[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
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
remove_circle_outline

Search Results (3,790)

Search Parameters:
Keywords = MSCs

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
13 pages, 2539 KiB  
Article
Differential Effects of Four Canonical Notch-Activating Ligands on c-Kit+ Cardiac Progenitor Cells
by Matthew Robeson, Steven L. Goudy and Michael E. Davis
Int. J. Mol. Sci. 2024, 25(20), 11182; https://doi.org/10.3390/ijms252011182 (registering DOI) - 17 Oct 2024
Abstract
Notch signaling, an important signaling pathway in cardiac development, has been shown to mediate the reparative functions of c-kit+ progenitor cells (CPCs). However, it is unclear how each of the four canonical Notch-activating ligands affects intracellular processes in c-kit+ cells when used as [...] Read more.
Notch signaling, an important signaling pathway in cardiac development, has been shown to mediate the reparative functions of c-kit+ progenitor cells (CPCs). However, it is unclear how each of the four canonical Notch-activating ligands affects intracellular processes in c-kit+ cells when used as an external stimulus. Neonatal c-kit+ CPCs were stimulated using four different chimeric Notch-activating ligands tethered to Dynabeads, and the resulting changes were assessed using TaqMan gene expression arrays, with subsequent analysis by principal component analysis (PCA). Additionally, functional outcomes were measured using an endothelial cell tube formation assay and MSC migration assay to assess the paracrine capacity to stimulate new vessel formation and recruit other reparative cell types to the site of injury. Gene expression data showed that stimulation with Jagged-1 is associated with the greatest pro-angiogenic gene response, including the expression of VEGF and basement membrane proteins, while the other canonical ligands, Jagged-2, Dll-1, and Dll-4, are more associated with regulatory and epigenetic changes. The functional assay showed differential responses to the four ligands in terms of angiogenesis, while none of the ligands produced a robust change in migration. These data demonstrate how the four Notch-activating ligands differentially regulate CPC gene expression and function. Full article
(This article belongs to the Special Issue Notch Signaling Pathways)
Show Figures

Figure 1

Figure 1
<p>Fc-Jagged functionalized beads bind and activate Notch in vitro. (<b>A</b>) Fc-Jagged1 functionalized beads aggregate on the surface of c-kit+ hCPCs compared to nonspecific human IgG functionalized beads. (<b>B</b>,<b>C</b>) Fluorescent YFP is expressed when cells are stimulated with Jagged-1 beads. The response is attenuated when the small molecule Notch inhibitor DAPT (10 µmol/L) is added to cell media. N = 6 per group. (***) <span class="html-italic">p</span> &lt; 0.005, one-way ANOVA.</p>
Full article ">Figure 2
<p>The four canonical Notch ligands drive substantially different gene expression profiles in c-kit+ hCPCs. (<b>A</b>) HES1 expression is upregulated in all four treatment groups, though Jag1 and Jag2 drive the strongest response. HEY1 expression is upregulated by Jag1, Jag2, and Dll4, but not Dll1. N = 4 per group, (**) <span class="html-italic">p</span> &lt; 0.005, (*) <span class="html-italic">p</span> &lt; 0.05, (n.s.) not significant, one-way ANOVA. (<b>B</b>) HES1 protein level increases upon activation of Notch by Jag1/Jag2/Dll1/Dll4. (<b>C</b>) Heat map demonstrating gene expression in response to all 4 ligands. Gene expression was normalized to IgG control beads and presented as fold change over control.</p>
Full article ">Figure 3
<p>PCA reveals clusters of genes that co-vary closely with particular Notch-activating ligands. (<b>A</b>) Genes associated with pro-angiogenic factors co-vary with Jag1 stimulation. (<b>B</b>) Genes associated with ECM remodeling co-vary with Jag1 stimulation. (<b>C</b>) Genes associated with Notch receptors and ligands co-vary with Jag2/Dll1/Dll4 stimulation. (<b>D</b>) Genes associated with Hitone remodeling co-vary with Jag2/Dll1/Dll4 stimulation.</p>
Full article ">Figure 4
<p>Endothelial cell tube formation is enhanced across all treatment groups compared to IgG, but differences between the four canonical ligands are not significant. (<b>A</b>) Representative fluorescent images of endothelial cell tubes after 6 h. (<b>B</b>) Quantified average tube length. N = 16 for each treatment group. (*) <span class="html-italic">p</span> &lt; 0.05, (**) <span class="html-italic">p</span> &lt; 0.005, (****) <span class="html-italic">p</span> &lt; 0.00005, (ns) Not significant, one-way ANOVA.</p>
Full article ">Figure 5
<p>MSC migration in response to conditioned media is upregulated across all treatment groups, though differences between groups are non-significant. (<b>A</b>) MSCs are seeded with serum-free media in the upper well of a transwell insert, with conditioned media positioned below. Migration is quantified after overnight incubation. (<b>B</b>) MSC migration quantified by fluorescence measurement.</p>
Full article ">Figure 6
<p>Illustrated design of the study. Protein G Dynabeads are functionalized overnight by incubation with Fc-conjugated Notch-activating ligands. Following functionalization, c-kit+ CPCs are incubated with functionalized beads in a well plate for 48 h. At the conclusion of this, RNA and conditioned media are collected to perform gene expression analysis via PCR array, endothelial cell tube formation assays, and a transwell assay to assess the migration of MSCs.</p>
Full article ">
19 pages, 1565 KiB  
Review
The Role of Mitochondrial Permeability Transition in Bone Metabolism, Bone Healing, and Bone Diseases
by Xiting Zhu, Ziqi Qin, Min Zhou, Chen Li, Junjun Jing, Wushuang Ye and Xueqi Gan
Biomolecules 2024, 14(10), 1318; https://doi.org/10.3390/biom14101318 - 17 Oct 2024
Abstract
Bone is a dynamic organ with an active metabolism and high sensitivity to mitochondrial dysfunction. The mitochondrial permeability transition pore (mPTP) is a low-selectivity channel situated in the inner mitochondrial membrane (IMM), permitting the exchange of molecules of up to 1.5 kDa in [...] Read more.
Bone is a dynamic organ with an active metabolism and high sensitivity to mitochondrial dysfunction. The mitochondrial permeability transition pore (mPTP) is a low-selectivity channel situated in the inner mitochondrial membrane (IMM), permitting the exchange of molecules of up to 1.5 kDa in and out of the IMM. Recent studies have highlighted the critical role of the mPTP in bone tissue, but there is currently a lack of reviews concerning this topic. This review discusses the structure and function of the mPTP and its impact on bone-related cells and bone-related pathological states. The mPTP activity is reduced during the osteogenic differentiation of mesenchymal stem cells (MSCs), while its desensitisation may underlie the mechanism of enhanced resistance to apoptosis in neoplastic osteoblastic cells. mPTP over-opening triggers mitochondrial swelling, regulated cell death, and inflammatory response. In particular, mPTP over-opening is involved in dexamethasone-induced osteoblast dysfunction and bisphosphonate-induced osteoclast apoptosis. In vivo, the mPTP plays a significant role in maintaining bone homeostasis, with many bone disorders linked to its excessive opening. Genetic deletion or pharmacological inhibition of the over-opening of mPTP has shown potential in enhancing bone injury recovery and alleviating bone diseases. Here, we review the findings on the relationship of the mPTP and bone at both the cellular and disease levels, highlighting novel avenues for pharmacological approaches targeting mitochondrial function to promote bone healing and manage bone-related disorders. Full article
(This article belongs to the Section Cellular Biochemistry)
Show Figures

Figure 1

Figure 1
<p>mPTP structure and the consequences of mPTP opening. The mitochondrial permeability transition pore (mPTP) has different conductance states. In its low-conductance state, the mPTP, potentially formed by the adenine nucleotide translocator (ANT), allows for the passage of ions and small metabolites. Ca<sup>2+</sup> efflux in this state limits the Ca<sup>2+</sup>-dependent tricarboxylic acid (TCA) cycle. In its high-conductance state, the mPTP is formed by a rearrangement of the F<sub>1</sub>F<sub>o</sub> ATP synthase complex. This state has more detrimental effects on cellular function. The release of mitochondrial DNA (mtDNA) through the mPTP triggers inflammatory responses. Extensive water influx causes mitochondrial swelling and subsequently induces outer membrane permeabilisation and the release of pro-apoptotic cofactors, leading to apoptosis or necrosis and TCA cycle collapse. ↓(orange downward arrow) represents downregulation of the TCA cycle.</p>
Full article ">Figure 2
<p>Diverse roles of mPTP opening in bone-related cells. The mitochondrial permeability transition pore (mPTP) is crucial in various functions within bone-related cells. In bone marrow-derived mesenchymal stem cells (BMSCs), cyclophilin D (CypD) downregulation and subsequent mPTP activity decline are required for osteogenic differentiation. In osteoblasts, mPTP over-opening is involved in several drug-induced regulated cell deaths (RCDs) in an ROS-dependent or ROS-independent manner. The mechanism by which osteosarcoma cells are resistant to apoptosis is, at least in part, due to desensitisation to the mPTP. Several studies have indicated that the mPTP is also involved in osteoclast apoptosis. ↑(black upward arrow) and ↓(black downward arrow) represents upregulation and downregulation of mPTP opening, respectively.</p>
Full article ">
18 pages, 901 KiB  
Systematic Review
Characterization of the Joint Microenvironment in Osteoarthritic Joints for In Vitro Strategies for MSC-Based Therapies: A Systematic Review
by Aline Silvestrini da Silva, Fernanda Campos Hertel, Fabrício Luciani Valente, Fabiana Azevedo Voorwald, Andrea Pacheco Batista Borges, Adriano de Paula Sabino, Rodrigo Viana Sepulveda and Emily Correna Carlo Reis
Appl. Biosci. 2024, 3(4), 450-467; https://doi.org/10.3390/applbiosci3040029 - 17 Oct 2024
Abstract
Osteoarthritis is a joint disease that causes pain, stiffness, and reduced joint function because the protective cushioning inside the joints, called cartilage, gradually wears away. This condition is caused by various factors and complex processes in the joint’s environment, involving different types of [...] Read more.
Osteoarthritis is a joint disease that causes pain, stiffness, and reduced joint function because the protective cushioning inside the joints, called cartilage, gradually wears away. This condition is caused by various factors and complex processes in the joint’s environment, involving different types of cells producing factors that can either maintain the joint health or contribute to osteoarthritis. This study aimed to understand the factors influencing both healthy and diseased joints in DDD strategies for the in vitro preconditioning of MSCs. An electronic search in the PubMed, Scopus, and Web of Science databases was carried out using the terms (cartilage OR chondr*) AND (repair OR regeneration OR healing) AND (niche OR microenvironment)) AND (“growth factor” OR GF OR cytokine). Researchers used various methods, including macroscopic examinations, histology, immunohistochemistry, and microCT. Molecules associated with joint inflammation were identified, like macrophage markers, MMP-13, TNF, apoptotic markers, and interleukins. Chondrogenesis-related factors such as aggrecan GAG, collagen type II, and TGF beta family were also identified. This study suggests that balancing certain molecules and ensuring the survival of joint chondrocytes could be crucial in improving the condition of osteoarthritic joints, emphasizing the importance of chondrocyte survival and activity. Future preconditioning methods for MSC- and EV-based therapies can find suitable strategies in the described microenvironments to explore co-culture systems and soluble or extracellular matrix factors. Full article
(This article belongs to the Special Issue Anatomy and Regenerative Medicine: From Methods to Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Flowchart of the systematic review mechanism.</p>
Full article ">Figure 2
<p>Risk of bias in assessing the methodological quality of the 20 articles included in this systematic review.</p>
Full article ">
20 pages, 1417 KiB  
Review
Molecular and Cellular Mechanisms of the Therapeutic Effect of Mesenchymal Stem Cells and Extracellular Vesicles in Corneal Regeneration
by Nina Kobal, Miha Marzidovšek, Petra Schollmayer, Elvira Maličev, Marko Hawlina and Zala Lužnik Marzidovšek
Int. J. Mol. Sci. 2024, 25(20), 11121; https://doi.org/10.3390/ijms252011121 - 16 Oct 2024
Viewed by 266
Abstract
The cornea is a vital component of the visual system, and its integrity is crucial for optimal vision. Damage to the cornea resulting from trauma, infection, or disease can lead to blindness. Corneal regeneration using mesenchymal stem cells (MSCs) and MSC-derived extracellular vesicles [...] Read more.
The cornea is a vital component of the visual system, and its integrity is crucial for optimal vision. Damage to the cornea resulting from trauma, infection, or disease can lead to blindness. Corneal regeneration using mesenchymal stem cells (MSCs) and MSC-derived extracellular vesicles (MSC-EVs) offers a promising alternative to corneal transplantation. MSCs are multipotent stromal cells that can differentiate into various cell types, including corneal cells. They can also secrete a variety of anti-inflammatory cytokines and several growth factors, promoting wound healing and tissue reconstruction. This review summarizes the current understanding of the molecular and cellular mechanisms by which MSCs and MSC-EVs contribute to corneal regeneration. It discusses the potential of MSCs and MSC-EV for treating various corneal diseases, including corneal epithelial defects, dry eye disease, and keratoconus. The review also highlights finalized human clinical trials investigating the safety and efficacy of MSC-based therapy in corneal regeneration. The therapeutic potential of MSCs and MSC-EVs for corneal regeneration is promising; however, further research is needed to optimize their clinical application. Full article
(This article belongs to the Special Issue Recent Advances in Molecular and Cellular Research in Ophthalmology)
Show Figures

Figure 1

Figure 1
<p>Schematic cross-section of a human eye with an expanded view of the cornea.</p>
Full article ">Figure 2
<p>Various corneal pathologies can lead to corneal edema and opacification (<b>A</b>,<b>B</b>), which is conventionally treated by surgical removal and donor corneal transplantation (penetrating (<b>C</b>) and lamellar keratoplasties (<b>D</b>)). (<b>A</b>) Bullous keratopathy and corneal scar years after penetrating injury and (<b>C</b>) 1 month after PK. (<b>B</b>) Bullous keratopathy years after cataract surgery and (<b>D</b>) 1 month after DMEK. (<b>E</b>) A schematic representation of different types of corneal transplantation techniques. The blue section represents the recipient cornea and the yellow section in the red square–dot line represents the transplanted donor corneal graft tissue. In PK, all corneal layers are transplanted, whereas in DALK, only the anterior corneal layers are transplanted. Posterior lamellar techniques involve selective removal of the patient’s Descemet membrane (DM) and endothelium, which is followed by either the transplantation of the donor corneal endothelium, the DM and a thin stromal layer in DSAEK or by the transplantation of only the donor DM and the endothelium in DMEK. Abbreviations: PK—penetrating keratoplasty; DALK—deep anterior lamellar keratoplasty; DSAEK—Descemet’s stripping automated endothelial keratoplasty; DMEK—Descemet’s membrane endothelial keratoplasty.</p>
Full article ">Figure 3
<p>Schematic representation of different regenerative corneal therapies. Abbreviations: LESC—limbal epithelial stem cell; CSSC—corneal stromal stem cell; CEnC—corneal endothelial cell; MSC—mesenchymal stem cell; SC—stem cell; iPSC—induced pluripotent stem cell; AM—amniotic membrane; MSC-EV—extracellular vesicle derived from mesenchymal stem cell; MSC(M)—bone marrow-derived mesenchymal stem cell; MSC(UC)—umbilical cord-derived mesenchymal stem cell.</p>
Full article ">Figure 4
<p>Schematic representation of mesenchymal stem cell-derived extracellular vesicles (such as exosomes, microvesicles and apoptotic bodies), their biogenesis and transfer from cell of origin to recipient cell. They are capable of transferring bioactive molecules to recipient cells through three different mechanisms: endocytosis, specific receptor–ligand interactions and direct fusion. Extracellular vesicles have various effects on corneal cells. Abbreviations: MSC—mesenchymal stem cell; ECM—extracellular matrix; SC—stem cell.</p>
Full article ">
21 pages, 9532 KiB  
Article
Dual-Function Femtosecond Laser: β-TCP Structuring and AgNP Synthesis via Photoreduction with Azorean Green Tea for Enhanced Osteointegration and Antibacterial Properties
by Marco Oliveira, Liliya Angelova, Liliana Grenho, Maria Helena Fernandes and Albena Daskalova
Materials 2024, 17(20), 5057; https://doi.org/10.3390/ma17205057 - 16 Oct 2024
Viewed by 307
Abstract
β-Tricalcium phosphate (β-TCP) is a well-established biomaterial for bone regeneration, highly regarded for its biocompatibility and osteoconductivity. However, its clinical efficacy is often compromised by susceptibility to bacterial infections. In this study, we address this limitation by integrating femtosecond (fs)-laser processing with the [...] Read more.
β-Tricalcium phosphate (β-TCP) is a well-established biomaterial for bone regeneration, highly regarded for its biocompatibility and osteoconductivity. However, its clinical efficacy is often compromised by susceptibility to bacterial infections. In this study, we address this limitation by integrating femtosecond (fs)-laser processing with the concurrent synthesis of silver nanoparticles (AgNPs) mediated by Azorean green tea leaf extract (GTLE), which is known for its rich antioxidant and anti-inflammatory properties. The fs laser was employed to modify the surface of β-TCP scaffolds by varying scanning velocities, fluences, and patterns. The resulting patterns, formed at lower scanning velocities, display organized nanostructures, along with enhanced roughness and wettability, as characterized by Scanning Electron Microscopy (SEM), optical profilometry, and contact angle measurements. Concurrently, the femtosecond laser facilitated the photoreduction of silver ions in the presence of GTLE, enabling the efficient synthesis of small, spherical AgNPs, as confirmed by UV–vis spectroscopy, Transmission Electron Microscopy (TEM), and Fourier Transform Infrared Spectroscopy (FTIR). The resulting AgNP-embedded β-TCP scaffolds exhibited a significantly improved cell viability and elongation of human bone marrow mesenchymal stem cells (hBM-MSCs), alongside significant antibacterial activity against Staphylococcus aureus (S. aureus). This study underscores the transformative potential of combining femtosecond laser surface modification with GTLE-mediated AgNP synthesis, presenting a novel and effective strategy for enhancing the performance of β-TCP scaffolds in bone-tissue engineering. Full article
Show Figures

