Review Open Access
Copyright ©The Author(s) 2021. Published by Baishideng Publishing Group Inc. All rights reserved.
World J Stem Cells. Dec 26, 2021; 13(12): 1863-1880
Published online Dec 26, 2021. doi: 10.4252/wjsc.v13.i12.1863
Dental mesenchymal stromal/stem cells in different microenvironments— implications in regenerative therapy
Ivana Okić-Đorđević, Hristina Obradović, Tamara Kukolj, Anđelija Petrović, Slavko Mojsilović, Diana Bugarski, Aleksandra Jauković, Laboratory for Experimental Hematology and Stem Cells, Institute for Medical Research, University of Belgrade, Belgrade 11129, Serbia
ORCID number: Ivana Okić-Đorđević (0000-0003-3552-1546); Hristina Obradović (0000-0003-4626-7184); Tamara Kukolj (0000-0002- 3174-4358); Anđelija Petrović (0000-0001-9011-8800); Slavko Mojsilović (0000-0002-4399-6720); Diana Bugarski (0000-0002-2629-5471); Aleksandra Jauković (0000-0003-2686-7481).
Author contributions: Okić Đorđević I designed and wrote the manuscript; Obradović H contributed to the article’s conception and writing; Kukolj T performed the literature review and article writing and editing; Petrović A contributed to the literature review and article writing; Mojsilović S performed the literature analysis and article writing; Bugarski D and Jauković A made critical revisions to the manuscript; all authors read and approved the final version of the manuscript.
Supported by the Ministry of Education, Science and Technological Development, Republic of Serbia, No. 451-03-9/2021-14/200015.
Conflict-of-interest statement: The authors declare no conflicts of interest.
Open-Access: This article is an open-access article that was selected by an in-house editor and fully peer-reviewed by external reviewers. It is distributed in accordance with the Creative Commons Attribution NonCommercial (CC BY-NC 4.0) license, which permits others to distribute, remix, adapt, build upon this work non-commercially, and license their derivative works on different terms, provided the original work is properly cited and the use is non-commercial. See: https://creativecommons.org/Licenses/by-nc/4.0/
Corresponding author: Ivana Okić-Đorđević, PhD, Assistant Professor, Laboratory for Experimental Hematology and Stem Cells, Institute for Medical Research, University of Belgrade, Dr Subotića 4, PO BOX 102, Belgrade 11129, Serbia. ivana.okic@imi.bg.ac.rs
Received: May 17, 2021
Peer-review started: May 17, 2021
First decision: June 5, 2021
Revised: June 15, 2021
Accepted: November 25, 2021
Article in press: November 25, 2021
Published online: December 26, 2021

Abstract

Current research data reveal microenvironment as a significant modifier of physical functions, pathologic changes, as well as the therapeutic effects of stem cells. When comparing regeneration potential of various stem cell types used for cytotherapy and tissue engineering, mesenchymal stem cells (MSCs) are currently the most attractive cell source for bone and tooth regeneration due to their differentiation and immunomodulatory potential and lack of ethical issues associated with their use. The microenvironment of donors and recipients selected in cytotherapy plays a crucial role in regenerative potential of transplanted MSCs, indicating interactions of cells with their microenvironment indispensable in MSC-mediated bone and dental regeneration. Since a variety of MSC populations have been procured from different parts of the tooth and tooth-supporting tissues, MSCs of dental origin and their achievements in capacity to reconstitute various dental tissues have gained attention of many research groups over the years. This review discusses recent advances in comparative analyses of dental MSC regeneration potential with regards to their tissue origin and specific microenvironmental conditions, giving additional insight into the current clinical application of these cells.

Key Words: Microenvironment, Dental mesenchymal stem cells, Modulation of regenerative potential, Tissue origin, Hypoxia microenvironment, Inflammatory microenvironment, Clinical application

Core Tip: This review discusses recent advances in comparative analyses of dental mesenchymal stem cell (MSC) regeneration potential. We have summarized the available research evidence concerning the effects of hypoxic and inflammatory microenvironmental factors on dental MSC differentiation capacity. Existing investigations indicate the very important aspect of the recipient microenvironment niche in terms of therapeutic efficacy of transplanted dental MSCs. However, some of the data for the same cell type (especially in hypoxic in vitro conditions) are conflicting, so it is important to point out that the biology of MSCs is not yet fully known, and further research in this area is needed.



INTRODUCTION

Bone defects and dental loss connected with either disease or trauma seriously influence the quality of life of the whole population, including emotional, physical, and financial load on the society. For the medical treatment of oral diseases, periodontal treatment, dental implants, and dental protheses are the gold standards[1]. Since these oral therapies can only maintain the current state and stop further complications of the disease, failing to influence complete tissue regeneration, new technologies are needed to overcome these limitations at various tissue regeneration steps[2]. Currently, tissue engineering represents a promising future approach for recovering the function and integrity of tooth’s hard tissue[3]. The microenvironment of dental tissues, containing dental immune cells, blood vessels, extracellular matrix (ECM), numerous secreted soluble mediators, and various stromal cells, essentially influences the healing process of diseased dental tissue[4,5]. In terms of regenerative features, mesenchymal stem cells (MSCs) are the most prominent among stromal cells. Their clinical application in the treatment of dental diseases is still at the beginning since the exact mechanism of their therapeutic properties is not yet clear.

MSCs are the population of multipotent stromal cells present in many adults, perinatal and fetal tissues, where they participate in homeostasis maintenance. They were initially isolated from bone marrow and characterized as fibroblast-like cells[6]. After a while, their presence has been demonstrated in various fetal and adult tissues, such as peripheral blood, umbilical cord, placenta, adipose tissue, and others[7,8]. In the past two decades, a variety of MSC populations have also been procured from different parts of the tooth and tooth-supporting tissues (Figure 1): Dental pulp stem cells (DPSCs), stem cells from human exfoliated deciduous teeth (SHEDs), periodontal ligament stem cells (PDLSCs), dental follicle stem cells (DFSCs), gingival MSCs (GMSCs), and stem cells from the dental apical papilla (SCAPs)[9]. These cells are particularly suitable for research, given the easy availability of tissues through non-invasive dental procedures and simple methods for their isolation. According to the International Society for Cellular Therapy, minimal criteria to characterize MSC population consider the positive expression of surface markers CD73, CD90, and CD105 and negative expression of CD11b, CD19, CD79α, CD34, CD31, CD45, and human leukocyte antigen-DR isotype, along with their self-renewal and multilineage differentiation capacity into cells of osteogenic, chondrogenic, and adipogenic lineages[10] (Figure 2). Yet, the defining characteristics of MSCs are inconsistent among researchers. In addition to their regenerative role, MSCs exert immunomodulatory properties by affecting cells of the innate and adaptive immune system through direct intercellular contacts and/or secretion of soluble mediators (Figure 2). Moreover, MSCs express multiple paracrine functions, thus modulating surrounding microenvironment response to numerous autoimmune and inflammatory diseases[11]. Previous research has shown that MSCs suppress activation and proliferation of CD4+ helper and CD8+ cytotoxic T lymphocytes, B lymphocytes[12,13], dendritic cells, and natural killer cells[14], while increasing the production of regulatory T-lymphocytes[15]. In addition, it has been determined that MSCs modulate microenvironment in the injured tissue by releasing anti-inflammatory and anti-apoptotic molecules[16].

Figure 1
Figure 1 Schematic drawing illustrating sources of human dental tissue-derived mesenchymal stem cells . DFSCs: Dental follicle stem cells; DPSCs: Dental pulp stem cells; GMSCs: Gingival mesenchymal stem cells; PDLSCs: Periodontal ligament stem cells; SCAPs: Stem cells from the dental apical papilla; SHEDs: Stem cells from human exfoliated deciduous teeth.
Figure 2
Figure 2 Properties of mesenchymal stem cells . mesenchymal stem cells are the population of multipotent stromal cells present in adult and perinatal tissues where they participate in maintaining of homeostasis. Due to their self-renewal capacity, differentiation potential into specialized cells of mesodermal origin and immunomodulatory features, these undifferentiated cells can be potentially applied in regenerative medicine and cell therapy.

It is well known that MSCs’ behavior depends on the context of the microenvironment in which they reside and function. These physiologically defined compartments, named stem cell niches, are the sites of external cues integration that determine the fate of MSCs and govern them towards specific functions[17]. Therefore, a better understanding of the effects of specific microenvironmental conditions on MSCs’ fate is of great importance for getting a real insight into their biology and optimizing the conditions for their successful use.

Oxygen (O2) represents an important factor in the stem cell niche necessary for cell activity and metabolism. Unlike laboratory standard conditions that correspond to the atmospheric (21%) O2 levels, physiological oxygenation in tissues is hypoxic, ranging from 1%-14% with an average of about 5%[18]. Response to O2 level changes in mammals is mainly regulated by hypoxia-inducible factor 1 (HIF-1)-α, an ubiquitously expressed transcription factor subunit that translocates to the nucleus under hypoxic conditions where it binds to HIF-1β to regulate target genes[19]. Different O2 levels affect various MSCs features[20,21]. Moreover, the hypoxia effects depend on cell type, oxygen concentration, and experimental design; thus, many studies gave contradictory results, especially considering short-term exposure to hypoxia[20]. In dental and dental-supporting tissue derived cells, low oxygen levels (1%) increase the formation of reactive oxygen species, leading to oxidative stress, specifically in periodontal ligament (PDL) cells[22]. The reactive oxygen species level multiplies when the bacterial inflammation occurs within a hypoxic environment[23]. A unique feature of the oral cavity is the presence of plenty of microorganisms such as bacteria, fungi, protozoa, or viruses, organized in the complex communities, termed as oral microbiome. These microorganisms coexist with the host in a symbiotic way. Depending on the mouth area, the microbiome's composition is diverse, with the dental tissues (including teeth and teeth supportive tissues) being the habitat of many microorganisms (dental plaque). As the composition of the oral microbiome changes dynamically[24,25], physiological balance can be disrupted, consequently encouraging the infection development in the host[26,27].

This review provides a detailed summary of currently available data concerning dental MSC regeneration potential in terms of the tissue origin and influence of hypoxic and proinflammatory microenvironments. Furthermore, it analyzes current evidence regarding clinical applicability of dental MSCs.

DENTAL MSCS REGENERATIVE PROPERTIES WITH REGARDS TO THEIR TISSUE ORIGIN

The term dental tissue refers to the tooth and tooth-supporting tissues (periodontium). There are two major constitutive parts of the tooth, a crown (mostly visible part of the tooth) and a root. The crown is formed of three hard tissues, enamel, dentin, and cementum, and one soft tissue, dental pulp[28]. Enamel, dentin, and cementum are tissues with limited or no possibility to regenerate[29], while the dental pulp is a loose connective tissue profusely vascularized and innervated. Structurally, dental pulp is divided into three main regions that form a continuum: The peripheral odontoblastic and the sub-odontoblastic layer and the central pulp[30,31]. In terms of cellularity, the most common dental pulp cell types are fibroblasts, peripheral odontoblasts, and DPSCs, whereas collagens I and III represent extracellular pulpal matrix key protein components[32]. Moreover, the proportion of cellularity to collagen with aging favors collagen[30].

