Next Article in Journal
Diagnosis of Myasthenia Gravis
Next Article in Special Issue
Skeletal, Dentoalveolar and Dental Changes after “Mini-Screw Assisted Rapid Palatal Expansion” Evaluated with Cone Beam Computed Tomography
Previous Article in Journal
Increased Creatine Kinase May Predict A Worse COVID-19 Outcome
Previous Article in Special Issue
Validity and Reproducibility of the Peer Assessment Rating Index Scored on Digital Models Using a Software Compared with Traditional Manual Scoring
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mechanistic Insight into Orthodontic Tooth Movement Based on Animal Studies: A Critical Review

Department of Orthodontics, School of Dental Medicine, University of Pennsylvania, 240 South 40th Street, Philadelphia, PA 19104-6030, USA
*
Author to whom correspondence should be addressed.
J. Clin. Med. 2021, 10(8), 1733; https://doi.org/10.3390/jcm10081733
Submission received: 9 March 2021 / Revised: 7 April 2021 / Accepted: 13 April 2021 / Published: 16 April 2021
(This article belongs to the Special Issue New Approaches and Technologies in Orthodontics)

Abstract

:
Alveolar bone remodeling in orthodontic tooth movement (OTM) is a highly regulated process that coordinates bone resorption by osteoclasts and new bone formation by osteoblasts. Mechanisms involved in OTM include mechano-sensing, sterile inflammation-mediated osteoclastogenesis on the compression side and tensile force-induced osteogenesis on the tension side. Several intracellular signaling pathways and mechanosensors including the cilia and ion channels transduce mechanical force into biochemical signals that stimulate formation of osteoclasts or osteoblasts. To date, many studies were performed in vitro or using human gingival crevicular fluid samples. Thus, the use of transgenic animals is very helpful in examining a cause and effect relationship. Key cell types that participate in mediating the response to OTM include periodontal ligament fibroblasts, mesenchymal stem cells, osteoblasts, osteocytes, and osteoclasts. Intercellular signals that stimulate cellular processes needed for orthodontic tooth movement include receptor activator of nuclear factor-κB ligand (RANKL), tumor necrosis factor-α (TNF-α), dickkopf Wnt signaling pathway inhibitor 1 (DKK1), sclerostin, transforming growth factor beta (TGF-β), and bone morphogenetic proteins (BMPs). In this review, we critically summarize the current OTM studies using transgenic animal models in order to provide mechanistic insight into the cellular events and the molecular regulation of OTM.

1. Introduction

Alveolar bone remodeling in orthodontic tooth movement (OTM) requires the coordinated action of different cell types, including periodontal ligament (PDL) fibroblasts, mesenchymal stem cells, inflammatory cells, osteoblasts, osteocytes, and osteoclasts. Generally, OTM is composed of three stages on the compression side; (i) a gradual compression of the PDL, which may last from about 4–7 days, (ii) the hyalinization period, when cell death due to lack of blood supply in the compressed area of the PDL occurs, which may last from 7–14 days or more, and (iii) the secondary period, which is characterized by direct bone resorption so that the tooth will continue to move [1,2,3]. On the tension side, the PDL is stretched and blood flow is activated, stimulating osteoblastic activity and osteoid deposition and mineralization. On mechanical force loading, cells around the tooth sense either compression or tension and release multiple cytokines and growth factors that stimulate subsequent biological responses. The process by which cells transmit mechanical forces and generate biological responses is essential for bone remodeling in OTM [4].
On the compression side, multinucleated osteoclasts initiate bone resorption to allow tooth movement to occur in the direction of the applied force, which is a rate limiting step in OTM. In addition, a sterile inflammatory response is induced by the generation of proinflammatory cytokines such as tissue necrosis factor (TNF), interleukin-1 (IL-1), prostaglandins, and IL-6, along with matrix metalloproteinases (MMPs) within a short time after the application of pressure [5,6,7,8,9,10]. The response to mechanical stress induces transitory inflammation that is pathogen-free. In addition, prostaglandins are secreted when cells are mechanically deformed and focal adhesion kinase, the mechanosensor in PDL cells, is known to be related with this process [11]. Therefore, both sterile inflammation and mechano-transduction are important for OTM [12]. Cells experiencing compressive forces induce osteoclastogenesis through up-regulation of receptor activator of nuclear factor kappa-Β ligand (RANKL) [13]. Proinflammatory cytokines induce RANKL expression to stimulate osteoclastogenesis, further contributing to bone resorption in OTM [14]. On the tension side, PDL cells are stretched and proliferate with increased PDL width, followed by new bone formation, eventually returning to a normal PDL width [15]. Progenitor cells in the PDL and alveolar bone proliferate and differentiate into osteoblasts to produce new bone. The osteogenic transcription factor Runx2 and bone matrix proteins osteocalcin and osteopontin are significantly up-regulated by tension forces [7]. The mechano-response, osteoclastogenesis and osteogenesis are important components of OTM as they represent simultaneous bone remodeling processes in response to mechanical loading.
Human studies examining changes induced by orthodontic forces have frequently examined gingival crevicular fluid after orthodontic force loading. While important, these studies do not establish the cause and effect relationships. Transgenic mouse models are ideal in delineating the molecular actions of specific genes, as they facilitate lineage-specific gene deletion to well-defined cell types [16]. In addition, inducible transgenic mice models are available, allowing the induction of a transgene or the deletion of an endogenous gene in a time- and tissue-specific manner to address limitations of global constitutive germ-line deletion [17].
In this review, we focus on the roles of various mechanosensory cells, cytokine expression and signaling pathways in OTM that have been identified by animal models and summarize the possible cellular mechanosensors. A better understanding of the cellular processes in OTM may one day benefit our patients by expediting tooth movement, preventing relapse and improving treatment stability through the modification of specific genes which are critical for the orthodontic bone remodeling.

2. Cytokines, Mechanosensory Cells, and Intracellular Signaling Pathways in OTM

2.1. Cytokines in OTM

2.1.1. RANKL

RANKL is a member of the TNF cytokine family and is critical for osteoclastogenesis [18]. During OTM, RANKL is highly expressed in periodontal tissue on the compression and tension side [13,19]. Numerous cell types in OTM have been shown to express RANKL including PDL fibroblasts, mesenchymal stem cells, lymphocytes, osteoblasts, and osteocytes, particularly in response to inflammatory cytokines [9,13,20,21,22].
Human gingival crevicular fluid (GCF) samples have been used for cytokine analysis during OTM as it is non-invasive and convenient. Human GCF isolated from the tooth 24 h after orthodontic forces application, the early phase of OTM, had shown a significant increase in the levels of RANKL, IL-1β, IL-6, and TNF-α, while the levels of osteoprotegerin (OPG) had remained significantly lower when compared to the control teeth [7,23,24,25]. Furthermore, Garlet et al. examined the cytokine expression on the PDL of extracted human teeth after OTM [7]. After 7 days of OTM, teeth were extracted and PDL cells on both compression and tension sides were collected for real-time PCR analysis. On the compression side, tumor necrosis factor alpha (TNF-α), RANKL, and matrix metalloproteinases (MMPs) were highly expressed. On the tension side, IL-10, tissue inhibitors of metalloproteinase 1 (TIMP-1),type I collagen, OPG, and osteocalcin were highly expressed. The same author tested chemokine expression on the extracted teeth after OTM and found that the compression side exhibited higher expression of monocyte chemoattractant protein-1 (MCP-1/CCL2), macrophage inflammatory protein-1α (MIP-1α/CCL3) and RANKL, which predominate bone resorption activity, while the tension side presented higher expression of osteocalcin.
Consistent with human studies, animal models demonstrated bone resorption activity with proinflammatory cytokines and osteoclastic markers on the compression side and bone formation activity with the osteogenic markers on the tension side. To further apply the findings to clinical orthodontics, several animal studies have examined the modulation of RANKL to accelerate OTM [26,27,28]. Injection of RANKL during OTM increases osteoclastogenesis and the rate of tooth movement [26]. Indeed, the rate of OTM is increased by 130% with RANKL injection [28]. Local RANKL gene transfer in animals with OTM increases RANKL protein expression and osteoclastogenesis without any systemic effects, accelerating the amount of tooth movement [27]. The authors proposed that local RANKL gene transfer might be a useful tool to accelerate orthodontic tooth movement, even the ankylosed teeth. Conversely, daily local RANKL antibody injection reduces the rate OTM by 70% [22]. In a comparison of RANKL gene transduction compared to periodontal accelerated osteogenic orthodontics, gene transduction led to more prolonged osteoclastogenesis and a greater rate of tooth movement during OTM [29].
In vitro compressive force causes an increase of RANKL expression and a decrease of OPG expression in human PDL cells, consistent with the human GCF and animal studies results. PDL fibroblasts are distorted under compressive force and express higher amounts of RANKL, TNF-α, MMPs, IL-1β and prostaglandins on the compression side [30]. Experimental compressive forces on the PDL resulted in a 16.7-fold increase in RANKL secretion and a 2.9-fold decrease in OPG secretion when compared to the control [23].
Taken together, studies with RANKL indicate that this cytokine is a central pro-osteoclastogenic factor that is expressed in response to mechanical forces. Interestingly, RANKL is also expressed on the tension side [19]. Thus, early induction of RANKL and osteoclastogenesis in response to tension may initiate a formation of bone remodeling that leads to increased bone formation on the tension side. This concept warrants further investigation.

2.1.2. Sclerostin

Sclerostin, encoded by the SOST gene and is primarily produced by mature osteocytes in response to OTM, promotes bone resorption and inhibits new bone formation [31,32]. Sclerostin stimulates RANKL expression by osteocytes, negatively regulates expression of BMP proteins and prevents canonical Wnt signaling [33,34]. Sclerostin expression is initially induced on the compression side in OTM models and gradually diminishes after 5–7 days, demonstrating their effect in the early phase of OTM [31,35]. During OTM, sclerostin KO mice have a 20% reduction in osteoclasts and reduced RANKL expression on the compression side with a reduced rate of tooth movement [35]. Local injection of sclerostin on the compression side doubles RANKL expression, reduces OPG expression by 30%, increases osteoclastogenesis by 150% and accelerates tooth movement [36]. In addition, in vitro studies showed that rhSCL-supplement enhanced the expression of RANKL and the RANKL/OPG ratio in osteocytes, supporting the in vivo finding. In addition, the intensity of sclerostin expression is closely related with the force magnitude [37]. On the tension side, sclerostin expression is immediately decreased and maintained at low levels during OTM, negating their negative effect on new bone formation [31]. These studies suggest that sclerostin can be a key factor in OTM by regulating both bone resorption and formation.

2.1.3. Bone Morphogenetic Proteins (BMPs)

It is well known that BMPs induce new bone formation and that the expression of BMPs increases on the tension side during OTM, stimulating the differentiation of mesenchymal stem cells to osteoblasts [30,38]. Noggin, an inhibitor of several bone morphogenetic proteins (BMPs), prevents mechanical force-induced osteoblast differentiation. BMP-3 expression is gradually increased on the tension side until 14 days in rodent models of OTM, the mid-stage in OTM [39]. One study examined the effect of BMP2 injection on tension side and found that local injection of BMP-2 on the tension side did not accelerate OTM, indicating that new bone formation per se is not a rate limiting step in OTM [40].

2.1.4. Transforming Growth Factor (TGF)-β

TGF-β signaling is involved in many cellular processes, including cell migration, proliferation, differentiation, and cellular homeostasis [41]. A previous study with extracted human teeth after OTM showed that TGF-β expression was similarly increased in both the compression and tension sides [7]. In OTM, its role on the compression side is complex as TGF-β has both positive and negative effects on osteoclastogenesis [42]. In some studies, TGF-β has been reported to inhibit osteoclastic activity. However, other studies found that TGF-β actually induces bone resorption, depending on the cell types involved, TGF-β concentration, and inducing mechanism [42,43]. Its expression on the tension side is significantly greater than that on the compression side [7,44,45,46]. TGF- β is generally known for its anabolic activity, regulating osteoblast differentiation from progenitors on the tension side [47]. Pretreatment with a TGF-β receptor inhibitor inhibits mechanical force-induced bone mineralization in vitro, suggesting that TGF-β could play a role in osteogenesis in response to tension forces during OTM.
Combining all, the findings from the animal studies could be the base foundation for the studies to expedite the OTM in humans. For example, RANKL or sclerostin can be given on the compression side to accelerate the osteoclastogenesis or BMPs can be given on the tension side to support the new bone formation. As previously mentioned, the new bone formation itself on the tension side cannot make the tooth movement faster while their effects are more important in the late phase in OTM. Therefore, many studies to speed up the velocity of OTM have been focused on the osteoclastogenic markers on the compression side.

2.2. The Mechanosensory Cells in OTM

PDL cells, osteocytes, and osteoblasts are the principal mechanosensory cells that produce various cytokines to regulate alveolar bone remodeling in OTM, by converting mechanical force into intracellular signals [48,49,50,51] (Figure 1). The role of mechanosensors in these cells during OTM will be reviewed later in this paper.

2.2.1. Periodontal Ligament Cells

Periodontal Ligament Fibroblasts

The periodontal ligament (PDL) is a fibrous tissue that connects teeth with alveolar bone and transmits mechanical stimuli [52]. The PDL comprises of heterogenous cell types including fibroblasts, progenitor cells, bone-lining cells, osteoclasts, endothelial cells, nerve cells and others [53]. PDL fibroblasts constitute 50–60% of the total PDL cellularity and contribute to bone resorption in OTM as a main source of RANKL [21]. Interestingly, PDL fibroblasts have some characteristics similar to those of osteoblasts, expressing a 2.3 kb regulatory unit of Col1α1 promoter typical of osteoblasts and osteocytes [54,55] and bone-associated proteins such as alkaline phosphatase [56]. Indeed, PDL fibroblasts are more similar to tendon cells than skin fibroblasts in many respects [30,57].
On the compression side, the experimental mice with RANKL deletion in PDL fibroblasts showed significantly less osteoclast formation with narrower PDL space compared to control mice, leading to severely impaired OTM [21]. In addition, our recent study found that this up-regulation of RANKL depends on NF-κB activation [58]. NF-κB inhibition in PDL fibroblasts blocked the OTM with significantly reduced osteoclastogenesis, narrower PDL width, higher bone volume fraction and reduced RANKL expression compared to wild type mice. Both studies support the critical role of PDL fibroblasts via NF-κB activation in OTM.

Periodontal Ligament Stem Cells

Mesenchymal stem cells reside in the PDL, giving rise to PDL, alveolar bone, and cementum during alveolar bone remodeling. Gli1+ cells have been identified as the multipotent stem cells in adult mouse PDL [59]. Complete removal of Gli1+ cells using the inhibitors or by the genetic ablation significantly reduce OTM by 60% and diminish osteoclast formation by more than 80% [60]. On the tension side, Gli1+ increases its cell number and differentiate into osteoblasts with increased Runx2 expression during OTM [60].
Mesenchymal stem cells reside in the PDL, giving rise to PDL, alveolar bone, and cementum function as mechanosensory cells during OTM [60]. Yes-associated protein (YAP) and the paralogue transcriptional coactivator with PDZ-binding motif (TAZ), the downstream effectors of the Hippo signaling pathway, have been identified as important regulators during mechanotransduction [61]. Recent rodent OTM studies showed that YAP and TAZ expression were up-regulated with nuclear translocation in the PDL cells on both compression and tension side [62,63]. Moreover, YAP and TAZ expression were proportional to the applied orthodontic force. A recent study investigated the role of Gli1+ cells through Yes-associated protein (YAP) activation in mouse OTM models [60].Lineage-specific deletion of the YAP in Gli1+ cells significantly reduced OTM by 50% with decreased osteoclast formation by more than 80% on compression side. On the tension side, the same transgenic mice with the YAP deletion in Gli1+ cells showed a significantly decreased proportion of Runx2+ cells by more than 80%. In vitro cyclic stretch promoted the osteogenic differentiation of human PDL cells [62]. Moreover, the nuclear translocation of YAP was significantly increased with increased expression of connective tissue growth factor (CTGF) and cysteine-rich angiogenic inducer 61 (CYR61) mRNA, the target gene of YAP. Furthermore, knockdown of YAP suppressed the cyclic stretch induced osteogenesis in human PDL cells, while overexpression of YAP enhanced osteogenesis. Both in vivo and in vitro data supportthe role of YAP as the mechanical sensor and important regulator of the osteogenic differentiation in PDL cells under tensile force. In addition, the level of tension is important in the osteogenic differentiation as the magnitude of tension differentially regulates osteogenic and osteoclastic process [64]. Tension with a magnitude of 12% could increase osteogenic differentiation and proliferation of mesenchymal stem cells whereas tension above 12% would up-regulate the function of mesenchymal stem cells to regulate osteoclast differentiation, demonstrating the importance of the light force during OTM especially for the patients with poor bony support such as periodontitis [65,66].
Mechanical force-induced hydrogen sulfide (H2S), produced by PDL mesenchymal stem cells, supports macrophage polarization toward an inflammatory, M1 phenotype and promotes osteoclast activity in OTM [67,68]. These cells express cystathionine-β-synthase that generates H2S. Treatment with an inhibitor of H2S reduces osteoclast formation and OTM by almost half. The generation of M1 macrophages is increased 5.6-fold after orthodontic force loading. An H2S blocker reduces M1 macrophage formation by 70%, and an H2S donor enhances it 1.4-fold. This shows that PDL mesenchymal stem cells can increase the expression of M1-macrophages, which are main source of several proinflammatory cytokines such as IL-1β, IL-6, and TNF-α, leading to the RANKL stimulation.

2.2.2. Osteocytes

Osteocytes are terminally differentiated from osteoblasts and embedded in the bone matrix. They are the most abundant cells in the adult skeleton, comprising 90–95% of all bone cells [12,51]. They are the primary mechanosensory cells in bone and regulate both osteoclast and osteoblast formation and function during mechanical force-induced bone remodeling [69,70]. They have dendritic processes that interact with other osteocytes and bone-lining cells. Mechanical loading stimulates dentin matrix protein 1 (DMP1) expression in osteocytes in vivo, which is a key molecule in regulating osteocyte formation, maturation, phosphate regulation and regulating mineralization [71,72]. In addition to RANKL, osteocytes produce sclerostin, M-CSF, OPG, and other cytokines during OTM.
As the early findings in OTM, osteocyte apoptosis peaks at 24 h on the compression side in mouse OTM models, as measured by TUNEL and caspase-3 immunofluorescence stain [73]. Osteoclastogenesis was evident after 72 h and continued to increase up to 7 days. Apoptotic osteocytes were preferentially located close to osteoclasts, suggesting that dying osteocytes produce active signaling to recruit osteoclasts [73,74,75].
Osteocytes can be also an important source of RANKL in OTM mouse models [22,76,77]. Osteocyte-deleted mice have a 60% reduction in osteoclasts and a 50% reduction in tooth movement compared to normal controls [78]. Under basal conditions osteocyte ablation negatively affects bone quality by increasing intracortical porosity, osteoblastic dysfunction, and adipose tissue proliferation in the marrow space [79]. These mice showed a severe osteopetrotic phenotype due to a lack of osteoclasts. The mice with lineage-specific RANKL deletion in osteocytes decreased OTM by 40% and osteoclast number by 60% compared with WT mice [22]. In vitro, osteocytes express a higher amount of RANKL and have a greater capacity to support osteoclastogenesis than osteoblasts and bone marrow stromal cells [77]. Interestingly, the osteoblast number on the tension side was significantly reduced in the same transgenic mice, possibly through a coupling mechanism.

2.2.3. Osteoblasts

Osteoblasts are bone forming cells, accounting for the 4–6% of total bone cells and differentiate from mesenchymal stem cells [80]. Runx2 and osterix are transcription factors that promote osteoblastic differentiation from mesenchymal stem cells. The fate of osteoblasts includes: (1) apoptosis, (2) become bone-lining cells or (3) form osteocytes. Bone-lining cells maintain homeostasis of bone and contain osteoblast progenitors [81]. Osteoblasts are mechanosensory cells and convert the mechanical signals into biological responses, producing various cytokines such as prostaglandin, OPG, RANKL and BMPs [82,83]. In OTM, bone-lining cells and osteoblasts express M-CSF and RANKL and produce other factors that positively influence osteoclastogenesis, including IL- 1β, IL-6 and TNF-α [84,85,86].
Osteoblast differentiation is an important process on the tension side during OTM. The initial response to OTM on the tension side is a proliferation of osteoblast progenitors that express α-SMA, which peaks at 2 days after initiating OTM while osteoid formation in mice peak at 4 days, the early phase of OTM, represented by osteopontin, osteocalcin, and bone sialoprotein in mouse OTM models [87,88]. Endothelin B receptors (ETB) play an important role in alveolar bone modeling in the late stage of OTM in the rat animal model [89]. To examine the role of osteoblasts in OTM, ETB knockout rats (ETB-KO) exhibited decreased OTM after 35 days, a late stage in OTM, by 27% compared to the ETB-WT mice. The alveolar bone volume in the ETB-KO appliance group was significantly less due to diminished osteoblast activity, but osteoclast volume was not significantly different compared to the ETB-WT appliance group. In addition, the expression levels of osteocalcin and DMP1, the osteoblast activity markers, were significantly down-regulated by 70% in the ETB-KO appliance group compared to the ETB-WT appliance group. However, the expression of cathepsin K, an osteoclast activity marker, did not show any statistical difference. In summary, ETB knockout rats (ETB-KO) have significantly lower osteoblast activity, unbalanced bone resorption/new bone formation, and reduced OTM with increased tooth mobility compared with control group, explaining the role of osteoblasts in the late stage of OTM.
Taken together, identifying the roles of each cells during OTM is critical and use of the transgenic mouse with each cell type-specific gene deletion can be a great tool for these studies.

2.3. Intracellular Signaling Pathways Stimulated by Mechanical Force

In OTM, various signaling pathways are activated, which mediate the response of mechanosensory cells that modulate bone resorption and formation. The function of Wnt/β-Catenin signaling, and Yes-associated protein and transcriptional coactivator with PDZ-binding motif (YAP/TAZ) signaling participate in bone remodeling in tooth movement.
Wnt/β-catenin signaling is critical for bone homeostasis [49,50]. β-catenin is a transcription factor that is activated by canonical Wnt signaling and translocates to the nucleus in osteoblasts lineage cells subjected to mechanical stimulation [90]. During OTM, Wnt/β-catenin signaling modulate expression of osteogenesis- and osteoclastogenesis-related factors in response to mechanotransduction [91,92,93]. Mice with global loss-of-function Lrp5, a Wnt receptor, have low bone mineral density and impaired osteogenic response to mechanical loading [94]. Conversely, mice with gain-of function mutations in the Lrp5 gene have significantly increase bone mineral density and bone mass in response to mechanical forces [95,96]. In OTM, a gain-of-function mutation in Lrp5 decreases orthodontic tooth movement by reducing osteoclast- mediated bone resorption and increasing alveolar bone mass [97]. Consistent with this, constitutive Wnt signaling increases osteogenic gene expression and reduces RANKL expression and osteoclast activity [98]. Conversely, viral transduction of DKK1, a Wnt inhibitor, increases osteoclast activity and reduces osteogenic markers, resulting in increased PDL width [98]. In a rat OTM model, the expression of Wnt3a, Wnt10b, and β-catenin is stronger on the tension side, consistent with Wnt induced bone formation observed under tension. In contrast, Dkk-1 levels are much higher on the compression side, consistent with reduced Wnt signaling and greater bone resorption on the compression side [99].
Yes-associated protein (YAP) and transcriptional coactivator with PDZ-binding motif (TAZ) play a key role in the mechanotransduction process [63,100]. YAP senses extracellular mechanical signals and translocates into nucleus to function as the coactivator of other transcription factors [62]. During OTM, YAP/TAZ signaling is observed in osteoblasts, osteocytes, osteoclasts and PDL fibroblasts and increases proportionally with the degree of orthodontic force [62,63]. Conditional deletion of YAP in PDL mesenchymal stem cells decreases osteoclast formation by 80% on the compression side and reduces tooth movement by half. In vitro cyclic stretch stimulates proliferation of PDL fibroblasts and osteoblast differentiation via YAP activation [62,101]. YAP knockdown suppresses mechanical forced-induced osteogenesis while overexpression of YAP enhances osteogenesis in PDL fibroblasts [62].

3. Possible Mechanosensors

Cells sense their mechanical environment through cell-cell or cell-matrix adhesions during physiologic growth and development and during mechanical loading. Mechanosensing occurs by mechanical force-induced conformational changes in cellular molecules, including force-activated cytoskeleton, integrins, ion channels and cell-cell adhesions, consequently affecting cellular gene expression and its function and regulating orthodontic bone remodeling (Figure 2).

3.1. Cytoskeleton

Cellular cytoskeletons provide structural frameworks for the cell and are largely comprised of microtubules, actin, and intermediate filaments [102]. Cytoskeletons play a role in the response to mechanical force and are responsible for cell motility [103]. For example, cilia and flagella are mainly composed of microtubules and move as a result of microtubules sliding. In OTM, PDL and alveolar bone cells are reconstructed and their cellular cytoskeleton changes stimulate the elaboration of multiple cytokines and growth factors, mediating the cell morphology, differentiation, and proliferation [102,104,105]. On the tension side, cytoskeletal reorganization influences the differentiation of osteoprogenitors to osteoblasts and bone formation, stressing the critical role of cytoskeleton to influence both compression and tension sides during OTM [106].

3.2. Focal Adhesions (FAs)

Focal adhesions are integrin-associated proteins that connect intracellular actin filaments and extracellular matrix proteins [51,107]. Orthodontic force-induced stress on the extracellular matrix can be transmitted to cells through focal adhesions to induce proliferation and differentiation of several cells in the PDL and alveolar bone, leading to the balanced bone remodeling in response to the applied force [104]. Focal adhesions are involved in mechanosensing and downstream signaling through focal adhesion kinase in osteoblasts [108] and osteocytes [109,110]. Gene deletion that results in loss of focal adhesions in osteoblasts reduce mechanical responses to fluid flow [108]. Mechanical forces through focal adhesion kinases stimulate Wnt/β-catenin signaling in osteocytes [109].

3.3. Primary Cilia

Primary cilia are non-motile protruding organelles from the cell membrane and are observed in chondrocytes, mesenchymal stem cells, osteoblasts and osteocytes as mechanosensors [111,112]. Changes in fluid flow stimulate numerous cells via primary cilia [113], which may be important in OTM. Blocking primary cilia formation inhibits the expression of osteopontin, prostaglandins and cyclooxygenase-2 in osteoblasts or osteocytes and reduces their response to fluid flow. Tensile forces promote the osteogenic differentiation and proliferation of PDL mesenchymal stem cells via primary cilia that are needed for osteoblast differentiation and bone formation [114]. Lineage-specific deletion of key ciliary proteins including the IFT80, IFT88, Kif3a, Evc and polycystin in osteoblasts or osteoblast precursors leads to cilia loss, impairs osteoblast differentiation, reduces osteoid formation, and inhibits bone mineralization in response to mechanical loading in vivo [114,115,116]. On this basis, it may have a role in bone formation on tension side during OTM.
A calcium channel complex composed of the polycystin-1 and polycystin-2 is located at the base of primary cilium and mediates the effect of cilia bending [117]. When the primary cilium is bent by dynamic fluid flow, a Ca2+ signal is transduced proportional to the degree of distortion. This bending motion opens Ca2+-permeable ion channels and stimulates formation of inositol (1,3,5)-trisphosphate (IP3) that is transmitted through gap junctions, thereby transmitting the ciliary signal to neighboring cells [4,118,119,120]. Loss of polycystin-1 function in vivo leads to reduced formation of osteoblasts, a reduced anabolic response to mechanical loading and the development of osteopenia [121]. Conditional deletion of polycystin-1 under the control of a regulatory element of the Wnt1 promoter has been used in OTM studies [122]. Conditional polycystin-1 deletion blocks the tooth movement with reduced osteoclast formation on the compression side. This study demonstrates that the calcium channels in primary cilia play an important role in the transduction of mechanical signals to induce bone resorption.

3.4. Gap Junctions: Connexins

Connexins are gap junction proteins that connect two neighboring cells [123,124]. Connexin 43 (Cx43) is the most abundant connexin in bone and modulates bone resorption and formation activity by regulating osteoprotegerin and sclerostin levels [74,125]. During OTM, Cx43 is strongly expressed in osteoclasts and PDL cells on the compression side and in osteoblasts and osteocytes on the tension side in vivo [124]. In vitro studies with PDL fibroblasts report that mechanical tension increases Cx43, up-regulating the expression of Runx2 and osterix, and down-regulating RANKL expression [126]. Suppression of Cx43 reduces the induction of osteogenic markers but promotes RANKL expression [126,127]. Given its function in regulating the response of osteoblasts and osteocytes to mechanical forces, it is reasonable to speculate that connexins play a role in OTM.

3.5. Ion Channels

Ion channels are pore-forming membrane proteins that facilitate direct ion passage through the cell membrane [51]. Mechanical force-activated ion channels increase membrane permeability and trigger the influx of extracellular calcium, demonstrating their role in mechanotransduction in osteocytes and PDL fibroblasts [51,128,129]. Piezo1 ion channel and transient receptor potential cation channel subfamily V member 4 (TRPV4) are key factors in the mechanotransduction of osteocytes and PDL fibroblasts under mechanical loading. Conditional deletion of Piezo1 in osteoblasts and osteocytes significantly reduced bone mass and strength in mice [130]. Conversely, administration of a Piezo1 agonist to adult mice increased bone mass in a way that mimicked the effects of mechanical loading, demonstrating that Piezo1 is a mechanosensitive ion channel by which osteoblast lineage cells sense and respond to changes in mechanical load. In vitro mechanical stimulation of mature osteocytes activates Piezo1, which rapidly activates Akt and down-regulates sclerostin [131]. Piezo1 and TRPV4 increase their expression 8 h after mechanical loading, followed by the increased expression of M-CSF, RANKL and COX2 [128]. However, pretreatment with the inhibitors of Piezo1 and TRPV4 suppressed the related cytokine expression. Fluid shear stress on osteocytes activates TRPV4 to rapidly increase intracellular Ca2+ levels, which activates Ca2+/calmodulin-dependent kinase (CaMK) II and down-regulates sclerostin [132,133]. This is functionally important as shown by in vivo and in vitro studies that conditional deletion of Piezo1 in osteoblasts and osteocytes reduces bone mass and strength [130], while administration of a Piezo1 agonist increases bone mass, mimicking the effects of mechanical loading. It is likely that ion channel proteins are important in OTM.

4. Conclusions

Orthodontic tooth movement is a highly coordinated process in which various cells, cytokines, and complex mechanisms are involved. To date, numbers of OTM studies have been performed, but many are in vitro studies or examined the global deletion of a specific gene or cell type. Transgenic animal studies with the cell type-specific gene deletion can provide the insight into the key cellular and molecular mechanisms in OTM by establishing the cause and effect relationships. Findings from those studies could be applied for our daily orthodontic practice in the future, accelerating osteoclastogenesis and reducing treatment time. Conversely, blocking osteoclastogenesis can be applied to prevent orthodontic relapse. In addition, increasing osteogenesis can greatly help the maxillary expansion procedure, reducing the retention period of 5 to 6 months. The RANKL gene transfer to expedite the OTM is just one example. Furthermore, the findings from the transgenic animal studies can contribute to the development of precision orthodontics in the future so that we can provide patient-specific orthodontic treatment.
One of the limitations of this review is that animal studies that specifically examined mechanosensors are rare and many of them were conducted in vitro. Mechanosensors play a critical role in the mechanotransduction process and further investigation is needed. In addition, several OTM studies used slightly different amount of orthodontic force and time points. The use of standardized OTM methods would greatly help compare the outcomes from multiple animal OTM studies. Lastly, applying the findings from rodent studies to humans does warrant some modification considering the species differences, for example when considering the time periods in OTM.

Author Contributions

Conceptualization, H.H.J.; writing—original draft preparation, H.H.J.; writing—review and editing, H.H.J., H.T. and A.T. All authors have read and agreed to the published version of the manuscript

Funding

This research was supported by the Schoenleber Pilot Grant from the University of Pennsylvania School of Dental Medicine (HHJ) and the Orthodontic Faculty Development Fellowship Award from the American Association of Orthodontists Foundation (HHJ).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We greatly appreciate Dana Graves for his helpful discussion and guidance.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Reitan, K. Clinical and histologic observations on tooth movement during and after orthodontic treatment. Am. J. Orthod. 1967, 53, 721–745. [Google Scholar] [CrossRef]
  2. Profit, W.R.; Fields, H.W.; Larson, B.E.; Sarver, D.M. Contemporary Orthodontics, 6th ed.; Mosby: Maryland Heights, MO, USA, 2018. [Google Scholar]
  3. von Bohl, M.; Kuijpers-Jagtman, A.M. Hyalinization during orthodontic tooth movement: A systematic review on tissue reactions. Eur. J. Orthod. 2009, 31, 30–36. [Google Scholar] [CrossRef]
  4. Temiyasathit, S.; Jacobs, C.R. Osteocyte primary cilium and its role in bone mechanotransduction. Ann. N. Y. Acad. Sci. 2010, 1192, 422–428. [Google Scholar] [CrossRef]
  5. Uematsu, S.; Mogi, M.; Deguchi, T. Interleukin (IL)-1 beta, IL-6, tumor necrosis factor-alpha, epidermal growth factor, and beta 2-microglobulin levels are elevated in gingival crevicular fluid during human orthodontic tooth movement. J. Dent. Res. 1996, 75, 562–567. [Google Scholar] [CrossRef] [PubMed]
  6. Ren, Y.; Hazemeijer, H.; de Haan, B.; Qu, N.; de Vos, P. Cytokine profiles in crevicular fluid during orthodontic tooth movement of short and long durations. J. Periodontol. 2007, 78, 453–458. [Google Scholar] [CrossRef]
  7. Garlet, T.P.; Coelho, U.; Silva, J.S.; Garlet, G.P. Cytokine expression pattern in compression and tension sides of the periodontal ligament during orthodontic tooth movement in humans. Eur. J. Oral Sci. 2007, 115, 355–362. [Google Scholar] [CrossRef]
  8. Rubartelli, A.; Lotze, M.T.; Latz, E.; Manfredi, A. Mechanisms of sterile inflammation. Front. Immunol. 2013, 4, 398. [Google Scholar] [CrossRef] [Green Version]
  9. Klein, Y.; Fleissig, O.; Polak, D.; Barenholz, Y.; Mandelboim, O.; Chaushu, S. Immunorthodontics: In vivo gene expression of orthodontic tooth movement. Sci. Rep. 2020, 10, 8172. [Google Scholar] [CrossRef]
  10. Meikle, M.C. The tissue, cellular, and molecular regulation of orthodontic tooth movement: 100 years after Carl Sandstedt. Eur. J. Orthod. 2006, 28, 221–240. [Google Scholar] [CrossRef]
  11. Kang, Y.G.; Nam, J.H.; Kim, K.H.; Lee, K.S. FAK pathway regulates PGE(2) production in compressed periodontal ligament cells. J. Dent. Res. 2010, 89, 1444–1449. [Google Scholar] [CrossRef] [PubMed]
  12. Krishnan, V.; Davidovitch, Z. On a path to unfolding the biological mechanisms of orthodontic tooth movement. J. Dent. Res. 2009, 88, 597–608. [Google Scholar] [CrossRef]
  13. Yamaguchi, M. RANK/RANKL/OPG during orthodontic tooth movement. Orthod. Craniofac. Res. 2009, 12, 113–119. [Google Scholar] [CrossRef] [PubMed]
  14. Glantschnig, H.; Fisher, J.E.; Wesolowski, G.; Rodan, G.A.; Reszka, A.A. M-CSF, TNFalpha and RANK ligand promote osteoclast survival by signaling through mTOR/S6 kinase. Cell Death Differ. 2003, 10, 1165–1177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Garlet, T.P.; Coelho, U.; Repeke, C.E.; Silva, J.S.; Cunha Fde, Q.; Garlet, G.P. Differential expression of osteoblast and osteoclast chemmoatractants in compression and tension sides during orthodontic movement. Cytokine 2008, 42, 330–335. [Google Scholar] [CrossRef] [PubMed]
  16. Cho, A.; Haruyama, N.; Kulkarni, A.B. Generation of transgenic mice. Curr. Protoc. Cell Biol. 2009, 19, 11. [Google Scholar] [CrossRef]
  17. Elefteriou, F.; Yang, X. Genetic mouse models for bone studies-strengths and limitations. Bone 2011, 49, 1242–1254. [Google Scholar] [CrossRef] [Green Version]
  18. Kim, T.; Handa, A.; Iida, J.; Yoshida, S. RANKL expression in rat periodontal ligament subjected to a continuous orthodontic force. Arch. Oral Biol. 2007, 52, 244–250. [Google Scholar] [CrossRef] [PubMed]
  19. Otero, L.; Garcia, D.A.; Wilches-Buitrago, L. Expression and Presence of OPG and RANKL mRNA and Protein in Human Periodontal Ligament with Orthodontic Force. Gene Regul. Syst. Biol. 2016, 10, 15–20. [Google Scholar] [CrossRef] [PubMed]
  20. Yan, Y.; Liu, F.; Kou, X.; Liu, D.; Yang, R.; Wang, X.; Song, Y.; He, D.; Gan, Y.; Zhou, Y. T Cells Are Required for Orthodontic Tooth Movement. J. Dent. Res. 2015, 94, 1463–1470. [Google Scholar] [CrossRef]
  21. Yang, C.Y.; Jeon, H.H.; Alshabab, A.; Lee, Y.J.; Chung, C.H.; Graves, D.T. RANKL deletion in periodontal ligament and bone lining cells blocks orthodontic tooth movement. Int. J. Oral Sci. 2018, 10, 3. [Google Scholar] [CrossRef] [Green Version]
  22. Shoji-Matsunaga, A.; Ono, T.; Hayashi, M.; Takayanagi, H.; Moriyama, K.; Nakashima, T. Osteocyte regulation of orthodontic force-mediated tooth movement via RANKL expression. Sci. Rep. 2017, 7, 8753. [Google Scholar] [CrossRef]
  23. Nishijima, Y.; Yamaguchi, M.; Kojima, T.; Aihara, N.; Nakajima, R.; Kasai, K. Levels of RANKL and OPG in gingival crevicular fluid during orthodontic tooth movement and effect of compression force on releases from periodontal ligament cells in vitro. Orthod. Craniofac. Res. 2006, 9, 63–70. [Google Scholar] [CrossRef]
  24. Ren, Y.; Vissink, A. Cytokines in crevicular fluid and orthodontic tooth movement. Eur. J. Oral Sci. 2008, 116, 89–97. [Google Scholar] [CrossRef]
  25. Meeran, N.A. Biological response at the cellular level within the periodontal ligament on application of orthodontic force—An update. J. Orthod. Sci. 2012, 1, 2–10. [Google Scholar] [CrossRef] [Green Version]
  26. Li, C.; Chung, C.J.; Hwang, C.J.; Lee, K.J. Local injection of RANKL facilitates tooth movement and alveolar bone remodelling. Oral Dis. 2019, 25, 550–560. [Google Scholar] [CrossRef] [PubMed]
  27. Kanzaki, H.; Chiba, M.; Arai, K.; Takahashi, I.; Haruyama, N.; Nishimura, M.; Mitani, H. Local RANKL gene transfer to the periodontal tissue accelerates orthodontic tooth movement. Gene Ther. 2006, 13, 678–685. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Chang, J.H.; Chen, P.J.; Arul, M.R.; Dutra, E.H.; Nanda, R.; Kumbar, S.G.; Yadav, S. Injectable RANKL sustained release formulations to accelerate orthodontic tooth movement. Eur. J. Orthod. 2020, 42, 317–325. [Google Scholar] [CrossRef]
  29. Iglesias-Linares, A.; Moreno-Fernandez, A.M.; Yanez-Vico, R.; Mendoza-Mendoza, A.; