[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 (120)

Search Parameters:
Keywords = melanosome

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 15119 KiB  
Article
Skin Anatomy, Bone Histology and Taphonomy of a Toarcian (Lower Jurassic) Ichthyosaur (Reptilia: Ichthyopterygia) from Luxembourg, with Implications for Paleobiology
by Ida Bonnevier Wallstedt, Peter Sjövall, Ben Thuy, Randolph G. De La Garza, Mats E. Eriksson and Johan Lindgren
Diversity 2024, 16(8), 492; https://doi.org/10.3390/d16080492 - 12 Aug 2024
Viewed by 372
Abstract
A partial ichthyosaur skeleton from the Toarcian (Lower Jurassic) bituminous shales of the ‘Schistes Carton’ unit of southern Luxembourg is described and illustrated. In addition, associated remnant soft tissues are analyzed using a combination of imaging and molecular techniques. The fossil (MNHNL TV344) [...] Read more.
A partial ichthyosaur skeleton from the Toarcian (Lower Jurassic) bituminous shales of the ‘Schistes Carton’ unit of southern Luxembourg is described and illustrated. In addition, associated remnant soft tissues are analyzed using a combination of imaging and molecular techniques. The fossil (MNHNL TV344) comprises scattered appendicular elements, together with a consecutive series of semi-articulated vertebrae surrounded by extensive soft-tissue remains. We conclude that TV344 represents a skeletally immature individual (possibly of the genus Stenopterygius) and that the soft parts primarily consist of fossilized skin, including the epidermis (with embedded melanophore pigment cells and melanosome organelles) and dermis. Ground sections of dorsal ribs display cortical microstructures reminiscent of lines of arrested growth (LAGs), providing an opportunity for a tentative age determination of the animal at the time of death (>3 years). It is further inferred that the exceptional preservation of TV344 was facilitated by seafloor dysoxia/anoxia with periodical intervals of oxygenation, which triggered phosphatization and the subsequent formation of a carbonate concretion. Full article
Show Figures

Figure 1

Figure 1
<p>Geological map of Luxembourg and its surroundings, with the study area indicated by a green star (modified from <a href="#diversity-16-00492-f001" class="html-fig">Figure 1</a> in Ref. [<a href="#B11-diversity-16-00492" class="html-bibr">11</a>]). White lines indicate national borders.</p>
Full article ">Figure 2
<p>TV344, a partial ichthyosaur from the ‘Schistes Carton’ unit of Luxembourg. (<b>A</b>): Photograph taken under polarized light. The sample site for the petrographic sections is marked with a red rectangle, whereas the samples used in our molecular analyses are demarcated by a green rectangle. The scale bar represents 5 cm. (<b>B</b>): Sketch map outlining the different components of the fossil and its surrounding matrix.</p>
Full article ">Figure 3
<p>Petrographic section produced from a set of ribs on the left-hand side of TV344. (<b>A</b>): Upper part of the section, showing a more proximally located section through a rib. Scale bar represents 1 mm. (<b>B</b>): Lower part of the petrographic slide, showing a distal cross-section of a rib. Scale bar represents 500 µm. (<b>C</b>): Magnification of (<b>B</b>), as indicated by red frame. Cortical bone, cancellous bone and the medullary cavity are all indicated. Circumferential lines in the cortex are highlighted by arrowheads. Lacunae are indicated by white lines. Scale bar represents 50 µm.</p>
Full article ">Figure 4
<p>Detailed photographs of TV344. (<b>A</b>): Ribs from the left side of the fossil covered by rippled skin (indicated by an arrowhead). (<b>B</b>): Close-up view of skin covering the ribs, showing the transition between the epidermis and underlying ridged layer (indicated by arrowheads).</p>
Full article ">Figure 5
<p>FEG-SEM and EDX micrographs of TV344 integument. Back-scattered electron image to the left and EDX elemental maps to the right. Note dermal ridges protruding into the phosphatized epidermis, and localized concentrations of iron and sulfur in the ridges. Scale bars represent 50 μm.</p>
Full article ">Figure 6
<p>Pigmentation in ichthyosaur fossils. (<b>A</b>): Epidermis of TV344 displaying dark, branching bodies (highlighted by arrowheads) with a diameter of ~10–20 μm. Scale bar represents 100 μm. (<b>B</b>): Epidermis of <span class="html-italic">Stenopterygius</span> specimen MH 432 similarly displaying dark, branching melanophores (see [<a href="#B3-diversity-16-00492" class="html-bibr">3</a>]). Melanophores highlighted by arrowheads. Scale bar represents 200 μm. (<b>C</b>,<b>D</b>): FEG-SEM micrographs of demineralized TV344 integument showing remnant melanosomes embedded in amorphous organic matter. Scale bars represent 1 μm.</p>
Full article ">Figure 7
<p>Principal sketch of the inferred layering of the integument in TV344.</p>
Full article ">Figure 8
<p>Model for the taphonomic conditions that enabled the fossilization of TV344. The carcass is covered by a microbial mat containing both sulphate-oxidizing and sulphate-reducing bacteria. During brief oxygenated periods (upper left half of figure), sulphate-oxidizing bacteria (black ovals) locally lower the pH, thereby enabling ions in the seawater and sediment to form a sheet of calcium phosphate on top of the carcass. During dysoxic/anoxic periods (upper right half of figure), sulphate-reducing bacteria (white ovals) locally elevated the pH, thereby facilitating the formation of a calcium carbonate nodule from ions in the seawater and sediment.</p>
Full article ">Figure A1
<p>TV344 photographed under ultraviolet light.</p>
Full article ">Figure A2
<p>(<b>A</b>): ToF-SIMS images of negative ions taken at Sample Point #5 on the right-hand side of the fossil (see <a href="#diversity-16-00492-f002" class="html-fig">Figure 2</a>A), representative of nitrogen-containing organics (CN<sup>−</sup>), epoxy (C<sub>4</sub>H<sub>5</sub>O<sub>2</sub><sup>−</sup>), silica (SiO<sub>2</sub><sup>−</sup>) and sulfate (SO<sub>3</sub><sup>−</sup>), respectively. The two panels in the bottom row depict an overlay of epoxy, CN<sup>−</sup> and silica together with the total signal intensity distribution. (<b>B</b>): ToF-SIMS images of positive ions acquired at Sample Point #5 on the right-hand side of the fossil (see <a href="#diversity-16-00492-f002" class="html-fig">Figure 2</a>A), representative of amino acid-containing organics (C<sub>x</sub>H<sub>y</sub>N<sup>+</sup>), aliphatic hydrocarbons (C<sub>x</sub>H<sub>y</sub><sup>+</sup>), potassium (K<sup>+</sup>) and silica (SiOH<sup>+</sup>). The two panels in the bottom row include an overlay of aliphatic hydrocarbons, potassium and silica, and the total signal intensity distribution.</p>
Full article ">Figure A3
<p>Overview of a complete petrographic section obtained from TV344. Scale bar represents 5 mm.</p>
Full article ">Figure A4
<p>FEG-SEM micrograph of TV344 integument, showing oblong microbodies clustered near the inner termination of the epidermis. Scale bar represents 2 μm.</p>
Full article ">
17 pages, 5733 KiB  
Article
Transcriptomic and Metabolomic Analyses Reveal Molecular Regulatory Networks for Pigmentation Deposition in Sheep
by Mancheng Zhang, Xiaoli Xu, Yuan Chen, Chengqi Wei, Siyuan Zhan, Jiaxue Cao, Jiazhong Guo, Dinghui Dai, Linjie Wang, Tao Zhong, Hongping Zhang and Li Li
Int. J. Mol. Sci. 2024, 25(15), 8248; https://doi.org/10.3390/ijms25158248 - 28 Jul 2024
Viewed by 481
Abstract
Domestic animals have multiple phenotypes of skin and coat color, which arise from different genes and their products, such as proteins and metabolites responsible with melanin deposition. However, the complex regulatory network of melanin synthesis remains to be fully unraveled. Here, the skin [...] Read more.
Domestic animals have multiple phenotypes of skin and coat color, which arise from different genes and their products, such as proteins and metabolites responsible with melanin deposition. However, the complex regulatory network of melanin synthesis remains to be fully unraveled. Here, the skin and tongue tissues of Liangshan black sheep (black group) and Liangshan semi-fine-wool sheep (pink group) were collected, stained with hematoxylin–eosin (HE) and Masson–Fontana, and the transcriptomic and metabolomic data were further analyzed. We found a large deposit of melanin granules in the epidermis of the black skin and tongue. Transcriptome and metabolome analysis identified 744 differentially expressed genes (DEGs) and 443 differentially expressed metabolites (DEMs) between the pink and black groups. Gene ontology (GO) and Kyoto encyclopedia of genes and genomes (KEGG) enrichment analyses revealed the DEGs and DEMs were mainly enriched in the pathways of secondary metabolic processes, melanin biosynthesis processes, melanin metabolism processes, melanosome membranes, pigment granule membranes, melanosome, tyrosine metabolism, and melanogenesis. Notably, we revealed the gene ENSARG00020006042 may be a family member of YWHAs and involved in regulating melanin deposition. Furthermore, several essential genes (TYR, TYRP1, DCT, PMEL, MLANA, SLC45A2) were significantly associated with metabolite prostaglandins and compounds involved in sheep pigmentation. These findings provide new evidence of the strong correlation between prostaglandins and related compounds and key genes that regulate sheep melanin synthesis, furthering our understanding of the regulatory mechanisms and molecular breeding of pigmentation in sheep. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Morphological and melanin deposition characteristics of skin and tongue of sheep. (<b>A</b>) The skin stained with HE and Masson–Fontana stains. (<b>B</b>) The tongue stained with HE and Masson–Fontana stains. Scale bars: 100 μm. Ep: epidermal layer, De: dermal layer, Hf: hair follicle, Sg: sebaceous gland, Ml: muscular layer. Purple arrow: melanin. (<b>C</b>) Statistical analysis of skin epidermis and dermis. (<b>D</b>) Statistical analysis of tongue epidermis. (<b>E</b>) Black skin and black tongue melanin diameter size. Section count: 15 fields of view per section were selected for counting (<span class="html-italic">n</span> = 5). In the figure: “ns” indicates no significant difference, “**” indicates <span class="html-italic">p</span> &lt; 0.01, “****” indicates <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 2
<p>DEGs in skin and tongue between pink and black sheep. (<b>A</b>) PCA plot of 4 groups of samples. (<b>B</b>) Violin plot of gene expression in skin and tongue samples. (<b>C</b>) Volcanic map of DEGs in skin group and tongue group. (<b>D</b>) Venn diagram of DEGs in skin group and tongue group. (<b>E</b>) Amino acid sequence comparison between <span class="html-italic">ENSOARG00020006042</span> and <span class="html-italic">YWHAs</span> family. Note: <span class="html-italic">ENSOARP00020028330</span> is the protein ID of <span class="html-italic">ENSOARG00020006042</span>.</p>
Full article ">Figure 3
<p>GO and KEGG enrichment analysis of DEGs associated with sheep pigmentation. (<b>A</b>) Top 20 GO enrichment terms for upregulated and downregulated genes in the skin. (<b>B</b>) Top 20 GO enrichment terms for upregulated and downregulated genes in the tongue. (<b>C</b>) Top 10 KEGG metabolic pathways for upregulated and downregulated genes in the skin and tongue tissues.</p>
Full article ">Figure 4
<p>Pigmentation-related metabolomics of sheep skin and tongue. PCA maps of skin and tongue tissue samples in positive (<b>A</b>) and negative ion modes (<b>B</b>). PLS-da score plots for skin and tongue tissue samples in positive (<b>C</b>) and negative ion mode (<b>D</b>). Volcano plots of DEMs in the Black_vs_Pink_skin (<b>E</b>) and Black_vs_Pink_tongue (<b>F</b>) groups, respectively. (<b>G</b>) Venn diagram of DEMs in the Black_vs_Pink_skin and Black_vs_Pink_tongue groups.</p>
Full article ">Figure 5
<p>KEGG functional enrichment analysis of DEMs. (<b>A</b>) Enrichment bubble plots of KEGG metabolic pathways for DEMs. (<b>B</b>) Heatmap of metabolites enriched in the tyrosine metabolic pathway across samples. (<b>C</b>) Boxplots of the content of several metabolites in different tissues. In the figure: “ns” indicates no significant difference, “*” indicates <span class="html-italic">p</span> &lt; 0.05, “**” indicates <span class="html-italic">p</span> &lt; 0.01, “***” indicates <span class="html-italic">p</span> &lt; 0.001, “****” indicates <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p>Integrative analysis for transcriptome and metabolome. (<b>A</b>) Joint pathway analysis of DEGs and DEMs in skin group and tongue group. (<b>B</b>) Network diagram of genes, metabolites, and metabolic pathways in the skin and tongue groups. (<b>C</b>) Heatmaps of genes and metabolites correlate in the skin group and tongue group.</p>
Full article ">Figure 7
<p>Network of pigmentation-associated genes and metabolites in the skin (<b>A</b>) and tongue group (<b>B</b>). (|r| &gt; 0.75, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>Data analysis flowchart.</p>
Full article ">
15 pages, 15132 KiB  
Article
Ceramide Ehux-C22 Targets the miR-199a-3p/mTOR Signaling Pathway to Regulate Melanosomal Autophagy in Mouse B16 Cells
by Jiyue Wan, Shumiao Zhang, Guiling Li, Shiying Huang, Jian Li, Zhengxiao Zhang and Jingwen Liu
Int. J. Mol. Sci. 2024, 25(15), 8061; https://doi.org/10.3390/ijms25158061 - 24 Jul 2024
Viewed by 411
Abstract
Melanosomes are specialized membrane-bound organelles where melanin is synthesized and stored. The levels of melanin can be effectively reduced by inhibiting melanin synthesis or promoting melanosome degradation via autophagy. Ceramide, a key component in the metabolism of sphingolipids, is crucial for preserving the [...] Read more.
Melanosomes are specialized membrane-bound organelles where melanin is synthesized and stored. The levels of melanin can be effectively reduced by inhibiting melanin synthesis or promoting melanosome degradation via autophagy. Ceramide, a key component in the metabolism of sphingolipids, is crucial for preserving the skin barrier, keeping it hydrated, and warding off the signs of aging. Our preliminary study indicated that a long-chain C22-ceramide compound (Ehux-C22) isolated from the marine microalga Emiliania huxleyi, reduced melanin levels via melanosomal autophagy in B16 cells. Recently, microRNAs (miRNAs) were shown to act as melanogenesis-regulating molecules in melanocytes. However, whether the ceramide Ehux-C22 can induce melanosome autophagy at the post-transcriptional level, and which potential autophagy-dependent mechanisms are involved, remains unknown. Here, miR-199a-3p was screened and identified as a novel upregulated miRNA in Ehux-C22-treated B16 cells. An in vitro high melanin expression model in cultured mouse melanoma cells (B16 cells) was established by using 0.2 μM alpha-melanocyte-stimulating hormone(α-MSH) and used for subsequent analyses. miR-199a-3p overexpression significantly enhanced melanin degradation, as indicated by a reduction in the melanin level and an increase in melanosome autophagy. Further investigation demonstrated that in B16 cells, Ehux-C22 activated miR-199a-3p and inhibited mammalian target of rapamycin(mTOR) level, thus activating the mTOR-ULK1 signaling pathway by promoting the expression of unc-51-like autophagy activating kinase 1 (ULK1), B-cell lymphoma-2 (Bcl-2), Beclin-1, autophagy-related gene 5 (ATG5), and microtubule-associated protein light chain 3 (LC3-II) and degrading p62. Therefore, the roles of Ehux-C22-regulated miR-199a-3p and the mTOR pathway in melanosomal autophagy were elucidated. This research may provide novel perspectives on the post-translational regulation of melanin metabolism, which involves the coordinated control of melanosomes. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Impact of Ehux-C22 on the regulation of autophagy-associated miRNAs and their respective target genes in B16 cells: (<b>A</b>) Analysis of miRNA expression levels; (<b>B</b>) Assessment of target gene mRNA levels. Error bars indicate means ± SD of three biological replicates. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 2
<p>The targeting relationship between miR-199a-3p and the target gene mTOR: (<b>A</b>) The miR-199a-3p target locus was within the 3′UTR of mTOR. Red represented the seed sequence, and green represented the mutant region. (<b>B</b>) The dual-luciferase reporter assay was employed to measure the Luciferase relative activity. Error bars indicate means ± SD of three biological replicates. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>The expression level of miR-199a-3p in B16 cells. After transfection with miR-199a-3p or miR-199a-3p inhibitor for 6 h, B16 cells were incubated in DMEM with 0.2 µM α-MSH, and Ehux-C22 was added according to the group, followed by incubation for 48 h. Error bars indicate means ± SD of three biological replicates. ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>Effect of miR-199a-3p on melanin levels in B16 cells: (<b>A</b>–<b>D</b>) Images were observed by optical microscopy at 100× magnification. (<b>A</b>) Blank control: 0.2 µM α-MSH; (<b>B</b>) 0.2 µM α-MSH+2 µM Ehux-C22; (<b>C</b>) 0.2 µM α-MSH+2 µM Ehux-C22 + miR-199a-3p inhibitor; (<b>D</b>) 0.2 µM α-MSH+miR-199a-3p. (<b>E</b>) The absorbance (490 nm) was measured to calculate the melanin level. Error bars indicate means ± SD of three biological replicates. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>Electron micrograph of melanosomes structure in B16 cells: (<b>A</b>) Blank control: 0.2 µM α-MSH; (<b>B</b>) 0.2 µM α-MSH+2 µM Ehux-C22; (<b>C</b>) 0.2 µM α-MSH + 2 µM Ehux-C22 + miR-199a-3p inhibitor; (<b>D</b>) 0.2 µM α-MSH+miR-199a-3p. The scale bars represent 5.0 µm (<b>A-1</b>–<b>D-1</b>), 1.0 µm (<b>A-2</b>–<b>D-2</b>) and 0.5 µm (<b>A-3</b>–<b>D-3</b>). M-I/II/III/IV: I/II/III/IV stage melanosome; V: vacuole; AS: autolysosome; MC: melanosome complex; N: nucleus.</p>
Full article ">Figure 6
<p>Impact of miR-199a-3p on autophagic flux in B16 cells: (<b>A</b>) Two hours after transfection with mRFP-GFP-LC3 adenovirus, B16 cells were treated with control: 0.2 µM α-MSH, 0.2 µM α-MSH+2 µM Ehux-C22, 0.2 µM α-MSH+miR-199a-3p or 0.2 µM α-MSH+Ehux-C22+miR-199a-3p inhibitor. After 36 h, fluorescent LC3-positive puncta were observed by confocal microscopy. (<b>B</b>) Autolysosomes (red puncta) were quantified as the number of puncta (red puncta/total puncta) in the merged images. (<b>C</b>) Autophagosomes (yellow puncta) were quantified as the number of puncta (yellow puncta/total puncta) in the merged images. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>Impact of miR-199a-3p on the expression of mTOR in B16 cells: (<b>A</b>) Representative band from a Western blot for mTOR; (<b>B</b>) quantitative analysis by ImageJ 1.54d 30. * <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 8
<p>Protein expression levels of key autophagy-related components of the mTOR-ULK1 signaling pathway: (<b>A</b>) Representative Western blot bands for unc-51-like autophagy activating kinase 1 (ULK1), Beclin-1, p-Beclin-1 (Ser15), p-Bcl-2 (Ser70), autophagy-related gene 5 (ATG5), p62, and microtubule-associated protein light chain 3 (LC3-II/I). (<b>B</b>) Relative quantitative analysis of proteins by ImageJ 1.54d 30. * <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.</p>
Full article ">Figure 9
<p>Signaling pathway through which Ehux-C22 regulates miR-199a-3p targeting mTOR to induce melanosome autophagy in B16 cells.</p>
Full article ">
13 pages, 2772 KiB  
Article
Daily Light Onset and Plasma Membrane Tethers Regulate Mitochondria Redistribution within the Retinal Pigment Epithelium
by Matilde V. Neto, Giulia De Rossi, Bruce A. Berkowitz, Miguel C. Seabra, Philip J. Luthert, Clare E. Futter and Thomas Burgoyne
Cells 2024, 13(13), 1100; https://doi.org/10.3390/cells13131100 - 25 Jun 2024
Viewed by 898
Abstract
The retinal pigment epithelium (RPE) is an essential component of the retina that plays multiple roles required to support visual function. These include light onset- and circadian rhythm-dependent tasks, such as daily phagocytosis of photoreceptor outer segments. Mitochondria provide energy to the highly [...] Read more.
The retinal pigment epithelium (RPE) is an essential component of the retina that plays multiple roles required to support visual function. These include light onset- and circadian rhythm-dependent tasks, such as daily phagocytosis of photoreceptor outer segments. Mitochondria provide energy to the highly specialized and energy-dependent RPE. In this study, we examined the positioning of mitochondria and how this is influenced by the onset of light. We identified a population of mitochondria that are tethered to the basal plasma membrane pre- and post-light onset. Following light onset, mitochondria redistributed apically and interacted with melanosomes and phagosomes. In a choroideremia mouse model that has regions of the RPE with disrupted or lost infolding of the plasma membrane, the positionings of only the non-tethered mitochondria were affected. This provides evidence that the tethering of mitochondria to the plasma membrane plays an important role that is maintained under these disease conditions. Our work shows that there are subpopulations of RPE mitochondria based on their positioning after light onset. It is likely they play distinct roles in the RPE that are needed to fulfil the changing cellular demands throughout the day. Full article
(This article belongs to the Section Mitochondria)
Show Figures

Figure 1

Figure 1
<p>Mitochondria are in contact with the plasma membrane at the basal surface and lateral border of the RPE. (<b>A</b>,<b>B</b>) Electron microscopy images of mouse and human RPE with the lateral border shown between the green dotted lines and the basal region between the red dotted line. (<b>C</b>–<b>F</b>) Higher magnification images of the basal surface and lateral border of mouse and human RPE. (<b>C</b>,<b>E</b>) Mitochondria are in contact with the basal infoldings (BIs) of the plasma membrane as shown by the red arrowheads. The basal infoldings are intricate extensions of the plasma membrane surface that are positioned against the Bruch’s membrane (BM). (<b>D</b>,<b>F</b>) Mitochondria also make contact with the plasma membrane at the lateral border of the RPE as shown by the red arrowheads. Scale bars: (<b>A</b>,<b>B</b>) 1 µm, (<b>C</b>–<b>F</b>) 200 nm.</p>
Full article ">Figure 2
<p>Electron tomography resolves tethers between the outer mitochondrial membrane and the plasma membrane at the basal surface of mouse RPE. (<b>A</b>) A slice from a tomogram with zoomed-in images shown in (<b>B</b>–<b>D</b>). (<b>D</b>) The outer mitochondrial membrane is shown in red, the plasma membrane in green, and the tethers in cyan. Scale bars: (<b>A</b>) 100 nm, (<b>B</b>) 50 nm, (<b>C</b>,<b>D</b>) 25 nm.</p>
Full article ">Figure 3
<p>Mitochondria move from the basal surface to the apical side of the RPE after light onset. (<b>A</b>) Electron microscopy image from a mouse eye prepared 1 h after light onset, showing how the RPE can be split into three regions, basal, mid, and apical. These regions are positioned between the basal infolding (BI) and the base of the apical processes (APs) of the RPE. High-magnification panels of the regions indicated by the white arrowheads show the membrane contact sites between mitochondria and the infolded plasma membrane (black arrowheads) at the basal surface. (<b>B</b>–<b>D</b>) Measurements of the proportion of mitochondria within different regions of the RPE before and after lights (n = 3 eyes and &gt;430 mitochondria analyzed per timepoint). After light onset, there is significant repositioning of mitochondria from the basal region through to the mid (at 1 h after light onset) and apical regions (6 h after light onset) of the RPE. (<b>E</b>) When looking at the proportion of mitochondria in contact with the basal surface of the RPE, there was no significant difference between the timepoints. Mean shown with the standard deviation and statistical significance was determined by one-way ANOVA and Tukey’s multiple comparisons test. Scale bars: (<b>A</b>) left panel 600 nm and right panels 200 nm.</p>
Full article ">Figure 4
<p>After light onset, there is increased interaction of mitochondria with melanosomes and phagosomes containing a photoreceptor outer segment (POS). Electron microscopy images showing membrane contact sites (white arrowheads) between (<b>A</b>) mitochondria and melanosomes, (<b>B</b>) POS and mitochondria, and (<b>C</b>) POS and melanosomes. At a timepoint of 6 h after light onset, there are significantly more mitochondria in contact with (<b>A</b>) melanosomes and (<b>B</b>) phagosomes compared to dark-adapted and 1 h after light onset. (<b>C</b>) There was no difference in the percentage of POS in contact with melanosomes at the different timepoints. n = 3 mouse eyes and &gt;310 mitochondria or &gt;110 POS were examined at each timepoint. Mean shown with the standard deviation and statistical significance was determined by one-way ANOVA and Tukey’s multiple comparisons test. Scale bar: 250 nm.</p>
Full article ">Figure 5
<p>Mitochondrial–PM membrane contact sites are not disrupted at the basal RPE in a choroideremia mouse model. (<b>A</b>) Flox control mice present a normal morphology of the basal infoldings. (<b>B</b>) The RPE-specific choroideremia mouse model has regions with basal infoldings that are present (top panel), absent (middle panel), or have basal deposits (shown by red asterisks in the bottom panels). In this model, there is thickening of the Bruch’s membrane (BM), as highlighted by the yellow asterisk. (<b>A</b>,<b>B</b>) In both models, the mitochondria can be seen in contact with the basal plasma membrane, as shown by the black arrowheads. (<b>C</b>–<b>E</b>) When comparing the proportion of mitochondria within different regions of the RPE, there are significantly fewer mitochondria within the basal region of the cell. (<b>F</b>) Between the two models, there is no significant difference in the number of mitochondria in contact with the basal plasma membrane. n = 3 eyes and &gt;650 mitochondria analyzed for each model. The eyes were taken 1 h after light onset. Mean shown with the standard deviation and statistical significance was determined by one-way ANOVA and Tukey’s multiple comparisons test. Scale bars: (<b>A</b>,<b>B</b>) righthand panels 600 nm and lefthand panels 200 nm.</p>
Full article ">
20 pages, 6547 KiB  
Article
Petanin Potentiated JNK Phosphorylation to Negatively Regulate the ERK/CREB/MITF Signaling Pathway for Anti-Melanogenesis in Zebrafish
by Jian Ouyang, Na Hu and Honglun Wang
Int. J. Mol. Sci. 2024, 25(11), 5939; https://doi.org/10.3390/ijms25115939 - 29 May 2024
Viewed by 3127
Abstract
Petanin, an acylated anthocyanin from the Solanaceae family, shows potential in tyrosinase inhibitory activity and anti-melanogenic effects; however, its mechanism remains unclear. Therefore, to investigate the underlying mechanism of petanin’s anti-melanogenic effects, the enzyme activity, protein expression and mRNA transcription of melanogenic and [...] Read more.
Petanin, an acylated anthocyanin from the Solanaceae family, shows potential in tyrosinase inhibitory activity and anti-melanogenic effects; however, its mechanism remains unclear. Therefore, to investigate the underlying mechanism of petanin’s anti-melanogenic effects, the enzyme activity, protein expression and mRNA transcription of melanogenic and related signaling pathways in zebrafish using network pharmacology, molecular docking and molecular dynamics simulation were combined for analysis. The results showed that petanin could inhibit tyrosinase activity and melanogenesis, change the distribution and arrangement of melanocytes and the structure of melanosomes, reduce the activities of catalase (CAT) and peroxidase (POD) and enhance the activity of glutathione reductase (GR). It also up-regulated JNK phosphorylation, inhibited ERK/RSK phosphorylation and down-regulated CREB/MITF-related protein expression and mRNA transcription. These results were consistent with the predictions provided through network pharmacology and molecular docking. Thus, petanin could inhibit the activity of tyrosinase and the expression of tyrosinase by inhibiting and negatively regulating the tyrosinase-related signaling pathway ERK/CREB/MITF through p-JNK. In conclusion, petanin is a good tyrosinase inhibitor and anti-melanin natural compound with significant market prospects in melanogenesis-related diseases and skin whitening cosmetics. Full article
(This article belongs to the Section Bioactives and Nutraceuticals)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The comparison of the melanin (<b>A</b>) and tyrosinase (<b>B</b>) inhibition rate of petanin in zebrafish (NC: water; PC: 11 mM arbutin; PtL: 0.38 mM petanin; PtM: 0.80 mM petanin; PtH: 1.6 mM petanin; *: <span class="html-italic">p</span> &lt; 0.05; **: <span class="html-italic">p</span> &lt; 0.01; ***: <span class="html-italic">p</span> &lt; 0.001; ns: not significant).</p>
Full article ">Figure 2
<p>Typical images of the petanin inhibition of melanin production in zebrafish (NC: water; PC: 11 mM arbutin; PtL: 0.38 mM petanin; PtM: 0.80 mM petanin; PtH: 1.6 mM petanin; Scale: 1×; Zebrafish larvae: incubated for 45 h; Dotted red line: region of melanin signal intensity analysis in zebrafish head).</p>
Full article ">Figure 3
<p>The effect of petanin on the distribution of melanocytes in zebrafish (HE, 20×). (<b>A</b>) NC group; (<b>B</b>) PC group; (<b>C</b>) PtL group; (<b>D</b>) PtM group; (<b>E</b>) PtH group; →: Melanocytes.</p>
Full article ">Figure 4
<p>The effect of petanin on the arrangement of melanocytes in zebrafish (TEM, minimum scale value = 2.0 μm, De: dermis; Me: melanocyte. (<b>A</b>) NC group; (<b>B</b>) PC group; (<b>C</b>) PtL group; (<b>D</b>) PtM group; (<b>E</b>) PtH group).</p>
Full article ">Figure 5
<p>The effect of petanin on the structure of melanosomes in zebrafish (TEM, minimum scale value = 50 nm. (<b>A</b>) NC group; (<b>B</b>) PC group; (<b>C</b>) PtL group; (<b>D</b>) PtM group; (<b>E</b>) PtH group).</p>
Full article ">Figure 6
<p>The effects of different treatments on oxidoreductase activity in zebrafish ((<b>A</b>) CAT (compared to NC group, ***: <span class="html-italic">p</span> &lt; 0.001, **: <span class="html-italic">p</span> &lt; 0.01, *: <span class="html-italic">p</span> &lt; 0.05, ns: not significant), (<b>B</b>) POD (compared to NC group, ***: <span class="html-italic">p</span> &lt; 0.001, *: <span class="html-italic">p</span> &lt; 0.05, ns: not significant), (<b>C</b>) GR (compared to NC group, ***: <span class="html-italic">p</span> &lt; 0.001, **: <span class="html-italic">p</span> &lt; 0.01, *: <span class="html-italic">p</span> &lt; 0.05, ns: not significant)).</p>
Full article ">Figure 7
<p>The network pharmacological analysis of the anti-melanogenesis effect of petanin ((<b>A</b>) important intersection targets; (<b>B</b>) important life processes; (<b>C</b>) key signaling pathways).</p>
Full article ">Figure 8
<p>The molecular docking of petanin with JNK ((<b>A</b>) 3D structure; (<b>B</b>) 3D diagram of hydrogen bonds; (<b>C</b>) 2D diagram of hydrogen bonds).</p>
Full article ">Figure 9
<p>The molecular dynamics’ simulation of petanin–JNK complex ((<b>A</b>) RMSD; (<b>B</b>) RMSF; (<b>C</b>) Rg; (<b>D</b>) H-bonds number; (<b>E</b>) SASA; (<b>F</b>) Gibbs energy landscape).</p>
Full article ">Figure 10
<p>The effects of petanin on the ERK/CREB/MITF signaling pathway ((<b>A</b>) protein expression level; (<b>B</b>) MITF; (<b>C</b>) CREB; (<b>D</b>) p-ERK; (<b>E</b>) p-RSK1; (<b>F</b>) p-JNK; *: <span class="html-italic">p</span> &lt; 0.05; **: <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 11
<p>The effect of petanin on the transcription of mRNA related to melanogenesis in zebrafish (NC: water; PC: 11 mM arbutin; PtL: 0.38 mM petanin; PtM: 0.80 mM petanin; PtH: 1.6 mM petanin; *: <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).</p>
Full article ">
16 pages, 4054 KiB  
Article
N-Acetylneuraminic Acid Inhibits Melanogenesis via Induction of Autophagy
by Kei Yoshikawa and Kazuhisa Maeda
Cosmetics 2024, 11(3), 82; https://doi.org/10.3390/cosmetics11030082 - 21 May 2024
Viewed by 907
Abstract
N-acetylneuraminic acid (Neu5Ac) is the predominant form of sialic acid present in the glossy swiftlet (Collocalia esculenta). It is also the only form of sialic acid detected in the human body. In this study, we investigated the mechanism underlying melanogenesis [...] Read more.
N-acetylneuraminic acid (Neu5Ac) is the predominant form of sialic acid present in the glossy swiftlet (Collocalia esculenta). It is also the only form of sialic acid detected in the human body. In this study, we investigated the mechanism underlying melanogenesis inhibition by Neu5Ac. We discovered that a reduction in tyrosinase protein levels led to an inhibition of melanin production by Neu5Ac. Additionally, the mRNA and protein levels of ubiquitin-specific protease (USP5) and microtubule-associated protein 1 light chain 3 (LC3)-II increased, while those of p62 decreased, indicating enhanced autophagic activity. Lysosomal cathepsin L2 protein levels also increased, and immunostaining revealed colocalization of lysosomal membrane protein (LAMP)-1 and tyrosinase. Additionally, levels of chaperonin containing T-complex polypeptide (CCT), implicated in increased autophagic flux, were elevated. Altogether, these findings suggest that tyrosinase-containing coated vesicles are transported by Neu5Ac into the autophagic degradation pathway, suppressing mature melanosome generation. This process involves increased USP5 levels preventing recognition of polyubiquitin by proteasomes. Furthermore, elevated CCT3 protein levels may enhance autophagic flux, leading to the incorporation of tyrosinase-containing coated vesicles into autophagosomes. These autophagosomes then fuse with lysosomes for cathepsin L2–mediated degradation. Thus, our findings suggest that Neu5Ac reduces tyrosinase activity and inhibits melanosome maturation by promoting selective autophagic degradation of abnormal proteins by p62. Full article
(This article belongs to the Special Issue 10th Anniversary of Cosmetics—Recent Advances and Perspectives)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of <span class="html-italic">N</span>-acetylneuraminic acid (Neu5Ac).</p>
Full article ">Figure 2
<p>Effect of Neu5Ac on cell proliferation and tyrosinase and melanin levels in B16 melanoma cells. (<b>a</b>) Effect of Neu5Ac on cell number and tyrosinase activity. (<b>b</b>) Effect of Neu5Ac on tyrosinase mRNA and protein levels. Mean ± SD, <span class="html-italic">n</span> = 4 or <span class="html-italic">n</span> = 3, *: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01. (<b>c</b>) Effect of Neu5Ac on melanin levels.</p>
Full article ">Figure 2 Cont.
<p>Effect of Neu5Ac on cell proliferation and tyrosinase and melanin levels in B16 melanoma cells. (<b>a</b>) Effect of Neu5Ac on cell number and tyrosinase activity. (<b>b</b>) Effect of Neu5Ac on tyrosinase mRNA and protein levels. Mean ± SD, <span class="html-italic">n</span> = 4 or <span class="html-italic">n</span> = 3, *: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01. (<b>c</b>) Effect of Neu5Ac on melanin levels.</p>
Full article ">Figure 3
<p>Effect of Neu5Ac on tyrosinase and LAMP-1 localization, using fluorescence immunostaining (multiple staining). Nuclei were stained with DAPI (blue), tyrosinase was labeled with an anti-tyrosinase antibody, and Alexa Fluor 488 secondary antibody (green), and LAMP-1 was labeled with an anti-LAMP1 antibody and Alexa Fluor 568 secondary antibody (red). Tyrosinase and LAMP-1 colocalization areas are indicated in yellow. The scale bar is 100 μm.</p>
Full article ">Figure 4
<p>Effect of Neu5Ac on cathepsin L2 levels, using fluorescence immunostaining (multiple staining method). The nuclei were stained with DAPI (blue). Cathepsin L2 was labeled with an anti-cathepsin L2 antibody and Alexa Fluor 488 secondary antibody (green). The scale bar is 100 μm.</p>
Full article ">Figure 5
<p>(<b>a</b>) Effect of Neu5Ac on LC3. (<b>b</b>) LC3-II/β-actin ratio was calculated by measuring the brightness of LC3-II and β-actin bands with ImageJ. mean ± SD, <span class="html-italic">n</span> = 3. (<b>c</b>) Effect of Neu5Ac on p62. (<b>d</b>) The brightness of p62 and β-actin bands was measured with ImageJ. The p62/β-actin ratio was calculated and plotted. mean ± SD, <span class="html-italic">n</span> = 3, *: <span class="html-italic">p</span> &lt; 0.05 vs. control, **: <span class="html-italic">p</span> &lt; 0.01 vs. control.</p>
Full article ">Figure 5 Cont.
<p>(<b>a</b>) Effect of Neu5Ac on LC3. (<b>b</b>) LC3-II/β-actin ratio was calculated by measuring the brightness of LC3-II and β-actin bands with ImageJ. mean ± SD, <span class="html-italic">n</span> = 3. (<b>c</b>) Effect of Neu5Ac on p62. (<b>d</b>) The brightness of p62 and β-actin bands was measured with ImageJ. The p62/β-actin ratio was calculated and plotted. mean ± SD, <span class="html-italic">n</span> = 3, *: <span class="html-italic">p</span> &lt; 0.05 vs. control, **: <span class="html-italic">p</span> &lt; 0.01 vs. control.</p>
Full article ">Figure 6
<p>Effect of Neu5Ac on p62 localization, using fluorescence immunostaining. The nuclei were stained with DAPI (blue). P62 was labeled with an anti-p62/SQSTM1antibody and Alexa Fluor 568 secondary antibody (red). The scale bar was 50 μm.</p>
Full article ">Figure 6 Cont.
<p>Effect of Neu5Ac on p62 localization, using fluorescence immunostaining. The nuclei were stained with DAPI (blue). P62 was labeled with an anti-p62/SQSTM1antibody and Alexa Fluor 568 secondary antibody (red). The scale bar was 50 μm.</p>
Full article ">Figure 7
<p>Effect of Neu5Ac on mRNA (<b>a</b>) and protein levels (<b>b</b>–<b>d</b>) of the ubiquitin-proteasome system-related factors USP5 and UBE1Y. Mean ± SD, <span class="html-italic">n</span> = 3, *: <span class="html-italic">p</span> &lt; 0.05 vs. control, **: <span class="html-italic">p</span> &lt; 0.01 vs. control.</p>
Full article ">Figure 7 Cont.
<p>Effect of Neu5Ac on mRNA (<b>a</b>) and protein levels (<b>b</b>–<b>d</b>) of the ubiquitin-proteasome system-related factors USP5 and UBE1Y. Mean ± SD, <span class="html-italic">n</span> = 3, *: <span class="html-italic">p</span> &lt; 0.05 vs. control, **: <span class="html-italic">p</span> &lt; 0.01 vs. control.</p>
Full article ">Figure 8
<p>Effect of Neu5Ac on mRNA levels (<b>a</b>) of molecular chaperone CCT and protein levels of CCT3 (<b>b</b>,<b>c</b>). Mean ± SD, <span class="html-italic">n</span> = 3, *: <span class="html-italic">p</span> &lt; 0.05 vs. control, **: <span class="html-italic">p</span> &lt; 0.01 vs. control.</p>
Full article ">Figure 8 Cont.
<p>Effect of Neu5Ac on mRNA levels (<b>a</b>) of molecular chaperone CCT and protein levels of CCT3 (<b>b</b>,<b>c</b>). Mean ± SD, <span class="html-italic">n</span> = 3, *: <span class="html-italic">p</span> &lt; 0.05 vs. control, **: <span class="html-italic">p</span> &lt; 0.01 vs. control.</p>
Full article ">
14 pages, 964 KiB  
Review
Role of Dermal Factors Involved in Regulating the Melanin and Melanogenesis of Mammalian Melanocytes in Normal and Abnormal Skin
by Tomohisa Hirobe
Int. J. Mol. Sci. 2024, 25(8), 4560; https://doi.org/10.3390/ijms25084560 - 22 Apr 2024
Viewed by 881
Abstract
Mammalian melanin is produced in melanocytes and accumulated in melanosomes. Melanogenesis is supported by many factors derived from the surrounding tissue environment, such as the epidermis, dermis, and subcutaneous tissue, in addition to numerous melanogenesis-related genes. The roles of these genes have been [...] Read more.
Mammalian melanin is produced in melanocytes and accumulated in melanosomes. Melanogenesis is supported by many factors derived from the surrounding tissue environment, such as the epidermis, dermis, and subcutaneous tissue, in addition to numerous melanogenesis-related genes. The roles of these genes have been fully investigated and the molecular analysis has been performed. Moreover, the role of paracrine factors derived from epidermis has also been studied. However, the role of dermis has not been fully studied. Thus, in this review, dermis-derived factors including soluble and insoluble components were overviewed and discussed in normal and abnormal circumstances. Dermal factors play an important role in the regulation of melanogenesis in the normal and abnormal mammalian skin. Full article
(This article belongs to the Special Issue Melanins and Melanogenesis 4.0: From Nature to Applications)
Show Figures

Figure 1

Figure 1
<p>Hypothesis of the mechanism of action of dermis-derived factors on the melanogenesis of mammalian epidermal melanocytes. Abbreviations: TYR, tyrosinase; TRP-1, TYR-related protein-1; DCT, dopachrome tautomerase; MITF, microphthalmia-associated transcription factor; MAPK, MAP kinase; PKA, protein kinase A; PKC, protein kinase C; GSK3β, glycogen synthase kinase-3β; ERK 1/2, extracellular signal-regulated kinase 1/2; p38 ERK 1/2, p38 extracellular signal-regulated kinase 1/2; ATP, adenosine triphosphate; cAMP, cyclic 3′,5′ adenosine monophosphate; TβRI/II, receptor for TGFβ-1 (transforming growth factor-β1); LRP6 Frizzled, receptor for DKK-1 (Dickkopf-related protein-1) and WIF-1 (Wnt inhibitory factor-1); PTRβ/ζ, receptor for PTN (pleiotrophin)/protein tyrosine phosphatase; FGFR, receptor for bFGF (basic fibroblast growth factor); MC1R, melanocortin 1 receptor/receptor for α-MSH (melanocyte stimulating hormone); ETBR, receptor for ET-1 (endothelin-1); c-Kit, receptor for SCF (stem cell factor); c-Met, receptor for HGF (hepatocyte growth factor); IL-1RI, receptor for IL-1α (interleukin-1α); Integrin α6β1, receptor for CCN-1 (cellular communication network factor-1); KGFR, receptor for KGF (keratinocyte growth factor); ERBB-3, v-erb-b2 avian erythroblastic leukemia viral oncogene homolog/receptor for sFRP-2 (secreted frizzled-related protein-2); NRG-1, neuregulin-1; VEGFR, receptor for VEGF (vascular endothelial growth factor); PDGFR, receptor for PDGF (platelet-derived growth factor); EBP, elastin binding protein; EF, elastin fiber; EP, elastin peptide; ↑, increase; and ↓, decrease.</p>
Full article ">
15 pages, 1420 KiB  
Review
Retinal Pigment Epithelium Pigment Granules: Norms, Age Relations and Pathology
by Alexander Dontsov and Mikhail Ostrovsky
Int. J. Mol. Sci. 2024, 25(7), 3609; https://doi.org/10.3390/ijms25073609 - 23 Mar 2024
Viewed by 1565
Abstract
The retinal pigment epithelium (RPE), which ensures the normal functioning of the neural retina, is a pigmented single-cell layer that separates the retina from the Bruch’s membrane and the choroid. There are three main types of pigment granules in the RPE cells of [...] Read more.
The retinal pigment epithelium (RPE), which ensures the normal functioning of the neural retina, is a pigmented single-cell layer that separates the retina from the Bruch’s membrane and the choroid. There are three main types of pigment granules in the RPE cells of the human eye: lipofuscin granules (LG) containing the fluorescent “age pigment” lipofuscin, melanoprotein granules (melanosomes, melanolysosomes) containing the screening pigment melanin and complex melanolipofuscin granules (MLG) containing both types of pigments simultaneously—melanin and lipofuscin. This review examines the functional role of pigment granules in the aging process and in the development of oxidative stress and associated pathologies in RPE cells. The focus is on the process of light-induced oxidative degradation of pigment granules caused by reactive oxygen species. The reasons leading to increased oxidative stress in RPE cells as a result of the oxidative degradation of pigment granules are considered. A mechanism is proposed to explain the phenomenon of age-related decline in melanin content in RPE cells. The essence of the mechanism is that when the lipofuscin part of the melanolipofuscin granule is exposed to light, reactive oxygen species are formed, which destroy the melanin part. As more melanolipofuscin granules are formed with age and the development of degenerative diseases, the melanin in pigmented epithelial cells ultimately disappears. Full article
(This article belongs to the Collection Feature Papers in “Molecular Biology”)
Show Figures

Figure 1

Figure 1
<p>Scheme of the structure of the fundus. RPE—retinal pigment epithelium; IPM—interphotoreceptor matrix [<a href="#B8-ijms-25-03609" class="html-bibr">8</a>].</p>
Full article ">Figure 2
<p>Schematic illustration of lipofuscin and bisretinoid’s roles in the development of photooxidative stress in the RPE cell.</p>
Full article ">Figure 3
<p>MLG from human RPE cells contain less melanin than MG. ESR spectra of MG (<b>A</b>) and MLG (<b>B</b>) from human RPE cells.</p>
Full article ">Figure 4
<p>Scheme of the mechanisms involved in melanin degradation in the melanolipofuscin granule. Abbreviations: PMD—melanin degradation products, ROS—reactive oxygen species, LIP-ox—oxidized lipofuscin. (<b>1.</b>) Light in the presence of oxygen activates ROS generation mediated by lipofuscin fluorophores. The resulting ROS can oxidize both melanin, causing its degradation and the formation of PMD, and lipofuscin, causing the formation of reactive dicarbonyls. (<b>2.</b>) The resulting photosensitive melanin degradation products (PMDs) generate ROS when exposed to light and, in turn, cause the further degradation of melanin and lipofuscin. (<b>3.</b>) In a granule containing all three components, namely melanin, lipofuscin (bisretinoids) and PMD, light and ROS activate the transition of melanin to a high-energy state (melanin*) in which the excited pigment causes the degradation of lipofuscin [<a href="#B164-ijms-25-03609" class="html-bibr">164</a>,<a href="#B165-ijms-25-03609" class="html-bibr">165</a>].</p>
Full article ">
14 pages, 3409 KiB  
Article
Molecular Modeling of the Multiple-Substrate Activity of the Human Recombinant Intra-Melanosomal Domain of Tyrosinase and Its OCA1B-Related Mutant Variant P406L
by Monika B. Dolinska and Yuri V. Sergeev
Int. J. Mol. Sci. 2024, 25(6), 3373; https://doi.org/10.3390/ijms25063373 - 16 Mar 2024
Viewed by 730
Abstract
Tyrosinase serves as the key enzyme in melanin biosynthesis, catalyzing the initial steps of the pathway, the hydroxylation of the amino acid L-tyrosine into L-3,4-dihydroxyphenylalanine (L-DOPA), followed by the subsequent oxidation of L-DOPA into dopaquinone (DQ), and it facilitates the conversion of 5,6-dihydroxyindole-2-carboxylic [...] Read more.
Tyrosinase serves as the key enzyme in melanin biosynthesis, catalyzing the initial steps of the pathway, the hydroxylation of the amino acid L-tyrosine into L-3,4-dihydroxyphenylalanine (L-DOPA), followed by the subsequent oxidation of L-DOPA into dopaquinone (DQ), and it facilitates the conversion of 5,6-dihydroxyindole-2-carboxylic acid (DHICA) into 5,6-indolequinone-2-carboxylic acid (IQCA) and 5,6-dihydroxy indole (DHI) into indolequinone (IQ). Despite its versatile substrate capabilities, the precise mechanism underlying tyrosinase’s multi-substrate activity remains unclear. Previously, we expressed, purified, and characterized the recombinant intra-melanosomal domain of human tyrosinase (rTyr). Here, we demonstrate that rTyr mimics native human tyrosinase’s catalytic activities in vitro and in silico. Molecular docking and molecular dynamics (MD) simulations, based on rTyr’s homology model, reveal variable durability and binding preferences among tyrosinase substrates and products. Analysis of root mean square deviation (RMSD) highlights the significance of conserved residues (E203, K334, F347, and V377), which exhibit flexibility during the ligands’ binding. Additionally, in silico analysis demonstrated that the OCA1B-related P406L mutation in tyrosinase substantially influences substrate binding, as evidenced by the decreased number of stable ligand conformations. This correlation underscores the mutation’s impact on substrate docking, which aligns with the observed reduction in rTyr activity. Our study highlights how rTyr dynamically adjusts its structure to accommodate diverse substrates and suggests a way to modulate rTyr ligand plasticity. Full article
Show Figures

Figure 1

Figure 1
<p>Multi-substrate enzymatic reactions of rTyr. (<b>A</b>) Reaction of monophenol and diphenol oxidase activity of Tyr (Top Panel) and enzymatic activity of rTyr (bottom panels). Spectra represent the absorbance profile recorded for wavelengths from 200 to 900 nm at time 0 (black) and after incubation at 37 °C with L-tyrosine (left graph) and L-DOPA (right graph) (blue). Arrows indicate the maximum absorbance of L-tyrosine/L-DOPA (λ<sub>max</sub>~280 nm), cyclodopa (λ<sub>max</sub> 310 nm), and dopachrome (λ<sub>max</sub> 475 nm). (<b>B</b>) DHICA oxidase activity of rTyr. Spectra represent the absorbance profile recorded at wavelengths from 200 to 900 nm at time 0 (black) and after incubation at 37 °C with DHICA (blue). Arrows indicate the maximum absorbance of DHICA (λ<sub>max</sub> 325 nm) and IQCA (λ<sub>max</sub> 560 nm). (<b>C</b>) DHI oxidase activity of rTyr. Spectra represent the absorbance profile recorded for wavelengths from 200 to 900 nm at time 0 (black) and after incubation at 37 °C with DHI (blue). Arrows indicate the maximum absorbance of DHI (λ<sub>max</sub> 305 nm) and IQ (λ<sub>max</sub> 327 nm). All small molecules were visualized using USCF Chimera. Nitrogen atoms are colored blue, oxygen atoms are colored red, hydrogen atoms are colored white, and carbon atoms are colored tan.</p>
Full article ">Figure 2
<p>Molecular docking of the substrates and products of the enzymatic reaction catalyzed by the rTyr and P406L mutant variant (<b>A</b>) Number of complex formations (total) found after clustering the 25 docking runs of the ligands—substrates (solid bars) and products (stripped bars)—to rTyr (blue) and P406L (red). (<b>B</b>) shows the number of ligands bound to the active site in a distance below 4.0 Å from the CuA or CuB atoms. (<b>C</b>) shows the number of ligands of which the oxygen of the hydroxyl or carbonyl group is directed toward the active site (proper orientation). (<b>D</b>) shows the number of all properly docked structures that remain in optimal alignment when measured throughout 20 ns of the MD simulations.</p>
Full article ">Figure 3
<p>Computational docking of the small-molecule substrates to rTyr. L-tyrosine (green sticks (<b>A</b>)), L-DOPA (purple sticks (<b>B</b>)), DHICA (green sea sticks (<b>C</b>)), and DHI (gray sticks (<b>D</b>)) were docked to the rTyr active site. Hydrogen bonds, hydrophobic contacts, and π–π interactions are shown as blue, green, and red lines, respectively. Other contacts for all molecules are shown as purple dashed lines. Residues making any interactions with the docking molecule are shown as light gray sticks. The distances between the docked molecule and the Cu atoms, shown as orange spheres, are colored yellow. Gray boxes show the values of binding energy (ΔG<sub>bind</sub>), dissociation constant (K<sub>D</sub>), and molecular contact surface (CS) between the ligand and receptor. Inserts (<b>a</b>) show how the ligand settled during the MD simulation at 0 ns (beige), 1 ns (blue), 10 ns (pink), and 20 ns (green). Inserts (<b>b</b>) show the docking of ligands with the surface.</p>
Full article ">Figure 4
<p>Flexibility of the rTyr/substrates contacting residues after MD simulations. (<b>A</b>) RMSD (Å) calculated for the residues involved in docking of all substrates to rTyr after 20 ns of MD simulations. Each dot represents a separate docking action. Histidine-coordinating coppers in the active site are shown in gray. RMSD values &lt; 2 Å (below the black dashed line) are considered stable. The most flexible residues are in the yellow (E203), green (K334), red (F347), and blue (V377) frames. (<b>B</b>) Alignment of tyrosinase proteins from fungus, chicken, mouse, human, and bovine. The most flexible residues after 20 ns of MD simulations are shown as yellow (E203), green (K334), red (F347), and blue (V377) bars. (<b>C</b>) The homology model of the rTyr domain shows the most flexible residues of E203 (yellow), K334 (green), F347 (red), and V377 (blue) (<b>a</b>). Two copper atoms in the active site, CuA and CuB, which are coordinated by histidines (gray), are shown in orange. The protein backbone structure is shown as a tan ribbon. Panels (<b>b</b>–<b>e</b>) show the movement of E203 (<b>b</b>), K334 (<b>c</b>), F347 (<b>d</b>), and V377 (<b>e</b>) residues after 20 ns of MD simulations. Residues in tan indicate the intact rTyr, and blue indicates the supposition of rTyr obtained after docking and 20 ns MD run.</p>
Full article ">
13 pages, 2129 KiB  
Article
Evaluation of Teneligliptin and Retagliptin on the Clearance of Melanosome by Melanophagy in B16F1 Cells
by Seong Hyun Kim, Ji-Eun Bae, Na Yeon Park, Joon Bum Kim, Yong Hwan Kim, So Hyun Kim, Gyeong Seok Oh, Hee Won Wang, Jeong Ho Chang and Dong-Hyung Cho
Cosmetics 2024, 11(2), 35; https://doi.org/10.3390/cosmetics11020035 - 1 Mar 2024
Viewed by 2052
Abstract
A specialized membrane-bound organelle, named the melanosome, is central to the storage and transport of melanin as well as melanin synthesis in melanocytes. Although previous studies have linked melanosomal degradation to autophagy, the precise mechanisms remain elusive. Autophagy, a complex catabolic process involving [...] Read more.
A specialized membrane-bound organelle, named the melanosome, is central to the storage and transport of melanin as well as melanin synthesis in melanocytes. Although previous studies have linked melanosomal degradation to autophagy, the precise mechanisms remain elusive. Autophagy, a complex catabolic process involving autophagosomes and lysosomes, plays a vital role in cellular constituent degradation. In this study, the role of autophagy in melanosomal degradation was explored, employing a cell-based screening system designed to unveil key pathway regulators. We identified specific dipeptidyl peptidase-4 inhibitors, such as teneligliptin hydrobromide and retagliptin phosphate, as novel agents inducing melanophagy through a comprehensive screening of a ubiquitination-related chemical library. We found that treatment with teneligliptin hydrobromide or retagliptin phosphate not only diminishes melanin content elevated by alpha-melanocyte-stimulating hormone (α-MSH) but also triggers autophagy activation within B16F1 cells. In addition, the targeted inhibition of unc-51-like kinase (ULK1) significantly attenuated both the anti-pigmentation effects and autophagy induced by teneligliptin hydrobromide and retagliptin phosphate in α-MSH-treated cells. Collectively, our data demonstrate a new frontier in understanding melanosomal degradation, identifying teneligliptin hydrobromide and retagliptin phosphate as promising inducers of melanophagy via autophagy activation. This study contributes essential insights into cellular degradation mechanisms and offers potential therapeutic avenues in the regulation of pigmentation. Full article
(This article belongs to the Collection Feature Papers in Cosmetics in 2023)
Show Figures

