, and neurons in addition to the liver and kidney. This suggest the possible of local synthesis of vitamin D metabolites in the brain and active signalling.</p>
Full article ">Figure 2
<p>Schematic model illustrating the genomic and non-genomic effects of VD<sub>3</sub> in the CNS. (<b>A</b>) Genomic action of calcitriol: This model focuses on VD<sub>3</sub> target genes and signalling pathways that have been identified within the brain or neural cells (neurons, astrocytes, oligodendrocytes, or microglia). Genomic actions primarily occur within the nucleus. In this simplified model, calcitriol binds to the VDR/RXR complex, leading to the release of co-repressors and the recruitment of co-activators at vitamin D response elements (VDREs) located in regulatory regions, which promotes the expression of specific calcitriol target genes. The listed genes are those whose expression is influenced by VD<sub>3</sub> within the brain and for which functional VDREs have been identified on relevant regulatory regions. (<b>B</b>) Non-genomic action of VD<sub>3</sub>: The non-genomic actions of <sub>3</sub> are potentially mediated through the classical VDR, protein disulphide isomerase A3 (PDIA3), or both of these proteins. Upon calcitriol binding, the rapid activation of protein kinases such as CaMII, PKA, and PI3K occurs, which in turn facilitates the influx of Ca<sup>2+</sup> ions via L-type voltage-gated calcium channels (L-VGCCs). Intracellular Ca<sup>2+</sup> then triggers the activation of p38MAPK, further modulating downstream signalling pathways. Both the genomic and non-genomic actions of VD<sub>3</sub> are likely to have an impact on brain development, function, and maintenance. These mechanisms play a role in shaping the intricate processes within the CNS.</p>
Full article ">Figure 3
<p>The connection between VD<sub>3</sub> and various neurological disorders is summarised in the figure. It highlights diseases in which molecular connections have been suggested between VD<sub>3</sub> signalling and gene regulatory or metabolic networks, such as the regulation of enzymes that facilitate the production of serotonin or melatonin. Additionally, a general neuroinflammation process has been proposed as a link to many neurological diseases, where VD<sub>3</sub> supplementation inhibits the expression of inflammatory cytokines and/or activates the expression of anti-inflammatory molecules. Throughout the figure, the following abbreviations have been used: glutathione (GSH), <span class="html-italic">interleukin-1β</span> (<span class="html-italic">IL1B</span>), <span class="html-italic">tumour necrosis factor</span> (<span class="html-italic">TNF</span>), <span class="html-italic">L-type voltage-sensitive Ca<sup>2+</sup></span> channel, <span class="html-italic">glial cell-derived neurotrophic factor</span> (<span class="html-italic">GDNF</span>), <span class="html-italic">neurotrophin 3</span> (<span class="html-italic">NTF3</span>), <span class="html-italic">nerve growth factor</span> (<span class="html-italic">NGF</span>), <span class="html-italic">tyrosine hydroxylase</span> (<span class="html-italic">TH</span>), <span class="html-italic">tryptophan hydroxylase</span> (<span class="html-italic">TPH1</span> and <span class="html-italic">TPH2</span>), <span class="html-italic">N-methyl-D-aspartate receptor</span> (<span class="html-italic">NMDAR</span>), <span class="html-italic">inducible nitric oxide synthase</span> (<span class="html-italic">iNOS</span>), <span class="html-italic">prostaglandin-endoperoxide synthase 2 or cyclooxygenase-2</span> (<span class="html-italic">COX2/PTGS2</span>), <span class="html-italic">insulin-degrading enzyme</span> (<span class="html-italic">IDE</span>), <span class="html-italic">low-density lipoprotein receptor-related protein 1</span> (<span class="html-italic">LRP1</span>), <span class="html-italic">vitamin D receptor</span> (<span class="html-italic">VDR</span>).</p>
Full article ">Figure 4
<p>Overview of the effects of VD<sub>3</sub> on individual neurons and the whole brain.</p>
Full article ">
15 pages, 1592 KiB  
Review
The “Sunshine Vitamin” and Its Antioxidant Benefits for Enhancing Muscle Function
by Cristina Russo, Rosa Santangelo, Lucia Malaguarnera and Maria Stella Valle
Nutrients 2024, 16(14), 2195; https://doi.org/10.3390/nu16142195 - 10 Jul 2024
Viewed by 1161
Abstract
Pathological states marked by oxidative stress and systemic inflammation frequently compromise the functional capacity of muscular cells. This progressive decline in muscle mass and tone can significantly hamper the patient’s motor abilities, impeding even the most basic physical tasks. Muscle dysfunction can lead [...] Read more.
Pathological states marked by oxidative stress and systemic inflammation frequently compromise the functional capacity of muscular cells. This progressive decline in muscle mass and tone can significantly hamper the patient’s motor abilities, impeding even the most basic physical tasks. Muscle dysfunction can lead to metabolic disorders and severe muscle wasting, which, in turn, can potentially progress to sarcopenia. The functionality of skeletal muscle is profoundly influenced by factors such as environmental, nutritional, physical, and genetic components. A well-balanced diet, rich in proteins and vitamins, alongside an active lifestyle, plays a crucial role in fortifying tissues and mitigating general weakness and pathological conditions. Vitamin D, exerting antioxidant effects, is essential for skeletal muscle. Epidemiological evidence underscores a global prevalence of vitamin D deficiency, which induces oxidative harm, mitochondrial dysfunction, reduced adenosine triphosphate production, and impaired muscle function. This review explores the intricate molecular mechanisms through which vitamin D modulates oxidative stress and its consequent effects on muscle function. The aim is to evaluate if vitamin D supplementation in conditions involving oxidative stress and inflammation could prevent decline and promote or maintain muscle function effectively. Full article
(This article belongs to the Special Issue Vitamins and Human Health: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Vitamin D’s role in skeletal muscle. Vitamin D can be acquired through skin synthesis or from diet. Both vitamin D3 and D2 undergo the same metabolic processes to produce their active forms. While the primary role of vitamin D is to regulate calcium levels and ensure skeletal and muscle health, it also serves as a powerful immunoregulator, influencing the inflammatory response, muscle damage, and aerobic capacity. Circulating 25(OH)D binds to the carrier protein DBP. PTH promotes renal Ca<sup>2+</sup> retention and activates the synthesis of active vitamin D, which, in conjunction with the vitamin D receptor (VDR), facilitates Ca<sup>2+</sup> and phosphate absorption. Vitamin D deficiency (VDD) or inadequate sun exposure can elevate PTH levels, leading to skeletal fragility. In skeletal muscle, the 25(OH)D-DBP complex is transported into target cells through the LRP2/CUBN transmembrane complex. Inside the cell, the D-DBP complex associates with cytoplasmic actin. 1,25(OH)2D triggers the expression of protein 1, affecting MyoD1 activation. Vitamin D also regulates the FOXO3 signaling pathways, enhancing myoblast self-renewal. VDR expression in skeletal muscle promotes muscle protein synthesis, is crucial for maintaining muscle mass, and aids in muscle regeneration. Abbreviations in alphabetical order: 1,25(OH)2D3—calcitriol; FOXO—forkhead family of transcription factors; LRP2/CUBN—megalin–cubilin transmembrane complex; MyoD1—myogenic determination factor 1; PTH—parathyroid hormone; DBP—vitamin D binding protein; vitamin D receptor (VDR).</p>
Full article ">Figure 2
<p>Relationship between oxidative stress and weakness of muscle. ROS generated from localized inflammation activate immune cells, initiating a harmful cycle characterized by the release of pro-inflammatory mediators, such as TNFα, IL-6, and CRP. Despite the presence of an impaired antioxidant system, including decreased levels of GSH and SOD, oxidative damage remains unchecked. Furthermore, inflammation disrupts mitochondrial function in muscle through the NO signaling pathway, triggering cell death via OMM permeabilization. Oxidative stress accelerates cellular senescence by activating FOXO and diminishing SIRT-1, leading to heightened MMP-9 and NF-κB activity. This escalated oxidative stress precipitates myocyte dysfunction and apoptosis, resulting in contractile dysfunction, fibrosis, hypertrophy, and impaired muscle remodeling. Additionally, there is a transition towards a type-IIx-oriented muscle phenotype with compromised oxygen distribution and utilization, ultimately impairing functionality. Abbreviations in alphabetical order: CRP—C reactive protein; FOXO—forkhead family of transcription factors; GSH—glutathione peroxidase; IL-6—Interleukin 6; MMP-9—matrix metallopeptidase 9; NF-κB—nuclear factor kappa-light-chain-enhancer of activated B cells; NO—nitrogen monoxide; OMM—outer mitochondrial membrane; RNS—reactive nitrogen species; ROS—reactive oxygen species; SIRT-1—sirtuin-1; SOD—superoxide dismutase; TNFα—tumor receptor necrosis factor alpha.</p>
Full article ">Figure 3
<p>Antioxidative role of vitamin D in muscle dysfunction. Blue arrow indicates activation; red arrow indicates inhibition. Vitamin D activates the VDR in satellite cells, enhancing their self-renewal, proliferation, and differentiation capabilities. Activation of the VDR also mitigates oxidative stress, promoting mitochondrial biogenesis and fusion while reducing oxidative damage and dysfunction. This process improves mitochondrial network structure through the regulation of MFN1/2, OPA1, and Drp1 expression. Vitamin D positively influences Sirt1 activity and enhances mitochondrial function. Supplementation with vitamin D activates Sirt1 and AMPK in skeletal muscle cells. Vitamin D further increases the expression of irisin precursors in muscle cells, contingent upon intact Sirt1 expression. Both AMPK and Sirt1 regulate PGC-1α activation and transcription, influencing irisin secretion in skeletal muscle cells. Vitamin D upregulates FOXO1 protein and suppresses atrogin-1 and MuRF1 expression. Additionally, VDS triggers VDR and induces the Nrf2-Keap1 antioxidant pathway. Abbreviations in alphabetical order: AMP—5′ AMP-activated protein kinase; 1,25(OH)2D3—calcitriol; Drp1—dynamin-related protein 1; IL-1β—interleukin-1 beta; MFN1/2—mitofusin; OPA—mitochondrial dynamin like GTPase; PGC-1α—peroxisome proliferator-activated receptor-gamma coactivator; PTH—parathyroid hormone; ROS—reactive oxygen species; SIRT-1—sirtuin-1; SOD—superoxide dis-mutase; TNFα—tumor receptor necrosis factor alpha; VDR—vitamin D receptor.</p>
Full article ">
28 pages, 41710 KiB  
Article
Putative Pharmacological Depression and Anxiety-Related Targets of Calcitriol Explored by Network Pharmacology and Molecular Docking
by Bruna R. Kouba, Glorister A. Altê and Ana Lúcia S. Rodrigues
Pharmaceuticals 2024, 17(7), 893; https://doi.org/10.3390/ph17070893 - 5 Jul 2024
Viewed by 1016
Abstract
Depression and anxiety disorders, prevalent neuropsychiatric conditions that frequently coexist, limit psychosocial functioning and, consequently, the individual’s quality of life. Since the pharmacological treatment of these disorders has several limitations, the search for effective and secure antidepressant and anxiolytic compounds is welcome. Vitamin [...] Read more.
Depression and anxiety disorders, prevalent neuropsychiatric conditions that frequently coexist, limit psychosocial functioning and, consequently, the individual’s quality of life. Since the pharmacological treatment of these disorders has several limitations, the search for effective and secure antidepressant and anxiolytic compounds is welcome. Vitamin D has been shown to exhibit neuroprotective, antidepressant, and anxiolytic properties. Therefore, this study aimed to explore new molecular targets of calcitriol, the active form of vitamin D, through integrated bioinformatic analysis. Calcitriol targets were predicted in SwissTargetPrediction server (2019 version). The disease targets were collected by the GeneCards database searching the keywords “depression” and “anxiety”. Gene ontology (GO) and the Kyoto Encyclopedia of Genes and Genomes (KEGG) were used to analyze the intersections of targets. Network analyses were carried out using GeneMania server (2023 version) and Cytoscape (V. 3.9.1.) software. Molecular docking predicted the main targets of the network and Ligplot predicted the main intermolecular interactions. Our study showed that calcitriol may interact with multiple targets. The main targets found are the vitamin D receptor (VDR), histamine H3 receptor (H3R), endocannabinoid receptors 1 and 2 (CB1 and CB2), nuclear receptor NR1H3, patched-1 (PTCH1) protein, opioid receptor NOP, and phosphodiesterase enzymes PDE3A and PDE5A. Considering the role of these targets in the pathophysiology of depression and anxiety, our findings suggest novel putative mechanisms of action of vitamin D as well as new promising molecular targets whose role in these disorders deserves further investigation. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic workflow showing the approach of this study. Genes modulated by calcitriol and genes related to depression and anxiety were searched in databases (Swiss Target Prediction and GeneCards). Subsequently, genes overlapping calcitriol, depression, and anxiety; genes overlapping calcitriol and depression; and genes overlapping calcitriol and anxiety were analyzed by Venn diagram, and then a pharmacological network using Genemania and Cytoscape software was built. Subsequently, gene ontology and KEGG analyses were carried out. Finally, molecular docking and interaction analyses of calcitriol with proteins were carried out using the Chimera, Dockthor, and LigPlot programs. To further validate the calcitriol targets, a STRING-DisGeNET network was constructed.</p>
Full article ">Figure 2
<p>Venn diagram illustrating the potential target genes related to depression, anxiety, and calcitriol.</p>
Full article ">Figure 3
<p>Network interactions among calcitriol, anxiety, and depression targets. (<b>A</b>) Complete network composed of 94 nodes. (<b>B</b>) Filtered network containing 59 most central and connected target genes shared among calcitriol, anxiety, and depression. The network was based on the degree center (DC) ≥ 21, betweenness centrality (BC) range of 0.004–0.046, and closeness centrality (CC) range of 0.477–0.780. The largest node indicates a higher degree in the network. The color of the edges refers to the types of interactions: physical interactions (green), co-expression (yellow), shared protein domains (pink), co-locations (blue), predictions (gray), genetic interactions (purple), and metabolic pathways (black).</p>
Full article ">Figure 4
<p>Network interactions between calcitriol and depression targets. (<b>A</b>) Complete network composed of 39 nodes. (<b>B</b>) Filtered network containing 17 genes was obtained using the parameters: degree center (DC) &gt; 9, betweenness centrality (BC) range of 2.000–0.335, and closeness centrality (CC) range of 0.300–6.000. The largest node indicates a higher degree in the network. The color of the edges refers to the types of interactions: physical interactions (green), co-expression (yellow), shared protein domains (pink), co-locations (blue), predictions (gray), genetic interactions (purple), and metabolic pathways (black).</p>
Full article ">Figure 5
<p>Network interactions between calcitriol and anxiety targets. (<b>A</b>) Complete network containing 21 nodes. (<b>B</b>) Filtered network composed of 5 nodes was obtained using the parameters: network based on the degree center (DC) &gt; 5, betweenness centrality (BC) range of 0.034–0.900, and closeness centrality (CC) range of 0.528–1.000. The largest node indicates a higher degree in the network. The color of the edges refers to the types of interactions: physical interactions (green), co-expression (yellow), shared protein domains (pink), co-locations (blue), predictions (gray), genetic interactions (purple), and metabolic pathways (black).</p>
Full article ">Figure 6
<p>GO and KEGG enrichment analysis of hub targets among genes overlapping calcitriol, depression, and anxiety. The GO enrichment analysis obtained 20 main terms related to (<b>A</b>) BP, (<b>B</b>) MF, and (<b>C</b>) CC. (<b>D</b>) The top 20 pathways were obtained by KEGG analysis. Abbreviations: BP, biological process; MF, molecular function; CC, cell composition.</p>
Full article ">Figure 7
<p>GO and KEGG enrichment analysis of hub targets among genes overlapping calcitriol and depression. The GO enrichment analysis obtained 20 main terms related to (<b>A</b>) BP, (<b>B</b>) MF, and (<b>C</b>) CC. (<b>D</b>) The top 20 pathways were obtained by KEGG analysis. Abbreviations: BP, biological process; MF, molecular function; CC, cell composition.</p>
Full article ">Figure 8
<p>GO and KEGG enrichment analysis of hub targets among genes overlapping calcitriol and anxiety. The GO enrichment analysis obtained 20 main terms related to (<b>A</b>) BP, (<b>B</b>) MF, and (<b>C</b>) CC. (<b>D</b>) The top 20 pathways were obtained by KEGG analysis. Abbreviations: BP, biological process; MF, molecular function; CC, cell composition.</p>
Full article ">Figure 9
<p>Molecular docking simulations in which calcitriol is colored in blue and the positive controls are colored in dark red. (<b>A</b>) VDR and calcitriol interacting in VDX binding site. (<b>B</b>) H3R and calcitriol in the 1IB binding site. (<b>C</b>) CB1 and calcitriol in the 9GF binding site. (<b>D</b>) CB2 and calcitriol in the WI5 binding site. (<b>E</b>) OPRL1/NOP and calcitriol in the DGV binding site. (<b>F</b>) NR1H3/LXR-α and calcitriol in the 965 binding site. (<b>G</b>) PDE3A and calcitriol in the X5M binding site. (<b>H</b>) PDE5A and calcitriol in the 5GP binding site. (<b>I</b>) PTCH1 and calcitriol in the Y01 binding site.</p>
Full article ">Figure 10
<p>The Ligplot + diagrams of the molecular docking results between the calcitriol and protein targets. Representative pose of calcitriol complexed with protein targets in which hydrogen bonds are represented as green dotted lines and hydrophobic interactions in red spoked arcs. Two-dimensional representation of protein–calcitriol interactions for the VDR (<b>A</b>), H3R (<b>B</b>), CB1 (<b>C</b>), CB2 (<b>D</b>), OPRL1/NOP (<b>E</b>), NR1H3/LXR-α (<b>F</b>), PDE3A (<b>G</b>), PDE5A (<b>H</b>), and PTCH1 (<b>I</b>). The amino acid residues highlighted (larger size and green color) are those shared between calcitriol and the respective positive controls.</p>
Full article ">Figure 11
<p>Bioinformatic validation networks. (<b>A</b>) Mental Depression PPI Network obtained in STRING and DisGENET. (<b>B</b>) Calcitriol top 10 targets PPI Network obtained in STRING. (<b>C</b>) PPI merged network constructed in STRING and Cytoscape of depression-related targets searched in DisGENET and 10 most relevant calcitriol potential targets obtained by molecular docking. Merged network is represented in circular layout by degree; the colors of nodes in green represent the highest betweenness centrality, the size of the node is represented by degree values, and the closeness centrality is represented by transparency. The edges were colored by experimentally determined interaction, with the intensity of the blue linked to the highest confidence value in this parameter. Of the top 10 promising targets of calcitriol in molecular docking (VDR, CB1, PDE5A, PDE3A, CB2, SMO, PTCH1, OPRL1, HRH3, and NR1H3), we were able to find 6 of these targets (CB1, PDE5A, PDE3A, CB2, SMO, and PTCH1 represented in black label) in the DisGeNET depression network. The proteins that interact with the target proteins of calcitriol are shown in blue and in smaller size.</p>
Full article ">
12 pages, 3470 KiB  
Article
Calcitriol in Sepsis—A Single-Centre Randomised Control Trial
by Siddhant Jeevan Thampi, Aneesh Basheer and Kurien Thomas
J. Clin. Med. 2024, 13(13), 3823; https://doi.org/10.3390/jcm13133823 - 29 Jun 2024
Viewed by 633
Abstract
Background/Objectives: Sepsis is a life-threatening organ dysfunction caused by a dysregulated host response to infection. Sepsis is a significant cause of hospital admission and the leading reason for admission to the ICU and is associated with high mortality. Vitamin D has shown [...] Read more.
Background/Objectives: Sepsis is a life-threatening organ dysfunction caused by a dysregulated host response to infection. Sepsis is a significant cause of hospital admission and the leading reason for admission to the ICU and is associated with high mortality. Vitamin D has shown promising immunomodulatory effects by upregulating the antimicrobial peptide, cathelicidin. However, previous studies analysing the use of calcitriol in sepsis have shown variable results and did not utilise APACHE II (Acute Physiology and Chronic Health Evaluation II) scores as endpoints. This study evaluates the efficacy of intramuscular calcitriol in patients admitted to the ICU with sepsis, focusing on its impact on APACHE II scores. The primary aim was to determine if intramuscular calcitriol improved APACHE II scores from day 1 to day 7 or discharge from the ICU, whichever was earlier. Secondary outcomes included 28-day mortality, ventilator days, vasopressor days, ICU stay length, adverse events, and hospital-acquired infections in ICU patients. Methods: This was a triple-blinded phase III randomised control trial. A total of 152 patients with suspected sepsis were block-randomised to receive either intramuscular calcitriol (300,000 IU) (n = 76) or a placebo (n = 76). The trial was registered with the Clinical Trials Registry—India (CTRI No: CTRI 2019/01/17066) following ethics committee approval and was not funded. Results: There was no significant difference in APACHE II scores between the calcitriol and placebo groups from day 1 to day 7 (p = 0.382). There were no significant changes in 28-day mortality (14.4% vs. 17%, p = 0.65), number of days on a ventilator (5 vs. 5, p = 0.84), number of days on vasopressors (3 vs. 3, p = 0.98), length of ICU stay (10 days vs. 11 days, p = 0.78), adverse events (27.6% vs. 19.7%, p = 0.25), and hospital-acquired infections (17.1% vs. 15.8%, p = 0.82). Conclusions: There was no effect of intramuscular calcitriol in patients admitted to the ICU with sepsis. Full article
(This article belongs to the Section Intensive Care)
Show Figures

Figure 1

Figure 1
<p>CONSORT flow diagram (ITT—intention to treat; TBM—tubercular meningitis).</p>
Full article ">Figure 2
<p>APACHE II scores from day 1 to day 2: From day 1 to day 2, APACHE II scores showed an improvement within both groups (<span class="html-italic">p</span> &lt; 0.001), but there was no significant difference between groups (<span class="html-italic">p</span> = 0.94). There was no interaction between the number of days and APACHE II scores (time*treatment interaction, <span class="html-italic">p</span> = 0.57), suggesting that the treatment does not differentially affect the rate of improvement in APACHE II scores compared to the placebo and standard care.</p>
Full article ">Figure 3
<p>APACHE II scores from day 1 to day 3: APACHE II scores showed an improvement within both groups (<span class="html-italic">p</span> &lt; 0.001), but there was no significant difference between groups (<span class="html-italic">p</span> = 0.42). There was no interaction between the number of days and APACHE II scores (time*treatment interaction, <span class="html-italic">p</span> = 0.71), suggesting that the treatment does not differentially affect the rate of improvement in APACHE II scores compared to the placebo and standard care.</p>
Full article ">Figure 4
<p>APACHE II scores from day 1 to day 4: APACHE II scores showed an improvement within both groups (<span class="html-italic">p</span> &lt; 0.001), but there was no significant difference between groups (<span class="html-italic">p</span> = 0.22). There was no interaction between the number of days and APACHE II scores (time*treatment interaction, <span class="html-italic">p</span> = 0.26), suggesting that the treatment does not differentially affect the rate of improvement in APACHE II scores compared to the placebo and standard care.</p>
Full article ">Figure 5
<p>APACHE II scores from day 1 to day 5: APACHE II scores showed an improvement within both groups (<span class="html-italic">p</span> &lt; 0.001), but there was no significant difference between groups (<span class="html-italic">p</span> = 0.39). There was no interaction between the number of days and APACHE II scores (time*treatment interaction, <span class="html-italic">p</span> = 0.43), suggesting that the treatment does not differentially affect the rate of improvement in APACHE II scores compared to the placebo and standard care.</p>
Full article ">Figure 6
<p>APACHE II scores from day 1 to day 6: APACHE II scores showed an improvement within both groups (<span class="html-italic">p</span> &lt; 0.001), but there was no significant difference between groups (<span class="html-italic">p</span> = 0.48). There was no interaction between the number of days and APACHE II scores (time*treatment interaction, <span class="html-italic">p</span> = 0.17) suggesting that the treatment does not differentially affect the rate of improvement in APACHE II scores compared to the placebo and standard care.</p>
Full article ">Figure 7
<p>APACHE II scores from day 1 to day 7: APACHE II scores showed an improvement within both groups (<span class="html-italic">p</span> &lt; 0.001), but there was no significant difference between groups (<span class="html-italic">p</span> = 0.38). There was no interaction between the number of days and APACHE II scores (time*treatment interaction, <span class="html-italic">p</span> = 0.08), suggesting that the treatment does not differentially affect the rate of improvement in APACHE II scores compared to the placebo and standard care.</p>
Full article ">Figure 8
<p>Kaplan–Meier survival analysis. There was no significant difference in mortality in both groups at 28 days.</p>
Full article ">
13 pages, 656 KiB  
Article
Vitamin D3 (Calcitriol) Monotherapy Decreases Tumor Growth, Increases Survival, and Correlates with Low Neutrophil-to-Lymphocyte Ratio in a Murine HPV-16-Related Cancer Model
by Alejandra E. Hernández-Rangel, Gustavo A. Hernandez-Fuentes, Daniel A. Montes-Galindo, Carmen A. Sanchez-Ramirez, Ariana Cabrera-Licona, Margarita L. Martinez-Fierro, Iram P. Rodriguez-Sanchez, Idalia Garza-Veloz, Janet Diaz-Martinez, Juan C. Casarez-Price, Jorge E. Plata-Florenzano, Hector Ochoa-Díaz-Lopez, Angel Lugo-Trampe and Iván Delgado-Enciso
Biomedicines 2024, 12(6), 1357; https://doi.org/10.3390/biomedicines12061357 - 18 Jun 2024
Viewed by 1018
Abstract
Vitamin D3 or calcitriol (VitD3) has been shown to have anticancer and anti-inflammatory activity in in vitro models and clinical studies. However, its effect on HPV-16-related cancer has been sparsely explored. In this study, we aimed to determine whether monotherapy or combination therapy [...] Read more.
