[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (2,723)

Search Parameters:
Keywords = osteoblasts

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
41 pages, 1772 KiB  
Review
Exploring the Role of Hormones and Cytokines in Osteoporosis Development
by Egemen Umur, Safiye Betül Bulut, Pelin Yiğit, Emirhan Bayrak, Yaren Arkan, Fahriye Arslan, Engin Baysoy, Gizem Kaleli-Can and Bugra Ayan
Biomedicines 2024, 12(8), 1830; https://doi.org/10.3390/biomedicines12081830 - 12 Aug 2024
Abstract
The disease of osteoporosis is characterized by impaired bone structure and an increased risk of fractures. There is a significant impact of cytokines and hormones on bone homeostasis and the diagnosis of osteoporosis. As defined by the World Health Organization (WHO), osteoporosis is [...] Read more.
The disease of osteoporosis is characterized by impaired bone structure and an increased risk of fractures. There is a significant impact of cytokines and hormones on bone homeostasis and the diagnosis of osteoporosis. As defined by the World Health Organization (WHO), osteoporosis is defined as having a bone mineral density (BMD) that is 2.5 standard deviations (SD) or more below the average for young and healthy women (T score < −2.5 SD). Cytokines and hormones, particularly in the remodeling of bone between osteoclasts and osteoblasts, control the differentiation and activation of bone cells through cytokine networks and signaling pathways like the nuclear factor kappa-B ligand (RANKL)/the receptor of RANKL (RANK)/osteoprotegerin (OPG) axis, while estrogen, parathyroid hormones, testosterone, and calcitonin influence bone density and play significant roles in the treatment of osteoporosis. This review aims to examine the roles of cytokines and hormones in the pathophysiology of osteoporosis, evaluating current diagnostic methods, and highlighting new technologies that could help for early detection and treatment of osteoporosis. Full article
(This article belongs to the Section Endocrinology and Metabolism Research)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of healthy bone (<b>A</b>), the two main cell types involved in bone remodeling: osteoblasts and osteoclasts (<b>B</b>) and osteoporotic bone (<b>C</b>). A healthy bone is characterized by a well-defined, thick trabecular structure, sufficient mineralization, and minimal signs of fractures or degradation. The structure of osteoporotic bone is characterized by advanced bone loss and weakening, as well as significant reductions in density and thickness of trabecular bone, which results in the appearance of porous and fragile bone. This figure was created using <a href="https://app.biorender.com/" target="_blank">https://app.biorender.com/</a>, accessed on 18 June 2024.</p>
Full article ">Figure 2
<p>Cells expressing cytokine receptors. This figure was created using <a href="https://app.biorender.com/" target="_blank">https://app.biorender.com/</a>, accessed on 18 June 2024.</p>
Full article ">Figure 3
<p>The upregulation and downregulation of cytokines in different cells during osteoporosis. (TNF-α can exhibit both upregulation and downregulation.) This figure was created using <a href="https://app.biorender.com/" target="_blank">https://app.biorender.com/</a>, accessed on 18 June 2024.</p>
Full article ">
4 pages, 191 KiB  
Editorial
Research on Bone Cells in Health and Disease
by Dávid S. Győri
Int. J. Mol. Sci. 2024, 25(16), 8758; https://doi.org/10.3390/ijms25168758 (registering DOI) - 12 Aug 2024
Viewed by 193
Abstract
Bone-forming osteoblasts, osteocytes, and bone-resorbing osteoclasts are responsible for life-long skeletal remodeling [...] Full article
(This article belongs to the Special Issue Research on Bone Cells in Health and Disease)
20 pages, 3114 KiB  
Review
The Manganese–Bone Connection: Investigating the Role of Manganese in Bone Health
by Gulaim Taskozhina, Gulnara Batyrova, Gulmira Umarova, Zhamilya Issanguzhina and Nurgul Kereyeva
J. Clin. Med. 2024, 13(16), 4679; https://doi.org/10.3390/jcm13164679 - 9 Aug 2024
Viewed by 440
Abstract
The complex relationship between trace elements and skeletal health has received increasing attention in the scientific community. Among these minerals, manganese (Mn) has emerged as a key element affecting bone metabolism and integrity. This review examines the multifaceted role of Mn in bone [...] Read more.
The complex relationship between trace elements and skeletal health has received increasing attention in the scientific community. Among these minerals, manganese (Mn) has emerged as a key element affecting bone metabolism and integrity. This review examines the multifaceted role of Mn in bone health, including its effects on bone regeneration, mineralization, and overall skeletal strength. This review article is based on a synthesis of experimental models, epidemiologic studies, and clinical trials of the mechanisms of the effect of Mn on bone metabolism. Current research data show that Mn is actively involved in the processes of bone remodeling by modulating the activity of osteoblasts and osteoclasts, as well as the main cells that regulate bone formation and resorption. Mn ions have a profound effect on bone mineralization and density by intricately regulating signaling pathways and enzymatic reactions in these cells. Additionally, Mn superoxide dismutase (MnSOD), located in bone mitochondria, plays a crucial role in osteoclast differentiation and function, protecting osteoclasts from oxidative damage. Understanding the nuances of Mn’s interaction with bone is essential for optimizing bone strategies, potentially preventing and managing skeletal diseases. Key findings include the stimulation of osteoblast proliferation and differentiation, the inhibition of osteoclastogenesis, and the preservation of bone mass through the RANK/RANKL/OPG pathway. These results underscore the importance of Mn in maintaining bone health and highlight the need for further research into its therapeutic potential. Full article
Show Figures

Figure 1

Figure 1
<p>The role of manganese (Mn) in bone cellular and molecular functions. The trace element Mn, with its various biochemical and physiological effects, participates in the synthesis of bone matrix, the inhibition of the formation of osteoclast-like cells, antioxidant function with the enzyme Mn superoxide dismutase (MnSOD), and mRNA expression of RANKL receptors; it also contributes to cell adhesion with extracellular matrix proteins, regulating osteoid formation. It also protects cartilage and stimulates chondrocyte growth through ZIP14. This is important for its integrin-activating functions, which contribute to the adhesion, integrity, and proliferation of osteoblasts.</p>
Full article ">Figure 2
<p>Manganese (Mn) superoxide dismutase (MnSOD) in the bone resorption [<a href="#B52-jcm-13-04679" class="html-bibr">52</a>]. RANKL-induced differentiation of macrophages into osteoclasts and the role of MnSOD in managing oxidative stress during bone resorption are depicted. RANKL binds to the RANK receptors on these cells, promoting their maturation. During bone resorption, superoxide (O<sub>2</sub><sup>−</sup>) is produced as a byproduct, and the mitochondrial enzyme MnSOD catalyzes the conversion of O<sub>2</sub><sup>−</sup> into hydrogen peroxide (H<sub>2</sub>O<sub>2</sub>) and oxygen (O<sub>2</sub>), thereby reducing oxidative stress. H<sub>2</sub>O<sub>2</sub> is subsequently converted into water (H<sub>2</sub>O), providing cellular protection. This process underscores the critical role of MnSOD in maintaining the functionality and integrity of osteoclasts during bone resorption.</p>
Full article ">Figure 3
<p>Manganese (Mn) and bone remodeling [<a href="#B72-jcm-13-04679" class="html-bibr">72</a>]. The role of Mn in bone remodeling highlights its dual impact on osteoclasts and osteoblasts. Mn promotes osteoclast differentiation by enhancing the RANKL/RANK signaling pathway, where RANKL binds to RANK receptors on osteoclast progenitor cells, leading to their maturation into osteoclasts. Mature osteoclasts resorb bone, a process associated with oxidative stress, during which O<sub>2</sub><sup>−</sup> is converted into less harmful molecules by the mitochondrial enzyme Mn superoxide dismutase (MnSOD). Concurrently, Mn inhibits the PI3K/AKT and WNT/β-catenin signaling pathways in mesenchymal stem cells (MSCs), thereby reducing the differentiation and activity of osteoblasts. This dual mechanism underscores the essential role of Mn in maintaining bone mass and integrity, ensuring effective bone regeneration and homeostasis by balancing bone resorption and formation.</p>
Full article ">Figure 4
<p>Molecular mechanisms of manganese (Mn) metabolism. The molecular pathways involved in Mn metabolism highlight its absorption, transport, and accumulation in the body. Mn ions (Mn<sup>2+</sup>) are absorbed in the intestines through the divalent metal transporter 1 (DMT1). After absorption, Mn<sup>2+</sup> ions enter the bloodstream and are transported in a complex with proteins. The liver, considered the central organ in Mn metabolism, plays a crucial role in processing and regulating Mn levels. Mn is then distributed from the liver to various tissues throughout the body, with a significant accumulation in the bones. This high accumulation in bones underscores the essential role of Mn in skeletal health.</p>
Full article ">Figure 5
<p>Manganese (Mn) hemostasis in the bone [<a href="#B90-jcm-13-04679" class="html-bibr">90</a>]. The cellular mechanisms involved in maintaining Mn homeostasis in bones reveal the key physiological functions of Mn transporters and regulators, including ZIP8, ZNT10, and ZIP14. The process begins with the intake of Mn from food, where ZIP8 facilitates the intracellular accumulation of Mn<sup>2+</sup> ions. These Mn<sup>2+</sup> ions enter the bloodstream and are transported to various tissues, including bones and liver hepatocytes. The transport of Mn<sup>2+</sup> ions into bones and other tissues is facilitated by the ubiquitously expressed ZIP14. Mn<sup>2+</sup> ions reach the liver, where they undergo further processing and regulation. The ZIP10 transporter acts as an apical exporter, transporting Mn from the blood to the lumen of the small intestine for excretion in feces. These intricate regulatory mechanisms ensure the balance of Mn, which is crucial for maintaining bone health and overall metabolic homeostasis.</p>
Full article ">
15 pages, 13327 KiB  
Article
Turning Portunus pelagicus Shells into Biocompatible Scaffolds for Bone Regeneration
by Louisa Candra Devi, Hendrik Satria Dwi Putra, Nyoman Bayu Wisnu Kencana, Ajiteru Olatunji and Agustina Setiawati
Biomedicines 2024, 12(8), 1796; https://doi.org/10.3390/biomedicines12081796 - 7 Aug 2024
Viewed by 295
Abstract
Bone tissue engineering (BTE) provides an alternative for addressing bone defects by integrating cells, a scaffold, and bioactive growth factors to stimulate tissue regeneration and repair, resulting in effective bioengineered tissue. This study focuses on repurposing chitosan from blue swimming crab (Portunus [...] Read more.
Bone tissue engineering (BTE) provides an alternative for addressing bone defects by integrating cells, a scaffold, and bioactive growth factors to stimulate tissue regeneration and repair, resulting in effective bioengineered tissue. This study focuses on repurposing chitosan from blue swimming crab (Portunus pelagicus) shell waste as a composite scaffold combined with HAP and COL I to improve biocompatibility, porosity, swelling, and mechanical properties. The composite scaffold demonstrated nearly 60% porosity with diameters ranging from 100–200 μm with an interconnected network that structurally mimics the extracellular matrix. The swelling ratio of the scaffold was measured at 208.43 ± 14.05%, 248.93 ± 4.32%, 280.01 ± 1.26%, 305.44 ± 20.71%, and 310.03 ± 17.94% at 1, 3, 6, 12, and 24 h, respectively. Thus, the Portunus pelagicus scaffold showed significantly lower degradation ratios of 5.64 ± 1.89%, 14.34 ± 8.59%, 19.57 ± 14.23%, and 29.13 ± 9.87% for 1 to 4 weeks, respectively. The scaffold supports osteoblast attachment and proliferation for 7 days. Waste from Portunus pelagicus shells has emerged as a prospective source of chitosan with potential application in tissue engineering. Full article
(This article belongs to the Special Issue Advances in 3D Printing and Biomaterials in Tissue Engineering)
Show Figures

Figure 1

Figure 1
<p>Work scheme of repurposing the shell of Portunus pelagicus for the engineering of bone tissue. This figure was prepared using BioRender.</p>
Full article ">Figure 2
<p><span class="html-italic">Portunus pelagicus</span> shell chitosan extraction and characterization. (<b>a</b>) The procedure for extracting chitosan from a <span class="html-italic">Portunus pelagicus</span> (PP) shell; (<b>b</b>) The chitosan from shell and control’s FTIR spectra (green for shells scaffold and black for control scaffold).</p>
Full article ">Figure 3
<p>Scaffold composite engineering of extracted chitosan. (<b>a</b>) Working design of scaffold composite synthesis; (<b>b</b>) Macroscopic profile <span class="html-italic">Portunus pelagicus’</span> shell (PP’s shell) and control scaffold; (<b>c</b>) FTIR spectra of engineered and control scaffolds (green for shells scaffold and black for control scaffold).</p>
Full article ">Figure 4
<p>Porosity and microscopic characteristics of engineered scaffolds. (<b>a</b>) Work design of porosity measurement method; (<b>b</b>) porosity of control and <span class="html-italic">Portunus pelagicus’</span> shell (PP’s shell) scaffolds; (<b>c</b>) pore diameter of the scaffolds; (<b>d</b>) SEM images of control and PP’s shell scaffold; (<b>e</b>) scheme of scaffold compressive strength measurement; (<b>f</b>) compressive strength of the scaffolds. All quantitative data are given in means ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 5
<p>The ratio of swelling and degradation ratio of the fabricated scaffold. (<b>a</b>) Swelling scaffold at 1 and 24 h in PBS, (<b>b</b>) The ratio of swelling of control and shell scaffolds at 0, 3, 6, 12 and 24 h; (<b>c</b>) Degradation morphology of the scaffold at 1- and 2-weeks ratio of the scaffolds, (<b>d</b>) The ratio of degradation of of control and shell scaffolds at 1, 2, 3 and 4 weeks. The data are presented as means ± SD (<span class="html-italic">n</span> = 4).</p>
Full article ">Figure 6
<p>Cell biocompatibility on chitosan derived <span class="html-italic">Portunus pelagicus</span>’ shell scaffolds. (<b>a</b>) Images of GFP-expressing fibroblasts on the scaffold at 0, 3, 5, and 7 days; (<b>b</b>) absorbance at 450 nm after MG-63 cells grown on control and shell scaffold and being treated with tetrazolium salt; (<b>c</b>) SEM profile of MG-63 osteoblast-like cell morphology on the scaffold after 7 days. The data are presented in means ± SD (<span class="html-italic">n</span> = 3). The statistical significance was computed using a one-way ANOVA followed by the Tukey Test; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
13 pages, 584 KiB  
Article
Pro-Osteogenic Effect of the Nutraceutical BlastiMin Complex® in Women with Osteoporosis or Osteopenia: An Open Intervention Clinical Trial
by Sofia Sabatelli, Emanuele-Salvatore Scarpa, Angelica Giuliani, Chiara Giordani, Jacopo Sabbatinelli, Maria Rita Rippo, Sara Cabodi, Barbara Petrini, Giancarlo Balercia and Gilberta Giacchetti
Int. J. Mol. Sci. 2024, 25(16), 8565; https://doi.org/10.3390/ijms25168565 - 6 Aug 2024
Viewed by 266
Abstract
Osteoporosis is a chronic disease that affects millions of patients worldwide and is characterized by low bone mineral density (BMD) and increased risk of fractures. Notably, natural molecules can increase BMD and exert pro-osteogenic effects. Noteworthily, the nutraceutical BlastiMin Complex® (Mivell, Italy, [...] Read more.
Osteoporosis is a chronic disease that affects millions of patients worldwide and is characterized by low bone mineral density (BMD) and increased risk of fractures. Notably, natural molecules can increase BMD and exert pro-osteogenic effects. Noteworthily, the nutraceutical BlastiMin Complex® (Mivell, Italy, European Patent Application EP4205733A1) can induce differentiation of human bone marrow mesenchymal stem cells (BM-MSCs) in osteoblasts and can exert in vitro pro-osteogenic and anti-inflammatory effects. Thus, the purpose of this study was to verify the effects of BlastiMin Complex® on bone turnover markers (BTMs) and BMD in patients with senile and postmenopausal osteopenia or osteoporosis. The efficacy of BlastiMin Complex® on BTMs in serum was evaluated through biochemical assays. BMD values were analyzed by dual-energy X-ray absorptiometry (DXA) and Radiofrequency Echographic Multi Spectrometry (R.E.M.S.) techniques, and the SNPs with a role in osteoporosis development were evaluated by PCR. Clinical data obtained after 12 months of treatment showed an increase in bone turnover index, a decrease in C-reactive protein levels, and a remarkable increase in P1NP levels, indicating the induction of osteoblast proliferation and activity in the cohort of 100% female patients recruited for the study. These findings show that the nutraceutical BlastiMin Complex® could be used as an adjuvant in combination with synthetic drugs for the treatment of osteoporosis pathology. Full article
(This article belongs to the Topic Bone as an Endocrine Organ)
Show Figures

Figure 1

Figure 1
<p>Flowchart diagram of the BlastiMin Complex<sup>®</sup> clinical trial.</p>
Full article ">
19 pages, 4612 KiB  
Article
Tibial Damage Caused by T-2 Toxin in Goslings: Bone Dysplasia, Poor Bone Quality, Hindered Chondrocyte Differentiation, and Imbalanced Bone Metabolism
by Wang Gu, Lie Hou, Qiang Bao, Qi Xu and Guohong Chen
Animals 2024, 14(15), 2281; https://doi.org/10.3390/ani14152281 - 5 Aug 2024
Viewed by 315
Abstract
T-2 toxin, the most toxic type A trichothecene, is widely present in grain and animal feed, causing growth retardation and tissue damage in poultry. Geese are more sensitive to T-2 toxin than chickens and ducks. Although T-2 toxin has been reported to cause [...] Read more.
T-2 toxin, the most toxic type A trichothecene, is widely present in grain and animal feed, causing growth retardation and tissue damage in poultry. Geese are more sensitive to T-2 toxin than chickens and ducks. Although T-2 toxin has been reported to cause tibial growth plate (TGP) chondrodysplasia in chickens, tibial damage caused by T-2 toxin in geese has not been fully demonstrated. This study aims to investigate the adverse effects of T-2 toxin on tibial bone development, bone quality, chondrocyte differentiation, and bone metabolism. Here, forty-eight one-day-old male Yangzhou goslings were randomly divided into four groups and daily gavaged with T-2 toxin at concentrations of 0, 0.5, 1.0, and 2.0 mg/kg body weight for 21 days, respectively. The development of gosling body weight and size was determined by weighing and taking body measurements after exposure to different concentrations of T-2 toxin. Changes in tibial development and bone characteristics were determined by radiographic examination, phenotypic measurements, and bone quality and composition analyses. Chondrocyte differentiation in TGP and bone metabolism was characterized by cell morphology, tissue gene-specific expression, and serum marker levels. Results showed that T-2 toxin treatment resulted in a lower weight, volume, length, middle width, and middle circumference of the tibia in a dose-dependent manner (p < 0.05). Moreover, decreased bone-breaking strength, bone mineral density, and contents of ash, Ca, and P in the tibia were observed in T-2 toxin-challenged goslings (p < 0.05). In addition, T-2 toxin not only reduced TGP height (p < 0.05) but also induced TGP chondrocytes to be disorganized with reduced numbers and indistinct borders. As expected, the apoptosis-related genes (CASP9 and CASP3) were significantly up-regulated in chondrocytes challenged by T-2 toxin with a dose dependence, while cell differentiation and maturation-related genes (BMP6, BMP7, SOX9, and RUNX2) were down-regulated (p < 0.05). Considering bone metabolism, T-2 toxin dose-dependently and significantly induced a decreased number of osteoblasts and an increased number of osteoclasts in the tibia, with inhibited patterns of osteogenesis-related genes and enzymes and increased patterns of osteoclast-related genes and enzymes (p < 0.05). Similarly, the serum Ca and P concentrations and parathyroid hormone, calcitonin, and 1, 25-dihydroxycholecalciferol levels decreased under T-2 toxin exposure (p < 0.05). In summary, 2.0 mg/kg T-2 toxin significantly inhibited tibia weight, length, width, and circumference, as well as decreased bone-breaking strength, density, and composition (ash, calcium, and phosphorus) in 21-day-old goslings compared to the control and lower dose groups. Chondrocyte differentiation in TGP was delayed by 2.0 mg/kg T-2 toxin owing to cell apoptosis. In addition, 2.0 mg/kg T-2 toxin promoted bone resorption and inhibited osteogenesis in cellular morphology, gene expression, and hormonal modulation patterns. Thus, T-2 toxin significantly inhibited tibial growth and development with a dose dependence, accompanied by decreased bone geometry parameters and properties, hindered chondrocyte differentiation, and imbalanced bone metabolism. Full article
Show Figures

Figure 1

Figure 1
<p>The schematic diagram and experimental strategy of tibial damage caused by T-2 toxin in goslings. T-2, T-2 toxin; TGP, tibial growth plate; HE, hematoxylin–eosin; AB, alcian blue; SF, safranin O-fast green; TRAP, tartrate-resistant acid phosphatase; RT-qPCR, real-time quantitative PCR; OB, osteoblast; OC, osteoclast.</p>
Full article ">Figure 2
<p>Phenotypic and radiographic observations of the tibia in goslings under T-2 toxin exposure through morphological image capture (<b>A</b>), X-ray examination (<b>B</b>), and CT scan (<b>C</b>). Ⅰ, 0 mg/kg T-2 toxin group; Ⅱ, 0.5 mg/kg T-2 toxin group; Ⅲ, 1.0 mg/kg T-2 toxin group; and Ⅳ, 2.0 mg/kg T-2 toxin group. The red arrow refers to the tibial growth plate. CT values in the selected area were detected, indicating bone mineral density in the bone plane.</p>
Full article ">Figure 3
<p>Morphological and structural changes in chondrocytes in the tibial growth plate (TGP) under T-2 toxin exposure. Ⅰ, 0 mg/kg T-2 toxin group; Ⅱ, 0.5 mg/kg T-2 toxin group; Ⅲ, 1.0 mg/kg T-2 toxin group; and Ⅳ, 2.0 mg/kg T-2 toxin group; (<b>A</b>) The phenotype of the tibial growth plate with a scale bar of 10 mm. Differences in TGP height were analyzed by one-way ANOVA analysis. <span class="html-italic">n</span> = 12, 9, 10, and 7, respectively. <sup>a–c</sup> Means within a row with no common superscript differ significantly (<span class="html-italic">p</span> &lt; 0.05). (<b>B</b>) Chondrocyte damage in the reserve zone (RZ), proliferative zone (PZ), hypertrophic zone (HZ), and calcified zone (CZ) of TGP through hematoxylin–eosin (HE), alcian blue (AB), and safranin O-fast green (SF) staining with a scale bar of 1000 μm. The representative images of higher magnifications of TGP with a scale bar of 20 μm in HE staining are shown in <a href="#app1-animals-14-02281" class="html-app">Supplementary Figure S1</a>. (<b>C</b>) Expression patterns of chondrocyte development-related genes in TGP chondrocytes. <span class="html-italic">BMP</span>, bone morphogenetic protein; <span class="html-italic">SOX9</span>, SRY-box transcription factor 9; <span class="html-italic">RUNX2</span>, RUNX family transcription factor 2. Differences in gene expressions were analyzed by one-way ANOVA analysis. <span class="html-italic">n</span> = 6. <sup>a–d</sup> Means within a row with no common superscript differ significantly (<span class="html-italic">p</span> &lt; 0.05). (<b>D</b>) Expression patterns of apoptosis-related genes in TGP chondrocytes. <span class="html-italic">BAK1</span>, BCL2 antagonist/killer 1; <span class="html-italic">BCL2</span>, BCL2 apoptosis regulator; <span class="html-italic">CASP</span>, caspase. Differences in gene expressions were analyzed by one-way ANOVA analysis. <span class="html-italic">n</span> = 6. <sup>a–c</sup> Means within a row with no common superscript differ significantly (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Effects of T-2 toxin on bone metabolism: bone formation (osteogenesis) and resorption (osteoclastogenesis) in the calcified zone of the tibia. (<b>A</b>) Changes in the number of osteoblasts and osteoclasts with a scale bar of 100 μm. Ⅰ, 0 mg/kg T-2 toxin group; Ⅱ, 0.5 mg/kg T-2 toxin group; Ⅲ, 1.0 mg/kg T-2 toxin group; and Ⅳ, 2.0 mg/kg T-2 toxin group; HE, hematoxylin–eosin; TRAP, tartrate-resistant acid phosphatase; OB, osteoblast (green arrow); OC, osteoclast (red arrow); C, chondrocyte; TB, trabecular bone; BM, bone marrow. Osteoblasts are polygonal or cuboidal in shape, arranged in a single layer on the surface of the newly formed bone matrix. Osteoclasts are large multinucleated cells (<span class="html-italic">n</span> ≥ 3) that usually show intense red or purple coloring in TRAP staining. Osteoblasts and osteoclasts were counted in five randomly selected fields of HE and TRAP staining, respectively, with a scale bar of 100 μm, and the mean values were used as a representative value for the individual sample within the group for subsequent significance analysis. Differences in cell numbers were analyzed by one-way ANOVA analysis. <span class="html-italic">n</span> = 12, 9, 10, and 7, respectively. <sup>a–c</sup> Means within a row with no common superscript differ significantly (<span class="html-italic">p</span> &lt; 0.05). The representative images of higher magnifications of osteoblasts and osteoclasts with a scale bar of 20 μm are shown in <a href="#app1-animals-14-02281" class="html-app">Supplementary Figure S2</a>. (<b>B</b>) Changes in the expression of genes associated with bone formation and bone resorption in the tibia. <span class="html-italic">BGP</span>, bone gamma-carboxyglutamate protein; <span class="html-italic">OPG</span>, osteoprotegerin; <span class="html-italic">RANKL</span>, receptor activator of nuclear factor kappa B ligand; <span class="html-italic">RANK</span>, receptor activator of nuclear factor kappa B. Differences in gene expressions were analyzed by one-way ANOVA analysis. <span class="html-italic">n</span> = 6. <sup>a–d</sup> Means within a row with no common superscript differ significantly (<span class="html-italic">p</span> &lt; 0.05). (<b>C</b>) Changes in the serum activity of enzymes associated with bone formation and bone resorption. ALP, alkaline phosphatase. Differences in enzyme activities were analyzed by one-way ANOVA analysis. <span class="html-italic">n</span> = 12, 9, 10, and 7, respectively. <sup>a–c</sup> Means within a row with no common superscript differ significantly (<span class="html-italic">p</span> &lt; 0.05). (<b>D</b>) Changes in the serum calcium (Ca) and phosphorus (P) concentrations and hormone levels associated with bone metabolism. PTH, parathyroid hormone; CT, calcitonin; 1,25-(OH)<sub>2</sub>-D<sub>3</sub>, 1,25-dihydroxyvitamin D<sub>3</sub>. Differences in Ca and P concentrations and enzyme activities were analyzed by one-way ANOVA analysis. <span class="html-italic">n</span> = 12, 9, 10, and 7, respectively. <sup>a–c</sup> Means within a row with no common superscript differ significantly (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
15 pages, 2997 KiB  
Article
Photothermal Antibacterial and Osteoinductive Polypyrrole@Cu Implants for Biological Tissue Replacement
by Tianyou Zhou, Zeyan Zhou and Yingbo Wang
Materials 2024, 17(15), 3882; https://doi.org/10.3390/ma17153882 - 5 Aug 2024
Viewed by 413
Abstract
The treatment of bone defects caused by disease or accidents through the use of implants presents significant clinical challenges. After clinical implantation, these materials attract and accumulate bacteria and hinder the integration of the implant with bone tissue due to the lack of [...] Read more.
The treatment of bone defects caused by disease or accidents through the use of implants presents significant clinical challenges. After clinical implantation, these materials attract and accumulate bacteria and hinder the integration of the implant with bone tissue due to the lack of osteoinductive properties, both of which can cause postoperative infection and even lead to the eventual failure of the operation. This work successfully prepared a novel biomaterial coating with multiple antibacterial mechanisms for potent and durable and osteoinductive biological tissue replacement by pulsed PED (electrochemical deposition). By effectively regulating PPy (polypyrrole), the uniform composite coating achieved sound physiological stability. Furthermore, the photothermal analysis showcased exceptional potent photothermal antibacterial activity. The antibacterial assessments revealed a bacterial eradication rate of 100% for the PPy@Cu/PD composite coating following a 24 h incubation. Upon the introduction of NIR (near-infrared) irradiation, the combined effects of multiple antibacterial mechanisms led to bacterial reduction rates of 99% for E. coli and 98% for S. aureus after a 6 h incubation. Additionally, the successful promotion of osteoblast proliferation was confirmed through the application of the osteoinductive drug PD (pamidronate disodium) on the composite coating’s surface. Therefore, the antimicrobial Ti-based coatings with osteoinductive properties and potent and durable antibacterial properties could serve as ideal bone implants. Full article
(This article belongs to the Special Issue Advanced Functional Nanomaterials for Biomedical Application)
Show Figures

Figure 1

Figure 1
<p>Characterization of PPy@Cu/PD. (<b>a</b>) The SEM image. (<b>b</b>) The diagram of particle size distribution. (<b>c</b>) XRD spectrum. (<b>d</b>) EDS energy spectrum analysis. (<b>e</b>) Surface scanning image. (<b>f</b>) FTIR spectrogram.</p>
Full article ">Figure 2
<p>Assessment of the physiological stability and antibacterial properties of the composite coating. (<b>a</b>) Qualitative analysis of <span class="html-italic">E. coli</span> and <span class="html-italic">S. aureus</span>. (<b>b</b>) The release kinetic curve of Cu<sup>2+</sup> (<span class="html-italic">n</span> = 3). (<b>c</b>) Qualitative analysis of <span class="html-italic">E. coli</span> and <span class="html-italic">S. aureus</span> antibacterial performance (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 3
<p>Analysis of photothermal properties of the composite coating. (<b>a</b>) UV-Vis absorption spectra of composite coatings. (<b>b</b>) Curves for temperature changes over time of PPy@Cu/PD composite coatings under different powers of 808 nm laser irradiation. (<b>c</b>) Variation curves of temperatures over time for different materials under 808 nm laser irradiation at a laser power of 1 W·cm<sup>−2</sup>. (<b>d</b>) Heating–cooling curves of PPy@Cu/PD composite coatings under 808 nm laser irradiation. (<b>e</b>) The relationship between cooling time and -ln(θ). (<b>f</b>) The cyclic heating curves of the composite coating under 808 laser irradiation at 1 W·cm<sup>−2</sup> (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 4
<p>Analysis of the photothermal properties of the composite coatings. (<b>a</b>) The changes in <sup>1</sup>O<sub>2</sub> generated by the composite coatings under 808 nm irradiation at 1 W·cm<sup>−2</sup> at different time points. (<b>b</b>) Variation in <sup>1</sup>O<sub>2</sub> in different composite coating groups at 1 W·cm<sup>−2</sup> under 808 nm laser irradiation for 10 min. (<b>c</b>) Qualitative antibacterial images of PPy@Cu/PD surfaces cultured with bacteria for 6 h with/without irradiation. (<b>d</b>) Quantitative analysis of the antibacterial effect of PPy@Cu/PD composite coating cultured with bacteria for 6 h with/without irradiation. (<b>e</b>) Antibacterial mechanisms scheme (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 5
<p>(<b>a</b>) SEM images of osteoblasts; (<b>b</b>) CCK−8 analysis of osteoblasts (<span class="html-italic">n</span> = 3). (<sup>★</sup> indicates <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Scheme 1
<p>Mechanism diagram of constructing composite coating on Ti surface via the PED method.</p>
Full article ">
14 pages, 3504 KiB  
Communication
Connexin 43 Modulation in Human Chondrocytes, Osteoblasts and Cartilage Explants: Implications for Inflammatory Joint Disorders
by Elena Della Morte, Chiara Giannasi, Alice Valenza, Francesca Cadelano, Alessandro Aldegheri, Luigi Zagra, Stefania Niada and Anna Teresa Brini
Int. J. Mol. Sci. 2024, 25(15), 8547; https://doi.org/10.3390/ijms25158547 - 5 Aug 2024
Viewed by 358
Abstract
Connexin 43 (Cx43) is crucial for the development and homeostasis of the musculoskeletal system, where it plays multifaceted roles, including intercellular communication, transcriptional regulation and influencing osteogenesis and chondrogenesis. Here, we investigated Cx43 modulation mediated by inflammatory stimuli involved in osteoarthritis, i.e., 10 [...] Read more.
Connexin 43 (Cx43) is crucial for the development and homeostasis of the musculoskeletal system, where it plays multifaceted roles, including intercellular communication, transcriptional regulation and influencing osteogenesis and chondrogenesis. Here, we investigated Cx43 modulation mediated by inflammatory stimuli involved in osteoarthritis, i.e., 10 ng/mL Tumor Necrosis Factor alpha (TNFα) and/or 1 ng/mL Interleukin-1 beta (IL-1β), in primary chondrocytes (CH) and osteoblasts (OB). Additionally, we explored the impact of synovial fluids from osteoarthritis patients in CH and cartilage explants, providing a more physio-pathological context. The effect of TNFα on Cx43 expression in cartilage explants was also assessed. TNFα downregulated Cx43 levels both in CH and OB (−73% and −32%, respectively), while IL-1β showed inconclusive effects. The reduction in Cx43 levels was associated with a significant downregulation of the coding gene GJA1 expression in OB only (−65%). The engagement of proteasome in TNFα-induced effects, already known in CH, was also observed in OB. TNFα treatment significantly decreased Cx43 expression also in cartilage explants. Of note, Cx43 expression was halved by synovial fluid in both CH and cartilage explants. This study unveils the regulation of Cx43 in diverse musculoskeletal cell types under various stimuli and in different contexts, providing insights into its modulation in inflammatory joint disorders. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>,<b>B</b>). Expression of Cx43 in TNFα and/or IL-1β-stimulated human articular chondrocytes (CHs) (<b>A</b>) and human osteoblasts (OBs) (<b>B</b>) at day 3, analyzed using Western blot. Specific bands were quantified through Image Lab Software v 6.1 (Bio-Rad, Milan, Italy) and data (<span class="html-italic">n</span> = 5 independent experiments/donors, each indicated by distinct dots) were normalized on ACTB and expressed as relative values (CTR = 1). The panels below show representative immunoblots. Statistical analysis was performed via one-way analysis of variance (ANOVA) using Tukey’s post hoc test. Data are shown as mean ± SD. Significances vs. CTR are shown as * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001; vs. IL # <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 2
<p>(<b>A</b>–<b>C</b>). Gene expression of GJA1 (<b>A</b>), MMP3 (<b>B</b>) and MMP13 (<b>C</b>) in TNFα-stimulated OB at day 1 analyzed using real-time PCR. Data (<span class="html-italic">n</span> = 6 independent experiments/donors, each indicated by distinct dots) are expressed as 2<sup>−ΔΔCt</sup> (TBP was used as a housekeeping gene). (<b>D</b>). Expression of Cx43 in OB pre-treated for 1 h with MG-132 and then stimulated with TNFα for 3 days, analyzed using Western blot. Specific bands were quantified through Image Lab Software v 6.1 (Bio-Rad, Milan, Italy) and data (<span class="html-italic">n</span> = 4 independent experiments/donors) were normalized on ACTB and expressed as relative values (CTR = 1). Statistical analyses were performed using paired <span class="html-italic">t</span>-test (<b>A</b>–<b>C</b>) or one-way analysis of variance (ANOVA) using Tukey’s post hoc test (<b>D</b>). Data are shown as mean ± SD. Significance vs. CTR is shown as * <span class="html-italic">p</span> ≤ 0.05 and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 3
<p>Laser scanning confocal microscopy of OB; Cx43 and β-Tubulin were revealed with an Alexa Fluor<sup>®</sup> 488 (Cx43) and 568 (β-tubulin) conjugated antibody (green and red respectively), while nuclei were stained with DAPI (blue) (magnification 63×). The scale bar indicates 10 µm and the orthogonal views (yellow dashed lines) were obtained using Fiji software (ImageJ 1.51).</p>
Full article ">Figure 4
<p>(<b>A</b>). Gene expression of GJA1 in CH treated with synovial fluid (SF50%) for 1 day analyzed using real-time PCR. Data (<span class="html-italic">n</span> = 4 independent experiments/donors, each indicated by distinct dots) are expressed as 2<sup>−ΔΔCt</sup> (TBP was used as a housekeeping gene). (<b>B</b>). Expression of Cx43 in CH treated with SF50% for 3 days analyzed using Western blot. A representative immunoblot is shown. Specific bands were quantified through Image Lab Software v 6.1 (Bio-Rad, Milan, Italy) and data (<span class="html-italic">n</span> = 4 independent experiments) were normalized on ACTB and expressed as relative values (CTR = 1). Data are shown as mean ± SD. Statistical analysis was performed using paired <span class="html-italic">t</span>-test. Significance vs. CTR is shown as * <span class="html-italic">p</span> ≤ 0.05 and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>(<b>A</b>,<b>B</b>). Expression of Cx43 in cartilage explants treated with TNFα (<b>A</b>) or SF50% (<b>B</b>) for 3 days analyzed using Western blot. Representative immunoblots are shown. Specific bands were quantified through Image Lab Software v 6.1 (Bio-Rad, Milan, Italy) and data (<span class="html-italic">n</span> = 8 and <span class="html-italic">n</span> = 5 independent experiments/donors, each indicated by distinct dots) were normalized on ACTB and expressed as relative values (CTR = 1). Data are shown as mean ± SD. Statistical analysis was performed using paired <span class="html-italic">t</span>-test. Significance vs. CTR is shown as * <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">
20 pages, 8416 KiB  
Article
In Vitro Biocompatibility Assessment of Bioengineered PLA-Hydrogel Core–Shell Scaffolds with Mesenchymal Stromal Cells for Bone Regeneration
by Federica Re, Luciana Sartore, Chiara Pasini, Matteo Ferroni, Elisa Borsani, Stefano Pandini, Andrea Bianchetti, Camillo Almici, Lorena Giugno, Roberto Bresciani, Silvia Mutti, Federica Trenta, Simona Bernardi, Mirko Farina and Domenico Russo
J. Funct. Biomater. 2024, 15(8), 217; https://doi.org/10.3390/jfb15080217 - 31 Jul 2024
Viewed by 614
Abstract
Human mesenchymal stromal cells (hMSCs), whether used alone or together with three-dimensional scaffolds, are the best-studied postnatal stem cells in regenerative medicine. In this study, innovative composite scaffolds consisting of a core–shell architecture were seeded with bone-marrow-derived hMSCs (BM-hMSCs) and tested for their [...] Read more.
Human mesenchymal stromal cells (hMSCs), whether used alone or together with three-dimensional scaffolds, are the best-studied postnatal stem cells in regenerative medicine. In this study, innovative composite scaffolds consisting of a core–shell architecture were seeded with bone-marrow-derived hMSCs (BM-hMSCs) and tested for their biocompatibility and remarkable capacity to promote and support bone regeneration and mineralization. The scaffolds were prepared by grafting three different amounts of gelatin–chitosan (CH) hydrogel into a 3D-printed polylactic acid (PLA) core (PLA-CH), and the mechanical and degradation properties were analyzed. The BM-hMSCs were cultured in the scaffolds with the presence of growth medium (GM) or osteogenic medium (OM) with differentiation stimuli in combination with fetal bovine serum (FBS) or human platelet lysate (hPL). The primary objective was to determine the viability, proliferation, morphology, and spreading capacity of BM-hMSCs within the scaffolds, thereby confirming their biocompatibility. Secondly, the BM-hMSCs were shown to differentiate into osteoblasts and to facilitate scaffold mineralization. This was evinced by a positive Von Kossa result, the modulation of differentiation markers (osteocalcin and osteopontin), an expression of a marker of extracellular matrix remodeling (bone morphogenetic protein-2), and collagen I. The results of the energy-dispersive X-ray analysis (EDS) clearly demonstrate the presence of calcium and phosphorus in the samples that were incubated in OM, in the presence of FBS and hPL, but not in GM. The chemical distribution maps of calcium and phosphorus indicate that these elements are co-localized in the same areas of the sections, demonstrating the formation of hydroxyapatite. In conclusion, our findings show that the combination of BM-hMSCs and PLA-CH, regardless of the amount of hydrogel content, in the presence of differentiation stimuli, can provide a construct with enhanced osteogenicity for clinically relevant bone regeneration. Full article
(This article belongs to the Special Issue Feature Papers in Bone Biomaterials)
Show Figures

Figure 1

Figure 1
<p>Geometry of lattice unit cells (<b>a</b>), dry specimens employed in in vitro experiments (<b>b</b>), and magnified details of wet specimens (<b>c</b>), relative to scaffolds hosting low (L), medium (M), or high (H) hydrogel content.</p>
Full article ">Figure 2
<p>(<b>a</b>) Live/dead staining of BM-hMSCs cultivated in PLA-CH (L), PLA-CH(M), and PLA-CH(H) for 28 days in GM hPL. Scale bar: 100 μm. (<b>b</b>) Three-dimensional culture proliferation of BM-hMSCs cultivated in PLA-CH (L), PLA-CH (M), and PLA-CH (H) in the GM FBS or GM hPL at 28 days measured by the CCK8 assay.</p>
Full article ">Figure 3
<p>BM-hMSC viable cells cultivated for 28 days in scaffolds in GM hPL, examined by hematoxylin–eosin staining. Scale bar: 100 μm.</p>
Full article ">Figure 4
<p>SEM images of PLA-CH(M) scaffolds. (inset (<b>a</b>)): the PLA structure; (inset (<b>b</b>)), detail of the hydrogel structure; (inset (<b>c</b>)), view of the entire structure. SEM images of cellular growth over the scaffolds with BM-hMSCs in GM hPL: PLA-CH(L), PLA-CH(M), and PLA-CH(H) in the inset (<b>d</b>–<b>f</b>) respectively.</p>
Full article ">Figure 5
<p>Calcium deposit distribution in hydrogels PLA-CH cultured with cells using von Kossa staining (brown/black dots). Microphotographs of scaffolds with differentiated (<b>a</b>,<b>c</b>) and undifferentiated (<b>b</b>,<b>d</b>) BM-hMSCs (400× magnification, scale bar: 40 µm) in (<b>A</b>) PLA-CH(L), (<b>B</b>) PLA-CH(M), and (<b>C</b>) PLA-CH(H). Calcium deposits appeared as dots in brown/black color. Quantification of percentage of positive area (<b>D</b>) within scaffold meshes; ° <span class="html-italic">p</span> &lt; 0.001 vs. respective GM; * <span class="html-italic">p</span> &lt; 0.001 vs. OM FBS. No statistical differences were observed when comparing the scaffolds with the same treatment.</p>
Full article ">Figure 6
<p>Micrographs of the hydrogels PLA-CH(M) with differentiated (OM) BM-hMSCs in hPL (<b>a</b>–<b>c</b>) and FBS (<b>d</b>–<b>f</b>) immunostained for OSC (<b>a</b>,<b>d</b>), OSP (<b>b</b>,<b>e</b>), and BMP2 (<b>c</b>,<b>f</b>) (brown color) with hematoxylin counterstaining (blue/violet) (400× magnification, bar 40 µm).</p>
Full article ">Figure 7
<p>SEM images and EDS investigation of calcium phosphate deposition in PLA-CH(M) with BM-hMSCs in the OM hPL and OM FBS at day 28. Upper part: the bright round particles observed on both samples (upper part of the figure) are ascribed to hydroxyapatite formation. Lower part: evaluation of calcium and phosphorous with SEM–EDS of PLA-CH(M) with BM-hMSCs in OM with FBS at day 28. The mapping of Ca and P in the boxed area (78 × 78 µ<sup>2</sup>) indicates the extended presence of these elements below the biological film visible at the surface.</p>
Full article ">
15 pages, 1991 KiB  
Article
Culture and Immunomodulation of Equine Muscle-Derived Mesenchymal Stromal Cells: A Comparative Study of Innovative 2D versus 3D Models Using Equine Platelet Lysate
by J. Duysens, H. Graide, A. Niesten, A. Mouithys-Mickalad, G. Deby-Dupont, T. Franck, J. Ceusters and D. Serteyn
Cells 2024, 13(15), 1290; https://doi.org/10.3390/cells13151290 - 31 Jul 2024
Viewed by 363
Abstract
Muscle-derived mesenchymal stromal cells (mdMSCs) hold great promise in regenerative medicine due to their immunomodulatory properties, multipotent differentiation capacity and ease of collection. However, traditional in vitro expansion methods use fetal bovine serum (FBS) and have numerous limitations including ethical concerns, batch-to-batch variability, [...] Read more.
Muscle-derived mesenchymal stromal cells (mdMSCs) hold great promise in regenerative medicine due to their immunomodulatory properties, multipotent differentiation capacity and ease of collection. However, traditional in vitro expansion methods use fetal bovine serum (FBS) and have numerous limitations including ethical concerns, batch-to-batch variability, immunogenicity, xenogenic contamination and regulatory compliance issues. This study investigates the use of 10% equine platelet lysate (ePL) obtained by plasmapheresis as a substitute for FBS in the culture of mdMSCs in innovative 2D and 3D models. Using muscle microbiopsies as the primary cell source in both models showed promising results. Initial investigations indicated that small variations in heparin concentration in 2D cultures strongly influenced medium coagulation with an optimal proliferation observed at final heparin concentrations of 1.44 IU/mL. The two novel models investigated showed that expansion of mdMSCs is achievable. At the end of expansion, the 3D model revealed a higher total number of cells harvested (64.60 ± 5.32 million) compared to the 2D culture (57.20 ± 7.66 million). Trilineage differentiation assays confirmed the multipotency (osteoblasts, chondroblasts and adipocytes) of the mdMSCs generated in both models with no significant difference observed. Immunophenotyping confirmed the expression of the mesenchymal stem cell (MSC) markers CD-90 and CD-44, with low expression of CD-45 and MHCII markers for mdMSCs derived from the two models. The generated mdMSCs also had great immunomodulatory properties. Specific immunological extraction followed by enzymatic detection (SIEFED) analysis demonstrated that mdMSCs from both models inhibited myeloperoxidase (MPO) activity in a strong dose-dependent manner. Moreover, they were also able to reduce reactive oxygen species (ROS) activity, with mdMSCs from the 3D model showing significantly higher dose-dependent inhibition compared to the 2D model. These results highlighted for the first time the feasibility and efficacy of using 10% ePL for mdMSC expansion in novel 2D and 3D approaches and also that mdMSCs have strong immunomodulatory properties that can be exploited to advance the field of regenerative medicine and cell therapy instead of using FBS with all its drawbacks. Full article
(This article belongs to the Collection Stem Cells in Tissue Engineering and Regeneration)
Show Figures

Figure 1

Figure 1
<p>Collection of ePL from an awake horse on which apheresis is being carried out using the COM.tec plasmapheresis device.</p>
Full article ">Figure 2
<p>Representative microphotographs by optical microscope (×100) of mdMSCs cultured after 7 days with 10% platelet lysate and 1.44 IU/mL of heparin (<b>left</b>) and 3 IU/mL of heparin (<b>right</b>) (Horse 1).</p>
Full article ">Figure 3
<p>Representative microphotographs by optical microscope (×100) of mdMSCs cultured after 4 days (<b>left</b>) and 8 days (<b>right</b>) with 10% platelet lysate and 1.44 IU/mL of heparin (Horse 3).</p>
Full article ">Figure 4
<p>Representative microphotographs by optical microscope (×100) of mdMSCs cultured in 3D after 4 days (<b>left</b>) and 8 days (<b>right</b>) with DMEM/Ham’s F12 medium supplemented with 10% platelet lysate (Horse 3).</p>
Full article ">Figure 5
<p>Total number of mdMSCs harvested with the 3D and 2D models at passage 3.</p>
Full article ">Figure 6
<p>Representative microphotographs obtained by optical microscope (×100) of trilineage differentiations of mdMSCs (from the 3D model at passage 3) with, respectively, for the upper line chondroblast (<b>left</b>), adipocyte (<b>between</b>), and osteoblast (<b>right</b>) cells cultured without differentiation media. Lower line is composed of differentiated chondroblasts (<b>left</b>), adipocytes (<b>between</b>), and osteoblasts (<b>right</b>).</p>
Full article ">Figure 7
<p>Effect of mdMSCs (passage 3) on the activity of equine MPO measured by SIEFED. Results from five independent experiments with two technical replicates for each concentration (<span class="html-italic">n</span> = 10). The means ± SD are shown as relative percentages compared to the MPO control, which was performed without mdMSCs and defined as 100% response.</p>
Full article ">Figure 8
<p>Effects of different concentrations of mdMSCs (passage 3) in Ringer lactate on the ROS production by neutrophils (<span class="html-italic">n</span> = 10). NA and A represent, respectively, non-activated and activated neutrophils alone. Means ± SD are shown in relative percentages. Stimulated neutrophils without mdMSCs (A) is defined as 100% response. Means ± SD are shown in relative percentages of A.</p>
Full article ">
15 pages, 5592 KiB  
Article
Apoptosis and Inflammation Involved with Fluoride-Induced Bone Injuries
by Miao Wang, Kangting Luo, Tongtong Sha, Qian Li, Zaichao Dong, Yanjie Dou, Huanxia Zhang, Guoyu Zhou, Yue Ba and Fangfang Yu
Nutrients 2024, 16(15), 2500; https://doi.org/10.3390/nu16152500 - 31 Jul 2024
Viewed by 486
Abstract
Background: Excessive fluoride exposure induces skeletal fluorosis, but the specific mechanism responsible is still unclear. Therefore, this study aimed to identify the pathogenesis of fluoride-induced bone injuries. Methods: We systematically searched fluoride-induced bone injury-related genes from five databases. Then, these genes were subjected [...] Read more.
Background: Excessive fluoride exposure induces skeletal fluorosis, but the specific mechanism responsible is still unclear. Therefore, this study aimed to identify the pathogenesis of fluoride-induced bone injuries. Methods: We systematically searched fluoride-induced bone injury-related genes from five databases. Then, these genes were subjected to enrichment analyses. A TF (transcription factor)–mRNA–miRNA network and protein–protein interaction (PPI) network were constructed using Cytoscape, and the Human Protein Atlas (HPA) database was used to screen the expression of key proteins. The candidate pharmacological targets were predicted using the Drug Signature Database. Results: A total of 85 studies were included in this study, and 112 osteoblast-, 35 osteoclast-, and 41 chondrocyte-related differential expression genes (DEGs) were identified. Functional enrichment analyses showed that the Atf4, Bcl2, Col1a1, Fgf21, Fgfr1 and Il6 genes were significantly enriched in the PI3K-Akt signaling pathway of osteoblasts, Mmp9 and Mmp13 genes were enriched in the IL-17 signaling pathway of osteoclasts, and Bmp2 and Bmp7 genes were enriched in the TGF-beta signaling pathway of chondrocytes. With the use of the TF–mRNA–miRNA network, the Col1a1, Bcl2, Fgfr1, Mmp9, Mmp13, Bmp2, and Bmp7 genes were identified as the key regulatory factors. Selenium methyl cysteine, CGS-27023A, and calcium phosphate were predicted to be the potential drugs for skeletal fluorosis. Conclusions: These results suggested that the PI3K-Akt signaling pathway being involved in the apoptosis of osteoblasts, with the IL-17 and the TGF-beta signaling pathways being involved in the inflammation of osteoclasts and chondrocytes in fluoride-induced bone injuries. Full article
(This article belongs to the Special Issue Nutritional Supplements for Bone Health)
Show Figures

Figure 1

Figure 1
<p>Flow diagram for the study selection procedure.</p>
Full article ">Figure 2
<p>The top 10 biological process (BP), cellular component (CC), and molecular function (MF) terms of the osteoblast-related differentially expressed genes (DEGs) (<b>A</b>), osteoclast-related DEGs (<b>B</b>), and chondrocyte-related DEGs (<b>C</b>) in the fluoride exposure group compared with the control group. <span class="html-italic">p</span> value &lt; 0.05 was considered significant.</p>
Full article ">Figure 3
<p>The top 8 enriched signaling pathways of the osteoblast-related DEGs (<b>A</b>), osteoclast-related DEGs (<b>B</b>), and chondrocyte-related DEGs (<b>C</b>) in the fluoride exposure group compared with the control group. Pathways with <span class="html-italic">p</span> value &lt; 0.05 were considered significant.</p>
Full article ">Figure 4
<p>Protein–protein interaction (PPI) network among the osteoblast-related DEGs (<b>A</b>), osteoclast-related DEGs (<b>B</b>), and chondrocyte-related DEGs (<b>C</b>) was constructed by STRING database and visualized by Cytoscape. The top 5 hub genes were identified via the MCC algorithm of the cyto-hubba plug-in. Orange represents the key genes.</p>
Full article ">Figure 5
<p>The transcription factor (TF)–mRNA–miRNA regulatory network (<b>A</b>) and immunohistochemistry images of the key genes from the HPA database (<b>B</b>). Green represents the key genes. Yellow represents the target TFs. Blue represents the target miRNAs. The scale bar is 200 µm.</p>
Full article ">Figure 6
<p>The regulatory mechanism of bone injuries induced by fluoride. The Bcl2, Col1a1, and Fgfr1 genes were involved in the apoptosis of osteoblasts via the PI3K-Akt signaling pathway, inhibiting bone formation. The Mmp13 and Mmp9 genes caused the inflammation of osteoclasts by activating the IL-17 signaling pathway, promoting bone resorption. The Bmp2 and Bmp7 genes caused the inflammation of chondrocytes by activating the TGF-beta signaling pathway, resulting in bone necrosis.</p>
Full article ">
21 pages, 4224 KiB  
Article
Arid1a Loss Enhances Disease Progression in a Murine Model of Osteosarcoma
by Kaniz Fatema, Yanliang Wang, Adriene Pavek, Zachary Larson, Christopher Nartker, Shawn Plyler, Amanda Jeppesen, Breanna Mehling, Mario R. Capecchi, Kevin B. Jones and Jared J. Barrott
Cancers 2024, 16(15), 2725; https://doi.org/10.3390/cancers16152725 - 31 Jul 2024
Viewed by 561
Abstract
Osteosarcoma is an aggressive bone malignancy, molecularly characterized by acquired genome complexity and frequent loss of TP53 and RB1. Obtaining a molecular understanding of the initiating mutations of osteosarcomagenesis has been challenged by the difficulty of parsing between passenger and driver mutations [...] Read more.
Osteosarcoma is an aggressive bone malignancy, molecularly characterized by acquired genome complexity and frequent loss of TP53 and RB1. Obtaining a molecular understanding of the initiating mutations of osteosarcomagenesis has been challenged by the difficulty of parsing between passenger and driver mutations in genes. Here, a forward genetic screen in a genetic mouse model of osteosarcomagenesis initiated by Trp53 and Rb1 conditional loss in pre-osteoblasts identified that Arid1a loss contributes to OS progression. Arid1a is a member of the canonical BAF (SWI/SNF) complex and a known tumor suppressor gene in other cancers. We hypothesized that the loss of Arid1a increases the rate of tumor progression and metastasis. Phenotypic evaluation upon in vitro and in vivo deletion of Arid1a validated this hypothesis. Gene expression and pathway analysis revealed a correlation between Arid1a loss and genomic instability, and the subsequent dysregulation of genes involved in DNA DSB or SSB repair pathways. The most significant of these transcriptional changes was a concomitant decrease in DCLRE1C. Our findings suggest that Arid1a plays a role in genomic instability in aggressive osteosarcoma and a better understanding of this correlation can help with clinical prognoses and personalized patient care. Full article
(This article belongs to the Special Issue Multimodality Management of Sarcomas)
Show Figures

Figure 1

Figure 1
<p>Forward genetic screen using <span class="html-italic">piggyBac</span> transposon identifies <span class="html-italic">Arid1a</span> as a tumor suppressor gene. (<b>a</b>) Schematic of <span class="html-italic">PBonc</span> design, random genomic integration of transposable elements from <span class="html-italic">Rosa26</span> locus via Cre-lox mediated recombination (PBIRL/PBIRR = piggyBac Inverted Repeat sequences left/right, SA/SD = Splice Acceptor/Donor, PBase = piggyBac transposase, IRES = Internal Ribosomal Entry Sites, Luc = Luciferase). (<b>b</b>) Histology of tamoxifen-induced skeletal tumors harvested from <span class="html-italic">piggyBac</span> inserted cohort shows similar appearance with osteosarcoma (osteoid matrix = bright pink, scale bar 100 and 10 μm). (<b>c</b>) Kaplan–Meier survival analysis of mice with (red) <span class="html-italic">piggyBac</span> mutagenesis showed a 5-month survival disadvantage compared to the ones without (blue); median overall survival was 10.0 months and 15.0 months, respectively (<span class="html-italic">p</span>-value &lt; 0.0001).</p>
Full article ">Figure 2
<p>In vivo <span class="html-italic">Arid1a</span> knockout resulted in a greater tumor burden and metastasis. (<b>a</b>) H&amp;E staining of <span class="html-italic">Arid1a</span> homozygous (<b>bottom</b>) and heterozygous (<b>top</b>) mouse tumor tissue shows an osteosarcoma-like appearance (osteoid matrix = bright pink, scale bar 100 μm). (<b>b</b>) Relative expression of <span class="html-italic">Arid1a</span> in wildtype, <span class="html-italic">Arid1a<sup>fl/fl</sup></span> and <span class="html-italic">Arid1a<sup>fl/wt</sup></span> mouse tumor samples (**** <span class="html-italic">p</span>-value &lt; 0.0001). (<b>c</b>) Pie chart shows the percent metastasis found in <span class="html-italic">Arid1a</span> in wildtype, <span class="html-italic">Arid1a<sup>fl/fl</sup></span> and <span class="html-italic">Arid1a<sup>fl/wt</sup></span> mice. (<b>d</b>) Kaplan–Meier tumor-free curve of <span class="html-italic">Arid1a</span> in wildtype (n = 142), <span class="html-italic">Arid1a<sup>fl/fl</sup></span> (n = 41), and <span class="html-italic">Arid1a<sup>fl/wt</sup></span> (n = 22), mice (<span class="html-italic">p</span>-value &lt; 0.0001). Black lines represent all the mice that were censored (no tumor). (<b>e</b>) Gross necropsy showing the common sites of primary and metastasized tumor formation in <span class="html-italic">Arid1afl/fl</span> and <span class="html-italic">Arid1a<sup>fl/wt</sup></span> mice (P = Primary, M = Metastatic tumors). (<b>f</b>) Micro-CT image showing the formation of multiple tumors (yellow arrows) on <span class="html-italic">Arid1a<sup>fl/fl</sup></span> mice at various ages—(<b>top</b>) <span class="html-italic">Arid1a</span> wildtype control, (<b>middle</b>) <span class="html-italic">Arid1a<sup>fl/fl</sup></span> mice at 217 days, (<b>bottom</b>) <span class="html-italic">Arid1a<sup>fl/fl</sup></span> mice at 91 days.</p>
Full article ">Figure 3
<p>Transcriptional comparison of <span class="html-italic">Arid1a</span> WT and <span class="html-italic">Arid1a</span> KO osteosarcomas. (<b>a</b>) Heatmap demonstrating the distribution of top 1000 significant differentially expressed genes (clustered by Z-scores) between <span class="html-italic">Arid1a</span> WT (Red) and <span class="html-italic">Arid1a</span> KO (Blue) mouse osteosarcomas. (<b>b</b>) Euclidean distance between samples. (<b>c</b>) Volcano plot representing most significant DE genes in both cohorts, X-axis representing log<sub>2</sub> FC and Y-axis presenting the –log<sub>10</sub> of the adjusted <span class="html-italic">p</span>-value.</p>
Full article ">Figure 4
<p>Molecular pathway analysis. (<b>a</b>) mRNA profiling by Almac categorized the differential expression profile based on the hallmarks of cancers. Heatmaps showing the topmost significantly correlated hallmarks, (<b>top</b>) EMT pathways and (<b>bottom</b>) genomic instability. Samples represented by columns and RNA expression signatures represented in rows. Each colored square indicates the expression of the multi-gene signature for that sample (Red = increased expression levels or activation of the hallmark, green = decreased expression levels or repression of the hallmark). (<b>b</b>) Principal Component Analysis of the gene expression data. (<b>c</b>) IPA analysis showing the top 20 significant signaling pathways correlating the differentially expressed genes between <span class="html-italic">Arid1a</span> WT and KO osteosarcomas (<span class="html-italic">p</span>-value &lt; 0.05). X-axis representing the −log<sub>10</sub> <span class="html-italic">p</span>-value. (<b>d</b>) IPA upstream regulator analysis based on activation Z-score and <span class="html-italic">p</span>-values.</p>
Full article ">Figure 5
<p>Impact of <span class="html-italic">Arid1a</span> loss on DNA damage repair pathways. (<b>a</b>) Volcano plot displaying the most significant differentially regulated genes involved in DNA repair pathways between <span class="html-italic">Arid1a</span> KO and WT mouse OS; the X-axis represents log<sub>2</sub> fold change (thresholds of −1 and 1) and the Y-axis shows the adjusted −log10 <span class="html-italic">p</span>-values (threshold 0.05). (<b>b</b>) Correlation between <span class="html-italic">DCLRE1C/Artemis</span> and <span class="html-italic">Arid1a</span> expression in all human cancer cell lines (n = 1450) and in OS cell lines (n = 16) from the CCLE database (R<sup>2</sup> = 0.435, <span class="html-italic">p</span>-value = 4.3 × 10<sup>−68</sup> and R<sup>2</sup> = 0.581, <span class="html-italic">p</span>-value = 0.0183, respectively).</p>
Full article ">Figure 6
<p>In vitro CRISPR/Cas9 knockout resulted in a more proliferative and chemoresistant cellular phenotype. (<b>a</b>) Endogenous expression of <span class="html-italic">Arid1a</span> in osteosarcoma cell lines as demonstrated by immunofluorescence staining showing nuclear localization of <span class="html-italic">Arid1a</span> (Actin = Red, Nuclei = Blue). (<b>b</b>) Schematic for targeting <span class="html-italic">Arid1a</span> with CRISPR/Cas9. (<b>c</b>) Relative mRNA expression of <span class="html-italic">Arid1a</span> after knockout in two biological replicates of U2-OS and SJSA-1. The data represent an average of three technical replicates and error bars represent the standard error of the mean (** <span class="html-italic">p</span>-value &lt; 0.05; *** <span class="html-italic">p</span>-value &lt; 0.005; **** <span class="html-italic">p</span>-value &lt; 0.0005). (<b>d</b>) Real-time cell proliferation assay for 48 h in U2-OS cell lines, error bar represents the standard error of the mean (**** <span class="html-italic">p</span>-value &lt; 0.0001; individual <span class="html-italic">t</span>-tests compared between groups at different time points), (n = 5). (<b>e</b>) Real-time cell migration assay at different time points (T1 = 0 h, T5 = 24 h), error bars represents the standard deviation of mean (**** <span class="html-italic">p</span>-value &lt; 0.0001), (n = 3). (<b>f</b>) Doxorubicin chemosensitivity assay at 72 h, IC<sub>50</sub> for KO U2-OS = 0.7 μM, and Control U2-OS 0.3 μM.</p>
Full article ">
12 pages, 6629 KiB  
Communication
Osteoclast-Driven Polydopamine-to-Dopamine Release: An Upgrade Patch for Polydopamine-Functionalized Tissue Engineering Scaffolds
by Lufei Wang, Huamin Hu and Ching-Chang Ko
J. Funct. Biomater. 2024, 15(8), 211; https://doi.org/10.3390/jfb15080211 - 29 Jul 2024
Viewed by 630
Abstract
Polydopamine, a mussel-inspired self-adherent polymer of dopamine, has impressive adhesive properties and thus is one of the most versatile approaches to functionalize tissue engineering scaffolds. To date, many types of polydopamine-functionalized scaffolds have been manufactured and extensively applied in bone tissue engineering at [...] Read more.
Polydopamine, a mussel-inspired self-adherent polymer of dopamine, has impressive adhesive properties and thus is one of the most versatile approaches to functionalize tissue engineering scaffolds. To date, many types of polydopamine-functionalized scaffolds have been manufactured and extensively applied in bone tissue engineering at the preclinical stage. However, how polydopamine is biodegraded and metabolized during the bone healing process and the side effects of its metabolite remain largely unknown. These issues are often neglected in the modern manufacture of polydopamine-functionalized materials and restrict them from stepping forward to clinical applications. In this study, using our bioinspired polydopamine-laced hydroxyapatite collagen calcium silicate material as a representative of polydopamine-functionalized tissue engineering scaffolds, we discovered that polydopamine can be metabolized to dopamine specifically by osteoclasts, which we termed “osteoclast-driven polydopamine-to-dopamine release”. The released dopamine showed an osteoinductive effect in vitro and promoted bone regeneration in calvarial critical-sized defects. The concept of “osteoclast-driven polydopamine-to-dopamine release” has considerable application potential. It could be easily adopted by other existing polydopamine-functionalized scaffolds: just by recruiting osteoclasts. Once adopted, scaffolds will obtain a dopamine-releasing function, which enables their modulation of osteoblast activity and hence elevates the osteoinductive effect. Thus, “osteoclast-driven polydopamine-to-dopamine release” serves as an upgrade patch, which is useful for many existing polydopamine-functionalized materials. Full article
Show Figures

Figure 1

Figure 1
<p>For (<b>A</b>–<b>D</b>), RAW cells were differentiated to OCs on PDHC disks. “1X OC” indicates 4 × 10<sup>4</sup> cells/cm<sup>2</sup> and “2X OC” indicates 8 × 10<sup>4</sup> cells/cm<sup>2</sup>. (<b>A</b>) OC morphology on PDHC coating, visualized by TRAP staining. Scale bar is 20 µm. (<b>B</b>) PDHC material degradation percentage during osteoclastic resorption. * <span class="html-italic">p</span> &lt; 0.05. (<b>C</b>) SEM images of PDHC surface after osteoclastic resorption. Images are shown on 12 d and the scale bar is 200 µm. (<b>D</b>) dopamine release curve of PDHC-OC constructs. * <span class="html-italic">p</span> &lt; 0.05 when “PDHC 1X OC” vs. “PDHC 2X OC”. (<b>E</b>) dopamine release when PDHC immersed in various solutions. N.D.: not detected. (<b>F</b>) dopamine release from PDHC-OC constructs after knocking down target genes. Time point: day 12. * <span class="html-italic">p</span> &lt; 0.05 compared to the Ctrl group.</p>
Full article ">Figure 2
<p>(<b>A</b>) Experiment design for (<b>B</b>–<b>D</b>): collect the conditioned medium of PDHC-OC or HCCS-OC construct at different time points and test their influences on MSC. (<b>B</b>) Effect on MSC viability. * <span class="html-italic">p</span> &lt; 0.05 compared to the matched time point “Ctrl” group. (<b>C</b>,<b>D</b>) Effect on MSC ALP, <span class="html-italic">Osx</span>, and <span class="html-italic">Ocn</span> expression. * <span class="html-italic">p</span> &lt; 0.05, # <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>(<b>A</b>) Schematic diagram of the study. (<b>B</b>,<b>C</b>) After 14d osteogenic induction under Transwell co-culture condition, MSC ALP activity and expression of osteogenic genes, <span class="html-italic">Osx</span> and <span class="html-italic">Ocn</span>. * <span class="html-italic">p</span> &lt; 0.05 compared to the Ctrl group. # <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>HCCS scaffold, PDHC scaffold, and exogenous-OC-seeded PDHC scaffold were implanted to repair rat calvarial critical-sized defects. Experiment period was 8 weeks; animal n = 4. (<b>A</b>) MicroCT images. The 8 mm in diameter red circle represents the defect site. (<b>B</b>) Quantification of new bone formation within the defect site. * <span class="html-italic">p</span> &lt; 0.05. (<b>C</b>) Dopamine concentration in the tissue fluid from the defect site. * <span class="html-italic">p</span> &lt; 0.05 when “PDHC” vs. “PDHC OC”.</p>
Full article ">Figure 5
<p>Concept map of OC-driven polydopamine-to-dopamine release.</p>
Full article ">
19 pages, 4260 KiB  
Article
Mangiferin Induces Post-Implant Osteointegration in Male Diabetic Rats
by Bünyamin Ongan, Ömer Ekici, Gökhan Sadi, Esra Aslan and Mehmet Bilgehan Pektaş
Medicina 2024, 60(8), 1224; https://doi.org/10.3390/medicina60081224 - 28 Jul 2024
Viewed by 445
Abstract
Background and Objectives: Hyperglycemia is known to undermine the osteointegration process of implants. In this study, the effects of mangiferin (MF) on the post-implant osteointegration process in a type-II diabetes model were investigated molecularly and morphologically. Materials and Methods: Sprague Dawley male rats [...] Read more.
Background and Objectives: Hyperglycemia is known to undermine the osteointegration process of implants. In this study, the effects of mangiferin (MF) on the post-implant osteointegration process in a type-II diabetes model were investigated molecularly and morphologically. Materials and Methods: Sprague Dawley male rats were divided into three groups: control, diabetes, and diabetes + MF. All animals were implanted in their tibia bones on day 0. At the end of the 3-month experimental period, the animals’ blood and the implant area were isolated. Biochemical measurements were performed on blood samples and micro-CT, qRT-PCR, histological, and immunohistochemical measurements were performed on tibia samples. Results: MF significantly improved the increased glucose, triglyceride-VLDL levels, and liver enzymes due to diabetes. By administering MF to diabetic rats, the osteointegration percentage and bone volume increased while porosity decreased. DKK1 and BMP-2 mRNA expressions and OPN, OCN, and OSN mRNA–protein expressions increased by MF administration in diabetic rats. Additionally, while osteoblast and osteoid surface areas increased with MF, osteoclast and eroded surface areas decreased. Conclusions: The findings of our study indicate that MF will be beneficial to the bone-repairing process and osteointegration, which are impaired by type-II diabetes. Full article
(This article belongs to the Section Orthopedics)
Show Figures

Figure 1

Figure 1
<p>Implant placement at proximal tibia of the rat.</p>
Full article ">Figure 2
<p>mRNA expressions of BMP-2 (<b>a</b>), DKK1 (<b>b</b>), IBSP (<b>c</b>), LRP5 (<b>d</b>), OCN (<b>e</b>), OPN (<b>f</b>), OSN (<b>g</b>), OSP (<b>h</b>), RANKL (<b>i</b>), and RUNX2 (<b>j</b>) at the bone–implant site of rats in the control, diabetes, and diabetes + MF groups. Data were normalized by GAPDH. Each bar represents the means of at least six rats. Values are expressed as mean ± SEM, <span class="html-italic">n</span> = 6–8. * Significantly different (<span class="html-italic">p</span> &lt; 0.05) compared to control group; # significantly different (<span class="html-italic">p</span> &lt; 0.05) compared to diabetes group.</p>
Full article ">Figure 3
<p>Imaging of connectivity densities (<b>a</b>), connectivity (<b>b</b>), bone volume (<b>c</b>), osteointegration percentage (<b>d</b>), porosity (<b>e</b>), and the pore volume (<b>f</b>) at the bone–implant site of rats in the control, diabetes, and diabetes + MF groups. Each bar represents the means of at least six rats. Values are expressed as mean ± SEM, <span class="html-italic">n</span> = 6–8. * Significantly different (<span class="html-italic">p</span> &lt; 0.05) compared to control group; # significantly different (<span class="html-italic">p</span> &lt; 0.05) compared to diabetes group.</p>
Full article ">Figure 4
<p>Representative imaging of the tibia bone–implant complex in the control, diabetes, and diabetes + MF groups.</p>
Full article ">Figure 5
<p>H-Score values for immunostaining by OPN (<b>a</b>), RANKL (<b>b</b>), OCN (<b>c</b>), OSN (<b>d</b>), and DCN (<b>e</b>) proteins at the bone–implant site of rats in the control, diabetes, and diabetes + MF groups. Each bar represents the means of at least six rats. Values are expressed as mean ± SEM, <span class="html-italic">n</span> = 6–8. * Significantly different (<span class="html-italic">p</span> &lt; 0.05) compared to control group; # significantly different (<span class="html-italic">p</span> &lt; 0.05) compared to diabetes group.</p>
Full article ">Figure 6
<p>Immunostaining by OPN, RANKL, OCN, OSN, and DCN proteins at the bone–implant site of rats in the control, diabetes, and diabetes + MF groups.</p>
Full article ">Figure 7
<p>Histochemical staining of osteoblast surface (Ob.S/BS) (<b>a</b>), osteoclast surface (Oc.S/BS) (<b>b</b>), eroded surface (ES/BS) (<b>c</b>), and osteoid surface (OS/BS) (<b>d</b>) at the bone–implant site in the control, diabetes, and diabetes + MF groups. Values are expressed as mean ± SEM, <span class="html-italic">n</span> = 6–12. * Significantly different (<span class="html-italic">p</span> &lt; 0.05) compared to control group; # significantly different (<span class="html-italic">p</span> &lt; 0.05) compared to diabetes group.</p>
Full article ">Figure 8
<p>Histochemical staining in the control, diabetes, and diabetes + MF groups.</p>
Full article ">
17 pages, 18198 KiB  
Article
Combined Effects of Fibroblast Growth Factor-2 and Carbonate Apatite Granules on Periodontal Healing: An In Vivo and In Vitro Study
by Naoki Miyata, Shinta Mori, Tasuku Murakami, Takahiro Bizenjima, Fumi Seshima, Kentaro Imamura and Atsushi Saito
Biomedicines 2024, 12(8), 1664; https://doi.org/10.3390/biomedicines12081664 - 25 Jul 2024
Viewed by 425
Abstract
The aim of this study was to investigate in vivo and in vitro the effectiveness of the use of fibroblast growth factor (FGF)-2 with carbonate apatite (CO3Ap) on periodontal healing. Periodontal defects created in the maxillary first molars in rats were [...] Read more.
The aim of this study was to investigate in vivo and in vitro the effectiveness of the use of fibroblast growth factor (FGF)-2 with carbonate apatite (CO3Ap) on periodontal healing. Periodontal defects created in the maxillary first molars in rats were treated with FGF-2, CO3Ap, FGF-2 + CO3Ap or left unfilled. Healing was evaluated using microcomputed tomography, histological, and immunohistochemical analyses. In vitro experiments were performed to assess cellular behaviors and the expression of osteoblastic differentiation markers in MC3T3-E1 cells. At 4 weeks, the bone volume fraction in the FGF-2 + CO3Ap group was significantly greater than that in the CO3Ap group, but there was no significant difference from the FGF-2 group. The FGF-2 + CO3Ap group demonstrated greater new bone compared with the FGF-2 or CO3Ap group. The FGF-2 + CO3Ap group showed greater levels of osteocalcin-positive cells compared with the CO3Ap group, but there was no significant difference from the FGF-2 group. In vitro, the FGF-2 + CO3Ap group exhibited a greater extent of cell attachment and more elongated cells compared with the CO3Ap group. Compared with the CO3Ap group, the FGF-2 + CO3Ap group showed significantly higher viability/proliferation, but the expressions of Runx2 and Sp7 were reduced. The results indicated that the use of FGF-2 with CO3Ap enhanced healing in the periodontal defects. FGF-2 promoted cell attachment to and proliferation on CO3Ap and regulated osteoblastic differentiation, thereby contributing to novel bone formation. Full article
(This article belongs to the Special Issue Periodontal Disease and Periodontal Tissue Regeneration)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Micro-CT images and quantification analysis. The color scale indicates bone mineral density (BMD): light blue and purple, low; yellow and green, medium; and red and orange, high. (<b>a</b>,<b>d</b>) Sagittal images from micro-CT (bar = 1000 µm). (<b>b</b>,<b>c</b>,<b>e</b>,<b>f</b>) Quantitative data. Bone volume (BV)/total volume (TV) within the ROI (<b>b</b>,<b>e</b>) and radiopaque volume of newly formed bone and CO<sub>3</sub>Ap particles (RV)/total volume (TV) (<b>c</b>,<b>f</b>) were compared between groups. Data are presented as box-and-whiskers plots with maximum, median, minimum, and 75th and 25th percentiles (<span class="html-italic">n</span> = 10). * <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 by ANOVA with Tukey post hoc test.</p>
Full article ">Figure 2
<p>Histopathological analysis. (<b>a</b>–<b>d</b>) Images at 2 weeks (original magnification ×25; bar = 500 µm; green arrowheads indicate the blood vessels; black arrows indicate the most coronal position of newly formed bone; yellow arrowheads indicate multinucleated giant cell; NB, newly formed bone; asterisk indicates CO<sub>3</sub>Ap particle). (<b>a</b>) The Unfilled group shows minimal new bone formation. (<b>b</b>–<b>d</b>) In the FGF-2, CO<sub>3</sub>Ap, and FGF-2 + CO<sub>3</sub>Ap groups, novel bone can be observed in the Root side of the intrabony defect. (<b>e</b>,<b>f</b>) Higher-magnification images of the framed area in the corresponding group (original magnification ×100; scale bar = 200 µm). Fibrous connective tissue surrounded the CO<sub>3</sub>Ap granules, with the presence of a multinucleated giant cell observed. (<b>g</b>–<b>j</b>) Images at 4 weeks (original magnification ×25; bar = 500 µm). (<b>e</b>) The Unfilled group exhibits limited new bone formation. (<b>h</b>–<b>j</b>) Newly formed bone in the FGF-2, CO<sub>3</sub>Ap, and FGF-2 + CO<sub>3</sub>Ap groups appears to be greater compared with the Unfilled group. (<b>k</b>,<b>l</b>) At 4 weeks, an enlarged image of the framed area was captured (original magnification ×100; scale bar = 200 µm).</p>
Full article ">Figure 3
<p>Immunohistochemistry for Osx. Positive cells are assessed in the Root side (<b>a</b>,<b>d</b>), Bone side (<b>b</b>,<b>e</b>), and Middle area (<b>c</b>,<b>f</b>). A brown coloration shows an Osx-positive reaction (arrowheads indicate the general location of the positive cells). At 2 weeks (<b>a</b>–<b>c</b>), the Osx-positive cells were observed in new bone and existing bone and around CO<sub>3</sub>Ap. At 4 weeks (<b>d</b>–<b>f</b>), the number of Osx-positive cells in the Root side and Bone side appears to be greater in the FGF-2 + CO<sub>3</sub>Ap group compared with the Unfilled group. (Osx and counterstaining with Mayer’s hematoxylin stain, original magnification ×200; bar = 50 µm; asterisk shows CO<sub>3</sub>Ap particles.)</p>
Full article ">Figure 4
<p>Immunohistochemical staining for OCN. Prevalence of OCN-positive cells is assessed in the Root side (<b>a</b>,<b>d</b>), Bone side (<b>b</b>,<b>e</b>), and Middle area (<b>c</b>,<b>f</b>). A brown coloration shows an OCN-positive reaction (arrowheads indicate the general location of the positive cells). At 2 weeks (<b>a</b>–<b>c</b>), the OCN-positive cells were observed in new bone, existing bone, and CO<sub>3</sub>Ap. At 4 weeks (<b>d</b>–<b>f</b>), the number of OCN-positive cells in the Root side and Middle area appears to be greater in the FGF-2 + CO<sub>3</sub>Ap groups compared with the Unfilled group. (OCN and counterstaining with Mayer’s hematoxylin stain, original magnification ×200; bar = 50 µm; asterisk indicates CO<sub>3</sub>Ap particles.)</p>
Full article ">Figure 5
<p>MC3T3-E1 cells cultured on the CO<sub>3</sub>Ap with/without FGF-2. SEM images show that MC3T3-E1 attached to the CO<sub>3</sub>Ap (<b>a</b>,<b>c</b>) and FGF-2 + CO<sub>3</sub>Ap (<b>b</b>,<b>d</b>) at 24 h. (<b>a</b>–<b>d</b>) A greater number of cells appear to be attached to the FGF-2-treated CO<sub>3</sub>Ap compared with CO<sub>3</sub>Ap. In the FGF-2 + CO<sub>3</sub>Ap group, cell protrusions are more evident compared with the CO<sub>3</sub>Ap group. (<b>a</b>,<b>b</b>) Original magnification ×130; bar = 400 μm. (<b>c</b>,<b>d</b>) Original magnification ×1000; bar = 50 μm. Yellow arrowheads indicate MC3T3-E1. CLSM images reveal cells stained for actin (green) and the nucleus (blue) at 24 h (<b>e</b>,<b>f</b>). Higher-magnification images are shown in the insets. Compared with the CO<sub>3</sub>Ap group (<b>e</b>), a greater number of attached cells are observed in the FGF-2 + CO<sub>3</sub>Ap group (<b>f</b>). (Original magnification ×100; bar = 100 μm.)</p>
Full article ">Figure 6
<p>Viability/proliferation of MC3T3-E1 cells. Cells were seeded onto the CO<sub>3</sub>Ap with/without FGF-2 in the culture media and allowed to grow for up to 5 days. The WST-8 assay was employed to assess cell viability and proliferation at the indicated time points. The reference absorbance at 450 nm was subtracted from the absorbance of each sample, and the resulting values were expressed relative to those at 0 h. The FGF-2 + CO<sub>3</sub>Ap group exhibited significantly higher viability/proliferation compared with the CO<sub>3</sub>Ap group at 1, 3, and 5 days. Data are shown as mean ± SD (<span class="html-italic">n</span> = 6). ** <span class="html-italic">p</span> &lt; 0.01, by Mann–Whitney U test. ††† <span class="html-italic">p</span> &lt; 0.001 significant difference from one day values by Kruskal–Wallis test with Dunn’s post hoc test.</p>
Full article ">Figure 7
<p>qRT-PCR assessment of the expression levels of <span class="html-italic">Runx2</span> and <span class="html-italic">Sp7</span>. Relative <span class="html-italic">Runx2</span> (<b>a</b>) and <span class="html-italic">Sp7</span> (<b>b</b>) expression levels in MC3T3-E1 cells on CO<sub>3</sub>Ap with/without FGF-2 at 7 days. The CO<sub>3</sub>Ap group demonstrated significantly higher expression levels of <span class="html-italic">Runx2</span> and <span class="html-italic">Sp7</span> compared with the FGF-2 + CO<sub>3</sub>Ap group at 7 days. The resulting values are expressed relative to the control group (control group; culture medium only). Data are presented as mean ± SD (<span class="html-italic">n</span> = 6). ** <span class="html-italic">p</span> &lt; 0.01, by Mann–Whitney U test.</p>
Full article ">
Back to TopTop