Figure 1

Figure 1
<p>Teneligliptin hydrobromide and retagliptin phosphate induce autophagy through TFEB translocation: (<b>A</b>,<b>B</b>) B16F1/GFP-LC3 cells were treated with either teneligliptin hydrobromide (TGN, 10, 100 μM), retagliptin phosphate (RGN, 10, 100 μM), or 3,4,5-trimethoxy cinnamate thymol ester (TCTE, 10 μg/mL). (<b>A</b>) Then, 24 h later, the treated cells were fixed to be imaged for green fluorescence. Cells in which autophagy was induced were analyzed by counting GFP-LC3 punctate dots under a confocal microscope. (<b>B</b>) The protein expression of LC3 was then examined by Western blotting. (<b>C</b>) B16F1 cells were treated with teneligliptin hydrobromide (TGN, 100 μM) or retagliptin phosphate (RGN, 100 μM), with or without bafilomycin A1 (Baf A1, 100 nM), for 6 h. The LC3 level was then examined by Western blotting. (<b>D</b>) B16F1/GFP-TFEB cells were exposed to teneligliptin hydrobromide (TGN, 100 μM) or retagliptin phosphate (RGN, 100 μM) for 24 h or Torin1 (0.25 μM) for 1 h. The cells were fixed for fluorescence imaging, and the nuclear translocalization of TFEB was assessed. The scale bar indicates 20 μm. (n = 3, ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 2
<p>Teneligliptin hydrobromide and retagliptin phosphate induce anti-pigmentation in B16F1 cells treated with α-MSH: (<b>A</b>) B16F1 cells were treated with teneligliptin hydrobromide (TGN, 10, 100 μM) or retagliptin phosphate (RGN, 10, 100 μM) for 24 h, then cell viability was determined by CCK-8 assay. (<b>B</b>) B16F1 cells were pre-treated with alpha-melanocyte-stimulating hormone (α-MSH, 0.5 μM) for 36 h and then additionally exposed to teneligliptin hydrobromide (TGN, 10, 100 μM), retagliptin phosphate (RGN, 10, 100 μM) or 3,4,5-trimethoxy cinnamate thymol ester (TCTE, 10 μg/mL) for 24 h. (<b>B</b>) Melanin content is shown in cell pellets. (<b>C</b>) The melanin content was measured by assessing absorbance at 405 nm through a microplate reader, as outlined in the Materials and Methods Section. (n = 3, * <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>Teneligliptin hydrobromide and retagliptin phosphate induce melanosomal degradation via autophagy: (<b>A</b>) First, 0.5 μM α-MSH-treated B16F1/TPC2-mRFP-EGFP cells were exposed to teneligliptin hydrobromide (TGN, 100 μM) or retagliptin phosphate (RGN, 100 μM) for 18 h, with or without 100nM bafilomycin (Baf A1) for 6 h. After fixation, the distribution of TPC2-mRFP-EGFP in the cells was imaged with confocal microscopy. (<b>B</b>) B16F1/TPC2-mRFP-EGFP cells were transfected with non-specific control siRNA (Sc) or siRNA targeting Atg5 (siAtg5). After 24 h transfection, the cells were further treated with 0.5 μM α-MSH for 36 h and then incubated with teneligliptin hydrobromide (TGN, 100 μM) or retagliptin phosphate (RGN, 100 μM). The number of RFP-only signals per cells was analyzed with merged images. The protein expression of Atg5 was then assessed by Western blotting. n = 3 and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>Inhibition of autophagy reverses the whitening effects of teneligliptin hydrobromide and retagliptin phosphate: (<b>A</b>,<b>B</b>) B16F1 cells were pre-treated with 0.5 μM α-MSH for 12 h and then the cells were incubated in the presence or absence of SBI-0206965 (SBI, 5 µM) for 24 h and additionally exposed to teneligliptin hydrobromide (TGN, 100 μM) or retagliptin phosphate (RGN, 100 μM) for 24 h. (<b>A</b>) Melanin content is shown in cell pellets. (<b>B</b>) Content in the cells was quantified by addressing absorbance at 405 nm using a microplate reader, as detailed in the Materials and Methods Section. (n = 3 and ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 5
<p>Inhibition of autophagy decreases melanosomal degradation by teneligliptin hydrobromide and retagliptin phosphate: (<b>A</b>,<b>B</b>) B16F1 cells or B16F1/TPC2-mRFP-EGFP cells were pre-treated with 0.5 μM α-MSH for 12 h. And the cells were further incubated with or without SBI-0206965 (SBI, 5 µM) for 24 h and additionally exposed to teneligliptin hydrobromide (TGN, 100 μM) or retagliptin phosphate (RGN, 100 μM) for an additional 24 h. (<b>A</b>) The protein expression of LC3 was then examined by Western blotting. (<b>B</b>) After fixation, the distribution of TPC2-mRFP-EGFP was imaged under a confocal microscope. The number of RFP-positive signal dots per cells was quantified from the merged cell images. The scale bar indicates 20 μm. (n = 3, * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">
14 pages, 1207 KiB  
Article
Runs of Homozygosity Detection and Selection Signature Analysis for Local Goat Breeds in Yunnan, China
by Chang Huang, Qian Zhao, Qian Chen, Yinxiao Su, Yuehui Ma, Shaohui Ye and Qianjun Zhao
Genes 2024, 15(3), 313; https://doi.org/10.3390/genes15030313 - 28 Feb 2024
Viewed by 1345
Abstract
Runs of Homozygosity (ROH) are continuous homozygous DNA segments in diploid genomes, which have been used to estimate the genetic diversity, inbreeding levels, and genes associated with specific traits in livestock. In this study, we analyzed the resequencing data from 10 local goat [...] Read more.
Runs of Homozygosity (ROH) are continuous homozygous DNA segments in diploid genomes, which have been used to estimate the genetic diversity, inbreeding levels, and genes associated with specific traits in livestock. In this study, we analyzed the resequencing data from 10 local goat breeds in Yunnan province of China and five additional goat populations obtained from a public database. The ROH analysis revealed 21,029 ROH segments across the 15 populations, with an average length of 1.27 Mb, a pattern of ROH, and the assessment of the inbreeding coefficient indicating genetic diversity and varying levels of inbreeding. iHS (integrated haplotype score) was used to analyze high-frequency Single-Nucleotide Polymorphisms (SNPs) in ROH regions, specific genes related to economic traits such as coat color and weight variation. These candidate genes include OCA2 (OCA2 melanosomal transmembrane protein) and MLPH (melanophilin) associated with coat color, EPHA6 (EPH receptor A6) involved in litter size, CDKAL1 (CDK5 regulatory subunit associated protein 1 like 1) and POMC (proopiomelanocortin) linked to weight variation and some putative genes associated with high-altitude adaptability and immune. This study uncovers genetic diversity and inbreeding levels within local goat breeds in Yunnan province, China. The identification of specific genes associated with economic traits and adaptability provides actionable insights for utilization and conservation efforts. Full article
(This article belongs to the Section Animal Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>The distributions of ROH statistics per individual for 15 goat populations, including ANG (n = 8), BER (n = 5), BEZ (n = 5), BLB (n = 6), FQ (n = 8), GS (n = 10), LL (n = 10), LP (n = 10), LSD (n = 4), MG (n = 10), ML (n = 6), NL (n = 10), WX (n = 10), YL (n = 10), and ZT (n = 10). (<b>A</b>) The length of ROHs per individual. (<b>B</b>) The number of ROHs per individual.</p>
Full article ">Figure 2
<p>Values of <span class="html-italic">F<sub>ROH</sub></span> for each population.</p>
Full article ">Figure 3
<p>Total number of ROH in 15 goat populations. (<b>A</b>) Number of ROH in chromosome with different size classes. (<b>B</b>) The number of ROH belonging to four size classes, including 500 kb to 1 Mb, 1 Mb to 2 Mb, 2 Mb to 4 Mb, and &gt;4 Mb for each of the different populations.</p>
Full article ">Figure 4
<p>The loci with strong selection signals, as identified by the iHS, were annotated to specific genes. (<b>A</b>) <span class="html-italic">EPHA6</span> gene in GS breed, (<b>B</b>) <span class="html-italic">OCA2</span> gene in ML breed, (<b>C</b>) <span class="html-italic">MLPH</span> gene in GS breed, (<b>D</b>) <span class="html-italic">CDKAL1</span> gene in FQ breed. The dashed line represents the top 5% SNPs with high frequency in the local goat breed.</p>
Full article ">
16 pages, 4363 KiB  
Article
Membrane-Associated Ubiquitin Ligase RING Finger Protein 152 Orchestrates Melanogenesis via Tyrosinase Ubiquitination
by Ryota Ueda, Rina Hashimoto, Yuki Fujii, José C. J. M. D. S. Menezes, Hirotaka Takahashi, Hiroyuki Takeda, Tatsuya Sawasaki, Tomonori Motokawa, Kenzo Tokunaga and Hideaki Fujita
Membranes 2024, 14(2), 43; https://doi.org/10.3390/membranes14020043 - 1 Feb 2024
Viewed by 2446
Abstract
Lysosomal degradation of tyrosinase, a pivotal enzyme in melanin synthesis, negatively impacts melanogenesis in melanocytes. Nevertheless, the precise molecular mechanisms by which lysosomes target tyrosinase have remained elusive. Here, we identify RING (Really Interesting New Gene) finger protein 152 (RNF152) as a membrane-associated [...] Read more.
Lysosomal degradation of tyrosinase, a pivotal enzyme in melanin synthesis, negatively impacts melanogenesis in melanocytes. Nevertheless, the precise molecular mechanisms by which lysosomes target tyrosinase have remained elusive. Here, we identify RING (Really Interesting New Gene) finger protein 152 (RNF152) as a membrane-associated ubiquitin ligase specifically targeting tyrosinase for the first time, utilizing AlphaScreen technology. We observed that modulating RNF152 levels in B16 cells, either via overexpression or siRNA knockdown, resulted in decreased or increased levels of both tyrosinase and melanin, respectively. Notably, RNF152 and tyrosinase co-localized at the trans-Golgi network (TGN). However, upon treatment with lysosomal inhibitors, both proteins appeared in the lysosomes, indicating that tyrosinase undergoes RNF152-mediated lysosomal degradation. Through ubiquitination assays, we found the indispensable roles of both the RING and transmembrane (TM) domains of RNF152 in facilitating tyrosinase ubiquitination. In summary, our findings underscore RNF152 as a tyrosinase-specific ubiquitin ligase essential for regulating melanogenesis in melanocytes. Full article
(This article belongs to the Section Biological Membranes)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Tyrosinase undergoes ubiquitination and degradation in lysosomes in B16 melanoma cells. (<b>A</b>,<b>B</b>) Cells were treated with either DMSO (control, Ctrl) or LPIs (leupeptin, pepstatin A, E64d; each 40 μM) for 24 h. Equal amounts of cell lysates, quantified using a commercial assay kit, were subjected to immunoblotting with either an anti-tyrosinase or anti-Tyrp-1 antibody. Additionally, lysates were immunoprecipitated with an anti-ubiquitin antibody (IP: anti-Ub) followed by immunoblotting using the anti-tyrosinase or anti-Tyrp-1 antibody. The expression levels of tyrosinase were quantified in the bands and presented as fold expression relative to the Ctrl (mean ± s.d., <span class="html-italic">n</span> = 3). *** <span class="html-italic">p</span> ≤ 0.001 compared with the Ctrl using two-tailed unpaired <span class="html-italic">t</span>-tests. ns, not significant. (<b>C</b>) Cells were treated with DMSO (control, Ctrl), LPIs, inulavosin (Inu; 15 μM), inulavosin combined with LP (Inu/LP), PTU (PTU; 100 μM), or PTU combined with LP (PTU/LP) for 24 h. Lysates were immunoprecipitated with the anti-ubiquitin antibody followed by immunoblotting with the anti-tyrosinase antibody (“IP: anti-Ub Blot: anti-tyrosinase”), or directly immunoblotted with the anti-tyrosinase or an anti-β-actin antibody. An asterisk indicates the presumed position of the mouse immunoglobulin G heavy chain.</p>
Full article ">Figure 2
<p>RNF152 modulation influences tyrosinase expression in B16 cells. (<b>A</b>) Cells were transfected with either an empty vector (Ctrl) or RNF152-myc. Lysates were immunoblotted with anti-tyrosinase, anti-Tyrp-1, anti-β-actin, and anti-myc antibodies. (<b>B</b>,<b>C</b>) The expression levels of Tyrp-1 (<b>B</b>) and tyrosinase (<b>C</b>) were quantified in the bands and presented as fold expression relative to the Ctrl (mean ± s.d., <span class="html-italic">n</span> = 3). ** <span class="html-italic">p</span> ≤ 0.01 compared with the Ctrl using two-tailed unpaired <span class="html-italic">t</span>-tests. ns, not significant. (<b>D</b>) Cells were transfected with either an empty vector (Ctrl) or RNF152-myc, and melanin content was measured using spectrophotometry. The data are shown as the percentage relative to the Ctrl (mean ± s.d., <span class="html-italic">n</span> = 3). ** <span class="html-italic">p</span> ≤ 0.01 compared with the Ctrl using two-tailed unpaired <span class="html-italic">t</span>-tests. (<b>E</b>) Cells transfected with RNF152-myc were subjected to immunofluorescence analysis with an anti-myc antibody and either an anti-tyrosinase or anti-Tyrp-1 antibody. Dotted white lines indicate transfected cells. Scale bar: 10 μm. (<b>F</b>) Cells were transfected with either control or RNF152 siRNA. Lysates were immunoblotted with anti-RNF152, anti-tyrosinase, or anti-β-actin antibody. Red asterisks indicate non-specific bands in the RNF152 blotting (see <a href="#app1-membranes-14-00043" class="html-app">Figure S1</a>). (<b>G</b>,<b>H</b>) The expression levels of RNF152 (<b>G</b>) and tyrosinase (<b>H</b>) were quantified in the bands and presented as fold expression relative to the Ctrl (mean ± s.d., <span class="html-italic">n</span> = 3). ** <span class="html-italic">p</span> ≤ 0.01, **** <span class="html-italic">p</span> &lt; 0.0001 compared with the Ctrl using two-tailed unpaired <span class="html-italic">t</span>-tests. (<b>I</b>) The melanin content data are shown as the percentage relative to the Ctrl (mean ± s.d., <span class="html-italic">n</span> = 3). *** <span class="html-italic">p</span> ≤ 0.001 compared with the Ctrl using two-tailed unpaired <span class="html-italic">t</span>-tests.</p>
Full article ">Figure 3
<p>RNF152 is co-localized with tyrosinase-HA in the TGN and degrades it in lysosomes. (<b>A</b>) HEK293T cells transfected with tyrosinase-HA (Tyr-HA) alone (lane 1), or RNF152-myc and Tyr-HA in combination (lanes 2, 3, and 4), were treated with DMSO (control; lanes 1 and 2), LPIs (leupeptin, pepstatin A, E64d; each 40 μM; lane 3), or bafilomycin A1 (BafA1, 5 μM; lane 4) for 14 h. Lysates were immunoblotted with anti-HA, anti-β-actin, and anti-myc antibodies. (<b>B</b>) HeLa cells transfected with RNF152-myc and Tyr-HA were treated with DMSO (Ctrl), LPIs, or bafilomycin A1 (BafA1) as in (A), fixed, and subjected to immunofluorescence analysis with anti-myc and anti-HA antibodies, together with either an anti-syntaxin 6 (TGN marker), or anti-LAMP1 (LEs/lysosome marker) antibody. Squares indicate magnified regions. Scale bar: 10 μm. (<b>C</b>) Quantitative analyses were performed using immunofluorescence data obtained in the experiments depicted in (B). The Pearson’s correlation coefficient between tyrosinase and LAMP1 in RNF152-myc-positive cells treated with DMSO (Ctrl), LPIs, or BafA1 is shown (<span class="html-italic">n</span> = 100 for each treatment). Data are presented in box-and-whisker plots with the minimum, maximum, sample median, and first versus third quartiles. **** <span class="html-italic">p</span> &lt; 0.0001, compared with Ctrl using one-way analysis of variance and Dunnett’s multiple-comparison tests.</p>
Full article ">Figure 4
<p>Interaction of RNF152 with tyrosinase. Lysates of HEK293T cells transfected with tyrosinase-HA (Tyr-HA) with an empty vector (lane 1), WT RNF152-myc (lane 2), C/S mutant RNF152-myc (lane 3), or ΔTM mutant (lane 4) were immunoblotted with anti-HA, anti-myc, or anti-β-actin antibody. Lysates were also immunoprecipitated with the anti-HA antibody, and immune complexes were immunoblotted with anti-myc antibody.</p>
Full article ">Figure 5
<p>Tyrosinase ubiquitination by RNF152. (<b>A</b>) Lysates from HEK293T cells transfected with 3×FLAG-ubiquitin (Ub) and tyrosinase-HA (Tyr-HA) with empty vector (lane 1), WT RNF152-myc (2), C/S mutant (lane 3), or ΔTM mutant (lane 4) were immunoblotted with anti-HA, anti-myc, anti-β-actin, and anti-FLAG antibodies. (<b>B</b>) Lysates were immunoprecipitated with the anti-HA antibody, and immune complexes were immunoblotted with the anti-FLAG antibody. It should be noted that, for the ubiquitin assay using 3×FLAG-ubiquitin, LPIs were added to the cell culture medium for the last 12 h to enhance the signal of ubiquitination.</p>
Full article ">Figure 6
<p>Ubiquitination of tyrosinase and Tyrp-1 by RNF152. (<b>A</b>) Lysates from HEK293T cells transfected with RNF152-myc with an empty vector (lane 1), tyrosinase-HA (Tyr-HA) (lane 2), or Tyrp-1-HA (lane 3) were immunoblotted with anti-HA, anti-myc, and anti-β-actin antibodies. Lysates were also immunoprecipitated with the anti-HA antibody, and immune complexes were immunoblotted with the anti-myc antibody. (<b>B</b>) Lysates from HEK293T cells transfected with 3×FLAG-ubiquitin and RNF152-myc with an empty vector (lane 1), Tyr-HA (lane 2), or Tyrp-1-HA (lane 3) were immunoblotted with anti-HA, anti-myc, anti-β-actin, or anti-FLAG antibody. (<b>C</b>) Lysates were processed as in <a href="#membranes-14-00043-f005" class="html-fig">Figure 5</a>B. Note that LPIs were added to the culture medium for the purpose described in the legend of <a href="#membranes-14-00043-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 7
<p>Putative mechanism of RNF152 interaction and ubiquitination of tyrosinase and Tyrp-1 (schematic). The TM domain of RNF152 interacts with the TM domains of tyrosinase and Tyrp-1. In addition, RNF152 interacts with an unknown E2 ubiquitin-conjugating enzyme, and the E2 conjugates ubiquitin to lysine residue(s) of CT domains of tyrosinase and Tyrp-1. Tyrosinase undergoes stronger ubiquitination than Tyrp-1, likely due to the cytoplasmic tail of mouse tyrosinase harboring seven lysine residues, whereas that of mouse Tyrp-1 has only one.</p>
Full article ">
29 pages, 5422 KiB  
Article
Scavenging of Cation Radicals of the Visual Cycle Retinoids by Lutein, Zeaxanthin, Taurine, and Melanin
by Malgorzata Rozanowska, Ruth Edge, Edward J. Land, Suppiah Navaratnam, Tadeusz Sarna and T. George Truscott
Int. J. Mol. Sci. 2024, 25(1), 506; https://doi.org/10.3390/ijms25010506 - 29 Dec 2023
Cited by 1 | Viewed by 1091
Abstract
In the retina, retinoids involved in vision are under constant threat of oxidation, and their oxidation products exhibit deleterious properties. Using pulse radiolysis, this study determined that the bimolecular rate constants of scavenging cation radicals of retinoids by taurine are smaller than 2 [...] Read more.
In the retina, retinoids involved in vision are under constant threat of oxidation, and their oxidation products exhibit deleterious properties. Using pulse radiolysis, this study determined that the bimolecular rate constants of scavenging cation radicals of retinoids by taurine are smaller than 2 × 107 M−1s−1 whereas lutein scavenges cation radicals of all three retinoids with the bimolecular rate constants approach the diffusion-controlled limits, while zeaxanthin is only 1.4–1.6-fold less effective. Despite that lutein exhibits greater scavenging rate constants of retinoid cation radicals than other antioxidants, the greater concentrations of ascorbate in the retina suggest that ascorbate may be the main protectant of all visual cycle retinoids from oxidative degradation, while α-tocopherol may play a substantial role in the protection of retinaldehyde but is relatively inefficient in the protection of retinol or retinyl palmitate. While the protection of retinoids by lutein and zeaxanthin appears inefficient in the retinal periphery, it can be quite substantial in the macula. Although the determined rate constants of scavenging the cation radicals of retinol and retinaldehyde by dopa-melanin are relatively small, the high concentration of melanin in the RPE melanosomes suggests they can be scavenged if they are in proximity to melanin-containing pigment granules. Full article
(This article belongs to the Special Issue The Role of Carotenoids in Health and Disease)
Show Figures

Figure 1

Figure 1
<p>Representative kinetics of the formation and decay of the transient species monitored at 610 nm after pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinyl palmitate (<b>A</b>) and transient absorption spectra at indicated times after the pulse radiolysis of that solution (<b>B</b>). (<b>C</b>) Representative kinetics of the formation and decay of the transient species monitored at 610 nm after pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinyl palmitate in the presence of 0.1 mM α-tocopherol. (<b>D</b>) Transient absorption spectra at indicated times after the pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinyl palmitate and 0.1 mM zeaxanthin, and representative kinetics of the formation and decay of transient species monitored at 610 nm (<b>E</b>) and 1000 nm (<b>F</b>) after pulse radiolysis of that solution. (<b>G</b>,<b>H</b>) Representative kinetics of the formation and decay of transient species monitored at 610 nm (<b>G</b>) and 950 nm (<b>H</b>) after pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinyl palmitate in the presence of 0.1 mM lutein.</p>
Full article ">Figure 2
<p>Transient absorption spectra after pulse radiolysis of aqueous solution saturated with N<sub>2</sub>O and containing 10 mM phosphate, pH 7, 0.1 M KBr, and 1 mM retinyl palmitate incorporated in 2% Triton X-100 micelles (<b>A</b>) and a representative kinetics of the formation and decay of retinyl palmitate cation radicals monitored at 590 nm after pulse radiolysis of that solution in the absence (<b>B</b>) and presence of 0.1 mM of ascorbate (<b>C</b>).</p>
Full article ">Figure 3
<p>Representative kinetics of the formation and decay of retinyl palmitate cation radicals monitored at 590 nm after pulse radiolysis of aqueous solution saturated with N<sub>2</sub>O and containing 10 mM phosphate, pH 7, 0.1 M KBr, and 1 mM retinyl palmitate incorporated in 2% Triton X-100 micelles, in the absence (<b>A</b>) and presence of 0.1 mM of taurine (<b>B</b>) or 0.1 mg/mL (equivalent to 0.67 mM monomers) dopa-melanin (<b>C</b>).</p>
Full article ">Figure 4
<p>(<b>A</b>) Representative kinetics of the formation and decay of the transient species monitored at 610 nm after pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinaldehyde. (<b>B</b>–<b>D</b>): Transient absorption spectra at indicated times after the pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinaldehyde and 0.1 mM lutein (<b>B</b>) and representative kinetics of the of the formation and decay of the transient species after the pulse radiolysis of that solution monitored at 610 nm (<b>C</b>) and 950 nm (<b>D</b>) after pulse radiolysis of that solution. (<b>E</b>,<b>F</b>) Representative kinetics of the formation and decay of transient species monitored at 610 nm (<b>E</b>) and 1000 nm (<b>F</b>) after pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinaldehyde in the presence of 0.1 mM zeaxanthin.</p>
Full article ">Figure 5
<p>Representative kinetics of the formation and decay of retinaldehyde cation radicals monitored at 590 nm after pulse radiolysis of aqueous solution saturated with N<sub>2</sub>O and containing 10 mM phosphate, pH 7, 0.1 M KBr, and 1 mM retinaldehyde incorporated in 2% Triton X-100 micelles, in the absence (<b>A</b>) and presence of 0.1 mM of taurine (<b>B</b>) or 0.1 mg/mL (equivalent to 0.67 mM monomers) dopa-melanin (<b>C</b>).</p>
Full article ">Figure 6
<p>(<b>A</b>) Representative kinetics of the formation and decay of the transient species monitored at 610 nm after pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinol. (<b>B</b>,<b>C</b>) Representative kinetics of the formation and decay of transient species monitored at 610 nm (<b>B</b>) and 950 nm (<b>C</b>) after pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinol in the presence of 0.1 mM lutein. (<b>D</b>,<b>E</b>) Representative kinetics of the formation and decay of transient species monitored at 610 nm (<b>D</b>) and 1000 nm (<b>E</b>) after pulse radiolysis of N<sub>2</sub>O-saturated benzene with solubilized 1 mM of retinaldehyde in the presence of 0.1 mM zeaxanthin.</p>
Full article ">Figure 7
<p>Representative kinetics of the formation and decay of retinol cation radicals monitored at 590 nm after pulse radiolysis of aqueous solution saturated with N<sub>2</sub>O and containing 10 mM phosphate, pH 7, 0.1 M KBr, and 1 mM retinol incorporated in 2% Triton X-100 micelles, in the absence (<b>A</b>) and presence of 0.1 mM of taurine (<b>B</b>) or 0.1 mg/mL (equivalent to 0.67 mM monomers) dopa-melanin (<b>C</b>).</p>
Full article ">
15 pages, 2781 KiB  
Article
Chemical Composition and Skin-Whitening Activities of Siegesbeckia glabrescens Makino Flower Absolute in Melanocytes
by Da Kyoung Lee, Kyung Jong Won, Do Yoon Kim, Yoon Yi Kim and Hwan Myung Lee
Plants 2023, 12(23), 3930; https://doi.org/10.3390/plants12233930 - 22 Nov 2023
Viewed by 1212
Abstract
Siegesbeckia glabrescens Makino (SGM) has been traditionally used to treat many disorders, including rheumatoid arthritis, hypertension, and acute hepatitis. However, the biological activities of SGM in skin remain unclear. The present study explored the effects of SGM flower absolute (SGMFAb) on skin-whitening-linked biological [...] Read more.
Siegesbeckia glabrescens Makino (SGM) has been traditionally used to treat many disorders, including rheumatoid arthritis, hypertension, and acute hepatitis. However, the biological activities of SGM in skin remain unclear. The present study explored the effects of SGM flower absolute (SGMFAb) on skin-whitening-linked biological activities in B16BL6 cells. SGMFAb was extracted using hexane, and its composition was analyzed through gas chromatography/mass spectrometry analysis. The biological effects of SGMFAb on B16BL6 melanoma cells were detected via WST and BrdU incorporation assays, ELISA, and immunoblotting. SGMFAb contained 14 compounds. In addition, SGMFAb was noncytotoxic, attenuated the serum-induced proliferation of, and inhibited melanin synthesis and tyrosinase activity in α-MSH-exposed B16BL6 cells. SGMFAb also reduced the expressions of MITF (microphthalmia-associated transcription factor), tyrosinase, tyrosinase-related protein (TRP)-1, and TRP-2 in α-MSH-exposed B16BL6 cells. Moreover, SGMFAb downregulated the activation of p38 MAPK, ERK1/2, and JNK in α-MSH-stimulated B16BL6 cells. In addition, SGMFAb reduced the expressions of three melanosome-transport-participating proteins (myosin Va, melanophilin, and Rab27a) in α-MSH-stimulated B16BL6 cells. These results indicate that SGMFAb positively influences skin whitening activities by inhibiting melanogenesis and melanosome-transport-related events in B16BL6 cells, and suggest that SGMFAb is a promising material for developing functional skin whitening agents. Full article
(This article belongs to the Special Issue Structural and Functional Analysis of Extracts in Plants IV)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>GC/MS total ion chromatogram of <span class="html-italic">Siegesbeckia glabrescens</span> Makino flower absolute. Numbers in brackets represent the compound numbers of the 14 compounds in <a href="#plants-12-03930-t001" class="html-table">Table 1</a>. Compound numbers and structures are shown.</p>
Full article ">Figure 2
<p>Effects of <span class="html-italic">Siegesbeckia glabrescens</span> Makino flower absolute on B16BL6 cell viability and proliferation. (<b>a</b>) B16BL6 cell viability. Cells were treated with or without <span class="html-italic">Siegesbeckia glabrescens</span> Makino flower absolute (SGMFAb; 0.1–20 μg/mL) in the presence of 2% FBS for 48 h, and then cell viability was quantified using the WST assay (<span class="html-italic">n</span> = 3). (<b>b</b>) B16BL6 cell proliferation. Cells were incubated with or without <span class="html-italic">Siegesbeckia glabrescens</span> Makino flower absolute (SGMFAb; 0.1–20 μg/mL) in MEM in the presence or absence of 2% FBS for 36 h, and the BrdU assay was conducted to test cell proliferation (<span class="html-italic">n</span> = 3). Cell viabilities (<b>a</b>) and proliferation (<b>b</b>) are expressed as percentages of 2% FBS-alone-treated and untreated controls, respectively. Results are presented as means ± S.E.Ms. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. 2% FBS-alone-treated controls.</p>
Full article ">Figure 3
<p>Effects of <span class="html-italic">Siegesbeckia glabrescens</span> Makino flower absolute on melanin synthesis and tyrosinase activity in B16BL6 cells treated with α-MSH. B16BL6 cells were incubated at 37 °C for 48 h with or without <span class="html-italic">Siegesbeckia glabrescens</span> Makino flower absolute (SGMFAb at 0.1–20 μg/mL in MEM with 2% FBS) in the presence or absence of 200 nM α-MSH. Melanin contents ((<b>a</b>); <span class="html-italic">n</span> = 3) and tyrosinase activities ((<b>b</b>); <span class="html-italic">n</span> = 3) were analyzed as described in <a href="#sec4-plants-12-03930" class="html-sec">Section 4</a>. The top picture in panel (<b>a</b>) indicates representative result. α-MSH represents a positive control. Data are shown as percentages of levels in FBS (2%)-alone-treated controls as means ± SEMs. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 vs. cells treated with α-MSH alone in the presence of FBS (2%). α-MSH, α-melanocyte-stimulating hormone.</p>
Full article ">Figure 4
<p>Effect of <span class="html-italic">Siegesbeckia glabrescens</span> Makino flower absolute on the expression of melanogenesis-related proteins in B16BL6 cells. (<b>a</b>) Representative images. B16BL6 cells were treated with or without <span class="html-italic">Siegesbeckia glabrescens</span> Makino flower absolute (SGMFAb at 0.1–20 μg/mL in MEM with 2% FBS) in the presence or absence of 200 nM α-MSH for 24 h. Cell lysates were immunoblotted with the indicated antibodies as described in <a href="#sec4-plants-12-03930" class="html-sec">Section 4</a>. (<b>b</b>–<b>e</b>) Relative expression levels of tyrosinase ((<b>b</b>); <span class="html-italic">n</span> = 3), tyrosinase-related protein-1 (TRP-1) ((<b>c</b>); <span class="html-italic">n</span> = 3), tyrosinase-related protein-1 (TRP-2) ((<b>d</b>); <span class="html-italic">n</span> = 3), and microphthalmia-associated transcription factor (MITF) ((<b>e</b>); <span class="html-italic">n</span> = 3). α-MSH indicates a positive control. Expressions are shown as percentages of levels in 2% FBS-alone-treated controls. Data are shown as means ± SEMs. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. cells treated with α-MSH alone in the presence of 2% FBS.</p>
Full article ">Figure 5
<p>Effect of <span class="html-italic">Siegesbeckia glabrescens</span> Makino flower absolute on the phosphorylations of MAPKs in B16BL6 cells. (<b>a</b>) Representative images. B16BL6 cells were treated with or without <span class="html-italic">Siegesbeckia glabrescens</span> Makino flower absolute (SGMFAb at 0.1–20 μg/mL in MEM with 2% FBS) in the presence or absence of 200 nM α-MSH for 5 min. Cell lysates were subjected to Western blotting with indicated antibodies as described in <a href="#sec4-plants-12-03930" class="html-sec">Section 4</a>. (<b>b</b>–<b>d</b>) Statistical results for phosphorylated ERK1/2 (<b>b</b>), p38 MAPK (<b>c</b>), and JNK levels (<b>d</b>) obtained from panel (<b>a</b>). α-MSH indicates a positive control. The phosphorylation levels of kinase are presented as percentages of levels in 2% FBS-alone-treated controls. Data are shown as means ± SEMs (<span class="html-italic">n</span> = 3 for each protein). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. cells exposed to α-MSH alone in the presence of 2% FBS. P-ERK1/2, phosphorylated ERK1/2; p-JNK, phosphorylated JNK; p-p38, phosphorylated p38 MAPK.</p>
Full article ">Figure 6
<p>Effect of SGMFAb on the expressions of melanosome transport proteins in B16BL6 cells. (<b>a</b>) Representative images. B16BL6 cells were incubated in the presence or absence of <span class="html-italic">Siegesbeckia glabrescens</span> Makino flower absolute (SGMFAb 0.1–20 μg/mL in MEM containing 2% FBS) with or without 200 nM of α-MSH for 24 h. Cell lysates were immunoblotted with indicated antibodies. (<b>b</b>–<b>d</b>) Statistical results for myosine Va (MyoVa; (<b>b</b>)), melanophillin (Mlph; (<b>c</b>)), and Rab27a (<b>d</b>) expression levels obtained from panel (<b>a</b>). α-MSH indicates a positive control. The phosphorylation levels of kinases are shown as percentages of levels in controls treated with 2% FBS alone. Results are expressed as means ± SEMs (<span class="html-italic">n</span> = 3 for each protein). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. cells treated with α-MSH alone in the presence of 2% FBS.</p>
Full article ">
14 pages, 1980 KiB  
Article
Oxidative Stress Induces Skin Pigmentation in Melasma by Inhibiting Hedgehog Signaling
by Nan-Hyung Kim and Ai-Young Lee
Antioxidants 2023, 12(11), 1969; https://doi.org/10.3390/antiox12111969 - 6 Nov 2023
Cited by 1 | Viewed by 2004
Abstract
There is growing evidence that oxidative stress plays a role in melasma and disrupts primary cilia formation. Additionally, primary cilia have been suggested to have an inhibitory role in melanogenesis. This study examined the potential link between oxidative stress, skin hyperpigmentation, and primary [...] Read more.
There is growing evidence that oxidative stress plays a role in melasma and disrupts primary cilia formation. Additionally, primary cilia have been suggested to have an inhibitory role in melanogenesis. This study examined the potential link between oxidative stress, skin hyperpigmentation, and primary cilia. We compared the expression levels of the nuclear factor E2-related factor 2 (NRF2), intraflagellar transport 88 (IFT88), and glioma-associated oncogene homologs (GLIs) in skin samples from patients with melasma, both in affected and unaffected areas. We also explored the roles of NRF2, IFT88, and GLIs in ciliogenesis and pigmentation using cultured adult human keratinocytes, with or without melanocytes. Our findings revealed decreased levels of NRF2, heme oxygenase-1, IFT88, and GLIs in lesional skin from melasma patients. The knockdown of NRF2 resulted in reduced expressions of IFT88 and GLI1, along with fewer ciliated cells. Furthermore, NRF2, IFT88, or GLI1 knockdown led to increased expressions in protease-activated receptor-2 (PAR2), K10, involucrin, tyrosinase, and/or melanin. These effects were reversed by the smoothened agonist 1.1. Calcium also upregulated these proteins, but not NRF2. The upregulation of involucrin and PAR2 after NRF2 knockdown was mitigated with a calcium chelator. In summary, our study suggests that oxidative stress in NRF2-downregulated melasma keratinocytes impedes ciliogenesis and related molecular processes. This inhibition stimulates keratinocyte differentiation, resulting in melanin synthesis and melanosome transfer, ultimately leading to skin hyperpigmentation. Full article
(This article belongs to the Special Issue Oxidative Stress and NRF2 in Health and Disease)
Show Figures

Figure 1

Figure 1
<p>NRF2 downregulation in the lesional epidermis of patients with melasma. (<b>A</b>) Reactive oxygen species concentrations at various time points in primary cultured keratinocytes following UVB irradiation. (<b>B</b>) Western blot analyses showing NRF2 protein level ratios after single and repeated UVB radiation. (<b>C</b>) Western blot analyses illustrating NRF2 and HO-1 protein level ratios over time in primary cultured normal human keratinocytes treated with different concentrations of H<sub>2</sub>O<sub>2</sub>. (<b>D</b>) Western blot analyses presenting HO-1 protein level ratios in cultured human keratinocytes with or without <span class="html-italic">NRF2</span> knockdown. β-actin served as the internal control for the Western blot analysis. The data are presented as means ± SD from four or eight independent experiments. (<b>E</b>,<b>F</b>) Representative immunofluorescence staining using anti-NRF2 (<b>E</b>) and anti-HO-1 antibodies (<b>F</b>) in the lesional (L) and non-lesional (N) epidermis of patients with melasma. The nuclei were counterstained with Hoechst 33258 (scale bar = 0.05 mm), and the intensities were quantified using ImageJ software 1.54d. * <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.</p>
Full article ">Figure 2
<p>Downregulation of NRF2 led to reduced expressions of IFT88 and Hh signaling molecules involved in ciliogenesis. (<b>A</b>) Western blot analyses depicting the ratios of PTCH, GLI1, and GLI2 levels in cultured keratinocytes subjected to <span class="html-italic">IFT88</span> knockdown. (<b>B</b>) Confocal microscopy images illustrating primary cilia stained with anti-acetylated α-tubulin (Ac α-tubulin) and/or ARL13b antibodies in cultured human keratinocytes and melanocytes, with or without <span class="html-italic">IFT88</span> knockdown (bar = 0.05 mm). The ciliated cell ratios were calculated by counting the number of ciliated cells among 30 cells. (<b>C</b>,<b>D</b>) Western blot analyses showing the ratios of NRF2, IFT88, and/or GLI1 levels in cultured keratinocytes with knockdowns of <span class="html-italic">NRF2</span> (<b>C</b>) or <span class="html-italic">IFT88</span> (<b>D</b>). (<b>E</b>) Real-time PCR results displaying the ratios of IFT88, PTCH1, and GLI1-3 mRNA levels in lesional compared to non-lesional skin specimens (seven sets) from melasma patients with downregulated NRF2. (<b>F</b>) Representative immunofluorescence staining for primary cilia using anti-NRF2 and anti-ARL13b antibodies in primary cultured human keratinocytes and melanocytes with or without <span class="html-italic">NRF2</span> knockdown (scale bar = 0.05 mm). β-actin and GAPDH served as internal controls for the Western blot analysis and real-time PCR, respectively. The data are presented as means ± SD from four independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>Enhancement of melanin pigmentation via <span class="html-italic">NRF2</span> knockdown involving IFT88 and GLI1. (<b>A</b>) Western blot analyses depicting varying levels of tyrosinase, PMEL, and PAR2, and assays showing tyrosinase activity and melanin contents in cultured keratinocytes with <span class="html-italic">NRF2</span> knockdown. (<b>B</b>–<b>E</b>) Western blot analyses revealing different ratios of NRF2, IFT88, GLI1, and/or tyrosinase levels in cultured keratinocytes with <span class="html-italic">NRF2</span> knockdown in the absence and presence of SAG (<b>B</b>), <span class="html-italic">IFT88</span> knockdown in the absence and presence of SAG (<b>C</b>), <span class="html-italic">IFT88</span> knockdown in the absence and presence of Shh 200 (<b>D</b>), and <span class="html-italic">GLI1</span> knockdown in the absence and presence of SAG (<b>E</b>). (<b>F</b>) Western blot analyses presenting different ratios of PTCH1, GLI1, GLI2, and tyrosinase levels in cultured keratinocyte–melanocyte cocultures treated with or without GANT61. β-actin served as an internal control. The data represent the means ± SD from four independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. control sgRNA, # <span class="html-italic">p</span> &lt; 0.05 vs. without SAG treatment.</p>
Full article ">Figure 4
<p>Effects of <span class="html-italic">NRF2</span>, <span class="html-italic">IFT88</span>, and <span class="html-italic">GLI1</span> knockdowns on keratinocyte differentiation and subsequent hyperpigmentation. (<b>A</b>–<b>C</b>) Western blot analyses illustrating the relative levels of K14, K10, and involucrin in cultured keratinocytes with or without knockdowns of <span class="html-italic">NRF2</span> (<b>A</b>), <span class="html-italic">IFT88</span> (<b>B</b>), or <span class="html-italic">GLI1</span> (<b>C</b>) in the absence and presence of SAG. (<b>D</b>) Representative immunofluorescence staining using anti-involucrin antibodies (<b>B</b>) in the lesional (L) and non-lesional (N) epidermis of seven patients with melasma. The nuclei were counterstained with Hoechst 33258 (bar = 0.05 mm), and the intensities were measured using ImageJ software 1.54d. (<b>E</b>,<b>F</b>) Western blot analyses for the ratios of tyrosinase levels in keratinocyte–melanocyte cocultures and the ratios of PAR2, K10, involucrin, NRF2, IFT88, and/or GLI1 levels in cultured keratinocytes, including those treated with calcium (<b>E</b>) and keratinocytes with or without NRF2 knockdown in the absence and presence of Bapta-AM (<b>F</b>). β-actin was used as an internal control for the Western blot analysis. The data are presented as the means ± SD from four independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. control sgRNA; # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01 vs. without SAG treatment.</p>
Full article ">Figure 5
<p>Schematic view of the role of <span class="html-italic">NRF2</span>-knockdown-induced ciliogenesis inhibition in skin hyperpigmentation. NRF2 downregulation caused by repeated UV exposure or melasma inhibited ciliogenesis and Hh signaling molecules, such as IFT88 and GLI1, stimulating keratinocyte differentiation with melanin synthesis and melanosome transfer to the keratinocytes, which resulted in skin hyperpigmentation.</p>
Full article ">
Back to TopTop