Vitamin D3 or calcitriol (VitD3) has been shown to have anticancer and anti-inflammatory activity in in vitro models and clinical studies. However, its effect on HPV-16-related cancer has been sparsely explored. In this study, we aimed to determine whether monotherapy or combination therapy with cisplatin (CP) reduces tumor growth and affects survival and systemic inflammation. Treatments were administered to C57BL/6 mice with HPV-16-related tumors (TC-1 cells) as follows: (1) placebo (100 µL vehicle, olive oil, orally administered daily); (2) VitD3 (3.75 µg/kg calcitriol orally administered daily); (3) CP (5 mg/kg intraperitoneally, every 7 days); and (4) VitD3+CP. Tumor growth was monitored for 25 days, survival for 60 days, and the neutrophil-to-lymphocyte ratio (NLR) was evaluated on days 1 (baseline), 7, and 14. VitD3+CP showed greater success in reducing tumor volume compared to CP monotherapy (p = 0.041), while no differences were observed between CP and VitD3 monotherapy (p = 0.671). Furthermore, VitD3+CP prolonged survival compared to CP (p = 0.036) and VitD3 (p = 0.007). Additionally, at day 14 the VitD3 and VitD3+CP groups showed significantly lower NLR values than the CP group (p < 0.05, for both comparisons). Vitamin D3 could be a promising adjuvant in the treatment of cervical cancer or solid tumors and deserves further investigation. Full article
(This article belongs to the Special Issue The Role of Inflammatory Cytokines in Cancer Progression 2.0)
Show Figures

Figure 1

Figure 1
<p>Tumor Growth and Survival Curves. (<b>A</b>) The tumor growth curve in response to different treatments is presented as mean ± standard error of the mean. All treatments significantly reduced tumor growth compared to the placebo (PB): PB vs. vitamin D3 (VitD3) <span class="html-italic">p</span> &lt; 0.05 from day 13 onwards; PB vs. cisplatin (CP) <span class="html-italic">p</span> &lt; 0.05 from day 15 onwards; and PB vs. VitD3+CP <span class="html-italic">p</span> &lt; 0.05 from day 15 onwards. The VitD3+CP group showed a lower growth rate compared to the groups treated with only cisplatin (<span class="html-italic">p</span> = 0.041, day 25) and only VitD3 (<span class="html-italic">p</span> = 0.043, day 25), but these differences were statistically significant only until day 25. No differences were observed between the VitD3 and CP groups during the 25-day follow-up period. (<b>B</b>) The cumulative survival of the groups with different treatments is shown. It was observed that the group with the lowest survival was the PB group (median 23 days, 95% CI 19.60–29.39), significantly lower than the rest of the groups (<span class="html-italic">p</span> &lt; 0.01 for all comparisons, log-rank test). The group with the longest survival was the VitD3+CP group (median of 41 days, 95% CI 25.88–56.10), significantly longer than the rest of the groups (<span class="html-italic">p</span> &lt; 0.05 for all comparisons, log-rank test). The CP and VitD3 groups, with medians of 32 days (95% CI 30.32–33.67) and 29 days (95% CI 26.76–31.23), respectively, showed no differences between them (<span class="html-italic">p</span> = 0.069, log-rank test).</p>
Full article ">
24 pages, 7574 KiB  
Review
Structure and the Anticancer Activity of Vitamin D Receptor Agonists
by Agnieszka Powała, Teresa Żołek, Geoffrey Brown and Andrzej Kutner
Int. J. Mol. Sci. 2024, 25(12), 6624; https://doi.org/10.3390/ijms25126624 - 16 Jun 2024
Cited by 1 | Viewed by 788
Abstract
Vitamin D is a group of seco-steroidal fat-soluble compounds. The two basic forms, vitamin D2 (ergocalciferol) and vitamin D3 (cholecalciferol), do not have biological activity. They are converted in the body by a two-step enzymatic hydroxylation into biologically active forms, 1α,25-dihydroxyvitamin [...] Read more.
Vitamin D is a group of seco-steroidal fat-soluble compounds. The two basic forms, vitamin D2 (ergocalciferol) and vitamin D3 (cholecalciferol), do not have biological activity. They are converted in the body by a two-step enzymatic hydroxylation into biologically active forms, 1α,25-dihydroxyvitamin D2 [ercalcitriol, 1,25(OH)2D2] and 1α,25-dihydroxyvitamin D3 [calcitriol, 1,25(OH)2D3], which act as classical steroid hormones. 1,25(OH)2D3 exerts most of its physiological functions by binding to the nuclear vitamin D receptor (VDR), which is present in most body tissues to provide support to a broad range of physiological processes. Vitamin D-liganded VDR controls the expression of many genes. High levels of 1,25(OH)2D3 cause an increase in calcium in the blood, which can lead to harmful hypercalcemia. Several analogs of 1,25(OH)2D3 and 1,25(OH)2D2 have been designed and synthesized with the aim of developing compounds that have a specific therapeutic function, for example, with potent anticancer activity and a reduced toxic calcemic effect. Particular structural modifications to vitamin D analogs have led to increased anticancer activity and reduced calcemic action with the prospect of extending work to provide future innovative therapies. Full article
(This article belongs to the Section Molecular Pharmacology)
Show Figures

Figure 1

Figure 1
<p>Chemical structures of the most active form of vitamin D<sub>2</sub>, 1,25-dihydroxyvitamin D<sub>2</sub> (<b>a</b>), and vitamin D<sub>3</sub>, 1,25-dihydroxyvitamin D<sub>3</sub> (<b>b</b>).</p>
Full article ">Figure 2
<p>The domain structure of the vitamin D receptor consists of the N-terminal domain (NTD), the DNA-binding domain, the hinge region, the insertion region, and the ligand-binding domain.</p>
Full article ">Figure 3
<p>Structure of hVDR LBD (RCSB PDB ID: 1DB1, adapted from Rochel N. et al. [<a href="#B63-ijms-25-06624" class="html-bibr">63</a>], created with DISCOVERY STUDIO v. 22 software (<a href="https://discover.3ds.com" target="_blank">https://discover.3ds.com</a>, accessed on 24 May 2024). (<b>a</b>) Ribbon representation of the VDR LBD bound to 1α,25(OH)<sub>2</sub>D<sub>3</sub> (green color), composed of 12 helices. (<b>b</b>) Binding mode of 1,25(OH)<sub>2</sub>D<sub>3</sub> in the VDR LBP. The volume of the LBP is shown as a hydrophobic surface. The interacting residues and their locations in the VDR structure are shown. (<b>c</b>) The residues anchored through specific H-bonds and hydrophobic interactions to 1,25(OH)<sub>2</sub>D<sub>3</sub> are depicted.</p>
Full article ">
31 pages, 3277 KiB  
Review
Physiology of Vitamin D—Focusing on Disease Prevention
by Sunil J. Wimalawansa
Nutrients 2024, 16(11), 1666; https://doi.org/10.3390/nu16111666 - 29 May 2024
Cited by 4 | Viewed by 2783
Abstract
Vitamin D is a crucial micronutrient, critical to human health, and influences many physiological processes. Oral and skin-derived vitamin D is hydroxylated to form calcifediol (25(OH)D) in the liver, then to 1,25(OH)2D (calcitriol) in the kidney. Alongside the parathyroid hormone, calcitriol [...] Read more.
Vitamin D is a crucial micronutrient, critical to human health, and influences many physiological processes. Oral and skin-derived vitamin D is hydroxylated to form calcifediol (25(OH)D) in the liver, then to 1,25(OH)2D (calcitriol) in the kidney. Alongside the parathyroid hormone, calcitriol regulates neuro-musculoskeletal activities by tightly controlling blood-ionized calcium concentrations through intestinal calcium absorption, renal tubular reabsorption, and skeletal mineralization. Beyond its classical roles, evidence underscores the impact of vitamin D on the prevention and reduction of the severity of diverse conditions such as cardiovascular and metabolic diseases, autoimmune disorders, infection, and cancer. Peripheral target cells, like immune cells, obtain vitamin D and 25(OH)D through concentration-dependent diffusion from the circulation. Calcitriol is synthesized intracellularly in these cells from these precursors, which is crucial for their protective physiological actions. Its deficiency exacerbates inflammation, oxidative stress, and increased susceptibility to metabolic disorders and infections; deficiency also causes premature deaths. Thus, maintaining optimal serum levels above 40 ng/mL is vital for health and disease prevention. However, achieving it requires several times more than the government’s recommended vitamin D doses. Despite extensive published research, recommended daily intake and therapeutic serum 25(OH)D concentrations have lagged and are outdated, preventing people from benefiting. Evidence suggests that maintaining the 25(OH)D concentrations above 40 ng/mL with a range of 40–80 ng/mL in the population is optimal for disease prevention and reducing morbidities and mortality without adverse effects. The recommendation for individuals is to maintain serum 25(OH)D concentrations above 50 ng/mL (125 nmol/L) for optimal clinical outcomes. Insights from metabolomics, transcriptomics, and epigenetics offer promise for better clinical outcomes from vitamin D sufficiency. Given its broader positive impact on human health with minimal cost and little adverse effects, proactively integrating vitamin D assessment and supplementation into clinical practice promises significant benefits, including reduced healthcare costs. This review synthesized recent novel findings related to the physiology of vitamin D that have significant implications for disease prevention. Full article
(This article belongs to the Section Nutrition and Public Health)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Illustrates the pathway for the generation of vitamin D<sub>3</sub> from 7-dehydrocholesterol (7-DHC) following exposure to UVB rays. The activation of vitamin D to its metabolites, 25(OH)D (calcifediol) and 1,25(OH)<sub>2</sub>D (calcitriol), is highlighted, including its 24-hydroxy, inactive catabolic metabolites. Critical organs responsible for vitamin D generation/metabolism and the parathyroid hormone (PTH)-mediated regulation of serum ionized calcium (Ca<sup>2+</sup>) levels are illustrated. The typical activation route for skin-derived and oral/dietary vitamin D forming 1,25(OH)<sub>2</sub>D is depicted. While 25-hydroxylase activity occurs mainly in the liver, the conversion of vitamin D and 25(OH)D to 1,25(OH)<sub>2</sub>D via the 1α-hydroxylase enzyme occurs in renal tubules and peripheral target cells of vitamin D. The figure also demonstrates the control of serum Ca<sup>2+</sup> levels through intestinal absorption, bone turnover, and PTH-mediated renal handling (+ upregulation and − downregulation) (Fibroblast growth factor-23 = FGF-23; UVB = Ultraviolet B rays; VDBP = Vitamin D binding protein).</p>
Full article ">Figure 2
<p>The essential pathways of acquiring vitamin D in humans are illustrated. The figure demonstrates key functional differences between the circulatory hormonal form of calcitriol (the hormonal form) that is generated in renal tubular cells vs. the intracellularly generated calcitriol in peripheral target cells, such as immune cells (modified from Wimalawansa, SJ., 2023 [<a href="#B7-nutrients-16-01666" class="html-bibr">7</a>]).</p>
Full article ">Figure 3
<p>Illustrates the pathways for the synthesis of vitamin D and its activation into 25(OH)D in the liver and 1,25(OH)<sub>2</sub>D in the kidneys, as well as the role of parathyroid hormone (PTH) in the maintenance of ionized calcium (Ca<sup>++</sup>) in the circulation (7-DHC, 7-dehydrocholesterol).</p>
Full article ">Figure 4
<p>The figure depicts the structures of the three most common vitamin D metabolites, highlighting sites of generation, specific hydroxylating enzymes, and their average concentrations in the bloodstream. With daily supplementation or sun exposure, D<sub>3</sub> and 25(OH)D concentrations remain similar and in equilibrium. Notably, while circulatory concentrations of D<sub>3</sub> and 25(OH)D<sub>3</sub> are in the nanomolar range, 1,25(OH)<sub>2</sub>D (calcitriol) is present in picomolar amounts—approximately 900-fold lower (modified from Bickel, D [<a href="#B186-nutrients-16-01666" class="html-bibr">186</a>] and Wimalawansa, 2022 [<a href="#B49-nutrients-16-01666" class="html-bibr">49</a>]).</p>
Full article ">Figure 5
<p>Broader functions of vitamin D (1,25(OH)<sub>2</sub>D): The figure illustrates both endocrine and non-endocrine functions that affect various cells and tissues in the body (modified from Wimalawansa, SJ, 2023; [<a href="#B15-nutrients-16-01666" class="html-bibr">15</a>]) ↑ = Increased activity; ↓ = Decreased activity; − = Negative (reduced); + = positive (enhanced).</p>
Full article ">Figure 6
<p>The illustration highlights the key activation steps and intricate yet vital interactions of vitamin D and its active metabolites, as well as their inherent feedback control systems. These mechanisms maintain circulatory ionized calcium concentrations, representing vitamin D’s fundamental endocrine function (Up arrow = up-regulated (increased activity); Down arrow = down-regulated (reduced functions)).</p>
Full article ">
22 pages, 3527 KiB  
Article
1,25-Dihydroxyvitamin D3 Suppresses Prognostic Survival Biomarkers Associated with Cell Cycle and Actin Organization in a Non-Malignant African American Prostate Cell Line
by Jabril R. Johnson, Rachel N. Martini, Yate-Ching Yuan, Leanne Woods-Burnham, Mya Walker, Greisha L. Ortiz-Hernandez, Firas Kobeissy, Dorothy Galloway, Amani Gaddy, Chidinma Oguejiofor, Blake Allen, Deyana Lewis, Melissa B. Davis, K. Sean Kimbro, Clayton C. Yates, Adam B. Murphy and Rick A. Kittles
Biology 2024, 13(5), 346; https://doi.org/10.3390/biology13050346 - 15 May 2024
Viewed by 2229
Abstract
Vitamin D3 is a steroid hormone that confers anti-tumorigenic properties in prostate cells. Serum vitamin D3 deficiency has been associated with advanced prostate cancer (PCa), particularly affecting African American (AA) men. Therefore, elucidating the pleiotropic effects of vitamin D on signaling [...] Read more.