Dental pulp MSCs, with their extensive proliferation and multipotential differentiation capability, have an intrinsic role in dental pulp regeneration potential (Figure 3). They were first discovered by Gronthos et al[33] in 2000 as MSCs derived from the pulp of the permanent, impacted third molars and supernumerary teeth, commonly considered as medical waste. Like other dental MSCs, DPSCs express osteoblastic markers such as alkaline phosphatase (ALP), collagen type I (COL1A1), and osteocalcin (OCN) and are able to differentiate into osteoblast-like cells. Furthermore, DPSCs are essential for postnatal tooth homeostasis through implementation of odontoblasts in the dentin restoration process. When dental injuries or odontoblast apoptosis occur, DPSCs rapidly proliferate, migrate, and differentiate into odontoblast cells. Moreover, being of neural crest origin, DPSCs can differentiate into functionally active neurons and glial cells[34]. Importantly, recent studies have revealed that DPSCs reside in neurovascular niche where they secrete an array of angiogenic regulatory factors and generate capillary-like structures demonstrating strong angiogenic ability[35]. Overall, neurovascular and MSC-like properties make DPSCs good candidates for bone and tooth regeneration.

Figure 3
Figure 3 Functional properties and differences among dental mesenchymal stem cells . Dental mesenchymal stem cells (MSCs) are involved in dental tissues regeneration which is influenced by local microenvironment of the tissues they reside in. Overall, all dental MSCs represent good candidates for tissue regeneration, however their capacities differ (shown in table on the right). > , < and ~ represent higher, lower or similar capacity/rate respectively; The numbers in square brackets indicate the references. DFSCs: Dental follicle stem cells; DPSCs: Dental pulp stem cells; GMSCs: Gingival mesenchymal stem cells; PDLSCs: Periodontal ligament stem cells; SCAPs: Stem cells from the dental apical papilla; SHEDs: Stem cells from human exfoliated deciduous teeth.

Stem cells from the pulp tissue of human exfoliated deciduous teeth (Figure 3), firstly isolated by Miura et al[36] in 2003, are capable of forming dentin-like structures. Moreover, they show higher proliferative, odontogenic, and osteogenic differentiation potential than DPSCs[37]. Furthermore, these cells have a higher doubling time[38] than DPSCs and higher expression of collagens I and III as well as of pluripotency markers such as octamer-binding transcription factor-3/4, sex determining region Y-box-2, and Nanog homeobox[39]. SHEDs are also able to differentiate into neural and glial cells under appropriate conditions. Being able to demonstrate regenerative potential even 2 years after cryopreservation[40] and, because of their easy accessibility, SHEDs represent good candidates for bone and tooth regeneration.

However, the possibility of dental pulp tissue regeneration is restricted by several factors. Since dental pulp has collateral blood circulation due to the anatomical organization of the pulp chamber, the efficacy of the immune response to infection is limited[41]. Thus, the localization of DPSCs in perivascular and perineural pulp regions and consequent contacts with different microenvironments, along with their high immunomodulatory activity[42], makes them good candidates for modulating the immune response to infection. Furthermore, when comparing cellular density in specific tooth regions of dental pulp tissue from different teeth, higher density was observed in the coronal region of deciduous and premolar teeth compared with supernumerary and third molars. In contrast, a high cell density was observed in the apical region of supernumeraries and premolars compared with third molar teeth[43]. Overall, supernumerary dental pulp as source for DPSCs has the best morphometric parameters, and its cell density is comparable to that of deciduous tooth pulp.

Periodontium is a specialized connective tooth-supporting tissue that surrounds the root of the tooth. It has the role of attaching the tooth to the jawbone, amortizing the mechanical pressure that occurs during chewing and speech but also in the formation and resorption of bone tissue. Periodontium involves two soft tissues—PDL and gums, and two hard tissues—alveolar bone and cementum[44]. The PDL is a connective tissue with high cellularity and amount of ECM components. As for cellular constituents, PDL consists mainly of fibroblasts, osteoblasts, osteoclasts, cementoblasts, and cementocytes. Other cells present in PDL include epithelial cell rests of Malassez, macrophages, nerve cells, endothelial cells, as well as MSCs. PDLSCs represent a unique population of somatic stem cells of mesenchymal origin with the regenerative potential reflected by self-renewal and multipotent differentiation ability as well as potency for the formation of tissues that support the teeth, including the PDL and cement (Figure 3). On the other hand, the rich ECM of PDL is formed of collagen type I, II, and XII fibers, proteoglycans, and a vast vascular network[45]. Interestingly, collagen type I is the most widely used scaffold material for dental pulp regeneration[46], and PDLSCs have shown higher expression of COL1A1 compared to DPSCs[47]. Taken together, these features qualify PDLSCs for use in the regeneration/reconstruction of tooth-supporting tissue in periodontal disease.

Like PDL, the gingival connective tissue also has a rich ECM. Collagen fibers and ground substances make up 60% and 35% of gingival ECM, respectively, and only about 5% of gingival connective tissue consists of various cells like fibroblasts, mast cells, macrophages, and inflammatory cells[48]. Structurally and functionally, the gingiva is different from PDL, displaying an even higher collagen turnover rate than PDL and having a distinct composition and organization of ECM. It was noticed that compared to PDL fibroblasts, gingival fibroblasts have a significantly lower level of ALP expression, an increased potential to stimulate epithelial growth, as well as a distinctive property of regulating cytokeratin expression by epithelial cells[49]. Furthermore, the gingival tissue is highly vascularized, thanks to the high number of anastomoses[50]. Moreover, due to the activities of local microenvironment factors, including transforming growth factor-α, transforming growth factor-β, insulin-like growth factor, nerve growth factor, epidermal growth factor, and fibroblast growth factor, gingiva has a particularly high wound-healing capacity. Other local factors, such as mechanical signals from the ECM to the cells, may be involved as well[49]. All of these structural characteristics provide gingival tissue with a great therapeutic potential in regenerative therapy especially in terms of GMSCs. In contrast to MSCs from other sources, GMSCs isolated from the gingival lamina propria are profuse and easily procured cells through minimally invasive cell isolation techniques, which make them suitable cells for regenerative purposes[51]. Compared to PDLSCs, GMSCs have a higher rate of proliferation while also exerting a higher capacity to proliferate, migrate, and form angiogenic tubules in comparison to DPSCs (Figure 3)[52].

The alveolar bone is located on the jaw bones that hold the teeth, and it arises in the process of immature tooth root development from the dental follicle (DF), one of the multipotent tissues[53]. The DF is made up of MSCs and fibers surrounding the enamel organ and dental papilla of a developing tooth[54]. It is a vascular fibrous sac of ectomesenchymal origin. Histologically, DF is characterized by fibrous connective tissue with variable amounts of lining epithelium, including enamel, cuboidal, squamous, and, rarely, respiratory epithelium. The type of lining epithelium seems to be related to the patient’s age[55]. DFSCs, originating from this developing tissue, possess higher plasticity than other dental stem cells[56]. Recently, isolated DFSCs were also found to have the ability to form salivary gland cells and ductal cells[56] (Figure 3).

Another immature dental tissue is the apical papilla, a tissue only present during root development before the tooth erupts into the oral cavity[57]. In comparison to the dental pulp, the apical papilla has less cellular and vascular components[57], and the tissue is more immature since it contains a higher number of MSCs than mature dental pulp tissue. Moreover, the apical papilla performs a key role in the differentiation of odontoblasts into cells capable of secreting the primary dentin matrix[58] . Furthermore, the localization of the apical papilla in the apical root of the tooth can benefit by its apical collateral circulation and thus survive during the pulp necrosis process, which additionally explains why immature teeth with necrotic pulps can undergo completion of root development[59]. SCAPs reside in the apical papilla of permanent immature teeth and appear to be the source of odontoblasts that are responsible for the formation of root dentin (Figure 3). Conservation of SCAPs when treating immature teeth may allow the continuous formation of the root to completion.

Overall, among adult tissues easily available through non-invasive dental procedures, SHEDs have a higher doubling time than DPSCs[39]. However, PDLSCs showed a significantly higher in vitro osteogenic differentiation potential than both SHEDs and DPSCs (Figure 3), as evidenced by functional studies and gene expression indicating the complex influence of stem cell origin on their regenerative potential[60].

MODULATION OF DENTAL MSCS REGENERATIVE PROPERTIES IN HYPOXIC MICROENVIRONMENT

In the oral cavity, O2 levels are shown to range from 0.7% and 3.8% in human periodontal pockets[61] and approximately 3% O2 in the pulp tissue of rats[62]; therefore, given the therapeutic potential of dental MSCs, a number of studies have focused on examining their functions in such microenvironments. According to the published data, dental MSCs have been intensively investigated with regard to different O2 levels, and detailed outcomes are shown in Table 1.

Table 1 Effects of different oxygen levels on regenerative potential of human dental mesenchymal stem cells.
Hypoxia level
Outcome
Ref.
DPSCs
< 1% O2Weak ALP activity, weak calcium depositionJanjić et al[75], 2019
2% O2Proliferation↑; odontogenic differentiation↑; angiogenesis↑; in vivo: Angiogenesis inside the pulp chamber↑, the formation of odontoblast-like cells lining along the dentin–pulp interface↑Kuang et al[76], 2016
No change in proliferation; calcium deposits↑; proteoglycan deposition↑; lipid droplets↑; PPAR𝛾2 mRNA↑Zhou et al[71], 2014
Proliferation↓; Runx2 mRNA expression↑ (both 2D and 3D conditions); Runx2 and Col1A mRNA expression, osteopontin in 2D culture↑, calcium deposition in 3D culture↑, calcium deposition in 2D culture↓Labedz-Maslowska et al[73], 2020
3% O2BMP2, OCN and RUNX2 protein expression↑; calcium deposits↑; RunX2 and Sp-7 mRNA expression↑; in vivo: In a mouse apical periodontitis bone destruction model, hDPSC recruitment and recovery of alveolar bone mass in infected periapical tissue↑, osteogenesis and bone mineralization↑Wu et al[77], 2016
5% O2Proliferation↑; mineralization↑Kwon et al[63], 2017
PDLSCs
1.5%-2% O2ALP activity↓; SPARC protein expression↓; ALP, OCN, and BMP-2 mRNA expression↓; proliferation↓Hou et al[79], 2009
2% O2Calcium deposition, proteoglycan deposition↑; lipid droplets↑; Runx2, Sox9 mRNA expression↑Zhou et al[71], 2014
ALP activity↑; Runx2 and Sp7 mRNA and protein expression↑; mineralization↑Wu et al[80], 2013
SPP1, RUNX2, SP7 mRNAs and protein expression↑Li et al[65], 2014
Proliferation↑; RUNX-2 and ALP protein expression↑; no effect on adipogenic differentiation; in vivo: Stronger bone regeneration region in male nude mice, more mineralized tissue in a periodontal defect modelYu et al[64], 2016
Proliferation↑; Runx2, osteopontin and osteocalcin mRNA expression↑; in vivo: After 12 wk of transplantation, hypoxia-treated cells differentiated into osteoblast-like cells that formed bone-like structuresZhang et al[66], 2014
ALP activity↑; Runx2 mRNA expression↑Chen et al[81], 2017
3% O2Proliferation↑; osteogenic and adipogenic differentiation↓; chondrogenic differentiation↑; preconditioning: Osteogenic and adipogenic differentiation↑Murabayashi et al[72], 2017
Runx2, Alp, Col1, and Ocn mRNA expression↑; RUNX2 protein ex vivo and in situXu et al[82], 2019
5% O2ALP activity↓Matsuda et al[78], 1998
5% O2; 1% O2Proliferation↓; ALP activity↑; Opn, Alp mRNA expression↑; Cemp1, Cap mRNA expression↑Xiao et al[74], 2017
8% O2Proliferation↑; Cemp1, Ocn mRNA expression↑; CEMP protein expression↑; mineral deposition↑; ALP activity↑; in vivo: CEMP1 protein expression in mouse PDL spaces↑Choi et al[67], 2014
SHEDs
1% O2Proliferation↑; mineralization↑; ALP activity↑; OPN and DMP1 protein expression↑; in vivo: After implantation in immunodeficient mice, the tissue-engineered constructs seeded with hypoxia primed SHED mediated faster intramembranous bone formation into critical size calvarial defectsNovais et al[69], 2019
2.3% O2Proliferation↑; no effect on adipogenic and osteogenic differentiationKanafi et al[68], 2013
SCAPs
1% O2No effect on proliferation; RunX2, Alp, TGF-1↑; neuronal differentiation (CNP, NSE, and SNAIL mRNA expression↑; angiogenesis (VEGF A and B)↑; adipocyte lipid binding protein (ALBP)↓Vanacker et al[70], 2014
3% O2ALP activity↑; mineralization↑; Dspp, Dmp1 and Bsp mRNA expression↑Yang et al[83], 2020
Chemical hypoxia
DPSCs
100 μM CoCl2No effect on proliferation; SOX9 and VCAN; no expression Col2a1, Acan↑, Col 10 mRNA expression↓; proteoglycans↓Teti et al[85], 2018
100 mM CoCl2Short term: RUNX2, ALP, OCN, COL1A1 mRNA and protein expression↑; long term: RUNX2, ALP, OCN, COL1A1 mRNA and protein expression↓Zheng et al[89], 2021
ALP activity↓; Alp, Ocn, and Runx2 mRNA expression↓; mineralization↓Osathanon et al[86], 2014
200 μM CoCl2ALP activity↓; Runx2, Alp, Ocn and Col-1 mRNA and protein expression↓; mineralization↓Song et al[87], 2017
PDLCs
200 μM; 400 μM CoCl2Proliferation↓; ALP, RUNX2, collagen I↓Dong et al[88], 2014
1 mM DMOGNo effect on proliferation; COL1, RUNX2 and CEMP1 protein expression↑Li et al[92], 2016
0, 5, 10, 20 μM deferoxamineProliferation↓; Runx2, Opn and Col1 mRNA expression↑; calcium deposition↑Mu et al[91], 2017
SHED
50 or 100 μM CoCl2ALP activity↓; calcium deposition↓; Alp, Runx2, and ColI mRNA expression↓Chen et al[90], 2019