Figure 1

Figure 1
<p>Representative image of four β-TCP samples prepared for fs-laser treatment.</p>
Full article ">Figure 2
<p>Schematic diagram illustrating the laser setup utilized for the surface treatment of β-TCP samples and the subsequent synthesis of AgNPs, each carried out as distinct steps.</p>
Full article ">Figure 3
<p>SEM micrographs illustrating the morphological changes in β-TCP samples induced by fs-laser treatment at various fluences (6.1 and 4.1 J/cm<sup>2</sup>), scanning velocities (1, 3.44, 5, 10, and 15 mm/s), and patterns (linear and crossed). Subfigures: (<b>A</b>) V = 1 mm/s, F = 4.1 J/cm<sup>2</sup>; (<b>B</b>) V = 3.44 mm/s, F = 4.1 J/cm<sup>2</sup>; (<b>C</b>) V = 5 mm/s, F = 4.1 J/cm<sup>2</sup>; (<b>D</b>) V = 1 mm/s, F = 6.1 J/cm<sup>2</sup>; (<b>E</b>) V = 3.44 mm/s, F = 6.1 J/cm<sup>2</sup>; (<b>F</b>) V = 5 mm/s, F = 6.1 J/cm<sup>2</sup>; (<b>G</b>) V = 10 mm/s, F = 4.1 J/cm<sup>2</sup>; (<b>H</b>) V = 15 mm/s, F = 4.1 J/cm<sup>2</sup>; (<b>I</b>) V = 1 mm/s (Crossed), F = 4.1 J/cm<sup>2</sup>; (<b>J</b>) V = 10 mm/s, F = 6.1 J/cm<sup>2</sup>; (<b>L</b>) V = 15 mm/s, F = 6.1 J/cm<sup>2</sup>. All micrographs were acquired with an acceleration of 20 kV and a magnification of 5000×.</p>
Full article ">Figure 4
<p>3D optical profilometry images illustrating the groove depth variations in β-TCP samples induced by fs-laser treatment at various fluences (6.1 and 4.1 J/cm<sup>2</sup>), scanning velocities (1, 3.44, 5, 10, and 15 mm/s), and patterns (linear and crossed), acquired with a magnification of 20×.</p>
Full article ">Figure 5
<p>Multivariable bubble plot (<b>A</b>) and biplot of PCA analysis (<b>B</b>) illustrating the effects of fluence, scanning velocity, and patterns on the surface roughness parameter Sa.</p>
Full article ">Figure 6
<p>(<b>A</b>) Graph illustrating the variation in contact angle over time for non-laser-treated samples, indicating changes in wettability. (<b>B</b>) Representative image of a water droplet on a non-laser-treated sample, demonstrating its wettability. (<b>C</b>) Comparison image of a water droplet on an fs-laser-treated sample, highlighting the increased wettability of the treated surface.</p>
Full article ">Figure 7
<p>UV–vis spectra showing the SPR peaks of synthesized AgNPs at laser fluences of 8.1 J/cm<sup>2</sup> (<b>A</b>) and 16.3 J/cm<sup>2</sup> (<b>B</b>). Graphs depicting the pseudo-first-order kinetics for the reduction of Ag<sup>+</sup> ions at fluences of 8.1 J/cm<sup>2</sup> (<b>C</b>) and 16.3 J/cm<sup>2</sup> (<b>D</b>).</p>
Full article ">Figure 8
<p>FTIR spectra showing the characteristic vibrational bands of the GTLE and the synthesized AgNPs.</p>
Full article ">Figure 9
<p>TEM micrographs illustrating the morphology of AgNPs synthesized at fluences of 8.1 J/cm<sup>2</sup> (<b>A</b>) and 16.3 J/cm<sup>2</sup> (<b>B</b>); corresponding histograms showing the size distribution of AgNPs synthesized at these fluences, 8.1 J/cm<sup>2</sup> (<b>C</b>) and 16.3 J/cm<sup>2</sup> (<b>D</b>).</p>
Full article ">Figure 10
<p>Metabolic activity of hBM-MSCs cultured over the fs-laser-treated β-TCP scaffolds for periods up to 12 days. Results are presented relative to the untreated samples (control, set up at 1.0, dotted line). Statistically different from control: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 11
<p>SEM images of human hBM-MSCs cultured over the fs-laser-treated β-TCP scaffolds for 12 days. Low (<b>A</b>–<b>C</b>) and high (<b>D</b>–<b>F</b>) magnification images (1000× and 5000×, respectively). Red arrows: examples of mineralized deposits.</p>
Full article ">Figure 12
<p>Antibacterial activity of the fs-laser-treated β-TCP scaffolds against sessile (<b>A</b>) and planktonic (<b>B</b>) growth of <span class="html-italic">S. aureus</span>. Results are presented relative to the untreated samples (control, set up at 1.0, dotted line). Statistically different from control: ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 13
<p>Representative SEM micrographs showing <span class="html-italic">S. aureus</span> cells (yellow) adhered to the non-laser-treated region of the β-TCP surface (<b>A</b>) and to the bottom of the laser-treated region (<b>B</b>).</p>
Full article ">
19 pages, 4508 KiB  
Article
Limited Adipogenic Differentiation Potential of Human Dental Pulp Stem Cells Compared to Human Bone Marrow Stem Cells
by Isaac Maximiliano Bugueno, Giuseppe Alastra, Anamaria Balic, Bernd Stadlinger and Thimios A. Mitsiadis
Int. J. Mol. Sci. 2024, 25(20), 11105; https://doi.org/10.3390/ijms252011105 - 16 Oct 2024
Viewed by 203
Abstract
Bone marrow and teeth contain mesenchymal stem cells (MSCs) that could be used for cell-based regenerative therapies. MSCs from these two tissues represent heterogeneous cell populations with varying degrees of lineage commitment. Although human bone marrow stem cells (hBMSCs) and human dental pulp [...] Read more.
Bone marrow and teeth contain mesenchymal stem cells (MSCs) that could be used for cell-based regenerative therapies. MSCs from these two tissues represent heterogeneous cell populations with varying degrees of lineage commitment. Although human bone marrow stem cells (hBMSCs) and human dental pulp stem cells (hDPSCs) have been extensively studied, it is not yet fully defined if their adipogenic potential differs. Therefore, in this study, we compared the in vitro adipogenic differentiation potential of hDPSCs and hBMSCs. Both cell populations were cultured in adipogenic differentiation media, followed by specific lipid droplet staining to visualise cytodifferentiation. The in vitro differentiation assays were complemented with the expression of specific genes for adipogenesis and osteogenesis–dentinogenesis, as well as for genes involved in the Wnt and Notch signalling pathways. Our findings showed that hBMSCs formed adipocytes containing numerous and large lipid vesicles. In contrast to hBMSCs, hDPSCs did not acquire the typical adipocyte morphology and formed fewer lipid droplets of small size. Regarding the gene expression, cultured hBMSCs upregulated the expression of adipogenic-specific genes (e.g., PPARγ2, LPL, ADIPONECTIN). Furthermore, in these cells most Wnt pathway genes were downregulated, while the expression of NOTCH pathway genes (e.g., NOTCH1, NOTCH3, JAGGED1, HES5, HEY2) was upregulated. hDPSCs retained their osteogenic/dentinogenic molecular profile (e.g., RUNX2, ALP, COLIA1) and upregulated the WNT-specific genes but not the NOTCH pathway genes. Taken together, our in vitro findings demonstrate that hDPSCs are not entirely committed to the adipogenic fate, in contrast to the hBMSCs, which are more effective to fully differentiate into adipocytes. Full article
Show Figures

Figure 1

Figure 1
<p><b>Experimental set-up of adipogenic induction in hBMSCs and hDPSCs.</b> hDPSCs and hBMSCs were cultured in 2D (6-well plates, 24-well plates, and μ-slide 4-well uncoated plates), and three time points were chosen for gene expression by RT qPCR (0, 7, and 21 days of induction) and six time points for lipid vesicle analysis (0, 7, 10, 14, and 21 days of induction). Bright-field and confocal scanning microscopy analyses before and after specific staining of lipid droplets (Nile red and LipidSpot<sup>TM</sup>) were performed up to 21 days of adipogenic induction.</p>
Full article ">Figure 2
<p><b>Adipogenic differentiation of hBMSCs and hDPSCs.</b> (<b>A</b>) Inverted bright-field microscopy on cells at 0 (a), 7 (b), 10 (c), 14 (d), and 21 (e) days of adipogenic induction in hBMSCs. (<b>B</b>) Inverted bright-field microscopy on cells at 0 (a), 7 (b), 10 (c), 14 (d), and 21 (e) days of adipogenic induction in hDPSCs. (<b>C</b>) Wide-field inverted fluorescence microscopy of hBMSCs at 0 (a,f), 7 (b,g), 10 (c,h), 14 (d,i), and 21 (e,j) days of differentiation (in yellow and red: Nile red staining). All scale bars represent 10 μm; n = 6. Below each photo, the isolated yellow channel allows better visualisation of the lipid droplets, indicated by the white arrows. (<b>D</b>) Wide-field inverted fluorescence microscopy of hDPSCs at 0 (a,f), 7 (b,g), 10 (c,h), 14 (d,i), and 21 (e,j) days of differentiation (in yellow and red: Nile red staining). All scale bars represent 10 μm; n = 6. Below each photo, the yellow channel allows better visualisation of the lipid droplets, indicated by the white arrows. (<b>E</b>,<b>F</b>) Confocal high-speed multispectral spinning-disk microscopy of lipid droplet staining of hBMSCs and hDPSCs at 0 (a), 7 (b), 10 (c), 14 (d), and 21 (e) days of differentiation (green colour: LipidSpot<sup>TM</sup> 488 staining; blue colour: DAPI staining). Scale bars: 10 μm.</p>
Full article ">Figure 3
<p><b>Comparison of the lipid vesicle formation between the hBMSC- and hDPSC-originated adipocytes</b>. (<b>A</b>) Quantification of fluorescence emitted by lipid droplet staining in fixed cells normalised with the cell number in three separate cultures of hBMSCs at 0, 7, 10, 14, and 21 days of differentiation; the curve shows the ratio of the green staining quantification versus the total cell number. (<b>B</b>) Histogram of fluorescence emitted by lipid droplet staining in live cells at 0, 7, 10, 14, and 21 days of adipogenic differentiation of hBMSCs and graphical representation of these values. (<b>C</b>) Quantification of fluorescence emitted by lipid droplet staining in fixed cells normalised with the cell number in three separate cultures of hDPSCs at 0, 7, 10, 14, and 21 days of differentiation; the curve shows the ratio of the green staining quantification versus the total cell number. (<b>D</b>) Histogram of fluorescence emitted by lipid droplet staining in live cells at 0, 7, 10, 14, and 21 days of adipogenic differentiation of hDPSCs and graphical representation of these values. (<b>E</b>) Lipid vesicles’ number quantification normalised by the cell number from live cell lipid droplets’ staining. (<b>F</b>) Lipid vesicles’ size quantification from live cell lipid droplets’ staining. One-way ANOVA followed by Dunnett’s post hoc test was used to compare time points for each cell type. Asterisks represent statistically significant differences between different time points and T0 control (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01, n = 6).</p>
Full article ">Figure 4
<p><b>Expression and comparison of adipogenic genes in cultured hBMSCs and hDPSCs under adipogenic conditions.</b> (<b>A</b>) Relative mRNA expression of the early adipogenic genes <span class="html-italic">PPARγ2</span> and <span class="html-italic">CEBPα</span> in cultured hBMSCs and hDPSCs for 0, 7, and 21 days under adipogenic conditions. (<b>B</b>) Expression of the characteristic late adipogenic genes <span class="html-italic">ADIPOQ</span>, <span class="html-italic">FABP4</span>, and <span class="html-italic">LPL</span> in cultured hBMSCs and hDPSCs at 0, 7, and 21 days of adipogenic induction was analysed by RT qPCR. The value of relative expression on comparison graphs is normalised to hBMSCs at T0 for each gene in statistical analysis. Data are presented as average values ± SD. One-way ANOVA followed by Dunnett’s post hoc test was used to compare time points for each cell type. Two-way ANOVA followed by Šídák post hoc test was used to compare hBMSCs and hDPSCs. Asterisks represent statistically significant differences between different time points and T0 control (* <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; **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 5
<p><b>Expression and comparison of stem cell and osteogenic genes in cultured hBMSCs and hDPSCs under adipogenic conditions</b>. (<b>A</b>) Relative mRNA expression of the stem cell gene <span class="html-italic">CD90</span> in cultured hBMSCs and hDPSCs at 0, 7, and 21 days under adipogenic conditions was analysed by RT qPCR. (<b>B</b>) Relative mRNA expression of the osteogenic genes <span class="html-italic">CD90</span>, <span class="html-italic">COLIII</span>, <span class="html-italic">ALPL</span>, <span class="html-italic">COLIA1</span>, and <span class="html-italic">RUNX2</span> in cultured hBMSCs and hDPSCs at 0, 7, and 21 days upon adipogenic induction was analysed by RT qPCR. The value of relative expression on comparison graphs is normalised to hBMSCs at T0 for the statistical analysis of each gene. Data are presented as average values ± SD. One-way ANOVA followed by Dunnett’s post hoc test was used to compare time points for each cell type. Two-way ANOVA followed by Šídák post hoc test was used to compare hBMSCs and hDPSCs. Asterisks represent statistically significant differences between different time points and T0 control (* <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; **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 6
<p><b>Expression and comparison of WNT and SWAT cell genes in hBMSCs and hDPSCs cultured under adipogenic conditions.</b> (<b>A</b>) Relative mRNA expression of the <span class="html-italic">WNT2</span> and <span class="html-italic">WNT10A</span> genes in cultured hBMSCs and hDPSCs at 0, 7, and 21 days under adipogenic conditions was analysed by RT qPCR. (<b>B</b>) Relative mRNA expression of the <span class="html-italic">PLIN1</span>, <span class="html-italic">DCN</span>, and <span class="html-italic">MFAP4</span> genes in cultured hBMSCs and hDPSCs at 0, 7, and 21 days of adipogenic induction was analysed by RT qPCR. The value of relative expression on comparison graphs is normalised to hBMSCs at T0 for each gene. Statistical analysis data are presented as average values ± SD. One-way ANOVA followed by Dunnett’s post hoc test was used to compare time points for each cell type. Two-way ANOVA followed by Šídák post hoc test was used for comparison between hBMSCs and hDPSCs. Asterisks represent statistically significant differences between different time points and T0 control (<b>*</b> <span class="html-italic">p</span> &lt; 0.05; <b>**</b> <span class="html-italic">p</span> &lt; 0.01; <b>***</b> <span class="html-italic">p</span> &lt; 0.001; <b>****</b> <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 7
<p><b>Expression and comparison of NOTCH pathway genes in hBMSCs and hDPSCs cultured in adipogenic conditions</b>. (<b>A</b>) Relative mRNA expression of <span class="html-italic">NOTCH1</span> and <span class="html-italic">NOTCH3</span> in cultured hBMSCs and hDPSCs at 0, 7, and 21 days of adipogenic induction. (<b>B</b>) Relative mRNA expression of <span class="html-italic">JAGGED1</span> and <span class="html-italic">DELTA-LIKE4</span> (<span class="html-italic">DLL4</span>) in cultured hBMSCs and hDPSCs at 0, 7, and 21 days of adipogenic induction. (<b>C</b>) Relative mRNA expression of the Notch pathway transcription factors <span class="html-italic">HES5</span> and <span class="html-italic">HEY2</span> in cultured hBMSCs and hDPSCs at 0, 7, and 21 days of adipogenic induction. The value of relative expression on comparison graphs is normalised to hBMSCs at T0 for each gene. Statistical analysis data are presented as average values ± SD. One-way ANOVA followed by Dunnett’s post hoc test was used to compare time points for each cell type. Two-way ANOVA followed by Šídák post hoc test was used to compare hBMSCs and hDPSCs. Asterisks represent statistically significant differences between different time points and T0 control (* <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; **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 8
<p><b>Schematic representation of the differentiation of cultured hBMSCs and hDPSCs under adipogenic conditions and the expression of various genes during the differentiation process.</b> Cultured hBMSCs form numerous large lipid droplets (yellow colour). Upregulation (green arrows) of adipogenic genes and concomitant downregulation (red arrows) of WNT, stem cell, and osteogenic genes indicate their unreserved commitment towards the adipogenic fate. By contrast, hDPSCs form tiny and less numerous lipid vesicles when compared to hBMSCs and demonstrate a less evident adipocyte morphology. Furthermore, hDPSCs keep high expression of most WNT, stem cell, and osteogenic genes, which indicates their partial commitment towards the adipogenic fate.</p>
Full article ">
20 pages, 4458 KiB  
Systematic Review
Stem Cells: Present Understanding and Prospects for Regenerative Dentistry
by Angelo Michele Inchingolo, Alessio Danilo Inchingolo, Paola Nardelli, Giulia Latini, Irma Trilli, Laura Ferrante, Giuseppina Malcangi, Andrea Palermo, Francesco Inchingolo and Gianna Dipalma
J. Funct. Biomater. 2024, 15(10), 308; https://doi.org/10.3390/jfb15100308 (registering DOI) - 15 Oct 2024
Viewed by 241
Abstract
Regenerative medicine in dentistry focuses on repairing damaged oral tissues using advanced tools like stem cells, biomaterials, and tissue engineering (TE). Mesenchymal stem cells (MSCs) from dental sources, such as dental pulp and periodontal ligament, show significant potential for tissue regeneration due to [...] Read more.
Regenerative medicine in dentistry focuses on repairing damaged oral tissues using advanced tools like stem cells, biomaterials, and tissue engineering (TE). Mesenchymal stem cells (MSCs) from dental sources, such as dental pulp and periodontal ligament, show significant potential for tissue regeneration due to their proliferative and differentiative abilities. This systematic review, following PRISMA guidelines, evaluated fifteen studies and identified effective strategies for improving dental, periodontal, and bone tissue regeneration through scaffolds, secretomes, and bioengineering methods. Key advancements include the use of dental pulp stem cells (DPSCs) and periodontal ligament stem cells (PDLSCs) to boost cell viability and manage inflammation. Additionally, pharmacological agents like matrine and surface modifications on biomaterials improve stem cell adhesion and promote osteogenic differentiation. By integrating these approaches, regenerative medicine and TE can optimize dental therapies and enhance patient outcomes. This review highlights the potential and challenges in this field, providing a critical assessment of current research and future directions. Full article
(This article belongs to the Section Biomaterials and Devices for Healthcare Applications)
Show Figures

Figure 1

Figure 1
<p>Dental pulp stem cells have the capacity to regenerate in various ways. Diagram illustrating the dental pulp stem cells’ capacity for multi-differentiation in the regeneration of periodontal and dentin–pulp complex tissues.</p>
Full article ">Figure 2
<p>PRISMA flowchart following guidelines.</p>
Full article ">Figure 3
<p>(<b>A</b>): Bias assessment; (<b>B</b>): legend (Xiaomeng Li et al., 2021 [<a href="#B81-jfb-15-00308" class="html-bibr">81</a>], Miao Yu et al., 2021 [<a href="#B82-jfb-15-00308" class="html-bibr">82</a>], Eleonora Cianci et al., 2015 [<a href="#B68-jfb-15-00308" class="html-bibr">68</a>], Wenyan Kang et al., 2019 [<a href="#B54-jfb-15-00308" class="html-bibr">54</a>], Junqing Liu et al., 2019 [<a href="#B83-jfb-15-00308" class="html-bibr">83</a>], Fa-Ming Chen et al., 2016 [<a href="#B84-jfb-15-00308" class="html-bibr">84</a>], Bin Ge et al., 2019 [<a href="#B85-jfb-15-00308" class="html-bibr">85</a>], Linglu Jia et al., 2019 [<a href="#B86-jfb-15-00308" class="html-bibr">86</a>], Hong Wang et al., 2019 [<a href="#B87-jfb-15-00308" class="html-bibr">87</a>], C-Y Lin et al., 2018 [<a href="#B88-jfb-15-00308" class="html-bibr">88</a>], Shuchen Li et al., 2018 [<a href="#B89-jfb-15-00308" class="html-bibr">89</a>], Jing Li et al., 2020 [<a href="#B90-jfb-15-00308" class="html-bibr">90</a>], Hynmin Choi et al., 2017 [<a href="#B74-jfb-15-00308" class="html-bibr">74</a>], Tara Gross et al., 2023 [<a href="#B91-jfb-15-00308" class="html-bibr">91</a>], Samer Hanna et al., 2023 [<a href="#B92-jfb-15-00308" class="html-bibr">92</a>]).</p>
Full article ">Figure 3 Cont.
<p>(<b>A</b>): Bias assessment; (<b>B</b>): legend (Xiaomeng Li et al., 2021 [<a href="#B81-jfb-15-00308" class="html-bibr">81</a>], Miao Yu et al., 2021 [<a href="#B82-jfb-15-00308" class="html-bibr">82</a>], Eleonora Cianci et al., 2015 [<a href="#B68-jfb-15-00308" class="html-bibr">68</a>], Wenyan Kang et al., 2019 [<a href="#B54-jfb-15-00308" class="html-bibr">54</a>], Junqing Liu et al., 2019 [<a href="#B83-jfb-15-00308" class="html-bibr">83</a>], Fa-Ming Chen et al., 2016 [<a href="#B84-jfb-15-00308" class="html-bibr">84</a>], Bin Ge et al., 2019 [<a href="#B85-jfb-15-00308" class="html-bibr">85</a>], Linglu Jia et al., 2019 [<a href="#B86-jfb-15-00308" class="html-bibr">86</a>], Hong Wang et al., 2019 [<a href="#B87-jfb-15-00308" class="html-bibr">87</a>], C-Y Lin et al., 2018 [<a href="#B88-jfb-15-00308" class="html-bibr">88</a>], Shuchen Li et al., 2018 [<a href="#B89-jfb-15-00308" class="html-bibr">89</a>], Jing Li et al., 2020 [<a href="#B90-jfb-15-00308" class="html-bibr">90</a>], Hynmin Choi et al., 2017 [<a href="#B74-jfb-15-00308" class="html-bibr">74</a>], Tara Gross et al., 2023 [<a href="#B91-jfb-15-00308" class="html-bibr">91</a>], Samer Hanna et al., 2023 [<a href="#B92-jfb-15-00308" class="html-bibr">92</a>]).</p>
Full article ">Figure 4
<p>Three key components are the focus of TE: biocompatible scaffolds, stem cells, and bioactive compounds such as drug or growth hormone delivery systems creating materials with strong regenerative potential for the purpose of repairing or regenerating damaged oral tissues in regenerative dentistry.</p>
Full article ">Figure 5
<p>Principles of TE: Various cells extracted from the oral cavity, e.g., pulp tissue, apical papilla, gingival tissue, and periodontal ligament tissue, are seeded on growth factor-soaked scaffolds. The required tissues are obtained after appropriate in vitro culture and finally implanted in vivo.</p>
Full article ">Figure 6
<p>Control of the differentiation of stem cells derived from oral tissues mediated by circRNA.</p>
Full article ">
18 pages, 3455 KiB  
Article
Isolation and Characterization of Canine Adipose-Derived Mesenchymal Stromal Cells: Considerations in Translation from Laboratory to Clinic
by Michael A. Rivera Orsini, Emine Berfu Ozmen, Alyssa Miles, Steven D. Newby, Nora Springer, Darryl Millis and Madhu Dhar
Animals 2024, 14(20), 2974; https://doi.org/10.3390/ani14202974 (registering DOI) - 15 Oct 2024
Viewed by 290
Abstract
In allogeneic MSC implantation, the cells are isolated from a donor different from the recipient. When tested, allogeneic MSCs have several advantages over autologous ones: faster cell growth, sufficient cell concentration, and readily available cells for clinics. To ensure the safe and efficient [...] Read more.
In allogeneic MSC implantation, the cells are isolated from a donor different from the recipient. When tested, allogeneic MSCs have several advantages over autologous ones: faster cell growth, sufficient cell concentration, and readily available cells for clinics. To ensure the safe and efficient use of allogeneic MSCs in clinics, the MSCs need to be first tested in vitro. With this study, we paved the way by addressing the in vitro aspects of canine adipose-derived MSCs, considering the limited studies on the clinical use of canine cells. We isolated cAD-MSCs from canine falciform ligament fat and evaluated their viability and proliferation using an MTS assay. Then, we characterized the MSC-specific antigens using immunophenotyping and immunofluorescence and demonstrated their potential for in vitro differentiation. Moreover, we established shipping and cryobanking procedures to lead the study to become an off-the-shelf therapy. During expansion, the cells demonstrated a linear increase in cell numbers, confirming their proliferation quantitatively. The cells showed viability before and after cryopreservation, demonstrating that cell viability can be preserved. From a clinical perspective, the established shipping conditions demonstrated that the cells retain their viability for up to 48 h. This study lays the groundwork for the potential use of allogeneic cAD-MSCs in clinical applications. Full article
Show Figures

Figure 1

Figure 1
<p>The proliferation of cAD-MSCs at passage 2 was assessed by MTS proliferation assay. Data were normalized using cell growth media alone as the control. Note the linear trend (R<sup>2</sup> = 0.9477) in proliferation with time.</p>
Full article ">Figure 2
<p>MSC surface marker expressions. Representative images of immunofluorescence show the expression of CD29 (<b>A</b>), CD44 (<b>B</b>), and CD90 (<b>C</b>). Cells were fixed and stained at 24 h post-seeding.</p>
Full article ">Figure 3
<p>In vitro differentiation of cAD-MSCs. Representative images demonstrating the in vitro potential of cAD-MSCs to undergo differentiation. (<b>A</b>) Adipogenic differentiation was examined using Oil Red O staining. Note that the cell morphology and the Oil Red O-stained cells appear at around day 9 and continue to progress with time. (<b>B</b>) Osteogenic differentiation was examined using alizarin red staining. Note that the cell morphology and the alizarin red-stained cells appear as nodules around day 17 and progress with time. Nodules that are rich in calcium are the hallmark features of osteogenic differentiation. Insets show the corresponding undifferentiated control cAD-MSCs. These cells were maintained in normal growth media without any inducers.</p>
Full article ">Figure 4
<p>Percentage of live to dead cells over a period of 75 h in shipping conditions (ambient temperature of transport vehicle). The viability of cAD-MSCs was maintained in shipping conditions for over 48 h. Viability decreases at 75 h post-preparation.</p>
Full article ">Figure 5
<p>Extracellular matrix (ECM) protein expressions by immunofluorescence. (<b>A</b>) Collagen Type I. (<b>B</b>) Collagen Type II. (<b>C</b>) Fibronectin. (<b>D</b>) Vimentin. (<b>E</b>) Vinculin. (<b>F</b>) F-actin. Even though the ECM expressions are not the same between all these images from different proteins, they demonstrate the expression of all the ECM proteins.</p>
Full article ">
11 pages, 1849 KiB  
Article
Early Effects of Porcine Placental Extracts and Stem Cell-Derived Exosomes on Aging Stress in Skin Cells
by Takaaki Matsuoka, Katsuaki Dan, Keita Takanashi and Akihiro Ogino
J. Funct. Biomater. 2024, 15(10), 306; https://doi.org/10.3390/jfb15100306 (registering DOI) - 15 Oct 2024
Viewed by 247
Abstract
The initial efficacy of placental extracts (Pla-Exts) and human mesenchymal stem-cell-derived exosomes (hMSC-Exo) against aging-induced stress in human dermal fibroblasts (HDFs) was examined. The effect of Pla-Ext alone, hMSC-Exo alone, the combined effect of Pla-Ext and hMSC-Exo, and the effect of hMSC-Exo (Pla/MSC-Exo) [...] Read more.
The initial efficacy of placental extracts (Pla-Exts) and human mesenchymal stem-cell-derived exosomes (hMSC-Exo) against aging-induced stress in human dermal fibroblasts (HDFs) was examined. The effect of Pla-Ext alone, hMSC-Exo alone, the combined effect of Pla-Ext and hMSC-Exo, and the effect of hMSC-Exo (Pla/MSC-Exo) recovered from cultures with Pla-Ext added to hMSC were verified using collagen, elastin, and hyaluronic acid synthase mRNA levels for each effect. Cells were subjected to photoaging (UV radiation), glycation (glycation end-product stimulation), and oxidation (H2O2 stimulation) as HDF stressors. Pla-Ext did not significantly affect normal skin fibroblasts with respect to intracellular parameters; however, a pro-proliferative effect was observed. Pla-Ext induced resistance to several stresses in skin fibroblasts (UV irradiation, glycation stimulation, H2O2 stimulation) and inhibited reactive oxygen species accumulation following H2O2 stimulation. Although the effects of hMSC-Exo alone or the combination of hMSC-Exo and Pla-Ext are unknown, pretreated hMSC-Exo stimulated with Pla-Ext showed changes that conferred resistance to aging stress. This suggests that Pla-Ext supplementation may cause some changes in the surface molecules or hMSC-Exo content (e.g., microRNA). In skin cells, the direct action of Pla-Ext and exosomes secreted from cultured hMSCs pretreated with Pla-Ext (Pla/MSC-Exo) also conferred resistance to early aging stress. Full article
(This article belongs to the Section Biomaterials and Devices for Healthcare Applications)
Show Figures

Figure 1

Figure 1
<p>Effects of Pla-Ext on (<b>a</b>) collagen, (<b>b</b>) elastin, and (<b>c</b>) hyaluronic acid synthase mRNA expression levels in normal human dermal fibroblasts. Pla-Ext, placental extract; hMSC-Exo, human mesenchymal stem-cell-derived exosomes; Pla extract-stimulated hMSC-Exo; placenta extract-stimulated human mesenchymal stem-cell-derived exosomes.</p>
Full article ">Figure 2
<p>Effects of Pla-Ext on normal human dermal fibroblast proliferation. Pla-Ext, placental extract; hMSC-Exo, human mesenchymal stem-cell-derived exosomes; Pla extract-stimulated hMSC-Exo; placenta extract-stimulated human mesenchymal stem-cell-derived exosomes. * <span class="html-italic">p</span> &lt; 0.05 vs. control (DW).</p>
Full article ">Figure 3
<p>Effects of Pla-Ext on (<b>a</b>) collagen, (<b>b</b>) elastin, and (<b>c</b>) hyaluronic acid synthase mRNA expression levels in UV-treated human dermal fibroblasts (5 or 25 min). Pla-Ext, placental extract; hMSC-Exo, human mesenchymal stem-cell-derived exosomes; Pla extract-stimulated hMSC-Exo; placenta extract-stimulated human mesenchymal stem-cell-derived exosomes. * <span class="html-italic">p</span> &lt; 0.001 vs. UV alone.</p>
Full article ">Figure 4
<p>Effects of Pla-Ext on (<b>a</b>) collagen, (<b>b</b>) elastin, and (<b>c</b>) hyaluronic acid synthase mRNA expression levels in human dermal fibroblasts treated with advanced glycation end-products (AGE). Pla-Ext, placental extract; hMSC-Exo, human mesenchymal stem-cell-derived exosomes; Pla extract-stimulated hMSC-Exo; Placenta extract-stimulated human mesenchymal stem-cell-derived exosomes. * <span class="html-italic">p</span> &lt; 0.001 vs. AGE alone.</p>
Full article ">Figure 5
<p>Effects of Pla-Ext on (<b>a</b>) collagen, (<b>b</b>) elastin, and (<b>c</b>) hyaluronic acid synthase in human dermal fibroblasts treated with H<sub>2</sub>O<sub>2</sub>. Pla-Ext, placental extract; hMSC-Exo, human mesenchymal stem-cell-derived exosomes; Pla extract-stimulated hMSC-Exo; Placenta extract-stimulated human mesenchymal stem-cell-derived exosomes. * <span class="html-italic">p</span> &lt; 0.001 vs. H<sub>2</sub>O<sub>2</sub> alone.</p>
Full article ">Figure 6
<p>Effects of Pla-Ext on the generation of reactive oxygen species (ROS) in human dermal fibroblasts treated with H<sub>2</sub>O<sub>2</sub>. Pla-Ext, placental extract; hMSC-Exo, human mesenchymal stem-cell-derived exosomes; Pla extract-stimulated hMSC-Exo; placenta extract-stimulated human mesenchymal stem-cell-derived exosomes. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. H<sub>2</sub>O<sub>2</sub> alone.</p>
Full article ">
10 pages, 433 KiB  
Article
Evaluating Synergistic Effects of Hyaluronic Acid, Human Umbilical Cord-Derived Mesenchymal Stem Cells, and Growth Hormones in Knee Osteoarthritis: A Multi-Arm Randomized Trial
by Ismail Hadisoebroto Dilogo, Anissa Feby Canintika, Bernadus Riyan Hartanto, Jacub Pandelaki and Irsa Gagah Himantoko
Biomedicines 2024, 12(10), 2332; https://doi.org/10.3390/biomedicines12102332 - 14 Oct 2024
Viewed by 396
Abstract
Background: Knee osteoarthritis (OA) significantly affects quality of life and imposes economic burdens due to its prevalence and the disability it causes. The efficacy of current treatments is limited to alleviating the symptoms, and they cannot be used for regenerative purposes. This study [...] Read more.
Background: Knee osteoarthritis (OA) significantly affects quality of life and imposes economic burdens due to its prevalence and the disability it causes. The efficacy of current treatments is limited to alleviating the symptoms, and they cannot be used for regenerative purposes. This study aims to evaluate the efficacy and safety of combining hyaluronic acid (HA), human umbilical cord-derived mesenchymal stem cells (hUC-MSCs), and synthetic human growth hormone (somatotropin) in the treatment of knee OA, assessing pain relief, functional improvement, and cartilage regeneration. Methods: A four-arm, double-blind randomized trial was conducted with 51 knees from 28 subjects aged ≥50 with primary knee OA. The treatments involved were HA alone, HA with hUC-MSCs, HA with somatotropin, and a combination of all three. Efficacy was measured through the International Knee Documentation Committee (IKDC) score, Western Ontario and McMaster Universities Osteoarthritis Index (WOMAC), and visual analog score (VAS), and MRI T2 mapping of cartilage was conducted on pre-implantation at the 6th and 12th month. Results: All treatment arms showed improvements in the VAS and WOMAC scores over 12 months, suggesting some pain relief and functional improvement. However, MRI T2 mapping showed no significant cartilage regeneration across the groups. Conclusions: While the combined use of HA, hUC-MSCs, and somatotropin improved symptoms of knee OA, it did not enhance cartilage regeneration significantly. This study highlights the potential of these combinations for symptom management but underscores the need for further research to optimize these therapies for regenerative outcomes. Full article
(This article belongs to the Special Issue Osteoarthritis: Molecular Pathways and Novel Therapeutic Strategies)
Show Figures

Figure 1

Figure 1
<p>Subject flowchart.</p>
Full article ">
13 pages, 2642 KiB  
Study Protocol
Evaluation of Safety and Efficacy of Cell Therapy Based on Osteoblasts Derived from Umbilical Cord Mesenchymal Stem Cells for Osteonecrosis of the Femoral Head: Study Protocol for a Single-Center, Open-Label, Phase I Clinical Trial
by Seung-Hoon Baek, Bum-Jin Shim, Heejae Won, Sunray Lee, Yeon Kyung Lee, Hyun Sook Park and Shin-Yoon Kim
Pharmaceuticals 2024, 17(10), 1366; https://doi.org/10.3390/ph17101366 - 13 Oct 2024
Viewed by 421
Abstract
Although mesenchymal stem cells (MSCs) insertion has gained recent attention as a joint-preserving procedure, no study has conducted direct intralesional implantation of human umbilical cord-derived MSCs (hUCMSCs) in patients with ONFH. This is a protocol for a phase 1 clinical trial designed to [...] Read more.
Although mesenchymal stem cells (MSCs) insertion has gained recent attention as a joint-preserving procedure, no study has conducted direct intralesional implantation of human umbilical cord-derived MSCs (hUCMSCs) in patients with ONFH. This is a protocol for a phase 1 clinical trial designed to assess the safety and exploratory efficacy of human umbilical cord-derived osteoblasts (hUC-Os), osteogenic differentiation-induced cells from hUCMSCs, in patients with early-stage ONFH. Nine patients with Association Research Circulation Osseous (ARCO) stage 1 or 2 will be assigned to a low-dose (1 × 107 hUC-O cells, n = 3), medium-dose (2 × 107 cells, n = 3), and high-dose group (4 × 107 cells, n = 3) in the order of their arrival at the facility, and, depending on the occurrence of dose-limiting toxicity, up to 18 patients can be enrolled by applying the 3 + 3 escalation method. We will perform hUC-O (CF-M801) transplantation combined with core decompression and follow-up for 12 weeks according to the study protocol. Safety will be determined through adverse event assessment, laboratory tests including a panel reactive antibody test, vital sign assessment, physical examination, and electrocardiogram. Efficacy will be explored through the change in pain visual analog scale, Harris hip score, Western Ontario and McMaster Universities Osteoarthritis Index, ARCO stage, and also size and location of necrotic lesion according to Japanese Investigation Committee classification before and after the procedure. Joint preservation is important, particularly in younger, active patients with ONFH. Confirmation of the safety and efficacy of hUC-Os will lead to a further strategy to preserve joints for those suffering from ONFH and improve our current knowledge of cell therapy. Full article
(This article belongs to the Special Issue New Advances in Mesenchymal Stromal Cells as Therapeutic Tools)
Show Figures

Figure 1

Figure 1
<p>A guide pin is inserted under fluoroscopic guidance towards the necrotic area for a core tract. (<b>A</b>) Anterior posterior view. (<b>B</b>) Lateral view.</p>
Full article ">Figure 2
<p>To guide a following hollow reamer, an entry hole is created using a cannulated solid reamer along the guide pin (<b>A</b>), and an autogenous bone chip is collected during reaming (<b>B</b>).</p>
Full article ">Figure 3
<p>A core tract is created for core decompression using a hollow biopsy cannula (<b>A</b>). During this procedure, a cylindrical autogenous bone block is collected, of which a proximal necrotic portion will be sent for pathologic evaluation and a distal viable portion will be implanted as a bone plug after stem cell insertion (<b>B</b>).</p>
Full article ">Figure 4
<p>A curette is inserted to remove the necrotic lesion (<b>A</b>), followed by washing the necrotic bone debris (<b>B</b>).</p>
Full article ">Figure 5
<p>A mixture of cell and collagen putty is implanted into the lesion (black arrow), and the autogenous bone chip and cylindrical bone block is inserted into the remaining space in the core tract (white arrow). If there is significant space remaining in the distal portion of the core tract, a cylindrical hydroxyapatite and tri-calcium phosphate block may be inserted (arrowhead).</p>
Full article ">
17 pages, 8054 KiB  
Article
Incorporation of Superparamagnetic Magnetic–Fluorescent Iron Oxide Nanoparticles Increases Proliferation of Human Mesenchymal Stem Cells
by Willian Pinheiro Becker, Juliana Barbosa Torreão Dáu, Wanderson de Souza, Rosalia Mendez-Otero, Rosana Bizon Vieira Carias and Jasmin
Magnetochemistry 2024, 10(10), 77; https://doi.org/10.3390/magnetochemistry10100077 - 12 Oct 2024
Viewed by 379
Abstract
Mesenchymal stem cells (MSCs) have significant therapeutic potential and their use requires in-depth studies to better understand their effects. Labeling cells with superparamagnetic iron oxide nanoparticles allows real-time monitoring of their location, migration, and fate post-transplantation. This study aimed to investigate the efficacy [...] Read more.
Mesenchymal stem cells (MSCs) have significant therapeutic potential and their use requires in-depth studies to better understand their effects. Labeling cells with superparamagnetic iron oxide nanoparticles allows real-time monitoring of their location, migration, and fate post-transplantation. This study aimed to investigate the efficacy and cytotoxicity of magnetic–fluorescent nanoparticles in human adipose tissue-derived mesenchymal stem cells (hADSCs). The efficacy of Molday ION rhodamine B (MIRB) labeling in hADSCs was evaluated and their biocompatibility was assessed using various techniques and differentiation assays. Prussian blue and fluorescence staining confirmed that 100% of the cells were labeled with MIRB and this labeling persisted for at least 3 days. Transmission electron microscopy revealed the internalization and clustering of the nanoparticles on the outer surface of the cell membrane. The viability assay showed increased cell viability 3 days after nanoparticle exposure. Cell counts were higher in the MIRB-treated group compared to the control group at 3 and 5 days and an increased cell proliferation rate was observed at 3 days post-exposure. Adipogenic, osteogenic, and chondrogenic differentiation was successfully achieved in all groups, with MIRB-treated cells showing an enhanced differentiation rate into adipocytes and osteocytes. MIRB was efficiently internalized by hADSCs but induced changes in cellular behavior due to the increased cell proliferation rate. Full article
Show Figures

Figure 1

Figure 1
<p>Characterization of MIRB NPs: (<b>A</b>) Transmission electron microscopy (TEM) image of MIRB in water. Images kindly provided by MAGTech Brazil; (<b>B</b>) hydrodynamic diameter (HD, nm) and polydispersity index (PdI) of MIRB NPs in water and cell culture medium determined by dynamic light scattering (DLS). (<b>C</b>) Surface charge analysis of the zeta potential of MIRB (ζ (mV)) in water and culture media. A statistically significant difference was found in the diameter (<span class="html-italic">p</span> = 0.004), PdI (<span class="html-italic">p</span> &lt; 0.0001) and zeta potential (* <span class="html-italic">p</span> = 0.009) between the groups.</p>
Full article ">Figure 2
<p>Labeling of hADSCs with MIRB for 18 h. (<b>A</b>) Control cells. (<b>B</b>) MIRB-treated cells were stained with Prussian blue. (<b>C</b>) Fluorescence optical microscopy image demonstrating the nuclei stained with DAPI (blue) and the nanoparticles (red). Scale bar = 100 μm (<b>A</b>–<b>C</b>). (<b>D</b>) Quantification of MIRB-positive hADSCs. (<b>E</b>) Transmission electron micrograph of control and (<b>F</b>) MIRB-interacting cells. Arrows show clustering of MIRB on the cell membrane, and arrowheads indicate intracellular endosomal vesicles with nanoparticles. (<b>G</b>) Endosomal vesicle at high magnification demonstrating the internalization of MIRB. Scale bar = 2 μm (<b>E</b>,<b>F</b>); 300 nm (<b>G</b>).</p>
Full article ">Figure 3
<p>Colorimetric cell viability assay after treatment with MIRB. The graph represents the cell viability analyzed 3 days after the start of exposure. *** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Quantification of hADSCs treated with MIRB for different durations and stained with Prussian blue and acid fuchsin. (<b>A</b>) Control cells on day 1. (<b>B</b>) MIRB-treated cells on day 1. (<b>C</b>) Control cells on day 2. (<b>D</b>) MIRB-treated cells on day 2. (<b>E</b>) Control cells on day 3. (<b>F</b>) MIRB-treated cells on day 3. Scale bars = 100 μm. (<b>G</b>) hADSCs were treated and quantified daily for up to 5 days after initial exposure. The black bar represents control hADSCs, and the gray bar represents hADSCs treated with MIRB. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>Immunofluorescence of hADSCs treated or not treated with MIRB. (<b>A</b>) Control cells stained with DAPI (blue) and (<b>A’</b>) anti-Ki67 antibody (green) and (<b>A”</b>) overlay of images. (<b>B</b>) Representative images of cells treated with MIRB and stained with DAPI (blue) and (<b>B’</b>) anti-Ki67 antibody (green). (<b>B”</b>) Overlay of images B and B’. (<b>B’”</b>) Overlay of images B and B’ together with rhodamine B (red) detection. Scale bar = 100 μm. (<b>C</b>) Quantification of Ki67-positive hADSCs on day 1 and day 3 after treatment. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Analysis of the adipogenic differentiation potential of MIRB-treated hADSCs revealed by Oil Red O staining. (<b>A</b>) Control cells not induced or (<b>B</b>) induced to adipogenesis. (<b>C</b>) MIRB-induced hADSCs. (<b>D</b>) MIRB-hADSCs were not induced to differentiate but were also stained with Prussian blue and acid fuchsin 1%; note the MIRB-positive cells 21 days after culture. (<b>E</b>) Differentiated MIRB cells stained with Prussian blue and (<b>E’</b>) higher magnification of the marked square in image E demonstrating lipid vacuoles and the presence of MIRB in the same cell after 21 days. (<b>F</b>) Percentage of hADSCs positive for oil red O dye. (<b>G</b>) Quantification of the number of control and MIRB cells. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.001. Scale bar = 100 μm.</p>
Full article ">Figure 7
<p>Osteogenic differentiation of hADSCs after 18 days of staining with Alizarin Red. (<b>A</b>) Quantification of Alizarin Red staining in control and induced cells. (<b>B</b>) Uninduced MIRB-treated cells. (<b>C</b>) Differentiated control hADSCs. (<b>D</b>) MIRB-treated hADSCs were induced to differentiate. ** <span class="html-italic">p</span> &lt; 0.01. Scale bar = 100 μm.</p>
Full article ">Figure 8
<p>Histological micromass sections of hADSCs induced into chondrocytes over 14 days. (<b>A</b>) Histological section of a micromass from induced control cells. (<b>B</b>) Histological section of micromass from induced MIRB-treated cells. (<b>C</b>) Higher magnification of a micromass from the MIRB group. (<b>D</b>) Histological section of micromass from hADSCs treated with MIRB stained with hematoxylin and eosin. Scale bars: 50 μm (<b>A</b>,<b>B</b>,<b>D</b>) and 20 μm (<b>C</b>).</p>
Full article ">
18 pages, 15413 KiB  
Article
Insights into Osteogenesis Induced by Crude Brassicaceae Seeds Extracts: A Role for Glucosinolates
by Laura Gambari, Eleonora Pagnotta, Luisa Ugolini, Laura Righetti, Emanuela Amore, Brunella Grigolo, Giuseppe Filardo and Francesco Grassi
Nutrients 2024, 16(20), 3457; https://doi.org/10.3390/nu16203457 - 12 Oct 2024
Viewed by 379
Abstract
Background/Objectives: Crude extracts from the Brassica genus have recently emerged as promising phytochemicals for preventing bone loss. While the most documented evidence suggests that their general biological activity is due to glucosinolates’ (GLSs’) hydrolysis products, the direct activity of GLSs is beginning [...] Read more.
Background/Objectives: Crude extracts from the Brassica genus have recently emerged as promising phytochemicals for preventing bone loss. While the most documented evidence suggests that their general biological activity is due to glucosinolates’ (GLSs’) hydrolysis products, the direct activity of GLSs is beginning to be uncovered. However, the contribution of GLSs to the bone-sparing activity of crude Brassicaceae extracts has seldom been addressed. Here, we aimed to gain insights into this gap by studying in the same in vitro model of human osteogenesis the effect of two Brassica seed extracts (Eruca sativa and Lepidium sativum) obtained from defatted seed meals, comparing them to the isolated GLSs most represented in their composition, glucoerucin (GER) and glucotropaeolin (GTL), for Eruca sativa and Lepidium sativum, respectively. Methods: Osteogenic differentiation of human mesenchymal stromal cells (hMSCs) was assessed by alizarin red staining assay and real-time PCR, respectively, evaluating mineral apposition and mRNA expression of specific osteogenic genes. Results: Both Brassica extracts and GLSs increased the osteogenic differentiation, indicating that the stimulating effect of Brassica extracts can be at least partially attributed to GLSs. Moreover, these data extend previous evidence of the effect of unhydrolyzed glucoraphanin (GRA) on osteogenesis to other types of GLSs: GER and GTL. Notably, E. sativa extract and GTL induced higher osteogenic stimulation than Lepidium sativum extract and GER, respectively. Conclusions: Overall, this study expands the knowledge on the possible application of Brassica-derived bioactive molecules as natural alternatives for the prevention and treatment of bone-loss pathologies. Full article
(This article belongs to the Special Issue Bioactive Ingredients in Plants Related to Human Health)
Show Figures

Figure 1

Figure 1
<p><b>Typical chromatogram of desulfated glucosinolates in the <span class="html-italic">E. sativa</span> extract.</b> Above UV spectra of internal standard (I.S.) desulfo sinigrin (SIN) and identified desulfoglucosinolates, glucoraphanin (GRA) and glucoerucin (GER) are reported with their retention times (r.t.). The insert on the right shows a chromatogram of purified GER and I.S. for the determination of purity on a weight basis.</p>
Full article ">Figure 2
<p><b>Typical chromatogram of desulfated glucosinolates in the <span class="html-italic">L. sativum</span> extract.</b> Above UV spectra of internal standard (I.S.) desulfo sinigrin (SIN) and glucotropeaolin (GTL) are reported with their retention times (r.t.). The insert on the right shows a chromatogram of isolated GTL and I.S. for the determination of purity on a weight basis.</p>
Full article ">Figure 3
<p><b>AR-S on ES- and GER-treated cells vs. CTRL cells during osteogenic stimulation.</b> Panels (<b>a</b>,<b>b</b>) show representative images at D14 ARS (<b>a</b>) and D21 (<b>b</b>). Panels c and d show histograms (mean ± SEM) of AR-S quantification obtained by 177 measurements in duplicate for each of the N = 8 donors comparing ES (red), GER (orange), CTRL (black) at D14 (<b>c</b>) and D21 (<b>d</b>). Friedman–Dunn’s multiple-comparisons test: *: Comparisons between CTRL and treatments (at each concentration); °: Comparisons between different concentrations of each treatment (10 μM vs. 30 μM; 10 μM vs. 100 μM); #: Comparisons between different treatments at the same concentration (es 10 μM ES vs. 10 μM GER). ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; ****, °°°°, #### <span class="html-italic">p</span> &lt; 0.0001, <span class="html-italic">ns</span> = non significant. ES extract concentration is expressed as GER content (μM).</p>
Full article ">Figure 4
<p><b>AR-S on LS- and GTL-treated cells vs. CTRL cells during osteogenic stimulation.</b> Panels a and b shows representative images at D14 ARS (<b>a</b>) and D21 (<b>b</b>). Panels (<b>c</b>,<b>d</b>) show histograms (mean ± SEM) of AR-S quantification obtained by 177 measurements in duplicate for each of the N = 8 donors comparing LS (green), GTL (blue), CTRL (black) at D14 (<b>c</b>) and D21 (<b>d</b>). Friedman–Dunn’s multiple-comparisons test: *: Comparisons between CTRL and treatments (at each concentration); °: Comparisons between different concentrations of each treatment (10 μM vs. 30 μM; 10 μM vs. 100 μM); #: Comparisons between different treatments at the same concentration (es 10 μM ES vs. 10 μM GER). **, ## <span class="html-italic">p</span> &lt; 0.01; °°° <span class="html-italic">p</span> &lt; 0.001; ****, °°°°, #### <span class="html-italic">p</span> &lt; 0.0001. LS extract concentration is expressed as GTL content (μM).</p>
Full article ">Figure 5
<p>AR-S on ES-, GER-, LS-, and GTL-treated cells vs. CTRL cells during osteogenic stimulation. Panels (<b>a</b>–<b>d</b>) show histograms (mean ± SEM) of AR-S quantification obtained by 177 measurements in duplicate for each of the N = 8 donors comparing (<b>a</b>) ES (red) vs. LS (green) D14; (<b>b</b>) ES vs. LS D21; (<b>c</b>) GER (orange) vs. GTL (blue) (D14); (<b>d</b>) GER vs. GTL (D21). Friedman–Dunn’s multiple-comparisons test: #: Comparisons between different treatments at the same concentration (es 10 μM ES vs. 10 μM GER). #### <span class="html-italic">p</span> &lt; 0.0001. ES and LS extract concentrations are expressed as GER and GTL content (μM), respectively.</p>
Full article ">Figure 6
<p><b>AR-S on ES-, GER-, LS-, and GTL-treated cells vs. CTRL cells during osteogenic stimulation in the total population and selected groups of donors at D14</b>. (<b>a</b>) Histograms (mean ± SEM) of AR-S quantification of 177 values in duplicate for each one of 8 donors, highlighting donors displaying different levels of mineral deposition. Arrows indicate donors displaying early high levels of mineral matrix deposition. (<b>b</b>) Representative images of AR-S at D0 and D14 (magnification 10×), highlighting the two subgroups of donors found at D14: non-producing mineral matrix (non-mineralizing) and highly producing mineral matrix (high-mineralizing). (<b>c</b>) Histograms (mean ± SEM) of AR-S of 177 values in duplicate for each one of 8 donors (TOTAL), non-mineralizing donors (N = 6) and high-mineralizing donors (N = 2) for ES (red), GER (orange), GTL (blue), and LS (green). Friedman–Dunn’s multiple-comparisons test: *: Comparisons between CTRL and treatments (at each concentration); °: Comparisons between different concentrations of each treatment (10 μM vs. 30 μM; 10 μM vs. 100 μM). °°° <span class="html-italic">p</span> &lt; 0.001; ****, °°°° <span class="html-italic">p</span> &lt; 0.0001. ES and LS extract concentrations are expressed as GER and GTL content (μM), respectively.</p>
Full article ">Figure 7
<p><b>AR-S on ES-, GER-, LS-, and GTL-treated cells vs. CTRL cells during osteogenic stimulation in the total population and selected groups of donors at D21.</b> (<b>a</b>) Histograms (mean ± SEM) of AR-S quantification of 177 values in duplicate for each of 8 donors, highlighting donors displaying different levels of mineral deposition. Arrows indicate donors displaying low levels of mineral matrix deposition. (<b>b</b>) Representative images of AR-S at D0 and D21 (magnification 10×), highlighting the two subgroups of donors found at D21: low-producing mineral matrix (low-mineralizing) and highly producing mineral matrix (high-mineralizing). (<b>c</b>) Histograms (mean ± SEM) of AR-S of 177 values in duplicate for each of 8 donors (TOTAL), low-mineralizing donors (N = 2) and high-mineralizing donors (N = 6) for ES, GER, GTL, and LS. Friedman–Dunn’s multiple-comparisons test: *: Comparisons between CTRL and treatments (at each concentration); °: Comparisons between different concentrations of each treatment (10 μM vs. 30 μM; 10 μM vs. 100 μM). *, ° <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; 0.0001. ES and LS extract concentrations are expressed as GER and GTL content (μM), respectively.</p>
Full article ">Figure 8
<p>mRNA expression of osteogenic markers in ES-, GER-, LS-, and GTL-treated cells vs. CTRL cells during osteogenic stimulation at D14. Histograms show mean ± SEM of mRNA expression in duplicate for each of 6 donors (for SMAD-1, WISP-1, BSP, and ALP) for ES (red), GER (orange), LS (green), GTL (blue). Two-way ANOVA and Tukey’s multiple-comparisons test were performed. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 9
<p>mRNA expression of osteogenic markers in ES-, GER-, LS-, and GTL-treated cells vs. CTRL cells during osteogenic stimulation at D21. Histograms show mean ± SEM of mRNA expression in duplicate for each of 6 donors (for SMAD-1, WISP-1, BSP, and ALP) for ES (red), GER (orange), LS (green), GTL (blue). Two-way ANOVA and Tukey’s multiple-comparisons test were performed. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
17 pages, 17639 KiB  
Article
Intravenous Infusion of Autologous Mesenchymal Stem Cells Expanded in Auto Serum for Chronic Spinal Cord Injury Patients: A Case Series
by Ryosuke Hirota, Masanori Sasaki, Satoshi Iyama, Kota Kurihara, Ryunosuke Fukushi, Hisashi Obara, Tsutomu Oshigiri, Tomonori Morita, Masahito Nakazaki, Takahiro Namioka, Ai Namioka, Rie Onodera, Yuko Kataoka-Sasaki, Shinichi Oka, Mitsuhiro Takemura, Ryo Ukai, Takahiro Yokoyama, Yuichi Sasaki, Tatsuro Yamashita, Masato Kobayashi, Yusuke Okuma, Reiko Kondo, Ryo Aichi, Satoko Ohmatsu, Noritaka Kawashima, Yoichi M. Ito, Masayoshi Kobune, Kohichi Takada, Sumio Ishiai, Toru Ogata, Atsushi Teramoto, Toshihiko Yamashita, Jeffery D. Kocsis and Osamu Honmouadd Show full author list remove Hide full author list
J. Clin. Med. 2024, 13(20), 6072; https://doi.org/10.3390/jcm13206072 - 11 Oct 2024
Viewed by 496
Abstract
Objective: The safety, feasibility, and potential functional improvement following the intravenous infusion of mesenchymal stem cells (MSCs) were investigated in patients with chronic severe spinal cord injury (SCI). Methods: The intravenous infusion of autologous MSCs cultured in auto-serum under Good Manufacturing Practices (GMP) [...] Read more.
Objective: The safety, feasibility, and potential functional improvement following the intravenous infusion of mesenchymal stem cells (MSCs) were investigated in patients with chronic severe spinal cord injury (SCI). Methods: The intravenous infusion of autologous MSCs cultured in auto-serum under Good Manufacturing Practices (GMP) was administered to seven patients with chronic SCI (ranging from 1.3 years to 27 years after the onset of SCI). In addition to evaluating feasibility and safety, neurological function was evaluated using the American Spinal Injury Association Impairment Scale (AIS), International Standards for Neurological Classification of Spinal Cord Injury (ISCSCI-92), and Spinal Cord Independence Measure III (SCIM-III). Results: No serious adverse events occurred. Neither CNS tumors, abnormal cell growth, nor neurological deterioration occurred in any patients. While this initial case series was not blinded, significant functional improvements and increased quality of life (QOL) were observed at 90 and 180 days post-MSC infusion compared to pre-infusion status. One patient who had an AIS grade C improved to grade D within six months after MSC infusion. Conclusions: This case series suggests that the intravenous infusion of autologous MSCs is a safe and feasible therapeutic approach for chronic SCI patients. Furthermore, our data showed significant functional improvements and better QOL after MSC infusion in patients with chronic SCI. A blind large-scale study will be necessary to fully evaluate this possibility. Full article
(This article belongs to the Section Clinical Neurology)
Show Figures

Figure 1

Figure 1
<p>Clinical protocol.</p>
Full article ">Figure 2
<p>Case 1. T2-weighted MRI. (<b>A</b>) Sagittal arrows indicate the high-intensity areas. (<b>B</b>) Axial images. The arrowhead indicates the high-intensity areas. (<b>C</b>) Sensory function (pre, post 6M), (<b>D</b>) motor function, (<b>E</b>) sensory function, and (<b>F</b>) SCIM-III score.</p>
Full article ">Figure 3
<p>Case 2. T2-weighted MRI. (<b>A</b>) Sagittal arrows indicate the high-intensity areas. (<b>B</b>) Axial images. The arrowhead indicates the high-intensity areas. (<b>C</b>) Sensory function (pre, post 6M), (<b>D</b>) motor function, (<b>E</b>) sensory function, and (<b>F</b>) SCIM-III score.</p>
Full article ">Figure 4
<p>Case 3. T2-weighted MRI. (<b>A</b>) Sagittal arrows indicate the high-intensity areas. (<b>B</b>) Axial images. The arrowhead indicates the high-intensity areas. (<b>C</b>) Sensory function (pre, post 6M), (<b>D</b>) motor function, (<b>E</b>) sensory function, and (<b>F</b>) SCIM-III score.</p>
Full article ">Figure 5
<p>Case 4. T2-weighted MRI. (<b>A</b>) Sagittal arrows indicate the high-intensity areas. (<b>B</b>) Axial images. The arrowhead indicates the high-intensity areas. (<b>C</b>) Sensory function (pre, post 6M), (<b>D</b>) motor function, (<b>E</b>) sensory function, and (<b>F</b>) SCIM-III score.</p>
Full article ">Figure 6
<p>Case 5. T2-weighted MRI. (<b>A</b>) Sagittal arrows indicate the high-intensity areas. (<b>B</b>) Axial images. The arrowhead indicates the high-intensity areas. (<b>C</b>) Sensory function (pre, post 6M), (<b>D</b>) motor function, (<b>E</b>) sensory function, and (<b>F</b>) SCIM-III score.</p>
Full article ">Figure 7
<p>Case 6. T2-weighted MRI. (<b>A</b>) Sagittal arrows indicate the high-intensity areas. (<b>B</b>) Axial images. The arrowhead indicates the high-intensity areas. (<b>C</b>) Sensory function (pre, post 6M), (<b>D</b>) motor function, (<b>E</b>) sensory function, and (<b>F</b>) SCIM-III score.</p>
Full article ">Figure 8
<p>Case 7. T2-weighted MRI. (<b>A</b>) Sagittal arrows indicate the high-intensity areas. (<b>B</b>) Axial images. The arrowhead indicates the high-intensity areas. (<b>C</b>) Sensory function (pre, post 6M), (<b>D</b>) motor function, (<b>E</b>) sensory function, and (<b>F</b>) SCIM-III score.</p>
Full article ">Figure 9
<p>Outcome measure scores according to AIS C classification ((<b>A</b>): motor; (<b>B</b>): sensory; (<b>C</b>): SCIM-III) prior to MSC infusion, 90 and 180 days post-MSC infusion.</p>
Full article ">Figure 10
<p>Comparison of outcome measure scores before infusion and six months post-MSC infusion based on AIS classification ((<b>A</b>,<b>D</b>): motor; (<b>B</b>,<b>E</b>): sensory; (<b>C</b>,<b>F</b>): SCIM-III).</p>
Full article ">
19 pages, 4062 KiB  
Article
Investigation of the Effect of High Shear Stress on Mesenchymal Stem Cells Using a Rotational Rheometer in a Small-Angle Cone–Plate Configuration
by Mario Mand, Olga Hahn, Juliane Meyer, Kirsten Peters and Hermann Seitz
Bioengineering 2024, 11(10), 1011; https://doi.org/10.3390/bioengineering11101011 - 11 Oct 2024
Viewed by 467
Abstract
Within the healthy human body, cells reside within the physiological environment of a tissue compound. Here, they are subject to constant low levels of mechanical stress that can influence the growth and differentiation of the cells. The liposuction of adipose tissue and the [...] Read more.
Within the healthy human body, cells reside within the physiological environment of a tissue compound. Here, they are subject to constant low levels of mechanical stress that can influence the growth and differentiation of the cells. The liposuction of adipose tissue and the subsequent isolation of mesenchymal stem/stromal cells (MSCs), for example, are procedures that induce a high level of mechanical shear stress. As MSCs play a central role in tissue regeneration by migrating into regenerating areas and driving regeneration through proliferation and tissue-specific differentiation, they are increasingly used in therapeutic applications. Consequently, there is a strong interest in investigating the effects of shear stress on MSCs. In this study, we present a set-up for applying high shear rates to cells based on a rotational rheometer with a small-angle cone–plate configuration. This set-up was used to investigate the effect of various shear stresses on human adipose-derived MSCs in suspension. The results of the study show that the viability of the cells remained unaffected up to 18.38 Pa for an exposure time of 5 min. However, it was observed that intense shear stress damaged the cells, with longer treatment durations increasing the percentage of cell debris. Full article
(This article belongs to the Special Issue Regenerative Technologies in Plastic and Reconstructive Surgery)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic illustration of the cone–plate geometry. As the size of the gap increases with the radius, the shear rate <math display="inline"><semantics> <mrow> <mover accent="true"> <mi>γ</mi> <mo>˙</mo> </mover> <mfenced> <mi>r</mi> </mfenced> </mrow> </semantics></math> is uniform in the measurement geometry in laminar flow conditions. Geometries are not shown to scale.</p>
Full article ">Figure 2
<p>Media density. Density measurements were performed at 25 °C and ambient pressure (mean with standard deviation, n = 3).</p>
Full article ">Figure 3
<p>Influence of shear rate on dynamic viscosity of cell media. Viscosity measured for DMEM (red), DMEM+G (purple), DMEM+G+FCS (green) and water (blue) at 25 °C via rotational rheometer in small-angle cone–plate configuration (mean with standard deviation, n = 6).</p>
Full article ">Figure 4
<p>Viscosity behavior of cell media over exposure time of 10 min at constant shear rate of 1 (blue), 2 (red) and 3 × 10<sup>4</sup> s<sup>−1</sup> (green). Viscosity measured for DMEM (circle), DMEM+G (triangle up), DMEM+G+FCS (triangle down) and water (square) at 25 °C via rotational rheometer in small-angle cone–plate configuration. The decline in apparent viscosity over time showed no correlation with the shear rate and remained constant for all media and shear rates.</p>
Full article ">Figure 5
<p>Agglomeration in center of the cone geometry during the rheological characterization of DMEM+G+FCS at 10 min exposure time. At shear rates of <math display="inline"><semantics> <mrow> <mover accent="true"> <mi>γ</mi> <mo>˙</mo> </mover> <mo>=</mo> <mn>3</mn> <mo>×</mo> <msup> <mrow> <mn>10</mn> </mrow> <mn>4</mn> </msup> <mo> </mo> <msup> <mi mathvariant="normal">s</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>, the proteins in the fetal calf serum appeared to agglomerate and adhered to the measurement geometry surface. The agglomeration led to the exclusion of DMEM+G+FCS as a possible carrier fluid.</p>
Full article ">Figure 6
<p>Characterization of viability and cell size after shearing. Quantification of cell viability (<b>A</b>,<b>B</b>) and relative cell size (<b>C</b>,<b>D</b>) after 5 min (<b>A</b>,<b>C</b>) or 10 min (<b>B</b>,<b>D</b>) shearing at different shear rates (1, 2, 2.5 and 3 × 10<sup>4</sup> s<sup>−1</sup>). Cell cultures without starting the shearing procedure served as control (Ctrl.) cultures (the data set was normalized to the control, depending on the normal data distribution (Shapiro–Wilk test); statistical significance was calculated using an ordinary one-way ANOVA with Dunnett’s multiple comparison test or by Kruskal–Wallis with Dunn’s multiple comparison test (* <span class="html-italic">p</span> &lt; 0.05 significant compared to the Ctrl., n = 6).</p>
Full article ">Figure 7
<p>Analysis of the cell debris fraction after shearing. Representative histograms of cell diameter (<b>A</b>) and quantification of cell debris fraction, defined as cell sizes &lt; 10 µm, (<b>B</b>) after 5 min or 10 min shearing with different shear rates (1, 2, 2.5 and 3 × 10<sup>4</sup> s<sup>−1</sup>). Cell cultures without incipient shearing procedure served as control cultures (data set was normalized to control, Shapiro–Wilk test indicated non-normal data distribution, statistical significance was calculated using a Kruskal–Wallis test with Dunn’s multiple comparison test, * <span class="html-italic">p</span> &lt; 0.05 significant compared to control cultures, n = 6).</p>
Full article ">Figure 8
<p>Effect of shear on adherence capacity and on the F-actin of AD-MSCs after 24 h. Representative images of adherent cell cultures (<b>A</b>) and F-actin (red) and nuclei (blue) staining (<b>B</b>) after 5 min or 10 min shearing with different shear rates (1, 2, 2.5 and 3 × 10<sup>4</sup> s<sup>−1</sup>). Cell cultures without incipient shearing procedure served as control cultures.</p>
Full article ">
Back to TopTop