Vitamin D3 is a steroid hormone that confers anti-tumorigenic properties in prostate cells. Serum vitamin D3 deficiency has been associated with advanced prostate cancer (PCa), particularly affecting African American (AA) men. Therefore, elucidating the pleiotropic effects of vitamin D on signaling pathways, essential to maintaining non-malignancy, may provide additional drug targets to mitigate disparate outcomes for men with PCa, especially AA men. We conducted RNA sequencing on an AA non-malignant prostate cell line, RC-77N/E, comparing untreated cells to those treated with 10 nM of vitamin D3 metabolite, 1α,25(OH)2D3, at 24 h. Differential gene expression analysis revealed 1601 significant genes affected by 1α,25(OH)2D3 treatment. Pathway enrichment analysis predicted 1α,25(OH)2D3- mediated repression of prostate cancer, cell proliferation, actin cytoskeletal, and actin-related signaling pathways (p < 0.05). Prioritizing genes with vitamin D response elements and associating expression levels with overall survival (OS) in The Cancer Genome Atlas Prostate Adenocarcinoma (TCGA PRAD) cohort, we identified ANLN (Anillin) and ECT2 (Epithelial Cell Transforming 2) as potential prognostic PCa biomarkers. Both genes were strongly correlated and significantly downregulated by 1α,25(OH)2D3 treatment, where low expression was statistically associated with better overall survival outcomes in the TCGA PRAD public cohort. Increased ANLN and ECT2 mRNA gene expression was significantly associated with PCa, and Gleason scores using both the TCGA cohort (p < 0.05) and an AA non-malignant/tumor-matched cohort. Our findings suggest 1α,25(OH)2D3 regulation of these biomarkers may be significant for PCa prevention. In addition, 1α,25(OH)2D3 could be used as an adjuvant treatment targeting actin cytoskeleton signaling and actin cytoskeleton-related signaling pathways, particularly among AA men. Full article
(This article belongs to the Special Issue Cancer and Signalling: Targeting Cellular Pathways)
Show Figures

Figure 1

Figure 1
<p>IPA and Pathway Studio Analysis of the DEGs between 1α,25(OH)<sub>2</sub>D<sub>3</sub>-treated and untreated AA non-malignant prostate cells. (<b>A</b>) Top significant canonical pathways and (<b>B</b>,<b>C</b>) disease and function terms enriched from 1601 DEGs between 1α,25(OH)<sub>2</sub>D<sub>3</sub>-treated and untreated AA non-malignant prostate cells. In figure (<b>A</b>), Z-score represents predicted activation (Z &gt; 0, orange) or inhibition (Z &lt; 0, blue), and bars in black have no predicted effect. <span class="html-italic">p</span>-value represents B–H adjusted <span class="html-italic">p</span>-value. In figure (<b>B</b>,<b>C</b>), proteins in an oval red color are downregulated, while those in an oval yellow color are upregulated, and their relationships with (<b>B</b>) oxidative stress, inflammation, cell motility, and cell cycle and (<b>C</b>) PCa, malignant transformation, prostate adenoma, DNA damage, and transcription activation are indicated. The raw data with the <span class="html-italic">p</span>-values of the relationship and pathways indicated in each of the generated panels are provided as supplementary raw data (<a href="#app1-biology-13-00346" class="html-app">Tables S3 and S4</a>).</p>
Full article ">Figure 2
<p>DEG Selection Algorithm. Selection algorithm designed to identify DEGs for validation and further downstream analyses.</p>
Full article ">Figure 3
<p><span class="html-italic">ANLN</span> and <span class="html-italic">ECT2</span> expression association with overall survival in the TCGA PRAD cohort. Kaplan–Meier curves showing overall survival outcomes of (<b>A</b>) <span class="html-italic">ANLN</span> low- versus high-expressing tumors based on the mean of mRNA expression of specific gene of interest in the TCGA PRAD cohort; (<b>B</b>) <span class="html-italic">ECT2</span> low- versus high-expressing tumors based on the mean of mRNA expression of specific gene of interests in the TCGA PRAD cohort (<span class="html-italic">n</span> = 489). Low expression is shown in red, and high expression is shown in black. N values of high- versus low-expressing and <span class="html-italic">p</span>-values generated using student’s <span class="html-italic">t</span>-test are reported.</p>
Full article ">Figure 4
<p><span class="html-italic">ANLN</span> and <span class="html-italic">ECT2</span> are significantly correlated and more highly expressed in prostate tumor tissue compared to non-malignant tissue. (<b>A</b>,<b>B</b>) Overexpression of <span class="html-italic">ANLN</span> and <span class="html-italic">ECT2</span> mRNA expression in tumor (<span class="html-italic">n</span> = 490) compared to normal (<span class="html-italic">n</span> = 52), measured in expression intensity. (<b>C</b>,<b>D</b>) Pearson correlation analysis of <span class="html-italic">ANLN</span> and <span class="html-italic">ECT2</span> expression in normal (<span class="html-italic">n</span> = 52) compared to tumor (<span class="html-italic">n</span> = 490), measured in expression intensity. (<b>E</b>,<b>F</b>) <span class="html-italic">ANLN</span> and <span class="html-italic">ECT2</span> association with sum Gleason score in the TCGA PRAD cohort (<span class="html-italic">p</span> &lt; 0.05), measured in expression intensity. (<b>G</b>,<b>H</b>) Overexpression of <span class="html-italic">ANLN</span> and <span class="html-italic">ECT2</span> mRNA expression in our AA clinical match tumor (<span class="html-italic">n</span> = 73)/non-malignant (<span class="html-italic">n</span> = 74) cohort, measured in TPM. (<b>I</b>,<b>J</b>) Pearson correlation analysis of <span class="html-italic">ANLN</span> and <span class="html-italic">ECT2</span> expression in non-malignant (<span class="html-italic">n</span> = 73) and tumor-matched tissue (<span class="html-italic">n</span> = 73), measured in TPM. (<b>K</b>,<b>L</b>) <span class="html-italic">ANLN</span> and <span class="html-italic">ECT2</span> association with sum Gleason score in an AA clinical cohort (<span class="html-italic">p</span> &lt; 0.05), measured in TPM. <span class="html-italic">p</span>-values and r-coefficients are reported on graph (<b>A</b>–<b>D</b>,<b>G</b>–<b>J</b>). * <span class="html-italic">p</span> &lt; 0.05; NS = not significant (<b>E</b>,<b>F</b>,<b>K</b>,<b>L</b>).</p>
Full article ">Figure 5
<p><span class="html-italic">ANLN</span> and <span class="html-italic">ECT2</span> are downregulated upon 1α,25(OH)<sub>2</sub>D<sub>3</sub> treatment in RC-77N/E and RWPE1 cell lines. Boxplots of (<b>A</b>) <span class="html-italic">ANLN</span> and (<b>B</b>) <span class="html-italic">ECT2</span> gene expression among untreated controls (<span class="html-italic">n</span> = 3) and 1α,25(OH)<sub>2</sub>D<sub>3</sub> treated (<span class="html-italic">n</span> = 2), at 24 h, in RC-77N/E cell line replicates. (<b>A</b>) <span class="html-italic">ANLN</span> and (<b>B</b>) <span class="html-italic">ECT2</span> are measured in TPM. In the validation microarray study, boxplots of (<b>C</b>) <span class="html-italic">ANLN</span> and (<b>D</b>) <span class="html-italic">ECT2</span> gene expression among untreated controls (<span class="html-italic">n</span> = 12) and 100 nM 1α,25(OH)<sub>2</sub>D<sub>3</sub> (<span class="html-italic">n</span> = 12) at (<span class="html-italic">n</span> = 4) 6 h (blue triangle), 24 h (red circle), and 48 h (purple diamond); RWPE1 cell line replicates. (<b>C</b>) <span class="html-italic">ANLN</span> and (<b>D</b>) <span class="html-italic">ECT2</span> are measured in log2 fold change intensity. For boxplots, student’s <span class="html-italic">t</span>-test <span class="html-italic">p</span>-values are reported on the graphs. * FDR <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>RT-qPCR validation of RNA–Seq analysis. (<b>A</b>) RT–qPCR confirmation results for the randomly relevant selected DEGs from the RC-77N/E-treated 1α,25(OH)<sub>2</sub>D<sub>3</sub> compared to the untreated control. (<b>B</b>) Regression analysis of the log2 fold change values between the RNA–Seq and RT–qPCR validation of the RC-77N/E-treated 1α,25(OH)<sub>2</sub>D<sub>3</sub> compared to the untreated control. The GAPDH gene was used as an internal reference control gene. RT–qPCR was performed with replicates, * FDR <span class="html-italic">p</span> &lt; 0.05 is significant.</p>
Full article ">Figure 7
<p>IPA and STRING PPI network analysis of DEGs associated with <span class="html-italic">ANLN</span> and <span class="html-italic">ECT2</span>. (<b>A</b>) 1α,25(OH)<sub>2</sub>D<sub>3</sub> treatment led to predicted repression of molecules and disease/function categories up- and downstream of <span class="html-italic">ANLN</span> and <span class="html-italic">ECT2</span>. Red (downregulated), green (upregulated), blue (predicted inhibition), and yellow (inconclusive). Log2 fold change and adjusted <span class="html-italic">p</span>-values for DEGs are reported in figure. (<b>B</b>) <span class="html-italic">ANLN</span> and <span class="html-italic">ECT2</span> Protein–Protein Interaction Network analysis using STRING database. Legend reported in figure. (<b>C</b>) Co-expression scores based on RNA expression patterns and on protein co-regulation provided by STRING ProteomeHD. Co-expression scale reported in figure.</p>
Full article ">
17 pages, 1713 KiB  
Opinion
Discordant Health Implications and Molecular Mechanisms of Vitamin D in Clinical and Preclinical Studies of Prostate Cancer: A Critical Appraisal of the Literature Data
by Annika Fendler, Carsten Stephan, Bernhard Ralla and Klaus Jung
Int. J. Mol. Sci. 2024, 25(10), 5286; https://doi.org/10.3390/ijms25105286 - 13 May 2024
Cited by 1 | Viewed by 1095
Abstract
Clinical and preclinical studies have provided conflicting data on the postulated beneficial effects of vitamin D in patients with prostate cancer. In this opinion piece, we discuss reasons for discrepancies between preclinical and clinical vitamin D studies. Different criteria have been used as [...] Read more.
Clinical and preclinical studies have provided conflicting data on the postulated beneficial effects of vitamin D in patients with prostate cancer. In this opinion piece, we discuss reasons for discrepancies between preclinical and clinical vitamin D studies. Different criteria have been used as evidence for the key roles of vitamin D. Clinical studies report integrative cancer outcome criteria such as incidence and mortality in relation to vitamin D status over time. In contrast, preclinical vitamin D studies report molecular and cellular changes resulting from treatment with the biologically active vitamin D metabolite, 1,25-dihydroxyvitamin D3 (calcitriol) in tissues. However, these reported changes in preclinical in vitro studies are often the result of treatment with biologically irrelevant high calcitriol concentrations. In typical experiments, the used calcitriol concentrations exceed the calcitriol concentrations in normal and malignant prostate tissue by 100 to 1000 times. This raises reasonable concerns regarding the postulated biological effects and mechanisms of these preclinical vitamin D approaches in relation to clinical relevance. This is not restricted to prostate cancer, as detailed data regarding the tissue-specific concentrations of vitamin D metabolites are currently lacking. The application of unnaturally high concentrations of calcitriol in preclinical studies appears to be a major reason why the results of preclinical in vitro studies hardly match up with outcomes of vitamin D-related clinical studies. Regarding future studies addressing these concerns, we suggest establishing reference ranges of tissue-specific vitamin D metabolites within various cancer entities, carrying out model studies on human cancer cells and patient-derived organoids with biologically relevant calcitriol concentrations, and lastly improving the design of vitamin D clinical trials where results from preclinical studies guide the protocols and endpoints within these trials. Full article
(This article belongs to the Section Molecular Endocrinology and Metabolism)
Show Figures

Figure 1

Figure 1
<p>Simplified schemes of (<b>A</b>) biosynthesis and degradation of vitamin D and (<b>B</b>) the calcitriol action as ligand of the vitamin D receptor and the subsequent epigenome and transcriptome changes. The metabolizing vitamin D hydroxylases are members of the cytochrome P450 family and are indicated abbreviated in parentheses. Abbreviations for subfigure (<b>A</b>): Vitamin D binding protein (VDBP); 25-hydroxylases encoded by cytochrome P450 family 2 subfamily R member 1 (<span class="html-italic">CYP2R1</span>) and cytochrome P450 family 27 subfamily A member 1 (<span class="html-italic">CYP27A1</span>); 1α-hydroxylase encoded by cytochrome P450 family 27 subfamily B member 1 (<span class="html-italic">CYP27B1</span>); 24-hydroxylase encoded by cytochrome P450 family 24 subfamily A member 1 (<span class="html-italic">CYP24A1</span>). (<b>B</b>) Vitamin D receptor (VDR); Retinoid X receptor (RXR); vitamin D response elements (VDREs). The different colored lines symbolize the two DNA strands.</p>
Full article ">Figure 2
<p>Gene expression of the metabolizing vitamin D hydroxylases of the cytochrome P450 family, the vitamin D receptor, and the LDL-receptor-related protein 2 in tissue samples of normal prostate and prostate cancer. Data from The Cancer Genome Atlas (TCGA) downloaded from UCSC Xena (<a href="https://xena.ucsc.edu/" target="_blank">https://xena.ucsc.edu/</a>, accessed on 28 March 2024) were used [<a href="#B42-ijms-25-05286" class="html-bibr">42</a>,<a href="#B43-ijms-25-05286" class="html-bibr">43</a>]. Further details are given in the <a href="#app1-ijms-25-05286" class="html-app">Supplementary Materials</a>. Abbreviations: Cytochrome P450 family 2 subfamily R member 1 (<span class="html-italic">CYP2R1</span>); cytochrome P450 family 27 subfamily A member 1 (<span class="html-italic">CYP27A1</span>); cytochrome P450 family 27 subfamily B member 1 (<span class="html-italic">CYP27B1</span>); cytochrome P450 family 24 subfamily A member 1 (<span class="html-italic">CYP24A1</span>); vitamin D receptor (<span class="html-italic">VDR</span>), LDL-receptor-related protein 2 (<span class="html-italic">LRP2</span>).</p>
Full article ">
Back to TopTop