Although most studies showed that hypoxia stimulated proliferation of DPSCs[63], PDLSCs[64-67], SHEDs[68,69], and SCAPs[70], some reported it had no[71,72] or a negative effect[73,74]. Also, results concerning the influence of hypoxia on differentiation capacity of MSCs derived from dental and dental supportive tissues are diverse. It was shown that under hypoxia, osteogenic differentiation of DPSCs was strongly suppressed compared to normoxia[75]; however, there are studies showing increased osteogenic[63,71], adipogenic and chondrogenic differentiation[71] at low O2 levels. In order to mimic the native microenvironment of DPSCs, Labedz-Maslowska et al[73] grew hydrogel-encapsulated cells in the presence of 2% O2. They showed up-regulation of osteogenic differentiation in hypoxic conditions, but calcium deposition was, in the case of two-dimensional culture, more prominent under normoxia. Another study seeded DPSCs into nanofibrous spongy microspheres and showed that their priming at 2% O2 prior to implantation significantly promoted the formation of odontoblast-like cells lining along the dentin–pulp interface of mice[76]. On the other hand, Wu et al[77] showed that preconditioning of cells at 3% O2 enhanced DPSC osteogenic differentiation in vitro and more importantly upregulated their recruitment in mouse apical periodontitis bone destruction model and enhanced osteogenesis and bone mineralization.

While there is evidence that hypoxia decreased osteogenic differentiation of PDLSCs[65,78,79], it was demonstrated that it can also increase it[66,71,74,80,81] as well as adipogenic and chondrogenic differentiation[71]. Interesting observations in some of these studies were that HIF1α promoted osteogenic differentiation of PDLSCs, while HIF1A antisense long noncoding RNA 2 had a negative effect on it[81] and that the stimulative effect of 2% O2 on osteogenic differentiation of PDLSCs was mediated by extracellular signal-regulated kinase and even more rapidly and vigorously by p38 mitogen-activated protein kinase[80]. The 24 h long pretreatment of PDLSCs under 2% O2 increased osteogenesis, whereas cotreatment with tumor necrosis factor (TNF)-α and interleukin (IL)-β significantly reduced this effect, and no significant effects on adipogenic differentiation were observed[64]. Importantly, the transplants containing hypoxic pretreated-PDLSCs led to significantly stronger bone regeneration when subcutaneously placed into the dorsal region of male nude mice as well as more mineralized tissue in a periodontal defect model[64]. Similarly, PDLSCs grown at 2% O2 differentiated into osteoblast-like cells that formed bone-like structures after transplantation into the backs of mice[66]. Cultivation of these cells at 3% O2 inhibited their differentiation capacity to osteogenic and adipogenic lineages, whereas it enhanced chondrogenic differentiation[72]. However, the authors indicated that subsequent switch of 2 wk hypoxic preconditioned PDLSCs to normoxia allowed successful differentiation into osteogenic and adipogenic lineages. In contrast, Xu et al[82] found that 3% O2 increased osteogenic markers expression in vitro and augmented runt-related transcription factor (RUNX) 2 protein expression ex vivo and in situ via HIF-1α-induced vascular endothelial growth factor, suggesting a positive role for HIF-1α in the early stage of osteogenesis of PDLSCs. Reduced O2 tension besides osteogenic, increased cementogenic differentiation capability of PDLSCs[67,74], probably via the Wnt/b-catenin signaling pathway[74]. It was found that HIF-1 activity is required to stimulate the differentiation response in vitro and, more importantly, cementum protein 1 expression in mouse PDL spaces in vivo[67].

Although no difference in adipogenic and osteogenic differentiation potential of SHEDs was detected at 2.3% O2 in comparison to normoxia[68], significantly higher osteogenesis was documented at 1% O2 compared to control in SHEDs incorporated into plastically compressed collagen hydrogels[69]. Moreover, after implantation in immunodeficient mice, these hypoxia-primed SHED constructs mediated faster intramembranous bone formation into critical size calvaria defects. Hypoxia significantly increased the osteogenic[70,83], neural, and angiogenic marker expression in SCAPs and suppressed their adipogenic differentiation[70].

When establishing hypoxic conditions in the laboratory, researchers sometimes encounter lots of technical difficulties (e.g., media changes), especially in long-term cultures, and therefore the use of chemical mimetic to induce hypoxic response has been an attractive alternative. Cobalt chloride (CoCl2) stabilizes HIF-1α and HIF-2α under normoxic conditions, and it is the most investigated hypoxia-mimetic agent[84]. Supplementation of DPSCs with CoCl2 had no effect on cell proliferation and reduced their chondrogenic[85] and osteogenic differentiation[86-88]. Interestingly, it was observed that enamel matrix proteins[87] or apigenin, an HIF-1a inhibitor[86], could reverse the effect of CoCl2 on osteogenic differentiation. Zheng et al[89] demonstrated that osteogenic differentiation of PDLSCs was activated by short-term exposure to CoCl2 but was inhibited following prolonged exposure, which might be mediated by circular RNA circCDK8. CoCl2 had the same osteo-reducing effect on SHEDs given the significantly decreased ALP activity, calcium deposition, and osteogenic marker messenger RNA expression[90]. It was confirmed that deferoxamine[91] and dimethyloxalylglycine[92] promoted HIF1α expression in PDLSCs, and it was demonstrated that while proliferation was inhibited by deferoxamine, osteogenic differentiation was significantly promoted by both agents. The Wnt signaling pathways might be involved in dimethyloxalylglycine-induced differentiation of cells[92].

These findings provide important insights into capacity of dental MSCs to adapt to physiological conditions of low oxygenation in vitro by changing their regenerative properties.

MODULATION OF MSC REGENERATIVE PROPERTIES IN THE INFLAMMATORY MICROENVIRONMENT

In the human population, dental chronic inflammatory diseases are very common. For example, caries that cause progressive destruction of dental hard tissue[93] are one of the most frequent conditions in the oral cavity[94], particularly in childhood[95], while periodontal diseases—oral infections of tooth supportive tissue (gingivitis and periodontitis)—affect 20%-50% of the world population[96]. If left untreated, these conditions can cause significant damage to the oral cavity and consequently cause major problems in the processes of chewing, swallowing, digestion, and speech and create aesthetic problems[93]. Therefore, to consider the possibilities for regeneration and recovery of damaged oral tissue, it is necessary to perceive the regenerative potential of dental MSCs in the context of the inflammatory microenvironment.

Viewed from the perspective of physiological or pathological conditions, endogenous, resident MSCs respond to factors present in the immediate vicinity[57]. Thus, inflammation caused by caries or periodontal disease may significantly affect regenerative capacity of dental MSCs, including their proliferation, migration, colony forming capacity, and differentiation. However, data related to the properties of MSCs isolated form inflamed dental tissues are not consistent (Table 2). Namely, recent results of Inostroza et al[97] showed no significant differences between immunophenotype, trilineage differentiation, colony-formation, and proliferation of the DPSCs derived from healthy and inflamed pulp, although immunomodulatory functions of DPSCs from inflamed pulp were altered—showing decreased capacity to suppress CD3 T cell proliferation. Absence of CD34 and CD45 markers, along with high expression of MSC-associated markers between DPSCs extracted from normal and diseased pulp, was also observed in Park et al[98]. Besides, in comparison to control cells, DPSCs from inflamed tissue manifested higher osteogenesis (stronger mineralization and expression of osteogenic markers OCN and RUNX2) but lower neurogenesis (decreased expression of neurogenic markers microtubule-associated protein 2, neuronal nuclear protein, and glutamate decarboxylase 6), along with higher level of IL-6 expression. This study also shows the significance of IL-6, as a strong inflammatory factor, to modulate DPSCs function. Namely, the IL-6 treatment of DPSCs derived from healthy tissue stimulated their osteogenic differentiation and reduced neurogenic differentiation, while IL-6 blocking in DPSCs of inflamed tissue annulled their osteogenic/neurogenic capacity[98].

Table 2 Effects of inflammatory microenvironment on regenerative potential of human dental mesenchymal stem cells.
Inflammation level
Outcome
Ref.
DPSCs
Irreversible pulpitisNo change in proliferation; no change in differentiation; not being able to supress CD3 proliferation; IDO activity↓Inostroza et al[97], 2020
Decayed and pain affected tissueNo change in immunophenotype; proliferation↓; Ca deposition↑, OCN and RunX2 protein expression↑; MAP2 and NeuN protein expression↓; IL-6 expression and IL-6- induced osteogenesis↑Park et al[98], 2019
Irreversible pulpitisPopulation doubling time↓; STRO‑1, CD90, CD105 and CD146 levels↑; Ca deposition↓; OCN and RUNX2 mRNA expression↓; LPL and PPARγ2 mRNA expression↓; in vivo: Retaining their stem cell potencyAlongi et al[99], 2010
Severe periodontal diseaseProliferation↑; Stro-1+, CD146, SSEA-4 levels↑; Ca deposition↑, OCN, RUNX2 and mRNA expression↑ (also upon stimulation with proinflammatory cytokines IL-1β and TNF-α)Tomasello et al[101], 2017
GMSCs
Severe periodontal diseaseProliferation↑; Stro-1+, CD146, SSEA-4 levels↑; Ca deposition↑, OCN, RUNX2 and mRNA expression↑ (also upon stimulation with proinflammatory cytokines IL-1β and TNF-α)Tomasello et al[101], 2017
Dental plaque-induced gingival hyperplasiaProliferation↑; Runx2 and OCN mRNA expression↓; ALP activity↓; PPARγ mRNA expression↓Li et al[102], 2013
PDLSCs
Chronic peridontitisNo change in Runx2, OCN and ALP mRNA expression; no change in PPARγ and aP2 mRNA expression; no change in proliferation; in vivo: Smaller newly formed cementumPark et al[105], 2011
Chronic periodontitisProliferation↑; migration↑; Ca deposition↓, Runx2, ALP and OCN mRNA expression↓; no change in lipid droplets level and PPARγ mRNA expressionTang et al[106], 2016
Periodontitis with alveolar bone lossProliferation↑; RUNX2 mRNA expression↓; PPARγ mRNA expression↓Liu et al[107], 2011
Chronic periodontitisviability and proliferation↓Soheilifar et al[108], 2016

Contrary to these studies, compared to the control cells, DPSCs isolated from inflamed pulp expressed higher levels of MSC markers, while their capacity to proliferate was increased with reduced osteo/dentinogenic differentiation potential in vitro. Still, as well as DPSCs from healthy pulp, DPSCs of inflamed tissue retained their capacity to form pulp/dentin complexes after transplantation into immunocompromised mice, i.e. their regenerative capacity was preserved[99]. In the context of optimal dental MSC isolation, Tsai et al[100] demonstrated that dental diseases represent significant factors that affect MSC isolation and quality, given that successful MSCs yield was less pronounced from carious deciduous tooth or tooth with pulpitis. As for dental MSCs isolated from inflamed tooth supportive tissue, Tomasello et al[101] showed that DPSCs and GMSCs derived from the tissue affected by periodontitis proliferate faster and possess more pronounced mineralization. Also, no changes in negative MSC marker expression or in high expression of a positive MSC marker (CD29) were detected, while expression of other positive MSCs markers, such as Stro-1, CD146, and stage-specific embryonic antigen-4, was higher in DPSCs and GMSCs of inflamed tissues. In another study, increased proliferation, but with reduced osteogenic and adipogenic differentiation, of GMSCs under inflammatory conditions was observed[102]. Also, higher levels of matrix metalloproteinases (MMP)-1, MMP-2, IL-1, IL-6, TNF-α and type COL1A1 were detected in GMSCs collected from inflamed gingival tissues, indicating the potential of inflammatory environment to shift differentiation capacity toward pro-fibrotic phenotype, thus representing possible mechanism of gingival hyperplasia development during inflammation[102].

A recently published article provided a comprehensive review of papers related to the characteristics of GMSCs from healthy and inflamed tissue, altogether suggesting that GMSCs of inflamed tissues could be a reliable source of MSCs when compared to the healthy gingiva[103]. Inflammatory environment can also alter properties of PDLSCs[104], however findings are not clear since other publications suggest that there are no significant differences between PDLSCs of healthy and diseased periodontal tissue[105]. Enhanced proliferation but reduced differentiation of dental MSCs derived from inflamed tissue has been detected for PDLSCs isolated from diseased tooth supportive tissue[106,107], while the study of Tang et al[106] also demonstrated increased migration of PDLSCs originating from inflamed periodontal tissue. Higher migratory potential with preserved osteogenic/cementogenic and adipogenic differentiation was reported in Park et al[105], and in an in vivo transplantation model, they showed that PDLSCs from inflamed periodontal tissue possess a preserved ability to form new cementum-like tissue and related periodontal fibers. However, similar to the findings of Tsai et al[100], related to the yield of MSCs derived from diseased tooth, Soheilifar et al[108] reported that viability and proliferation rate of PDLSCs isolated from the periodontitis-affected teeth were significantly lower in comparison to the control PDLSCs.

Having in mind that the regenerative ability of transplanted dental MSCs strongly depend on the donor/recipient microenvironment niche, gaining successful therapeutic effect in diseased microenvironment is the biggest challenge[56]. Further understanding the influence of the diseased tissue microenvironment on MSC regenerative potential would help the establishment of healing procedures.

CLINICAL RELEVANCE OF DENTAL MSCS

Numerous in vivo studies on animals and human clinical trials with various types of dental stem cells show a way to encourage novel tissue engineering strategies for therapies of dental diseases[109]. However, a long road is still ahead, as clinical trials are in their early phases. Despite abundance of preclinical studies, only a few clinical trials have been completed and published. One of the main reasons that stalls successful clinical application of dental stem cells resides in the fact that the exact mechanism of their therapeutic properties is not yet clear. The other challenges for clinical use of dental stem cells are strict regulations, high costs of cell processing, and lack of uniformity in approaches for isolation, expansion, and application of these cells. The therapeutic efficacy of transplanted dental MSCs is also compromised by the diseased microenvironment of the recipients[56].

DPSCs, as the first discovered dental tissue-derived stem cells, are also the most studied stem cells for dental tissue reparation[110]. Even though there are many studies in large animal models showing promising results in dental tissue regeneration, especially in dentin and dental pulp regeneration[110], results of clinical studies are less conclusive. The first study to achieve successful reparation of alveolar bone defect in humans was done by d’Aquino et al[111]. They used autologous DPSCs from third maxillar molars seeded onto a collagen sponge scaffold to fill the space left after the extraction of an impacted mandibular molar from the same patient (7 patients in total). A contralateral extraction site filled with sponge without cells served as a control (a split-mouth study design). Clinical and radiographic assessment after 3 mo and 1 year revealed optimal vertical repair and complete restoration of periodontal tissue[111]. In a pilot clinical study, Nakashima et al[112] demonstrated a safe and efficacious method for complete pulp regeneration in 5 patients with irreversible pulpitis using autologous DPSCs transplanted with granulocyte colony-stimulating factor in atelocollagen into pulpectomized teeth. However, a split-mouth randomized clinical trial on 32 patients, aiming to assess the efficacy of autologous DPSCs delivered in a collagen matrix for post-extraction socket healing, failed to show significant reduction in the socket bone resorption in the treated group compared to control[113]. It is noteworthy that, unlike in the other cited studies, in this trial a proprietary medical device and protocol (Rigenera®) for direct isolation of DPSCs from dental pulp was used, without prior expansion in culture.

DPSCs from deciduous teeth are also frequently used in dental tissue regeneration studies. In a randomized control clinical trial, Xuan et al[114] isolated autologous SHEDs from deciduous canine tooth pulp and implanted them in the form of aggregates, into an injured incisor of the patient with pulp necrosis secondary to trauma. Control patients were treated with standard apexification (a procedure that induces tooth root development and closure of the root apex through hard tissue deposition). After 12 mo follow-up, SHED implantation treatment led to regeneration of functional dental pulp with blood vessels and sensory nerves. In another study, Tanikawa et al[115] used SHEDs associated with HA-collagen sponge for closing alveolar defects in patients with unilateral alveolar cleft defects and demonstrated that this therapy leads to bone regeneration with dental eruption and reduced morbidity compared to traditional iliac crest bone grafting and rhBMP-2.

Among other dental stem cells, alveolar bone derived MSCs also represent great promise in regenerative therapy. A pilot clinical trial evaluated treatment of maxillary radicular cysts in 9 patients using autologous alveolar bone-derived MSCs seeded onto a glutaraldehyde-cross-linked autologous serum scaffold and subjected to osteogenic differentiation[116]. It demonstrated a significant promotion of bone growth in all MSC-treated cysts.

In addition to osteal defects and pulp necrosis, periodontal diseases make up a large proportion of dental ailments targeted for regenerative therapy. A retrospective pilot study examined feasibility and safety of reconstructing the periodontal intrabony defects in 16 teeth of 3 patients with implantation of autologous PDLSCs mixed with a HA-based bone-grafting material[117]. After 32-72 mo follow-up period, clinical examination indicated improvement of probing depth, clinical attachment level, and gingival recession. In a randomized controlled study, Ferrarotti et al[118] used a biocomplex of autologous dental pulp micrografts (Rigenera®) with collagen sponge to fill intrabony defects of 29 chronic periodontal patients randomly allocated to test (n = 15) and control (n = 14) groups[118]. This treatment significantly improved clinical parameters of periodontal regeneration 1 year after the procedure. A novel approach using cell sheets of cultured autologous PDLSCs to treat periodontitis was assessed in a case series study involving 10 patients with chronic periodontitis[119]. Triple layered PDL-derived cell sheets with PGA mesh were transplanted on the root surface, and β-tricalcium phosphate (β-TCP) granules were used to fill in bony defects. In all 10 cases, clinical as well as radiographic endpoint parameters improved 6 mo after the treatment. On the other hand, a randomized clinical trial on 30 patients, with 41 teeth in total, using autologous PDLSCs in combination with grafting materials to treat periodontal intraosseus defects, revealed radiologically and clinically greater, but not statistically significant, regeneration of alveolar bone in cell-treated group compared to control group[120].

A recent pilot trial evaluated the safety and efficacy of autologous PDLSC transplanted with a commercial xenogeneic (porcine) bone substitute as a matrix for the regeneration of intrabony defects of 19 patients with chronic periodontitis[121]. The study confirmed the safety of the treatment, but the results have not demonstrated a significant additional clinical benefit compared to control, after 12 mo follow-up. In a randomized clinical trial designed to investigate the use of gingival fibroblasts and GMSCs in the treatment of intrabony periodontal defects, a total of 20 patients with periodontitis were evenly assigned into two groups[122]. Experimental group received cultured autologous gingival fibroblasts/GMSCs on β-TCP scaffold, covered by a collagen membrane, and the control group received β-TCP without the cells. After 6 mo, the study showed significant improvement of clinical and radiological parameters in comparison to the control group.

CONCLUSION

Dental tissues represent valuable sources of MSCs for possible use in regenerative therapy, especially of diseases associated with bone defects and dental loss. Given the different structures and compositions of the tissues they reside in, various dental MSCs exhibit diverse biological and functional features. Therefore, in this review, we have summarized the available research evidence concerning the effects of hypoxic and inflammatory microenvironmental factors on dental MSCs differentiation capacity. We can conclude that the existing investigations indicate very important aspects of the recipient microenvironment niche in terms of therapeutic efficacy of transplanted dental MSCs. Moreover, the therapeutic potential of dental MSCs in in vitro conditions mimicking native hypoxic and inflammatory microenvironment can lead to significant development of cell-based therapies.

However, some of the data for the same cell type (especially in hypoxic in vitro conditions) are conflicting, which is a trend noticed for other MSCs as well[20]. Therefore, it is important to point out that the biology of MSCs is not yet fully known. As MSC populations exhibit functional heterogeneity and a hierarchy in the terms of proliferative and differentiation potential and metabolic properties of the cells composing the tissues[123], they can behave differently even within the same tissue of healthy individuals. Moreover, the cultivation conditions affect their characteristics, and not all studies covered by this review used the same cell isolation methods and/or the same cell passages. It should be also taken into account that inflammation doesn’t solely contribute to these contradictory findings, given the lack of data on specificity of the inflamed tissues in the terms of proinflammatory factors involved. In addition, the reason for inconsistencies in results related to the same O2 levels can be found in different experimental settings such as the duration of the hypoxic treatment. Taken together, we suggest further research in this area, but with synchronized cell isolation methods, cultivation conditions, and experimental designs among research groups.

Footnotes

Provenance and peer review: Invited article; Externally peer reviewed.

Peer-review model: Single blind

Specialty type: Cell and tissue engineering

Country/Territory of origin: Serbia

Peer-review report’s scientific quality classification

Grade A (Excellent): 0

Grade B (Very good): B, B

Grade C (Good): C

Grade D (Fair): 0

Grade E (Poor): 0

P-Reviewer: Huang (Academic Editor) YC, Naserian S, Rojas A S-Editor: Chang KL L-Editor: A P-Editor: Chang KL

References
1.  Petersen PE, Bourgeois D, Ogawa H, Estupinan-Day S, Ndiaye C. The global burden of oral diseases and risks to oral health. Bull World Health Organ. 2005;83:661-669.  [PubMed]  [DOI]  [Cited in This Article: ]
2.  Mao AS, Mooney DJ. Regenerative medicine: Current therapies and future directions. Proc Natl Acad Sci U S A. 2015;112:14452-14459.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 607]  [Cited by in F6Publishing: 458]  [Article Influence: 50.9]  [Reference Citation Analysis (2)]
3.  Han F, Wang J, Ding L, Hu Y, Li W, Yuan Z, Guo Q, Zhu C, Yu L, Wang H, Zhao Z, Jia L, Li J, Yu Y, Zhang W, Chu G, Chen S, Li B. Tissue Engineering and Regenerative Medicine: Achievements, Future, and Sustainability in Asia. Front Bioeng Biotechnol. 2020;8:83.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 108]  [Cited by in F6Publishing: 90]  [Article Influence: 22.5]  [Reference Citation Analysis (0)]
4.  Cekici A, Kantarci A, Hasturk H, Van Dyke TE. Inflammatory and immune pathways in the pathogenesis of periodontal disease. Periodontol 2000. 2014;64:57-80.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 556]  [Cited by in F6Publishing: 698]  [Article Influence: 77.6]  [Reference Citation Analysis (0)]
5.  Shah D, Lynd T, Ho D, Chen J, Vines J, Jung HD, Kim JH, Zhang P, Wu H, Jun HW, Cheon K. Pulp-Dentin Tissue Healing Response: A Discussion of Current Biomedical Approaches. J Clin Med 20209.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 29]  [Cited by in F6Publishing: 34]  [Article Influence: 8.5]  [Reference Citation Analysis (0)]
6.  Pittenger MF, Mackay AM, Beck SC, Jaiswal RK, Douglas R, Mosca JD, Moorman MA, Simonetti DW, Craig S, Marshak DR. Multilineage potential of adult human mesenchymal stem cells. Science. 1999;284:143-147.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 15372]  [Cited by in F6Publishing: 14773]  [Article Influence: 590.9]  [Reference Citation Analysis (0)]
7.  Hass R, Kasper C, Böhm S, Jacobs R. Different populations and sources of human mesenchymal stem cells (MSC): A comparison of adult and neonatal tissue-derived MSC. Cell Commun Signal. 2011;9:12.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1018]  [Cited by in F6Publishing: 1115]  [Article Influence: 85.8]  [Reference Citation Analysis (0)]
8.  Marquez-Curtis LA, Janowska-Wieczorek A, McGann LE, Elliott JA. Mesenchymal stromal cells derived from various tissues: Biological, clinical and cryopreservation aspects. Cryobiology. 2015;71:181-197.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 219]  [Cited by in F6Publishing: 222]  [Article Influence: 24.7]  [Reference Citation Analysis (0)]
9.  Huang GT, Gronthos S, Shi S. Mesenchymal stem cells derived from dental tissues vs. those from other sources: their biology and role in regenerative medicine. J Dent Res. 2009;88:792-806.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1184]  [Cited by in F6Publishing: 1215]  [Article Influence: 81.0]  [Reference Citation Analysis (0)]
10.  Dominici M, Le Blanc K, Mueller I, Slaper-Cortenbach I, Marini F, Krause D, Deans R, Keating A, Prockop Dj, Horwitz E. Minimal criteria for defining multipotent mesenchymal stromal cells. The International Society for Cellular Therapy position statement. Cytotherapy. 2006;8:315-317.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 11055]  [Cited by in F6Publishing: 11794]  [Article Influence: 693.8]  [Reference Citation Analysis (1)]
11.  Weiss ARR, Dahlke MH. Immunomodulation by Mesenchymal Stem Cells (MSCs): Mechanisms of Action of Living, Apoptotic, and Dead MSCs. Front Immunol. 2019;10:1191.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 287]  [Cited by in F6Publishing: 401]  [Article Influence: 80.2]  [Reference Citation Analysis (0)]
12.  Jones S, Horwood N, Cope A, Dazzi F. The antiproliferative effect of mesenchymal stem cells is a fundamental property shared by all stromal cells. J Immunol. 2007;179:2824-2831.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 183]  [Cited by in F6Publishing: 196]  [Article Influence: 11.5]  [Reference Citation Analysis (0)]
13.  Beldi G, Khosravi M, Abdelgawad ME, Salomon BL, Uzan G, Haouas H, Naserian S. TNFα/TNFR2 signaling pathway: an active immune checkpoint for mesenchymal stem cell immunoregulatory function. Stem Cell Res Ther. 2020;11:281.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 35]  [Cited by in F6Publishing: 42]  [Article Influence: 10.5]  [Reference Citation Analysis (0)]
14.  Klopp AH, Gupta A, Spaeth E, Andreeff M, Marini F 3rd. Concise review: Dissecting a discrepancy in the literature: do mesenchymal stem cells support or suppress tumor growth? Stem Cells. 2011;29:11-19.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 391]  [Cited by in F6Publishing: 408]  [Article Influence: 31.4]  [Reference Citation Analysis (0)]
15.  Bergfeld SA, DeClerck YA. Bone marrow-derived mesenchymal stem cells and the tumor microenvironment. Cancer Metastasis Rev. 2010;29:249-261.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 240]  [Cited by in F6Publishing: 256]  [Article Influence: 18.3]  [Reference Citation Analysis (0)]
16.  Le Blanc K, Ringdén O. Immunomodulation by mesenchymal stem cells and clinical experience. J Intern Med. 2007;262:509-525.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 530]  [Cited by in F6Publishing: 487]  [Article Influence: 28.6]  [Reference Citation Analysis (0)]
17.  Vishwakarma A, Rouwkema J, Jones PA, Karp JM.   The Need to Study, Mimic, and Target Stem Cell Niches. In: Vishwakarma A. Biology and Engineering of Stem Cell Niches. Amsterdam: Elsevier Inc, 2017: 3–13.  [PubMed]  [DOI]  [Cited in This Article: ]
18.  Carreau A, El Hafny-Rahbi B, Matejuk A, Grillon C, Kieda C. Why is the partial oxygen pressure of human tissues a crucial parameter? J Cell Mol Med. 2011;15:1239-1253.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 746]  [Cited by in F6Publishing: 809]  [Article Influence: 62.2]  [Reference Citation Analysis (0)]
19.  Chee NT, Lohse I, Brothers SP. mRNA-to-protein translation in hypoxia. Mol Cancer. 2019;18:49.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 58]  [Cited by in F6Publishing: 54]  [Article Influence: 10.8]  [Reference Citation Analysis (0)]
20.  Buravkova LB, Andreeva ER, Gogvadze V, Zhivotovsky B. Mesenchymal stem cells and hypoxia: where are we? Mitochondrion. 2014;19 Pt A:105-112.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 86]  [Cited by in F6Publishing: 91]  [Article Influence: 9.1]  [Reference Citation Analysis (0)]
21.  Ejtehadifar M, Shamsasenjan K, Movassaghpour A, Akbarzadehlaleh P, Dehdilani N, Abbasi P, Molaeipour Z, Saleh M. The Effect of Hypoxia on Mesenchymal Stem Cell Biology. Adv Pharm Bull. 2015;5:141-149.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 113]  [Cited by in F6Publishing: 132]  [Article Influence: 14.7]  [Reference Citation Analysis (0)]
22.  Gölz L, Memmert S, Rath-Deschner B, Jäger A, Appel T, Baumgarten G, Götz W, Frede S. LPS from P. gingivalis and hypoxia increases oxidative stress in periodontal ligament fibroblasts and contributes to periodontitis. Mediators Inflamm. 2014;2014:986264.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 80]  [Cited by in F6Publishing: 93]  [Article Influence: 9.3]  [Reference Citation Analysis (0)]
23.  Gölz L, Memmert S, Rath-Deschner B, Jäger A, Appel T, Baumgarten G, Götz W, Frede S. Hypoxia and P. gingivalis synergistically induce HIF-1 and NF-κB activation in PDL cells and periodontal diseases. Mediators Inflamm. 2015;2015:438085.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 65]  [Cited by in F6Publishing: 76]  [Article Influence: 8.4]  [Reference Citation Analysis (0)]
24.  Sultan AS, Kong EF, Rizk AM, Jabra-Rizk MA. The oral microbiome: A Lesson in coexistence. PLoS Pathog. 2018;14:e1006719.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 58]  [Cited by in F6Publishing: 66]  [Article Influence: 11.0]  [Reference Citation Analysis (0)]
25.  Deo PN, Deshmukh R. Oral microbiome: Unveiling the fundamentals. J Oral Maxillofac Pathol. 2019;23:122-128.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 109]  [Cited by in F6Publishing: 258]  [Article Influence: 51.6]  [Reference Citation Analysis (0)]
26.  Lu M, Xuan S, Wang Z. Oral microbiota: A new view of body health. Food Sci Hum Well. 2019;8:8-15.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 73]  [Cited by in F6Publishing: 77]  [Article Influence: 15.4]  [Reference Citation Analysis (0)]
27.  Ribeiro AA, Arnold RR.   Dysbiosis of the Oral Microbiome. In: Azcarate-Peril M., Arnold R., Bruno-Bárcena J. (eds) How Fermented Foods Feed a Healthy Gut Microbiota. Switzerland: Springer Nature, 2019.  [PubMed]  [DOI]  [Cited in This Article: ]
28.  Li J, Parada C, Chai Y. Cellular and molecular mechanisms of tooth root development. Development. 2017;144:374-384.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 115]  [Cited by in F6Publishing: 149]  [Article Influence: 21.3]  [Reference Citation Analysis (0)]
29.  Huang GT. Dental pulp and dentin tissue engineering and regeneration: advancement and challenge. Front Biosci (Elite Ed). 2011;3:788-800.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 135]  [Cited by in F6Publishing: 118]  [Article Influence: 9.1]  [Reference Citation Analysis (0)]
30.  Baker A, Karpagaselvi K, Kumaraswamy J, Ranjini MR, Gowher J. Role of dental pulp in age estimation: A quantitative and morphometric study. J Forensic Dent Sci. 2019;11:95-102.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3]  [Cited by in F6Publishing: 3]  [Article Influence: 0.8]  [Reference Citation Analysis (0)]
31.  Frank RM, Nalbandian J.   Structure and Ultrastructure of the Dental Pulp. In: Teeth. Handbook of Microscopic Anatomy (Continuation of Handbuch der mikroskopischen Anatomie des Menschen). Berlin, Heidelberg: Springer 1989: 249-307.  [PubMed]  [DOI]  [Cited in This Article: ]
32.  Abrahão IJ, Martins MD, Katayama E, Antoniazzi JH, Segmentilli A, Marques MM. Collagen analysis in human tooth germ papillae. Braz Dent J. 2006;17:208-212.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 19]  [Cited by in F6Publishing: 22]  [Article Influence: 1.2]  [Reference Citation Analysis (0)]
33.  Gronthos S, Mankani M, Brahim J, Robey PG, Shi S. Postnatal human dental pulp stem cells (DPSCs) in vitro and in vivo. Proc Natl Acad Sci U S A. 2000;97:13625-13630.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3174]  [Cited by in F6Publishing: 3128]  [Article Influence: 130.3]  [Reference Citation Analysis (0)]
34.  Arthur A, Rychkov G, Shi S, Koblar SA, Gronthos S. Adult human dental pulp stem cells differentiate toward functionally active neurons under appropriate environmental cues. Stem Cells. 2008;26:1787-1795.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 411]  [Cited by in F6Publishing: 399]  [Article Influence: 24.9]  [Reference Citation Analysis (0)]
35.  Sui B, Chen C, Kou X, Li B, Xuan K, Shi S, Jin Y. Pulp Stem Cell-Mediated Functional Pulp Regeneration. J Dent Res. 2019;98:27-35.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 66]  [Cited by in F6Publishing: 79]  [Article Influence: 15.8]  [Reference Citation Analysis (0)]
36.  Miura M, Gronthos S, Zhao M, Lu B, Fisher LW, Robey PG, Shi S. SHED: stem cells from human exfoliated deciduous teeth. Proc Natl Acad Sci U S A. 2003;100:5807-5812.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1922]  [Cited by in F6Publishing: 1852]  [Article Influence: 88.2]  [Reference Citation Analysis (0)]
37.  Kunimatsu R, Nakajima K, Awada T, Tsuka Y, Abe T, Ando K, Hiraki T, Kimura A, Tanimoto K. Comparative characterization of stem cells from human exfoliated deciduous teeth, dental pulp, and bone marrow-derived mesenchymal stem cells. Biochem Biophys Res Commun. 2018;501:193-198.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 84]  [Cited by in F6Publishing: 113]  [Article Influence: 18.8]  [Reference Citation Analysis (0)]
38.  Govindasamy V, Abdullah AN, Ronald VS, Musa S, Ab Aziz ZA, Zain RB, Totey S, Bhonde RR, Abu Kasim NH. Inherent differential propensity of dental pulp stem cells derived from human deciduous and permanent teeth. J Endod. 2010;36:1504-1515.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 108]  [Cited by in F6Publishing: 98]  [Article Influence: 7.0]  [Reference Citation Analysis (0)]
39.  Suchánek J, Visek B, Soukup T, El-Din Mohamed SK, Ivancaková R, Mokrỳ J, Aboul-Ezz EH, Omran A. Stem cells from human exfoliated deciduous teeth--isolation, long term cultivation and phenotypical analysis. Acta Medica (Hradec Kralove). 2010;53:93-99.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 38]  [Cited by in F6Publishing: 42]  [Article Influence: 3.0]  [Reference Citation Analysis (0)]
40.  Ma L, Makino Y, Yamaza H, Akiyama K, Hoshino Y, Song G, Kukita T, Nonaka K, Shi S, Yamaza T. Cryopreserved dental pulp tissues of exfoliated deciduous teeth is a feasible stem cell resource for regenerative medicine. PLoS One. 2012;7:e51777.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 104]  [Cited by in F6Publishing: 109]  [Article Influence: 9.1]  [Reference Citation Analysis (0)]
41.  Demarco FF, Conde MC, Cavalcanti BN, Casagrande L, Sakai VT, Nör JE. Dental pulp tissue engineering. Braz Dent J. 2011;22:3-13.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 75]  [Cited by in F6Publishing: 89]  [Article Influence: 6.8]  [Reference Citation Analysis (0)]
42.  Andrukhov O, Behm C, Blufstein A, Rausch-Fan X. Immunomodulatory properties of dental tissue-derived mesenchymal stem cells: Implication in disease and tissue regeneration. World J Stem Cells. 2019;11:604-617.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 99]  [Cited by in F6Publishing: 107]  [Article Influence: 21.4]  [Reference Citation Analysis (0)]
43.  Guerrero-Jiménez M, Nic-Can GI, Castro-Linares N, Aguilar-Ayala FJ, Canul-Chan M, Rojas-Herrera RA, Peñaloza-Cuevas R, Rodas-Junco BA. In vitro histomorphometric comparison of dental pulp tissue in different teeth. PeerJ. 2019;7:e8212.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1]  [Cited by in F6Publishing: 2]  [Article Influence: 0.4]  [Reference Citation Analysis (0)]
44.  Könönen E, Gursoy M, Gursoy UK. Periodontitis: A Multifaceted Disease of Tooth-Supporting Tissues. J Clin Med. 2019;8.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 199]  [Cited by in F6Publishing: 297]  [Article Influence: 59.4]  [Reference Citation Analysis (0)]
45.  Jiang N, Guo W, Chen M, Zheng Y, Zhou J, Kim SG, Embree MC, Songhee Song K, Marao HF, Mao JJ. Periodontal Ligament and Alveolar Bone in Health and Adaptation: Tooth Movement. Front Oral Biol. 2016;18:1-8.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 55]  [Cited by in F6Publishing: 76]  [Article Influence: 8.4]  [Reference Citation Analysis (0)]
46.  Zhang X, Li H, Sun J, Luo X, Yang H, Xie L, Yang B, Guo W, Tian W. Cell-derived micro-environment helps dental pulp stem cells promote dental pulp regeneration. Cell Prolif. 2017;50.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 30]  [Cited by in F6Publishing: 28]  [Article Influence: 4.0]  [Reference Citation Analysis (0)]
47.  Hakki SS, Kayis SA, Hakki EE, Bozkurt SB, Duruksu G, Unal ZS, Turaç G, Karaoz E. Comparison of mesenchymal stem cells isolated from pulp and periodontal ligament. J Periodontol. 2015;86:283-291.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 35]  [Cited by in F6Publishing: 35]  [Article Influence: 3.5]  [Reference Citation Analysis (0)]
48.  Koller A, Sapra A.   Anatomy, Head and Neck, Oral Gingiva. 2021 Aug 11. In: StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing; 2021 Jan-.  [PubMed]  [DOI]  [Cited in This Article: ]
49.  Häkkinen L, Larjava H, Fournier BP. Distinct phenotype and therapeutic potential of gingival fibroblasts. Cytotherapy. 2014;16:1171-1186.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 44]  [Cited by in F6Publishing: 54]  [Article Influence: 5.4]  [Reference Citation Analysis (0)]
50.  Penmetsa GS, Baddam S, Manyam R, Dwarakanath CD. Comparison of the number of gingival blood vessels between type 2 diabetes mellitus and chronic periodontitis patients: An immunohistological study. J Indian Soc Periodontol. 2015;19:164-168.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 5]  [Cited by in F6Publishing: 6]  [Article Influence: 0.7]  [Reference Citation Analysis (0)]
51.  Venkatesh D, Kumar KPM, Alur JB. Gingival mesenchymal stem cells. J Oral Maxillofac Pathol. 2017;21:296-298.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 22]  [Cited by in F6Publishing: 23]  [Article Influence: 3.3]  [Reference Citation Analysis (0)]
52.  Angelopoulos I, Brizuela C, Khoury M. Gingival Mesenchymal Stem Cells Outperform Haploidentical Dental Pulp-derived Mesenchymal Stem Cells in Proliferation Rate, Migration Ability, and Angiogenic Potential. Cell Transplant. 2018;27:967-978.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 22]  [Cited by in F6Publishing: 36]  [Article Influence: 6.0]  [Reference Citation Analysis (0)]
53.  Bastos VC, Gomez RS, Gomes CC. Revisiting the human dental follicle: From tooth development to its association with unerupted or impacted teeth and pathological changes. Dev Dyn. 2021;.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 4]  [Cited by in F6Publishing: 12]  [Article Influence: 4.0]  [Reference Citation Analysis (1)]
54.  Bi R, Lyu P, Song Y, Li P, Song D, Cui C, Fan Y. Function of Dental Follicle Progenitor/Stem Cells and Their Potential in Regenerative Medicine: From Mechanisms to Applications. Biomolecules. 2021;11.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 18]  [Cited by in F6Publishing: 15]  [Article Influence: 5.0]  [Reference Citation Analysis (0)]
55.  Meleti M, van der Waal I. Clinicopathological evaluation of 164 dental follicles and dentigerous cysts with emphasis on the presence of odontogenic epithelium in the connective tissue. The hypothesis of "focal ameloblastoma". Med Oral Patol Oral Cir Bucal. 2013;18:e60-e64.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 12]  [Cited by in F6Publishing: 13]  [Article Influence: 1.2]  [Reference Citation Analysis (0)]
56.  Zheng C, Chen J, Liu S, Jin Y. Stem cell-based bone and dental regeneration: a view of microenvironmental modulation. Int J Oral Sci. 2019;11:23.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 94]  [Cited by in F6Publishing: 112]  [Article Influence: 22.4]  [Reference Citation Analysis (0)]
57.  Kang J, Fan W, Deng Q, He H, Huang F. Stem Cells from the Apical Papilla: A Promising Source for Stem Cell-Based Therapy. Biomed Res Int. 2019;2019:6104738.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 42]  [Cited by in F6Publishing: 82]  [Article Influence: 16.4]  [Reference Citation Analysis (0)]
58.  Pelissari C, Paris AFC, Mantesso A, Trierveiler M. Apical Papilla Cells Are Capable of Forming a Pulplike Tissue with Odontoblastlike Cells without the Use of Exogenous Growth Factors. J Endod. 2018;44:1671-1676.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 10]  [Cited by in F6Publishing: 8]  [Article Influence: 1.6]  [Reference Citation Analysis (0)]
59.  Nada OA, El Backly RM. Stem Cells From the Apical Papilla (SCAP) as a Tool for Endogenous Tissue Regeneration. Front Bioeng Biotechnol. 2018;6:103.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 51]  [Cited by in F6Publishing: 74]  [Article Influence: 12.3]  [Reference Citation Analysis (0)]
60.  Winning L, El Karim IA, Lundy FT. A Comparative Analysis of the Osteogenic Potential of Dental Mesenchymal Stem Cells. Stem Cells Dev. 2019;28:1050-1058.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 31]  [Cited by in F6Publishing: 32]  [Article Influence: 6.4]  [Reference Citation Analysis (0)]
61.  Mettraux GR, Gusberti FA, Graf H. Oxygen tension (pO2) in untreated human periodontal pockets. J Periodontol. 1984;55:516-521.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 118]  [Cited by in F6Publishing: 120]  [Article Influence: 3.0]  [Reference Citation Analysis (0)]
62.  Yu CY, Boyd NM, Cringle SJ, Alder VA, Yu DY. Oxygen distribution and consumption in rat lower incisor pulp. Arch Oral Biol. 2002;47:529-536.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 36]  [Cited by in F6Publishing: 39]  [Article Influence: 1.8]  [Reference Citation Analysis (0)]
63.  Kwon SY, Chun SY, Ha YS, Kim DH, Kim J, Song PH, Kim HT, Yoo ES, Kim BS, Kwon TG. Hypoxia Enhances Cell Properties of Human Mesenchymal Stem Cells. Tissue Eng Regen Med. 2017;14:595-604.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 58]  [Cited by in F6Publishing: 49]  [Article Influence: 7.0]  [Reference Citation Analysis (0)]
64.  Yu Y, Bi CS, Wu RX, Yin Y, Zhang XY, Lan PH, Chen FM. Effects of short-term inflammatory and/or hypoxic pretreatments on periodontal ligament stem cells: in vitro and in vivo studies. Cell Tissue Res. 2016;366:311-328.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 28]  [Cited by in F6Publishing: 23]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
65.  Li L, Han MX, Li S, Xu Y, Wang L. Hypoxia regulates the proliferation and osteogenic differentiation of human periodontal ligament cells under cyclic tensile stress via mitogen-activated protein kinase pathways. J Periodontol. 2014;85:498-508.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 18]  [Cited by in F6Publishing: 18]  [Article Influence: 1.8]  [Reference Citation Analysis (0)]
66.  Zhang QB, Zhang ZQ, Fang SL, Liu YR, Jiang G, Li KF. Effects of hypoxia on proliferation and osteogenic differentiation of periodontal ligament stem cells: an in vitro and in vivo study. Genet Mol Res. 2014;13:10204-10214.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 25]  [Cited by in F6Publishing: 25]  [Article Influence: 2.5]  [Reference Citation Analysis (0)]
67.  Choi H, Jin H, Kim JY, Lim KT, Choung HW, Park JY, Chung JH, Choung PH. Hypoxia promotes CEMP1 expression and induces cementoblastic differentiation of human dental stem cells in an HIF-1-dependent manner. Tissue Eng Part A. 2014;20:410-423.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 11]  [Article Influence: 1.1]  [Reference Citation Analysis (0)]
68.  Kanafi MM, Ramesh A, Gupta PK, Bhonde RR. Influence of hypoxia, high glucose, and low serum on the growth kinetics of mesenchymal stem cells from deciduous and permanent teeth. Cells Tissues Organs. 2013;198:198-208.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 31]  [Cited by in F6Publishing: 32]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
69.  Novais A, Lesieur J, Sadoine J, Slimani L, Baroukh B, Saubaméa B, Schmitt A, Vital S, Poliard A, Hélary C, Rochefort GY, Chaussain C, Gorin C. Priming Dental Pulp Stem Cells from Human Exfoliated Deciduous Teeth with Fibroblast Growth Factor-2 Enhances Mineralization Within Tissue-Engineered Constructs Implanted in Craniofacial Bone Defects. Stem Cells Transl Med. 2019;8:844-857.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 51]  [Cited by in F6Publishing: 34]  [Article Influence: 6.8]  [Reference Citation Analysis (0)]
70.  Vanacker J, Viswanath A, De Berdt P, Everard A, Cani PD, Bouzin C, Feron O, Diogenes A, Leprince JG, des Rieux A. Hypoxia modulates the differentiation potential of stem cells of the apical papilla. J Endod. 2014;40:1410-1418.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 47]  [Cited by in F6Publishing: 43]  [Article Influence: 4.3]  [Reference Citation Analysis (0)]
71.  Zhou Y, Fan W, Xiao Y. The effect of hypoxia on the stemness and differentiation capacity of PDLC and DPC. Biomed Res Int. 2014;2014:890675.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 29]  [Cited by in F6Publishing: 33]  [Article Influence: 3.3]  [Reference Citation Analysis (0)]
72.  Murabayashi D, Mochizuki M, Tamaki Y, Nakahara T. Practical methods for handling human periodontal ligament stem cells in serum-free and serum-containing culture conditions under hypoxia: implications for regenerative medicine. Hum Cell. 2017;30:169-180.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 12]  [Cited by in F6Publishing: 12]  [Article Influence: 1.7]  [Reference Citation Analysis (0)]
73.  Labedz-Maslowska A, Bryniarska N, Kubiak A, Kaczmarzyk T, Sekula-Stryjewska M, Noga S, Boruczkowski D, Madeja Z, Zuba-Surma E. Multilineage Differentiation Potential of Human Dental Pulp Stem Cells-Impact of 3D and Hypoxic Environment on Osteogenesis In Vitro. Int J Mol Sci. 2020;21.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 14]  [Cited by in F6Publishing: 15]  [Article Influence: 3.8]  [Reference Citation Analysis (0)]
74.  Xiao Z, Han Y, Zhang Y, Zhang X. Hypoxia-regulated human periodontal ligament cells via Wnt/β-catenin signaling pathway. Medicine (Baltimore). 2017;96:e6562.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 20]  [Cited by in F6Publishing: 18]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
75.  Janjić K, Samiei M, Moritz A, Agis H. The Influence of Pro-Inflammatory Factors on Sclerostin and Dickkopf-1 Production in Human Dental Pulp Cells Under Hypoxic Conditions. Front Bioeng Biotechnol. 2019;7:430.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3]  [Cited by in F6Publishing: 3]  [Article Influence: 0.6]  [Reference Citation Analysis (0)]
76.  Kuang R, Zhang Z, Jin X, Hu J, Shi S, Ni L, Ma PX. Nanofibrous spongy microspheres for the delivery of hypoxia-primed human dental pulp stem cells to regenerate vascularized dental pulp. Acta Biomater. 2016;33:225-234.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 83]  [Cited by in F6Publishing: 62]  [Article Influence: 7.8]  [Reference Citation Analysis (0)]
77.  Wu Y, Huang F, Zhou X, Yu S, Tang Q, Li S, Wang J, Chen L. Hypoxic Preconditioning Enhances Dental Pulp Stem Cell Therapy for Infection-Caused Bone Destruction. Tissue Eng Part A. 2016;22:1191-1203.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 10]  [Cited by in F6Publishing: 10]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
78.  Matsuda N, Morita N, Matsuda K, Watanabe M. Proliferation and differentiation of human osteoblastic cells associated with differential activation of MAP kinases in response to epidermal growth factor, hypoxia, and mechanical stress in vitro. Biochem Biophys Res Commun. 1998;249:350-354.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 163]  [Cited by in F6Publishing: 152]  [Article Influence: 5.8]  [Reference Citation Analysis (0)]
79.  Hou JX, Loomer PM. [Hypoxia regulates osteogenic differentiation of human periodontal ligament stem cells]. Zhonghua Kou Qiang Yi Xue Za Zhi. 2009;44:543-547.  [PubMed]  [DOI]  [Cited in This Article: ]
80.  Wu Y, Cao H, Yang Y, Zhou Y, Gu Y, Zhao X, Zhang Y, Zhao Z, Zhang L, Yin J. Effects of vascular endothelial cells on osteogenic differentiation of noncontact co-cultured periodontal ligament stem cells under hypoxia. J Periodontal Res. 2013;48:52-65.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 44]  [Cited by in F6Publishing: 39]  [Article Influence: 3.5]  [Reference Citation Analysis (0)]
81.  Chen D, Wu L, Liu L, Gong Q, Zheng J, Peng C, Deng J. Comparison of HIF1A-AS1 and HIF1A-AS2 in regulating HIF1α and the osteogenic differentiation of PDLCs under hypoxia. Int J Mol Med. 2017;40:1529-1536.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 28]  [Cited by in F6Publishing: 39]  [Article Influence: 5.6]  [Reference Citation Analysis (0)]
82.  Xu Q, Liu Z, Guo L, Liu R, Li R, Chu X, Yang J, Luo J, Chen F, Deng M. Hypoxia Mediates Runt-Related Transcription Factor 2 Expression via Induction of Vascular Endothelial Growth Factor in Periodontal Ligament Stem Cells. Mol Cells. 2019;42:763-772.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 9]  [Reference Citation Analysis (0)]
83.  Yang H, Li G, Han N, Zhang X, Cao Y, Fan Z. Secreted frizzled-related protein 2 promotes the osteo/odontogenic differentiation and paracrine potentials of stem cells from apical papilla under inflammation and hypoxia conditions. Cell Prolif. 2020;53:e12694.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 10]  [Cited by in F6Publishing: 12]  [Article Influence: 2.4]  [Reference Citation Analysis (0)]
84.  Muñoz-Sánchez J, Chánez-Cárdenas ME. The use of cobalt chloride as a chemical hypoxia model. J Appl Toxicol. 2019;39:556-570.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 139]  [Cited by in F6Publishing: 149]  [Article Influence: 24.8]  [Reference Citation Analysis (0)]
85.  Teti G, Focaroli S, Salvatore V, Mazzotti E, Ingra' L, Mazzotti A, Falconi M. The Hypoxia-Mimetic Agent Cobalt Chloride Differently Affects Human Mesenchymal Stem Cells in Their Chondrogenic Potential. Stem Cells Int. 2018;2018:3237253.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 22]  [Cited by in F6Publishing: 24]  [Article Influence: 4.0]  [Reference Citation Analysis (0)]
86.  Osathanon T, Vivatbutsiri P, Sukarawan W, Sriarj W, Pavasant P, Sooampon S. Cobalt chloride supplementation induces stem-cell marker expression and inhibits osteoblastic differentiation in human periodontal ligament cells. Arch Oral Biol. 2015;60:29-36.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 39]  [Cited by in F6Publishing: 39]  [Article Influence: 4.3]  [Reference Citation Analysis (0)]
87.  Song ZC, Li S, Dong JC, Sun MJ, Zhang XL, Shu R. Enamel matrix proteins regulate hypoxia-induced cellular biobehavior and osteogenic differentiation in human periodontal ligament cells. Biotech Histochem. 2017;92:606-618.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 7]  [Cited by in F6Publishing: 7]  [Article Influence: 1.0]  [Reference Citation Analysis (0)]
88.  Dong JC, Song ZC, Shu R, Li S, Lin ZK, Zhang XL. [The effect of hypoxia on proliferation and osteogenic differentiation of periodontal ligament cells]. Shanghai Kou Qiang Yi Xue. 2014;23:397-401.  [PubMed]  [DOI]  [Cited in This Article: ]
89.  Zheng J, Zhu X, He Y, Hou S, Liu T, Zhi K, Hou T, Gao L. CircCDK8 regulates osteogenic differentiation and apoptosis of PDLSCs by inducing ER stress/autophagy during hypoxia. Ann N Y Acad Sci. 2021;1485:56-70.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 30]  [Cited by in F6Publishing: 45]  [Article Influence: 11.3]  [Reference Citation Analysis (0)]
90.  Chen Y, Zhao Q, Yang X, Yu X, Yu D, Zhao W. Effects of cobalt chloride on the stem cell marker expression and osteogenic differentiation of stem cells from human exfoliated deciduous teeth. Cell Stress Chaperones. 2019;24:527-538.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 20]  [Cited by in F6Publishing: 9]  [Article Influence: 1.8]  [Reference Citation Analysis (0)]
91.  Mu S, Guo S, Wang X, Zhan Y, Li Y, Jiang Y, Zhang R, Zhang B. Effects of deferoxamine on the osteogenic differentiation of human periodontal ligament cells. Mol Med Rep. 2017;16:9579-9586.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 10]  [Cited by in F6Publishing: 11]  [Article Influence: 1.6]  [Reference Citation Analysis (0)]
92.  Li S, Shao J, Zhou Y, Friis T, Yao J, Shi B, Xiao Y. The impact of Wnt signalling and hypoxia on osteogenic and cementogenic differentiation in human periodontal ligament cells. Mol Med Rep. 2016;14:4975-4982.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 15]  [Cited by in F6Publishing: 17]  [Article Influence: 2.1]  [Reference Citation Analysis (0)]
93.  Gao L, Xu T, Huang G, Jiang S, Gu Y, Chen F. Oral microbiomes: more and more importance in oral cavity and whole body. Protein Cell. 2018;9:488-500.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 382]  [Cited by in F6Publishing: 366]  [Article Influence: 61.0]  [Reference Citation Analysis (0)]
94.  Bowen WH, Burne RA, Wu H, Koo H. Oral Biofilms: Pathogens, Matrix, and Polymicrobial Interactions in Microenvironments. Trends Microbiol. 2018;26:229-242.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 581]  [Cited by in F6Publishing: 494]  [Article Influence: 82.3]  [Reference Citation Analysis (0)]
95.  Corrêa-Faria P, Paixão-Gonçalves S, Paiva SM, Pordeus IA. Incidence of dental caries in primary dentition and risk factors: a longitudinal study. Braz Oral Res. 2016;30.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 22]  [Cited by in F6Publishing: 25]  [Article Influence: 3.1]  [Reference Citation Analysis (0)]
96.  Nazir MA. Prevalence of periodontal disease, its association with systemic diseases and prevention. Int J Health Sci (Qassim). 2017;11:72-80.  [PubMed]  [DOI]  [Cited in This Article: ]
97.  Inostroza C, Vega-Letter AM, Brizuela C, Castrillón L, Saint Jean N, Duran CM, Carrión F. Mesenchymal Stem Cells Derived from Human Inflamed Dental Pulp Exhibit Impaired Immunomodulatory Capacity In Vitro. J Endod. 2020;46:1091-1098.e2.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3]  [Cited by in F6Publishing: 3]  [Article Influence: 0.8]  [Reference Citation Analysis (0)]
98.  Park YT, Lee SM, Kou X, Karabucak B. The Role of Interleukin 6 in Osteogenic and Neurogenic Differentiation Potentials of Dental Pulp Stem Cells. J Endod. 2019;45:1342-1348.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 12]  [Cited by in F6Publishing: 13]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
99.  Alongi DJ, Yamaza T, Song Y, Fouad AF, Romberg EE, Shi S, Tuan RS, Huang GT. Stem/progenitor cells from inflamed human dental pulp retain tissue regeneration potential. Regen Med. 2010;5:617-631.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 189]  [Cited by in F6Publishing: 189]  [Article Influence: 13.5]  [Reference Citation Analysis (0)]
100.  Tsai AI, Hong HH, Lin WR, Fu JF, Chang CC, Wang IK, Huang WH, Weng CH, Hsu CW, Yen TH. Isolation of Mesenchymal Stem Cells from Human Deciduous Teeth Pulp. Biomed Res Int. 2017;2017:2851906.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 12]  [Cited by in F6Publishing: 17]  [Article Influence: 2.4]  [Reference Citation Analysis (0)]
101.  Tomasello L, Mauceri R, Coppola A, Pitrone M, Pizzo G, Campisi G, Pizzolanti G, Giordano C. Mesenchymal stem cells derived from inflamed dental pulpal and gingival tissue: a potential application for bone formation. Stem Cell Res Ther. 2017;8:179.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 60]  [Cited by in F6Publishing: 63]  [Article Influence: 9.0]  [Reference Citation Analysis (0)]
102.  Li N, Liu N, Zhou J, Tang L, Ding B, Duan Y, Jin Y. Inflammatory environment induces gingival tissue-specific mesenchymal stem cells to differentiate towards a pro-fibrotic phenotype. Biol Cell. 2013;105:261-275.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 28]  [Cited by in F6Publishing: 28]  [Article Influence: 2.5]  [Reference Citation Analysis (0)]
103.  Al-Qadhi G, Al-Rai S, Hafed L. The Therapeutic Potential of Inflamed Gingiva-Derived Mesenchymal Stem Cells in Preclinical Studies: A Scoping Review of a Unique Biomedical Waste. Stem Cells Int. 2021;2021:6619170.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 6]  [Cited by in F6Publishing: 5]  [Article Influence: 1.7]  [Reference Citation Analysis (0)]
104.  Ramenzoni LL, Russo G, Moccia MD, Attin T, Schmidlin PR. Periodontal bacterial supernatants modify differentiation, migration and inflammatory cytokine expression in human periodontal ligament stem cells. PLoS One. 2019;14:e0219181.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 8]  [Cited by in F6Publishing: 11]  [Article Influence: 2.2]  [Reference Citation Analysis (0)]
105.  Park JC, Kim JM, Jung IH, Kim JC, Choi SH, Cho KS, Kim CS. Isolation and characterization of human periodontal ligament (PDL) stem cells (PDLSCs) from the inflamed PDL tissue: in vitro and in vivo evaluations. J Clin Periodontol. 2011;38:721-731.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 182]  [Cited by in F6Publishing: 189]  [Article Influence: 14.5]  [Reference Citation Analysis (0)]
106.  Tang HN, Xia Y, Yu Y, Wu RX, Gao LN, Chen FM. Stem cells derived from "inflamed" and healthy periodontal ligament tissues and their sheet functionalities: a patient-matched comparison. J Clin Periodontol. 2016;43:72-84.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 56]  [Cited by in F6Publishing: 64]  [Article Influence: 8.0]  [Reference Citation Analysis (1)]
107.  Liu N, Shi S, Deng M, Tang L, Zhang G, Liu N, Ding B, Liu W, Liu Y, Shi H, Liu L, Jin Y. High levels of β-catenin signaling reduce osteogenic differentiation of stem cells in inflammatory microenvironments through inhibition of the noncanonical Wnt pathway. J Bone Miner Res. 2011;26:2082-2095.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 144]  [Cited by in F6Publishing: 154]  [Article Influence: 11.8]  [Reference Citation Analysis (0)]
108.  Soheilifar S, Amiri I, Bidgoli M, Hedayatipanah M. Comparison of Periodontal Ligament Stem Cells Isolated from the Periodontium of Healthy Teeth and Periodontitis-Affected Teeth. J Dent (Tehran). 2016;13:271-278.  [PubMed]  [DOI]  [Cited in This Article: ]
109.  Fagalde P, Reininger D. Oral tissues regeneration using intraoral mesenchymal stem cells. J Clin Exp Dent. 2021;13:e268-e277.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3]  [Cited by in F6Publishing: 3]  [Article Influence: 1.0]  [Reference Citation Analysis (0)]
110.  Yoshida S, Tomokiyo A, Hasegawa D, Hamano S, Sugii H, Maeda H. Insight into the Role of Dental Pulp Stem Cells in Regenerative Therapy. Biology (Basel). 2020;9.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 19]  [Cited by in F6Publishing: 22]  [Article Influence: 5.5]  [Reference Citation Analysis (0)]
111.  d'Aquino R, De Rosa A, Lanza V, Tirino V, Laino L, Graziano A, Desiderio V, Laino G, Papaccio G. Human mandible bone defect repair by the grafting of dental pulp stem/progenitor cells and collagen sponge biocomplexes. Eur Cell Mater. 2009;18:75-83.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 303]  [Cited by in F6Publishing: 284]  [Article Influence: 18.9]  [Reference Citation Analysis (0)]
112.  Nakashima M, Iohara K, Murakami M, Nakamura H, Sato Y, Ariji Y, Matsushita K. Pulp regeneration by transplantation of dental pulp stem cells in pulpitis: a pilot clinical study. Stem Cell Res Ther. 2017;8:61.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 199]  [Cited by in F6Publishing: 218]  [Article Influence: 31.1]  [Reference Citation Analysis (0)]
113.  Barbier L, Ramos E, Mendiola J, Rodriguez O, Santamaria G, Santamaria J, Arteagoitia I. Autologous dental pulp mesenchymal stem cells for inferior third molar post-extraction socket healing: A split-mouth randomised clinical trial. Med Oral Patol Oral Cir Bucal. 2018;23:e469-e477.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 9]  [Cited by in F6Publishing: 11]  [Article Influence: 1.8]  [Reference Citation Analysis (0)]
114.  Xuan K, Li B, Guo H, Sun W, Kou X, He X, Zhang Y, Sun J, Liu A, Liao L, Liu S, Liu W, Hu C, Shi S, Jin Y. Deciduous autologous tooth stem cells regenerate dental pulp after implantation into injured teeth. Sci Transl Med. 2018;10.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 198]  [Cited by in F6Publishing: 243]  [Article Influence: 48.6]  [Reference Citation Analysis (0)]
115.  Tanikawa DYS, Pinheiro CCG, Almeida MCA, Oliveira CRGCM, Coudry RA, Rocha DL, Bueno DF. Deciduous Dental Pulp Stem Cells for Maxillary Alveolar Reconstruction in Cleft Lip and Palate Patients. Stem Cells Int. 2020;2020:6234167.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 14]  [Cited by in F6Publishing: 23]  [Article Influence: 5.8]  [Reference Citation Analysis (0)]
116.  Redondo LM, García V, Peral B, Verrier A, Becerra J, Sánchez A, García-Sancho J. Repair of maxillary cystic bone defects with mesenchymal stem cells seeded on a cross-linked serum scaffold. J Craniomaxillofac Surg. 2018;46:222-229.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 26]  [Cited by in F6Publishing: 23]  [Article Influence: 3.8]  [Reference Citation Analysis (0)]
117.  Feng F, Akiyama K, Liu Y, Yamaza T, Wang TM, Chen JH, Wang BB, Huang GT, Wang S, Shi S. Utility of PDL progenitors for in vivo tissue regeneration: a report of 3 cases. Oral Dis. 2010;16:20-28.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 209]  [Cited by in F6Publishing: 211]  [Article Influence: 15.1]  [Reference Citation Analysis (0)]
118.  Ferrarotti F, Romano F, Gamba MN, Quirico A, Giraudi M, Audagna M, Aimetti M. Human intrabony defect regeneration with micrografts containing dental pulp stem cells: A randomized controlled clinical trial. J Clin Periodontol. 2018;45:841-850.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 66]  [Cited by in F6Publishing: 90]  [Article Influence: 15.0]  [Reference Citation Analysis (0)]
119.  Iwata T, Yamato M, Washio K, Yoshida T, Tsumanuma Y, Yamada A, Onizuka S, Izumi Y, Ando T, Okano T, Ishikawa I. Periodontal regeneration with autologous periodontal ligament-derived cell sheets - A safety and efficacy study in ten patients. Regen Ther. 2018;9:38-44.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 96]  [Cited by in F6Publishing: 124]  [Article Influence: 20.7]  [Reference Citation Analysis (0)]
120.  Chen FM, Gao LN, Tian BM, Zhang XY, Zhang YJ, Dong GY, Lu H, Chu Q, Xu J, Yu Y, Wu RX, Yin Y, Shi S, Jin Y. Treatment of periodontal intrabony defects using autologous periodontal ligament stem cells: a randomized clinical trial. Stem Cell Res Ther. 2016;7:33.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 160]  [Cited by in F6Publishing: 196]  [Article Influence: 24.5]  [Reference Citation Analysis (0)]
121.  Sánchez N, Fierravanti L, Núñez J, Vignoletti F, González-Zamora M, Santamaría S, Suárez-Sancho S, Fernández-Santos ME, Figuero E, Herrera D, García-Sanz JA, Sanz M. Periodontal regeneration using a xenogeneic bone substitute seeded with autologous periodontal ligament-derived mesenchymal stem cells: A 12-month quasi-randomized controlled pilot clinical trial. J Clin Periodontol. 2020;47:1391-1402.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 14]  [Cited by in F6Publishing: 34]  [Article Influence: 11.3]  [Reference Citation Analysis (0)]
122.  Abdal-Wahab M, Abdel Ghaffar KA, Ezzatt OM, Hassan AAA, El Ansary MMS, Gamal AY. Regenerative potential of cultured gingival fibroblasts in treatment of periodontal intrabony defects (randomized clinical and biochemical trial). J Periodontal Res. 2020;55:441-452.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 19]  [Cited by in F6Publishing: 20]  [Article Influence: 5.0]  [Reference Citation Analysis (0)]
123.  Phinney DG. Functional heterogeneity of mesenchymal stem cells: implications for cell therapy. J Cell Biochem. 2012;113:2806-2812.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 279]  [Cited by in F6Publishing: 295]  [Article Influence: 24.6]  [Reference Citation Analysis (0)]