[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 (101)

Search Parameters:
Keywords = keloid

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
19 pages, 4691 KiB  
Article
Different Shades of Desmoid-Type Fibromatosis (DTF): Detection of Noval Mutations in the Clinicopathologic Analysis of 32 Cases
by Rana Ajabnoor
Diagnostics 2024, 14(19), 2161; https://doi.org/10.3390/diagnostics14192161 - 28 Sep 2024
Viewed by 367
Abstract
Background: Desmoid-type fibromatosis (DTF) is a locally aggressive myofibroblastic/fibroblastic neoplasm with a high risk of local recurrence. It has a variety of histologic features that might confuse diagnosis, especially when detected during core needle biopsy. The Wnt/β-catenin pathway is strongly linked to the [...] Read more.
Background: Desmoid-type fibromatosis (DTF) is a locally aggressive myofibroblastic/fibroblastic neoplasm with a high risk of local recurrence. It has a variety of histologic features that might confuse diagnosis, especially when detected during core needle biopsy. The Wnt/β-catenin pathway is strongly linked to the pathogenesis of DT fibromatosis. Method: This study examined 33 desmoid-type fibromatoses (DTFs) from 32 patients, analyzing its clinical characteristics, histologic patterns, occurrence rates, relationship with clinical outcomes, immunohistochemical and molecular findings. Results: The DTFs exhibit a range of 1 to 7 histologic patterns per tumor, including conventional, hypercellular, myxoid, hyalinized/hypocellular, staghorn/hemangiopericytomatous blood vessels pattern, nodular fasciitis-like, and keloid-like morphology. No substantial association was found between the existence of different histologic patterns and the clinical outcome. All thirty-three (100%) samples of DTF had a variable percentage of cells that were nuclear positive for β-catenin. An NGS analysis detected novel non-CTNNB1 mutations in two DTFs, including BCL10, MPL, and RBM10 gene mutations. Conclusions: This study reveals a diverse morphology of DTFs that could result in misdiagnosis. Therefore, surgical pathologists must comprehend this thoroughly. Also, the importance of the newly identified non-CTNNB1 gene mutations is still unclear. More research and analyses are needed to completely grasp the clinical implications of these mutations. Full article
(This article belongs to the Special Issue Histopathology in Cancer Diagnosis and Prognosis)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of the study method. DTF, desmoid-type fibromatosis; FAP, familia adenomatous polyposis; H&amp;E, hematoxylin and eosin; IHC, immunohistochemistry; NGS, next-generation sequencing.</p>
Full article ">Figure 2
<p>Macroscopically, DTF usually presents as a large infiltrative firm lesion with a trabeculated cut surface.</p>
Full article ">Figure 3
<p>The histologic pattern of desmoid-type fibromatosis. (<b>A</b>,<b>B</b>) The conventional pattern is characterized by the presence of long, sweeping fascicles of thin fibroblasts/myofibroblasts that are uniformly spaced and exhibit minimal or no cell-to-cell contact. (<b>C</b>,<b>D</b>) The hypercellular pattern has a higher cellular density compared with the conventional pattern, accompanied by an increase in nuclear overlaps.</p>
Full article ">Figure 4
<p>Continued histologic pattern of desmoid-type fibromatosis. (<b>A</b>) The hyalinized/hypocellular pattern is characterized by fibroblasts, and myofibroblasts are sparsely distributed within a heavily hyalinized collagenous background. (<b>B</b>,<b>C</b>) The myxoid pattern is distinguished by a profuse myxoid background including loosely arranged hypocellular regions of spindle cells, resembling cellular myxoma. (<b>D</b>) The keloid-like pattern characterized by the presence of collagenized bands that vary in size and exhibit a bright eosinophilic appearance.</p>
Full article ">Figure 5
<p>Continued histologic pattern of desmoid-type fibromatosis. (<b>A</b>) Nodular fasciitis-like pattern characterized by loose short fascicles of spindle to stellate cells with an edematous background and extravasated RBCs. (<b>B</b>) Hemangiopericytomatous-like patten characterized by the presence of thin-walled, branched blood vessels with staghorn-shaped vessels. Additional histological features: (<b>C</b>) the presence of lymphoid aggregation, mainly at the periphery of the tumor. (<b>D</b>) Perivascular edema.</p>
Full article ">Figure 6
<p>Additional histological features of DTF. (<b>A</b>,<b>B</b>) Spindle cells with mild to moderate atypia characterized by nuclear hyperchromasia. (<b>C</b>,<b>D</b>) Two cases showed focal chondromyxoid-like metaplasia.</p>
Full article ">Figure 7
<p>Immunohistochemistry of DTF. (<b>A</b>) β-catenin shows diffuse nuclear positivity in spindle cells. (<b>B</b>) SMA shows focal positivity in the spindle cells. (<b>C</b>) S100 shows rare positivity in spindle cells. (<b>D</b>) Desmin shows focal positivity in the spindle cells. (<b>E</b>) CD34 is negative in the spindle cells.</p>
Full article ">Figure 8
<p>A non-CTNNB1 mutation of BCL10 mutation variant c.136del p.(Ile46Tyrfs 24) (Frequency: 5.9% of 1649 NGS reads) causes a disruption in the reading frame, beginning at codon 46.</p>
Full article ">Figure 9
<p>A non-CTNNB1 mutation of MPL gene mutation variant c.317C&gt;T p.(Pro106Leu) (Frequency: 48.4% of 1544 NGS reads).</p>
Full article ">
13 pages, 5726 KiB  
Article
Increased Susceptibility to Mechanical Stretch Drives the Persistence of Keloid Fibroblasts: An Investigation Using a Stretchable PDMS Platform
by Jihee Kim, Chihyeong Won, Seoyoon Ham, Heetak Han, Sungsik Shin, Jieun Jang, Sanghyeon Lee, Chaebeen Kwon, Sungjoon Cho, Hyeonjoo Park, Dongwon Lee, Won Jai Lee, Taeyoon Lee and Ju Hee Lee
Biomedicines 2024, 12(10), 2169; https://doi.org/10.3390/biomedicines12102169 - 24 Sep 2024
Viewed by 560
Abstract
Background: Keloids are a common fibrotic disease of the skin, with the pathological hallmark of excessive extracellular matrix synthesis due to abnormal fibroblast activity. Since keloids clinically arise in areas of high mechanical tension, the mechanotransductory pathway may be attributed to its pathogenesis. [...] Read more.
Background: Keloids are a common fibrotic disease of the skin, with the pathological hallmark of excessive extracellular matrix synthesis due to abnormal fibroblast activity. Since keloids clinically arise in areas of high mechanical tension, the mechanotransductory pathway may be attributed to its pathogenesis. We aimed to establish a preclinical platform to elucidate the underlying mechanism of keloid development and its clinical persistence. Methods: We fabricated a mechanically stretchable polydimethylsiloxane cell culture platform; with its mimicry of the in vivo cyclic stretch of skeletal muscles, cells showed higher proliferation compared with conventional modalities. Results: In response to mechanical strain, TGF-β and type 1 collagen showed significant increases, suggesting possible TGF-β/Smad pathway activation via mechanical stimulation. Protein candidates selected by proteomic analysis were evaluated, indicating that key molecules involved in cell signaling and oxidative stress were significantly altered. Additionally, the cytoskeletal network of keloid fibroblasts showed increased expression of its components after periodic mechanical stimulation. Conclusions: Herein, we demonstrated and validated the existing body of knowledge regarding profibrotic mechanotransduction signaling pathways in keloid fibroblasts. Cyclic stretch, as a driving force, could help to decipher the tension-mediated biomechanical processes, leading to the development of optimized therapeutic targets. Full article
(This article belongs to the Special Issue Wound Healing: From Basic to Clinical Research)
Show Figures

Figure 1

Figure 1
<p>Fabrication of the high-throughput, mechanically stretchable, PDMS cell culture platform (<b>A</b>) A static overview of the mechanically stretchable cell culture platform; (<b>B</b>) the PDMS platform consisted of 12-wells (<b>iii</b>), with two major components: a stretchable culture plate (<b>i</b>), and a stretch device controlled by vacuum pressure (<b>ii</b>) (scale bar: 20 mm); (<b>C</b>) mechanical stimulation was provided by adjusting the pneumatic pressure; (<b>D</b>) fabrication of the replica modeling on the stretchable platform; initial length (L0) and stretch length (L1) were compared (scale bar: 4 mm); (<b>E</b>) finite-element method to compare the applied pneumatic pressure to the mechanical strain of the PDMS platform.</p>
Full article ">Figure 2
<p>Characterization of HDFs and KFs in response to cyclic mechanical stretch. (<b>A</b>) HDFs and KFs demonstrated increased proliferation rates, as well as morphological changes, in response to mechanical strain. Cell morphology and proliferation of HDFs and KFs on the PDMS plate were analyzed via MTT assay after mechanical tension. The images were captured using a light microscope, and the scale bar represents 100 μm. Proliferation of KFs was observed at a lower degree of mechanical strain (3%) than that of HDFs (5%) after three (<b>B</b>) and seven days (<b>C</b>) of cell culture. We used HDF at passage 6~9 and KF at passage 6. All experiments were repeated three times. Error bars, SD ** <span class="html-italic">p</span> &lt; 0.005; Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 3
<p>Mechanical stretch induced the increased expression of fibrotic markers and ECM components in HDFs and KFs. Gene expression profiles of TGF-β signaling and subsequent ECM accumulation on the PDMS plate after mechanical tension of RT-PCR. Mechanical tension induced increased expression of TGF-β and COL1A1 in HDFs and KFs. The expression of molecular markers was more significant with lower degrees of mechanical strain (3%) in KFs compared with HDFs (5%). We used HDFs at passage 6~9 and KFs at passage 6. All experiments were repeated three times. Error bars, SD. *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.005, * <span class="html-italic">p</span> &lt; 0.05; Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 4
<p>Quantitative proteomic analysis using the TMT-labeling method demonstrated target proteins subject to increased mechanosensitivity in KFs. The degree of mechanical strain was 3% for KFs and 5% for HDFs. Western blot protein expression profiles of the cell proliferation (PPP2R5D, PDLIM5), ROS (AKR1B1, SOD2), and fibrosis (COL1A1, TGF-β) markers on the PDMS plate after mechanical tension. Quantitative protein analysis identified the target proteins subject to increased mechanosensitivity in KFs. The original image of the Western blot is in <a href="#app1-biomedicines-12-02169" class="html-app">Figure S6</a>. We used HDFs at passage 6~9 and KFs at passage 6. All experiments were repeated three times. Error bars, s.e.m. * <span class="html-italic">p</span> &lt; 0.05; Student’s <span class="html-italic">t</span>-test. Western blot images are cropped for serial comparison between each molecule.</p>
Full article ">Figure 5
<p>Immunohistochemical staining of KFs demonstrated an increase and change in cytoskeletal composition due to cyclic mechanical stretch. (<b>A</b>) Immunofluorescent staining images of protein markers involved in actin filament formation after being exposed to mechanical tension (×20 magnification, scale bar = 50 μm); (<b>B</b>) after experiencing mechanical tension, expression of the actin filament formation-related proteins increased, and the cells were regularly arranged. We used KFs at passage 6. All experiments were repeated three times. Error bars, s.e.m. ** <span class="html-italic">p</span> &lt; 0.005, * <span class="html-italic">p</span> &lt; 0.05; Student’s <span class="html-italic">t</span>-test.</p>
Full article ">
13 pages, 4566 KiB  
Review
Reflectance Confocal Microscopy and Dermoscopy for the Diagnosis of Solitary Hypopigmented Pink Lesions: A Narrative Review
by Luca Ambrosio, Anna Pogorzelska-Antkowiak, Chiara Retrosi, Giovanni Di Lella, Marco Spadafora, Iris Zalaudek, Caterina Longo, Giovanni Pellacani and Claudio Conforti
Cancers 2024, 16(17), 2972; https://doi.org/10.3390/cancers16172972 - 26 Aug 2024
Viewed by 557
Abstract
Diagnosing solitary pink skin lesions poses a significant challenge due to the scarcity of specific clinical and dermoscopic criteria. Several benign lesions, such as cherry angioma, clear cell acanthoma, dermal nevus, keloid, hypertrophic scar, and Spitz nevus, often exhibit similar clinical and dermoscopic [...] Read more.
Diagnosing solitary pink skin lesions poses a significant challenge due to the scarcity of specific clinical and dermoscopic criteria. Several benign lesions, such as cherry angioma, clear cell acanthoma, dermal nevus, keloid, hypertrophic scar, and Spitz nevus, often exhibit similar clinical and dermoscopic features. This similarity extends to some malignant lesions, including basal cell carcinoma, actinic keratosis, and amelanotic melanoma, making differentiation difficult. Recent studies highlight the enhanced diagnostic accuracy of reflectance confocal microscopy (RCM), which offers increased sensitivity and specificity compared to dermoscopy alone for diagnosing skin cancer. This study aims to summarize the application of dermoscopy and RCM in distinguishing between benign and malignant pinkish–reddish skin lesions. The integration of RCM with traditional dermoscopic techniques improves the ability to accurately identify and differentiate these lesions. However, it is crucial to note that for any suspicious lesions, a final diagnosis must be confirmed through surgical excision and histopathological evaluation. This comprehensive approach ensures accurate diagnosis and appropriate treatment, highlighting the importance of combining advanced imaging techniques in clinical practice. Full article
(This article belongs to the Special Issue Dermoscopy in Skin Cancer)
Show Figures

Figure 1

Figure 1
<p>Cherry angioma. A 41-year-old woman presented with an erythematous papular lesion of the trunk. Dermoscopic image showing a well-demarcated pinkish lesion with white septa inside the lesion and red lacunes (<b>a</b>). RCM image with dark lacunes separated by bright septa (red arrow) (<b>b</b>). Courtesy, Anna Pogorzelska-Antkowiak, MD.</p>
Full article ">Figure 2
<p>Clear cell acanthoma. A 38-year-old woman presented with an erythematous papular lesion on the left leg. Dotted vessels resembling a necklace and a scaly peripheral collarette in dermoscopy (<b>a</b>). At the level of the epidermis, the RCM image reveals acanthosis with papillomatosis (red arrow) in a well-defined lesion with a clear border (<b>b</b>). Courtesy, Prof. Caterina Longo.</p>
Full article ">Figure 3
<p>Dermal nevus. A 54-year-old male with a nodular lesion of the abdomen characterized by comma-like vessels and residual pigmentation on dermoscopy (<b>a</b>). RCM image showing clusters of homogenous, dense, and sparse nests at the DEJ (red arrow) (<b>b</b>). Courtesy, Prof. Caterina Longo.</p>
Full article ">Figure 4
<p>Dermatofibroma. A 36-year-old woman with a solitary, firm papule of the right deltoid. Small, central white scar-like area surrounded by a pinkish structureless zone in dermoscopy (<b>a</b>). RCM image showing thick reticulated collagen fibers in the central part of the lesion (red arrow) surrounded by edged and slightly bright papillae at the DEJ (red asterisk) (<b>b</b>). Courtesy, Prof. Caterina Longo.</p>
Full article ">Figure 5
<p>Keloid. A 43-year-old male presented with a firm, nodular lesion on the right thigh. Dermoscopic image showing irregular vessels on a homogenous pinkish background (<b>a</b>). RCM image reveals numerous coarse collagen fibers (red arrow) (<b>b</b>). Courtesy, Prof. Caterina Longo.</p>
Full article ">Figure 6
<p>Spitz nevus. A 14-year-old boy presented with a pinkish macular lesion on the left arm. Dermoscopy is characterized by an inverse pigment network, dotted vessels, and pigment remnants (<b>a</b>). (<b>b</b>) In RCM, the DEJ reveals edged and non-edged papillae with spindle-atypical cells and homogeneous nests (red arrow). Courtesy, Anna Pogorzelska-Antkowiak, MD.</p>
Full article ">Figure 7
<p>Basal cell carcinoma. A 65-year-old man presented with a pinkish macular lesion of the dorsum. Dermoscopy reveals a pinkish, structureless area with thin telangiectatic vessels (<b>a</b>). RCM image showing a thick, elongated vessel (red asterisk) and a dark tumor island (red arrow) with palisading of nuclei and peritumoral clefts (<b>b</b>). Courtesy, Anna Pogorzelska-Antkowiak, MD.</p>
Full article ">Figure 8
<p>Actinic keratosis. A 72-year-old male presented with an erythematous–desquamative macular lesion of the frontal region. The dermoscopic image shows a red pseudonetwork or strawberry pattern with white–yellow scales (<b>a</b>). RCM image showing an irregular honeycombed pattern of the epidermis indicative of moderate-grade dyskeratosis (red arrow) (<b>b</b>). Courtesy, Anna Pogorzelska-Antkowiak, MD.</p>
Full article ">Figure 9
<p>Amelanotic melanoma. A 53-year-old woman presented with a hypopigmented-pink macular lesion with minimal residual pigment of the upper trunk. Dermoscopy shows linear irregular vessels and remnants of pigmentation (<b>a</b>). RCM at the level of the epidermis reveals numerous roundish (red arrow) and dendritic (red asterisk) pagetoid cells within a disarranged epidermis (<b>b</b>). Courtesy, Anna Pogorzelska-Antkowiak, MD.</p>
Full article ">
23 pages, 1237 KiB  
Review
Comprehensive Insights into Keloid Pathogenesis and Advanced Therapeutic Strategies
by Hyun Jee Kim and Yeong Ho Kim
Int. J. Mol. Sci. 2024, 25(16), 8776; https://doi.org/10.3390/ijms25168776 - 12 Aug 2024
Cited by 1 | Viewed by 1348
Abstract
Keloid scars, characterized by abnormal fibroproliferation and excessive extracellular matrix (ECM) production that extends beyond the original wound, often cause pruritus, pain, and hyperpigmentation, significantly impacting the quality of life. Keloid pathogenesis is multifactorial, involving genetic predisposition, immune response dysregulation, and aberrant wound-healing [...] Read more.
Keloid scars, characterized by abnormal fibroproliferation and excessive extracellular matrix (ECM) production that extends beyond the original wound, often cause pruritus, pain, and hyperpigmentation, significantly impacting the quality of life. Keloid pathogenesis is multifactorial, involving genetic predisposition, immune response dysregulation, and aberrant wound-healing processes. Central molecular pathways such as TGF-β/Smad and JAK/STAT are important in keloid formation by sustaining fibroblast activation and ECM deposition. Conventional treatments, including surgical excision, radiation, laser therapies, and intralesional injections, yield variable success but are limited by high recurrence rates and potential adverse effects. Emerging therapies targeting specific immune pathways, small molecule inhibitors, RNA interference, and mesenchymal stem cells show promise in disrupting the underlying mechanisms of keloid pathogenesis, potentially offering more effective and lasting treatment outcomes. Despite advancements, further research is essential to fully elucidate the precise mechanisms of keloid formation and to develop targeted therapies. Ongoing clinical trials and research efforts are vital for translating these scientific insights into practical treatments that can markedly enhance the quality of life for individuals affected by keloid scars. Full article
Show Figures

Figure 1

Figure 1
<p>Keloid pathogenesis.</p>
Full article ">
16 pages, 922 KiB  
Review
Dry Needling and Acupuncture for Scars—A Systematic Review
by Robert Trybulski, Adam Kawczyński, Jarosław Muracki, Nicola Lovecchio and Adrian Kużdżał
J. Clin. Med. 2024, 13(14), 3994; https://doi.org/10.3390/jcm13143994 - 9 Jul 2024
Viewed by 2122
Abstract
Objectives: This research aims to synthesize existing data on the evidence gap in scar treatment and evaluate the effectiveness of acupuncture and dry needling in treating scars and related symptoms. Methods: The article adhered to the PRISMA 2020 statement for recommended [...] Read more.
Objectives: This research aims to synthesize existing data on the evidence gap in scar treatment and evaluate the effectiveness of acupuncture and dry needling in treating scars and related symptoms. Methods: The article adhered to the PRISMA 2020 statement for recommended reporting elements in systematic reviews. The inclusion criteria followed the PICO methodology. The literature search was conducted using databases including PubMed, Cochrane Library, Semantic Scholar, Europe PubMed Central, and Google Scholar. Studies on acupuncture and dry needling for scar treatment were included. Because of the diversity of the studies’ results and methodologies, a systematic review was conducted to organize and describe the findings without attempting a numerical synthesis. Results: Nineteen studies relevant to the article’s theme were identified, with eleven selected for detailed review. The studies included two case reports on dry needling, one case series on dry needling, five case reports on acupuncture, two randomized controlled trials on acupuncture, and one case report on Fu’s subcutaneous needling. A quality assessment was conducted using the JBI CAT and PEDro scales. Four case reports scored 7 points, one case scored 8 points, three cases were rated 6 points or lower, the case series was rated 6 points, and the randomized controlled trials scored 8 and 5 points. Most studies demonstrated a desired therapeutic effect in scar treatment with acupuncture and dry needling, but the level of evidence varied across studies. The analysis does not conclusively support the use of acupuncture and dry needling to improve scar conditions. Conclusions: Although dry-needling and acupuncture techniques are popular in physiotherapy, adequate scientific evidence is currently not available to support their effectiveness in scar treatment. There are gaps in the research methodology, a lack of randomized trials, and significant heterogeneity in the assessment of effects. Full article
(This article belongs to the Special Issue Skin Rehabilitation: Recent Advances and Future Perspectives)
Show Figures

Figure 1

Figure 1
<p>Application of acupuncture in scar treatment. Source: Tuckey et al. (2022) [<a href="#B3-jcm-13-03994" class="html-bibr">3</a>].</p>
Full article ">Figure 2
<p>Application of dry needling in scar treatment. Source: Rozenfeld et al. (2020) [<a href="#B16-jcm-13-03994" class="html-bibr">16</a>].</p>
Full article ">Figure 3
<p>The PRISMA flowchart for the selection of the literature meeting the study’s criteria, according to M. Page et al. [<a href="#B22-jcm-13-03994" class="html-bibr">22</a>].</p>
Full article ">
28 pages, 14514 KiB  
Review
Disturbances in the Skin Homeostasis: Wound Healing, an Undefined Process
by Montserrat Férnandez-Guarino, Jorge Naharro-Rodriguez and Stefano Bacci
Cosmetics 2024, 11(3), 90; https://doi.org/10.3390/cosmetics11030090 - 4 Jun 2024
Cited by 1 | Viewed by 1607
Abstract
This review was written with the aim of examining the effects that cause an insult, such as a wound, to an organ, such as the skin. Before examining the cellular mechanisms relating to wound healing, the reader is invited to read about the [...] Read more.
This review was written with the aim of examining the effects that cause an insult, such as a wound, to an organ, such as the skin. Before examining the cellular mechanisms relating to wound healing, the reader is invited to read about the structure of the skin as a necessary basis for understanding the final aim of this review. The structure of the skin as a basis for understanding the phenomena relating to wound healing is addressed, taking into account the updated literature that addresses the numerous problems of the skin microenvironment. Starting from this awareness, the paragraphs dedicated to wound healing become complicated when this phenomenon is not implemented and therefore while the problems of chronic wounds, keloids, and hypertrophic scars are addressed, these are pathologies that are still difficult to understand and treat today. Full article
(This article belongs to the Special Issue 10th Anniversary of Cosmetics—Recent Advances and Perspectives)
Show Figures

Figure 1

Figure 1
<p>The skin: morphological differences between the epidermis and the dermis. Hematoxylin eosin, bar = 10 microns. The photograph is a property of one of the authors (SB).</p>
Full article ">Figure 2
<p>The skin: morphological differences between the epidermis and the dermis. The epidermis and the various layers that compose it. The underlying dermis has various cell types, a sebaceous gland, and a hair follicle. Hematoxylin eosin, bar = 10 microns. The photograph is a property of one of the authors (SB).</p>
Full article ">Figure 3
<p>Dermis: different features of the collagen fibers which histologically define the type of tissue considered Mallory Azan = 10 microns. The photograph is a property of one of the authors (SB).</p>
Full article ">Figure 4
<p>Dermis: indirect immunofluorescence for the detection of neuronal enolase in the skin. Confocal microscopy, scale bar <b>=</b> 100 microns. The photograph is a property of one of the authors (SB).</p>
Full article ">Figure 5
<p>Epidermis: indirect immunofluorescence for the detection of CD1a-positive Langerhans cells. Fluorescence Microscopy, scale bar = 10 microns. The photograph is a property of one of the authors (SB).</p>
Full article ">Figure 6
<p>Dermis: on the left (<b>a</b>) affinity cytochemistry for mast cell localization (stained with avidin in red), Fluorescence Microscopy, scale bar = 10 microns; on the right (<b>b</b>) ultrastructure of a mast cell, electron microscopy, scale bar 1 micron. The photographs are a property of one of the authors (SB).</p>
Full article ">Figure 7
<p>Dermis: indirect immunofluorescence and <span class="html-italic">affinity cytochemistry</span> for the colocalization of granulocytes (stained with Ly6g antibody in green) and MCs (stained with avidin in red). Fluorescence Microscopy, scale bar = 10 microns. The photograph is a property of one of the authors (SB).</p>
Full article ">Figure 8
<p>Dermis: <span class="html-italic">affinity cytochemistry</span> for the colocalization of vessels (stained with Griffonia (Bainderaea) simplicifolia) and MCs (stained with avidin in red). The photograph is a property of one of the authors (SB).</p>
Full article ">Figure 9
<p>Dermis: indirect immunofluorescence for the localization of fibroblasts (stained with HSP47g antibody in green). The photograph is a property of one of the authors (SB).</p>
Full article ">
15 pages, 4973 KiB  
Article
Ethyl Pyruvate Decreases Collagen Synthesis and Upregulates MMP Activity in Keloid Fibroblasts and Keloid Spheroids
by Wooyeol Baek, Seonghyuk Park, Youngdae Lee, Hyun Roh, Chae-Ok Yun, Tai Suk Roh and Won Jai Lee
Int. J. Mol. Sci. 2024, 25(11), 5844; https://doi.org/10.3390/ijms25115844 - 28 May 2024
Viewed by 944
Abstract
Keloids, marked by abnormal cellular proliferation and excessive extracellular matrix (ECM) accumulation, pose significant therapeutic challenges. Ethyl pyruvate (EP), an inhibitor of the high-mobility group box 1 (HMGB1) and TGF-β1 pathways, has emerged as a potential anti-fibrotic agent. Our research evaluated EP’s effects [...] Read more.
Keloids, marked by abnormal cellular proliferation and excessive extracellular matrix (ECM) accumulation, pose significant therapeutic challenges. Ethyl pyruvate (EP), an inhibitor of the high-mobility group box 1 (HMGB1) and TGF-β1 pathways, has emerged as a potential anti-fibrotic agent. Our research evaluated EP’s effects on keloid fibroblast (KF) proliferation and ECM production, employing both in vitro cell cultures and ex vivo patient-derived keloid spheroids. We also analyzed the expression levels of ECM components in keloid tissue spheroids treated with EP through immunohistochemistry. Findings revealed that EP treatment impedes the nuclear translocation of HMGB1 and diminishes KF proliferation. Additionally, EP significantly lowered mRNA and protein levels of collagen I and III by attenuating TGF-β1 and pSmad2/3 complex expression in both human dermal fibroblasts and KFs. Moreover, metalloproteinase I (MMP-1) and MMP-3 mRNA levels saw a notable increase following EP administration. In keloid spheroids, EP induced a dose-dependent reduction in ECM component expression. Immunohistochemical and western blot analyses confirmed significant declines in collagen I, collagen III, fibronectin, elastin, TGF-β, AKT, and ERK 1/2 expression levels. These outcomes underscore EP’s antifibrotic potential, suggesting its viability as a therapeutic approach for keloids. Full article
(This article belongs to the Special Issue Wound Healing and Hypertrophic Scar)
Show Figures

Figure 1

Figure 1
<p>Effects of ethyl pyruvate on HMGB1 expression in LPS-treated HDFs. (<b>A</b>) Western blotting confirmation of the nucleus and cytosolic HMGB1 expression in LPS-treated HDFs. α-tubulin was used as endogenous control for cytosolic protein; histone was used as endogenous control for nuclei protein. (<b>B</b>) Quantitative analysis of western blotting. Although not statistically significant, nuclear protein expression of HMGB1 was increased by treatment of ethyl pyruvate in LPS-treated HDFs. In contrast, cytosolic protein expression of HMGB1 was relatively decreased by ethyl pyruvate treatment. (<b>C</b>) ELISA analysis of extracellular HMGB1 concentration. EP treatment significantly reduced extracellular HMGB1 compared to LPS-treated HDFs (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Effects of ethyl pyruvate on cellular viability of TGF-β1-treated HDFs (<b>A</b>) and KFs (<b>B</b>). MTT cell proliferation assay shows that ethyl pyruvate significantly reduced the proliferation of both TGF-β1-treated HDFs and KFs in a dose-dependent manner (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Effects of ethyl pyruvate on the expression of fibrotic and cytostatic markers in HDFs and KFs. (<b>A</b>) Treatment of TGF-β significantly increased the expression of collagen I/III mRNA. Treating ethyl pyruvate significantly decreased mRNA expression as the dose increased. (<b>B</b>) Similarly, treating ethyl pyruvate to KFs showed reduced collagen I/III mRNA expression (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Effect of ethyl pyruvate on expression of fibrosis-related protein. (<b>A</b>) Western blot analysis of collagen I, III, TGF-β, and p-SMAD2/3 expression was performed; (<b>B</b>) upon densitometric analysis, collagen I, III, TGF-β, and p-SMAD2/3 protein expression levels decreased when KF cells were treated with 20mM of ethyl pyruvate (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Effects of ethyl pyruvate on the mRNA expression of MMP1 and MMP3 mRNA levels in TGF-β1-treated HDFs and KFs. (<b>A</b>) MMP1 and MMP3 mRNA levels in TGF-β1-treated HDFs increased sequentially according to the concentration of ethyl pyruvate (0 mM, 10 mM, and 20 mM). (<b>B</b>) MMP1 mRNA levels in KFs increased by treatment of ethyl pyruvate. The ratio of MMP1/TIMP1 was significantly increased by EP (20 Mm) treatment in the TGF-β1-treated HDFs and KFs (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Immunohistochemical analysis of collagen I and III, fibronectin, and elastin in ethyl pyruvate-treated keloid spheroids (<span class="html-italic">n</span> = 3). (<b>A</b>) Representative images of collagen I and III, fibronectin, and elastin IHC staining of keloid spheroids treated with ethyl pyruvate (10, 20, and 40 mM). (<b>B</b>) Semi-quantification of protein expression is shown. Collagen I and III, fibronectin, and elastin were significantly decreased in keloid spheroids following ethyl pyruvate application. Asterisks (*) over each column signify a statistically significant difference compared to the control (0 mM, white bar) group (* <span class="html-italic">p</span> &lt; 0.05). (<b>C</b>) Reduction in collagen I protein expression can be seen with western blot analysis.</p>
Full article ">Figure 7
<p>Effects of ethyl pyruvate on the expression of profibrotic factors in keloid spheroids. (<b>A</b>) Representative images of HMGB1, ERK 1/2, and AKT IHC staining of keloid spheroids treated with ethyl pyruvate (10, 20, and 40 mM). (<b>B</b>) Semi-quantification of protein expression is shown. HMGB1, ERK1/2, and AKT expression were significantly decreased after treatment with ethyl pyruvate. Asterisks (*) over each column signify a statistically significant difference compared to the control (0 mM, white bar) group. (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
15 pages, 6115 KiB  
Article
Profibrotic Inflammatory Cytokines and Growth Factors Are Predicted as the Key Targets of Uncaria gambir (Hunter) Roxb. in Keloids: An Epistatic and Molecular Simulation Approach
by Sri Suciati Ningsih, Fadilah Fadilah, Sri Widia A. Jusman, Rahimi Syaidah and Takashi Yashiro
Pharmaceuticals 2024, 17(6), 662; https://doi.org/10.3390/ph17060662 - 21 May 2024
Viewed by 1032
Abstract
Keloid is characterized as the fibrotic tissue resulting from the increase of fibroblast activity. Uncaria gambir (Hunter) Roxb. possesses bioactive compounds that have potential as antifibrotic agents, while the mechanism of action in keloid has not yet been elucidated. The aim of this [...] Read more.
Keloid is characterized as the fibrotic tissue resulting from the increase of fibroblast activity. Uncaria gambir (Hunter) Roxb. possesses bioactive compounds that have potential as antifibrotic agents, while the mechanism of action in keloid has not yet been elucidated. The aim of this study was to investigate the interaction of gambir bioactive compounds with keloid target proteins using an epistatic and molecular simulation approach. The known bioactive compounds of gambir targets and keloid-related protein targets were screened using databases. The network was constructed and analyzed to obtain the core protein targets. The targets were enriched to describe the Gene Ontology (GO) and pathway related to the proteins. Eleven targets were defined as the main targets of gambir bioactive compounds related to keloid disease. Gambiriin C, Isogambirine, and Procyanidin B1 were identified as the most promising compounds with the highest binding energy to transforming growth factor beta 1 (TGFβ1), AKT serine/threonine kinase 1 (AKT1), and matrix metallopeptidase 1 (MMP1) as the target proteins. GO enrichment and pathway analysis found that gambir bioactive compounds may act on keloid-related target proteins to regulate cell proliferation, migration, transcription, and signal transduction activity via profibrotic cytokine and growth factor signaling pathways. This study provides a reference for potential targets, compounds, and pathways to explain the mechanism of gambir against keloid. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>PPI network for the candidate targets. Black node indicates target protein and the correlation is represented as connecting colors.</p>
Full article ">Figure 2
<p>(<b>A</b>) Intersection Venn diagram between the predicted target proteins of gambir bioactive compounds, keloid-related target proteins, and high degree of connection values of keloid-related target proteins. (<b>B</b>) Compound–target network of gambir bioactive compounds targeted to keloid disease.</p>
Full article ">Figure 3
<p>Pathway prediction of the target proteins using KEGG. The red nodes are potential target proteins of <span class="html-italic">Uncaria gambir</span>, while the green nodes are relevant targets in the pathway.</p>
Full article ">Figure 4
<p>Three best molecular docking results: TGFB1 and Gambiriin C (<b>A</b>), AKT1 and Isogambirine (<b>B</b>), and MMP1 and ProcyanidinB1 (<b>C</b>).</p>
Full article ">
11 pages, 285 KiB  
Communication
No Association of Polymorphisms in the Genes Encoding Interleukin-6 and Interleukin-6 Receptor Subunit Alpha with the Risk of Keloids in Polish Patients
by Andrzej Dmytrzak, Klaudyna Lewandowska, Agnieszka Boroń, Beata Łoniewska, Natalie Grzesch, Andrzej Brodkiewicz, Jeremy S. C. Clark, Andrzej Ciechanowicz and Dorota Kostrzewa-Nowak
Int. J. Mol. Sci. 2024, 25(10), 5284; https://doi.org/10.3390/ijms25105284 - 13 May 2024
Viewed by 881
Abstract
A keloid is a benign fibroproliferative hypertrophy of scar tissue that extends outside the original wound and invades adjacent healthy skin. Keloid formation is thought to be a complex process including overactivity of the interleukin-6 signaling pathway and genetic susceptibility. The aim of [...] Read more.
A keloid is a benign fibroproliferative hypertrophy of scar tissue that extends outside the original wound and invades adjacent healthy skin. Keloid formation is thought to be a complex process including overactivity of the interleukin-6 signaling pathway and genetic susceptibility. The aim of the study was to investigate possible associations between rs1800797, rs1800796, and rs1800795 polymorphisms in the promoter of the IL6 gene encoding interleukin-6 and the rs2228145 polymorphism in the IL6R gene encoding the interleukin-6 receptor subunit alpha with the predisposition to keloids in Polish patients. The genetic polymorphisms were identified either using Polymerase Chain Reaction-Restriction Fragment Length Polymorphism (PCR-RFLP) or sequencing of samples of genomic DNA extracted from blood leukocytes of 86 adult patients with keloids and 100 newborns comprising a control group. No significant differences in the distributions of IL6 or IL6R alleles or genotypes were found between keloid patients and newborn controls. There were also no significant differences between both groups in the distribution of IL6 haplotypes. The IL6 rs1800797, rs1800796 and rs1800795 and IL6R rs2228145 polymorphisms were not found to predispose individuals in the study group to keloids. IL6 promoter haplotypes were not found to be associated with a higher risk of keloids in the studied group. Full article
(This article belongs to the Special Issue The Role of Cytokines in Diseases)
12 pages, 1749 KiB  
Review
Nodular/Keloidal Scleroderma with No Systemic Involvement—A Case Report and a Review of the Literature
by Ioana Irina Trufin, Loredana Ungureanu, Salomea-Ruth Halmágyi, Adina Patricia Apostu and Simona Corina Șenilă
J. Clin. Med. 2024, 13(9), 2662; https://doi.org/10.3390/jcm13092662 - 1 May 2024
Viewed by 1390
Abstract
Nodular or keloidal scleroderma is a rare condition with unclear cause and sporadic mentions in the medical literature. It was first recognized in the 19th century, yet its classification is still debated due to the limited number of reported cases. This rare variant [...] Read more.
Nodular or keloidal scleroderma is a rare condition with unclear cause and sporadic mentions in the medical literature. It was first recognized in the 19th century, yet its classification is still debated due to the limited number of reported cases. This rare variant of scleroderma is associated with either progressive systemic sclerosis or localized morphea. Clinically, it presents with asymptomatic nodules or plaques, resembling spontaneous keloid formation, often found on the trunk and proximal extremities. Recent literature reviews show a predominance of women with a mean age of 44 years. Diagnosis relies on clinical and histopathological findings, which usually show overlapping features of both scleroderma and true keloids, secondarily to an excessive fibrosing reaction attributed to collagen formation. We present an unusual case of a 70-year-old female patient who displayed the coexistence of two distinct subtypes of morphea (nodular/keloidal and linear), and exclusive skin involvement, which contrasts with the typical presentation of nodular/keloidal scleroderma, often associated with organ-specific disease. However, recent publications have diverged from previous ones regarding systemic sclerosis, with no systemic involvement reported between 2018 and 2024, which we evaluated in our descriptive literature review. With less than 50 cases reported in total, our case underlines the importance of recognizing this rare disease, ensuring appropriate evaluation, treatment, and follow-up. Full article
Show Figures

Figure 1

Figure 1
<p>Keloidal plaque before (<b>left</b>) and after (<b>right</b>) treatment (following two 10-day-cycles of UVA phototherapy with local application of Methoxalen solution 0.05%, separated by a one-month interval with only daily topical application of potent corticosteroids).</p>
Full article ">Figure 2
<p>Linear plaque before (<b>left</b>) and after (<b>right</b>) treatment (following two 10-day-cycles of UVA phototherapy with local application of Methoxalen solution 0.05%, separated by a one-month interval with only daily topical application of potent corticosteroids).</p>
Full article ">Figure 3
<p>Histology of keloidal scleroderma, showing hyalinized, thickened collagen bundles and spindled fibroblasts in the reticular dermis (H&amp;E staining, 5× magnification—<b>left</b>, and 10×—<b>right</b>).</p>
Full article ">
61 pages, 805 KiB  
Review
Pharmacotherapy for Keloids and Hypertrophic Scars
by Teruo Murakami and Sadayuki Shigeki
Int. J. Mol. Sci. 2024, 25(9), 4674; https://doi.org/10.3390/ijms25094674 - 25 Apr 2024
Cited by 5 | Viewed by 3287
Abstract
Keloids (KD) and hypertrophic scars (HTS), which are quite raised and pigmented and have increased vascularization and cellularity, are formed due to the impaired healing process of cutaneous injuries in some individuals having family history and genetic factors. These scars decrease the quality [...] Read more.
Keloids (KD) and hypertrophic scars (HTS), which are quite raised and pigmented and have increased vascularization and cellularity, are formed due to the impaired healing process of cutaneous injuries in some individuals having family history and genetic factors. These scars decrease the quality of life (QOL) of patients greatly, due to the pain, itching, contracture, cosmetic problems, and so on, depending on the location of the scars. Treatment/prevention that will satisfy patients’ QOL is still under development. In this article, we review pharmacotherapy for treating KD and HTS, including the prevention of postsurgical recurrence (especially KD). Pharmacotherapy involves monotherapy using a single drug and combination pharmacotherapy using multiple drugs, where drugs are administered orally, topically and/or through intralesional injection. In addition, pharmacotherapy for KD/HTS is sometimes combined with surgical excision and/or with physical therapy such as cryotherapy, laser therapy, radiotherapy including brachytherapy, and silicone gel/sheeting. The results regarding the clinical effectiveness of each mono-pharmacotherapy for KD/HTS are not always consistent but rather scattered among researchers. Multimodal combination pharmacotherapy that targets multiple sites simultaneously is more effective than mono-pharmacotherapy. The literature was searched using PubMed, Google Scholar, and Online search engines. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">
12 pages, 1876 KiB  
Article
Differential Photosensitivity of Fibroblasts Obtained from Normal Skin and Hypertrophic Scar Tissues
by Junya Kusumoto, Masaya Akashi, Hiroto Terashi and Shunsuke Sakakibara
Int. J. Mol. Sci. 2024, 25(4), 2126; https://doi.org/10.3390/ijms25042126 - 9 Feb 2024
Cited by 1 | Viewed by 1178
Abstract
It is unclear whether normal human skin tissue or abnormal scarring are photoreceptive. Therefore, this study investigated photosensitivity in normal skin tissue and hypertrophic scars. The expression of opsins, which are photoreceptor proteins, in normal dermal fibroblasts (NDFs) and hypertrophic scar fibroblasts (HSFs) [...] Read more.
It is unclear whether normal human skin tissue or abnormal scarring are photoreceptive. Therefore, this study investigated photosensitivity in normal skin tissue and hypertrophic scars. The expression of opsins, which are photoreceptor proteins, in normal dermal fibroblasts (NDFs) and hypertrophic scar fibroblasts (HSFs) was examined. After exposure to blue light (BL), changes in the expression levels of αSMA and clock-related genes, specifically PER2 and BMAL1, were examined in both fibroblast types. Opsins were expressed in both fibroblast types, with OPN3 exhibiting the highest expression levels. After peripheral circadian rhythm disruption, BL induced rhythm formation in NDFs. In contrast, although HSFs showed changes in clock-related gene expression levels, no distinct rhythm formation was observed. The expression level of αSMA was significantly higher in HSFs and decreased to the same level as that in NDFs upon BL exposure. When OPN3 knocked-down HSFs were exposed to BL, the reduction in αSMA expression was inhibited. This study showed that BL exposure directly triggers peripheral circadian synchronization in NDFs but not in HSFs. OPN3-mediated BL exposure inhibited HSFs. Although the current results did not elucidate the relationship between peripheral circadian rhythms and hypertrophic scars, they show that BL can be applied for the prevention and treatment of hypertrophic scars and keloids. Full article
(This article belongs to the Special Issue Molecular Advances in Skin Diseases 2.0)
Show Figures

Figure 1

Figure 1
<p>Gene expression levels of opsins in fibroblasts derived from normal skin tissue and hypertrophic scars. (<b>a</b>) Normal dermal fibroblasts (NDFs) derived from normal skin tissue. The expression level of OPN3 was significantly higher than that of other opsins (n = 4; relative to the expression level of β-Actin; Tukey’s test, * <span class="html-italic">p</span> &lt; 0.001). (<b>b</b>) Hypertrophic scar fibroblasts (HSFs) derived from hypertrophic scars. The expression level of opsin-3 (OPN3) was significantly higher than that of other opsins (n = 4; relative to the expression level of β-Actin; Tukey’s test, * <span class="html-italic">p</span> &lt; 0.001). Opsin-5 was not detected (ND, not detectable).</p>
Full article ">Figure 2
<p>Impact of blue light (BL) on peripheral circadian rhythms in NDFs and HSFs. (<b>a</b>) In NDFs, real-time quantitative reverse transcription PCR (qRT-PCR) analysis of brain and muscle Arnt-like protein-1 (<span class="html-italic">BMAL1</span>) and period2 (<span class="html-italic">PER2</span>) mRNA reveals the disappearance of the rhythm after 6 d of culture without light (n = 3; day 7 value as the reference; cosinor method; <span class="html-italic">BMAL1</span>: <span class="html-italic">p</span> = 0.527; <span class="html-italic">PER2</span>: <span class="html-italic">p</span> = 0.116). Adding dexamethasone (Dex) on day 7 restored the rhythm for both genes (n = 3; day 7 value as the reference; cosinor method; <span class="html-italic">BMAL1</span>: <span class="html-italic">p</span> = 0.015; <span class="html-italic">PER2</span>: <span class="html-italic">p</span> &lt; 0.001). DD, constant darkness. (<b>b</b>) In NDFs, qRT-PCR of <span class="html-italic">BMAL1</span> and <span class="html-italic">PER2</span> mRNA shows that BL irradiation forms the rhythm for restoring clock-related genes (n = 3; day 7 value as the reference; cosinor method; <span class="html-italic">BMAL1</span>, <span class="html-italic">PER2</span>: <span class="html-italic">p</span> &lt; 0.001). (<b>c</b>) In HSFs, qRT-PCR analysis of the <span class="html-italic">BMAL1</span> and <span class="html-italic">PER2</span> mRNA reveals the disappearance of the rhythm after 6 d of culture without light as well (n = 3; day 7 value as the reference; cosinor method; <span class="html-italic">BMAL1</span>: <span class="html-italic">p</span> = 0.947; <span class="html-italic">PER2</span>: <span class="html-italic">p</span> = 0.982). Despite BL irradiation, considerable variability was observed at each time point, and no statistically significant rhythm formation was observed (n = 3; day 7 value as the reference; cosinor method; <span class="html-italic">BMAL1</span>: <span class="html-italic">p</span> = 0.429; <span class="html-italic">PER2</span>: <span class="html-italic">p</span> = 0.075).</p>
Full article ">Figure 3
<p>Effect of BL and OPN3 on fibroblasts derived from normal skin tissue and hypertrophic scars. (<b>a</b>) Knockdown of <span class="html-italic">OPN3</span> in NDFs and HSFs. The knockdown efficacy was &gt;70%. (<b>b</b>) The impact of BL on <span class="html-italic">αSMA</span> expression in NDFs and HSFs. The expression level of <span class="html-italic">αSMA</span> was significantly higher in HSFs compared to that in NDFs (n = 7; Student’s t-test; <span class="html-italic">p</span> = 0.003; * <span class="html-italic">p</span> &lt; 0.05). BL irradiation significantly reduced <span class="html-italic">αSMA</span> expression in HSFs (n = 7; Tukey’s test; <span class="html-italic">p</span> = 0.025; * <span class="html-italic">p</span> &lt; 0.05). The reduction in <span class="html-italic">αSMA</span> expression caused by BL irradiation was not statistically significant following <span class="html-italic">OPN3</span> knockdown (siRNA) (n = 7; Tukey’s test; <span class="html-italic">p</span> = 0.725). (<b>c</b>) The effect of BL on <span class="html-italic">BMAL1</span> expression in NDFs and HSFs. BMAL1 expression was significantly higher in NDFs compared to that in HSFs (n = 7; Student’s t-test; <span class="html-italic">p</span> = 0.014, <span class="html-italic">p</span> = 0.006, and <span class="html-italic">p</span> = 0.011 under dark, BL, and BL+ siRNA conditions, respectively; * <span class="html-italic">p</span> &lt; 0.05). No significant changes in expression were observed following BL irradiation (n = 7; analysis of variance; NDFs, <span class="html-italic">p</span> = 0.669; HSFs, <span class="html-italic">p</span> = 0.072). (<b>d</b>) The effect of BL on <span class="html-italic">PER2</span> expression in NDFs and HSFs. <span class="html-italic">PER2</span> expression was significantly higher in NDFs compared to that in HSFs, under dark conditions, and BL irradiation following <span class="html-italic">OPN3</span> knockdown (siRNA) (n = 7; Student’s t-test; <span class="html-italic">p</span> = 0.010, <span class="html-italic">p</span> = 0.059, and <span class="html-italic">p</span> = 0.006 under dark, BL, and BL+ siRNA conditions, respectively; * <span class="html-italic">p</span> &lt; 0.05). No significant changes in expression were observed following BL irradiation (n = 7; analysis of variance; NDF, <span class="html-italic">p</span> = 0.862; HSF, <span class="html-italic">p</span> = 0.407).</p>
Full article ">Figure 4
<p>Impact of BL on cell viability in fibroblasts derived from normal skin tissue and hypertrophic scars. (<b>a</b>) Effect of BL on cell viability in NDFs. BL irradiation significantly increased cell proliferation in NDFs (n = 7; dark condition is used as the reference; Student’s <span class="html-italic">t</span>-test; <span class="html-italic">p</span> = 0.017; * <span class="html-italic">p</span> &lt; 0.05). Knockdown of <span class="html-italic">OPN3</span> (siRNA; efficacy: 90.3%) followed by BL irradiation did not show any increase in cell proliferation (n = 7; dark condition is used as the reference; Student’s t-test; <span class="html-italic">p</span> = 0.345). (<b>b</b>) Effect of BL on cell viability in HSFs. BL irradiation significantly reduced cell proliferation in HSFs (n = 7; dark condition is used as the reference; Student’s t-test; <span class="html-italic">p</span> &lt; 0.001; * <span class="html-italic">p</span> &lt; 0.05). Knockdown of <span class="html-italic">OPN3</span> (siRNA; efficacy: 74.3%) followed by BL irradiation resulted in a significant increase in cell proliferation (n = 7; dark condition is used as the reference; Student’s t-test; <span class="html-italic">p</span> &lt; 0.001; * <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
13 pages, 3541 KiB  
Review
Review of Leishmaniasis Treatment: Can We See the Forest through the Trees?
by Moshe Shmueli and Shalom Ben-Shimol
Pharmacy 2024, 12(1), 30; https://doi.org/10.3390/pharmacy12010030 - 8 Feb 2024
Cited by 7 | Viewed by 2745
Abstract
There are three known clinical syndromes of leishmaniasis: cutaneous (CL), mucocutaneous (MCL), and visceral disease (VL). In MCL and VL, treatment must be systemic (either oral or intravenous), while CL treatment options vary and include observation-only localized/topical treatment, oral medications, or parenteral drugs. [...] Read more.
There are three known clinical syndromes of leishmaniasis: cutaneous (CL), mucocutaneous (MCL), and visceral disease (VL). In MCL and VL, treatment must be systemic (either oral or intravenous), while CL treatment options vary and include observation-only localized/topical treatment, oral medications, or parenteral drugs. Leishmaniasis treatment is difficult, with several factors to be considered. First, the efficacy of treatments varies among different species of parasites prevalent in different areas on the globe, with each species having a unique clinical presentation and resistance profile. Furthermore, leishmaniasis is a neglected tropical disease (NTD), resulting in a lack of evidence-based knowledge regarding treatment. Therefore, physicians often rely on case reports or case series studies, in the absence of randomized controlled trials (RCT), to assess treatment efficacy. Second, defining cure, especially in CL and MCL, may be difficult, as death of the parasite can be achieved in most cases, while the aesthetic result (e.g., scars) is hard to predict. This is a result of the biological nature of the disease, often diagnosed late in the course of disease (with possible keloid formation, etc.). Third, physicians must consider treatment ease of use and the safety profile of possible treatments. Thus, topical or oral treatments (for CL) are desirable and promote adherence. Fourth, the cost of the treatment is an important consideration. In this review, we aim to describe the diverse treatment options for different clinical manifestations of leishmaniasis. For each currently available treatment, we will discuss the various considerations mentioned above (efficacy, ease of use, safety, and cost). Full article
Show Figures

Figure 1

Figure 1
<p>Visceral leishmaniasis in a 4-year-old patient. Hepatosplenomegaly (<b>left</b>) and amastigotes in bone marrow biopsy (<b>right</b>) are notable.</p>
Full article ">Figure 2
<p>Cutaneous leishmaniasis in a 24-year-old student diagnosed with <span class="html-italic">L. braziliensis</span> without adequate systemic treatment; patient is at risk for developing MCL.</p>
Full article ">Figure 3
<p>Cutaneous leishmaniasis in a 6-month-old infant, diagnosed with <span class="html-italic">L. major</span> infection (<b>left</b>). After one week of systemic treatment with IV Liposomal amphotericin B, the patient is showing clinical improvement (<b>right</b>).</p>
Full article ">Figure 4
<p>Cutaneous leishmaniasis in a 54-year-old patient (<b>left</b>, <b>middle</b>). Following miltefosine treatment, the patient showed improvement (<b>right</b>).</p>
Full article ">
27 pages, 2425 KiB  
Review
Insights into How Plant-Derived Extracts and Compounds Can Help in the Prevention and Treatment of Keloid Disease: Established and Emerging Therapeutic Targets
by Yong Chool Boo
Int. J. Mol. Sci. 2024, 25(2), 1235; https://doi.org/10.3390/ijms25021235 - 19 Jan 2024
Cited by 3 | Viewed by 2376
Abstract
Keloid is a disease in which fibroblasts abnormally proliferate and synthesize excessive amounts of extracellular matrix, including collagen and fibronectin, during the healing process of skin wounds, causing larger scars that exceed the boundaries of the original wound. Currently, surgical excision, cryotherapy, radiation, [...] Read more.
Keloid is a disease in which fibroblasts abnormally proliferate and synthesize excessive amounts of extracellular matrix, including collagen and fibronectin, during the healing process of skin wounds, causing larger scars that exceed the boundaries of the original wound. Currently, surgical excision, cryotherapy, radiation, laser treatment, photodynamic therapy, pressure therapy, silicone gel sheeting, and pharmacotherapy are used alone or in combinations to treat this disease, but the outcomes are usually unsatisfactory. The purpose of this review is to examine whether natural products can help treat keloid disease. I introduce well-established therapeutic targets for this disease and various other emerging therapeutic targets that have been proposed based on the phenotypic difference between keloid-derived fibroblasts (KFs) and normal epidermal fibroblasts (NFs). We then present recent studies on the biological effects of various plant-derived extracts and compounds on KFs and NFs. Associated ex vivo, in vivo, and clinical studies are also presented. Finally, we discuss the mechanisms of action of the plant-derived extracts and compounds, the pros and cons, and the future tasks for natural product-based therapy for keloid disease, as compared with existing other therapies. Extracts of Astragalus membranaceus, Salvia miltiorrhiza, Aneilema keisak, Galla Chinensis, Lycium chinense, Physalis angulate, Allium sepa, and Camellia sinensis appear to modulate cell proliferation, migration, and/or extracellular matrix (ECM) production in KFs, supporting their therapeutic potential. Various phenolic compounds, terpenoids, alkaloids, and other plant-derived compounds could modulate different cell signaling pathways associated with the pathogenesis of keloids. For now, many studies are limited to in vitro experiments; additional research and development are needed to proceed to clinical trials. Many emerging therapeutic targets could accelerate the discovery of plant-derived substances for the prevention and treatment of keloid disease. I hope that this review will bridge past, present, and future research on this subject and provide insight into new therapeutic targets and pharmaceuticals, aiming for effective keloid treatment. Full article
(This article belongs to the Special Issue New Insights in Natural Bioactive Compounds 2.0)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The biological effects of several existing treatments and quercetin on keloid-derived fibroblasts (KFs). Sharp arrows indicate stimulation or upregulation and blunted arrows indicate inhibition or downregulation. Black lines indicate the interaction between cellular factors and events (black letters) and red lines indicate the effects of various treatments (blue letters).</p>
Full article ">Figure 2
<p>Chemical structures of natural products.</p>
Full article ">Figure 2 Cont.
<p>Chemical structures of natural products.</p>
Full article ">Figure 3
<p>Cell cycle arrest by plant-derived extracts and compounds. The cell cycle consists of gap (G<sub>1</sub>) phase, synthesis (S) phase, (G<sub>2</sub>) phase, mitosis (M) phase, and (G<sub>0</sub>) phase. Different colors are used to represent different phases of cell cycle arrest induced by the treatments with matching colors.</p>
Full article ">Figure 4
<p>Potential mechanisms for cell cycle arrest and apoptosis induction of several plant-derived extracts and compounds in KFs. Sharp arrows indicate stimulation or upregulation and blunted arrows indicate inhibition or downregulation. Black lines indicate the interaction between cellular factors and events (black letters) and red lines indicate the effects of various treatments (blue letters).</p>
Full article ">Figure 5
<p>Effects of plant-derived extracts and compounds on cell signaling pathways in KFs. Sharp arrows indicate stimulation or upregulation and blunted arrows indicate inhibition or downregulation. Black lines indicate the interaction between cellular factors and events (black and white letters) and red lines indicate the effects of various plant-derived extracts and compounds (different color or grey letters).</p>
Full article ">
16 pages, 5127 KiB  
Article
Derazantinib Inhibits the Bioactivity of Keloid Fibroblasts via FGFR Signaling
by Shuqia Xu, Yongkang Zhu, Peng Wang, Shaohai Qi and Bin Shu
Biomedicines 2023, 11(12), 3220; https://doi.org/10.3390/biomedicines11123220 - 5 Dec 2023
Viewed by 1321
Abstract
Keloids are common benign cutaneous pathological fibrous proliferation diseases, which are difficult to cure and easily recur. Studies have shown that fibroblast growth factor receptor-1 (FGFR1) was enhanced in pathological fibrous proliferation diseases, such as cirrhosis and idiopathic pulmonary fibrosis (IPF), suggesting the [...] Read more.
Keloids are common benign cutaneous pathological fibrous proliferation diseases, which are difficult to cure and easily recur. Studies have shown that fibroblast growth factor receptor-1 (FGFR1) was enhanced in pathological fibrous proliferation diseases, such as cirrhosis and idiopathic pulmonary fibrosis (IPF), suggesting the FGFR1 pathway has potential for keloid treatment. Derazantinib is a selective FGFR inhibitor with antiproliferative activity in in vitro and in vivo models. The present study determined the effects of derazantinib on human keloid fibroblasts (KFs). Cell viability assay, migration assay, invasion assay, immunofluorescence staining, quantitative polymerase chain reaction, Western blot analysis, HE staining, Masson staining, and immunohistochemical analysis were used to analyze the KFs and keloid xenografts. In this study, we found that derazantinib inhibited the proliferation, migration, invasion, and collagen production of KFs in vitro. The transcription and expression of plasminogen activator inhibitor-1 (PAI-1), which is closely related to collagen deposition and tissue fibrosis, was significantly inhibited. Also, derazantinib inhibited the expression of FGFR1 and PAI-1 and reduced the weight of the implanted keloid from the xenograft mice model. These findings suggest that derazantinib may be a potent therapy for keloids via FGFR signaling. Full article
(This article belongs to the Section Cell Biology and Pathology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Immunofluorescence staining of keloids and normal skin. The length of the scale is 50 μm. (<b>B</b>) The CCK-8 assay demonstrated a significant inhibitory effect of derazantinib on the proliferation of keloid fibroblasts (KFs). (<b>C</b>) Ki67 immunofluorescence staining showed that Ki67 expression was decreased in the experiment group. The length of the scale is 50 μm. (<b>D</b>) The Calcein AM-PI staining showed that Calcein-AM-positive cells were decreased and PI-positive cells were increased in the experiment group. The length of the scale is 100 μm. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 1 Cont.
<p>(<b>A</b>) Immunofluorescence staining of keloids and normal skin. The length of the scale is 50 μm. (<b>B</b>) The CCK-8 assay demonstrated a significant inhibitory effect of derazantinib on the proliferation of keloid fibroblasts (KFs). (<b>C</b>) Ki67 immunofluorescence staining showed that Ki67 expression was decreased in the experiment group. The length of the scale is 50 μm. (<b>D</b>) The Calcein AM-PI staining showed that Calcein-AM-positive cells were decreased and PI-positive cells were increased in the experiment group. The length of the scale is 100 μm. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>(<b>A</b>,<b>B</b>) The scratch assay showed that derazantinib inhibited the migration of KFs. The length of the scale is 500 μm. * <span class="html-italic">p</span> &lt; 0.05 compared with the control group, # <span class="html-italic">p</span> &lt; 0.05 compared with the 0.31 μM group, <span>$</span> <span class="html-italic">p</span> &lt; 0.05 compared with the 0.63 μM group, and &amp; <span class="html-italic">p</span> &lt; 0.05 compared with the 1.25 μM group. (<b>C</b>) The transwell assay showed that derazantinib inhibited the invasion of KFs. The length of the scale is 100 μm.</p>
Full article ">Figure 3
<p>KFs were treated with 2.5 μmol derazantinib or without derazantinib for 48 h. (<b>A</b>) The PCR showed that the expression of fibrotic genes was suppressed in the experiment group. (<b>B</b>–<b>D</b>) The immunofluorescence staining showed that collagen I, α-SMA, and PAI-1 expression were decreased in the derazantinib group. The length of the scale is 50 μm. (<b>E</b>,<b>F</b>) The WB showed that the protein production of a-SMA, collagen I, PAI-1, and FGFR1 in the derazantinib group was suppressed. (<b>G</b>) The PCR showed that the expression of PI3K and JNK genes was suppressed in the experiment group. (<b>H</b>) The WB showed that the expression of ERK, p-ERK, AKT, TGF-β, p-AKT, and PI3K in the derazantinib group was suppressed. No significant effect was found on the SMAD expression levels. * <span class="html-italic">p</span> &lt; 0.05 compared to control group.</p>
Full article ">Figure 3 Cont.
<p>KFs were treated with 2.5 μmol derazantinib or without derazantinib for 48 h. (<b>A</b>) The PCR showed that the expression of fibrotic genes was suppressed in the experiment group. (<b>B</b>–<b>D</b>) The immunofluorescence staining showed that collagen I, α-SMA, and PAI-1 expression were decreased in the derazantinib group. The length of the scale is 50 μm. (<b>E</b>,<b>F</b>) The WB showed that the protein production of a-SMA, collagen I, PAI-1, and FGFR1 in the derazantinib group was suppressed. (<b>G</b>) The PCR showed that the expression of PI3K and JNK genes was suppressed in the experiment group. (<b>H</b>) The WB showed that the expression of ERK, p-ERK, AKT, TGF-β, p-AKT, and PI3K in the derazantinib group was suppressed. No significant effect was found on the SMAD expression levels. * <span class="html-italic">p</span> &lt; 0.05 compared to control group.</p>
Full article ">Figure 3 Cont.
<p>KFs were treated with 2.5 μmol derazantinib or without derazantinib for 48 h. (<b>A</b>) The PCR showed that the expression of fibrotic genes was suppressed in the experiment group. (<b>B</b>–<b>D</b>) The immunofluorescence staining showed that collagen I, α-SMA, and PAI-1 expression were decreased in the derazantinib group. The length of the scale is 50 μm. (<b>E</b>,<b>F</b>) The WB showed that the protein production of a-SMA, collagen I, PAI-1, and FGFR1 in the derazantinib group was suppressed. (<b>G</b>) The PCR showed that the expression of PI3K and JNK genes was suppressed in the experiment group. (<b>H</b>) The WB showed that the expression of ERK, p-ERK, AKT, TGF-β, p-AKT, and PI3K in the derazantinib group was suppressed. No significant effect was found on the SMAD expression levels. * <span class="html-italic">p</span> &lt; 0.05 compared to control group.</p>
Full article ">Figure 4
<p>(<b>A</b>) In an athymic nude mouse model (<span class="html-italic">n</span> = 12), keloid tissue from each patient (<span class="html-italic">n</span> = 3) was transplanted subcutaneously into 4 mice, respectively. The weight loss of the keloids transplanted into the nude mice was statistically different between the control group and the derazantinib group. * <span class="html-italic">p</span> &lt; 0.05 compared with the control group. (<b>B</b>) Masson staining and HE staining showed less abundant collagen and micro-vessels in the derazantinib groups. The length of the scale is 500 μm. (<b>C</b>) The immunofluorescence staining showed that the expression of FGFR1 and type I collagen was decreased in the derazantinib group compared with the control group. The length of the scale is 50 μm. (<b>D</b>) The immunofluorescence staining showed that the expression of PAI-1 and a-SMA was consistently reduced in the derazantinib group compared with the control group. The length of the scale is 50 μm. (DZB—derazantinib; BMT—betamethasone).</p>
Full article ">Figure 4 Cont.
<p>(<b>A</b>) In an athymic nude mouse model (<span class="html-italic">n</span> = 12), keloid tissue from each patient (<span class="html-italic">n</span> = 3) was transplanted subcutaneously into 4 mice, respectively. The weight loss of the keloids transplanted into the nude mice was statistically different between the control group and the derazantinib group. * <span class="html-italic">p</span> &lt; 0.05 compared with the control group. (<b>B</b>) Masson staining and HE staining showed less abundant collagen and micro-vessels in the derazantinib groups. The length of the scale is 500 μm. (<b>C</b>) The immunofluorescence staining showed that the expression of FGFR1 and type I collagen was decreased in the derazantinib group compared with the control group. The length of the scale is 50 μm. (<b>D</b>) The immunofluorescence staining showed that the expression of PAI-1 and a-SMA was consistently reduced in the derazantinib group compared with the control group. The length of the scale is 50 μm. (DZB—derazantinib; BMT—betamethasone).</p>
Full article ">
Back to TopTop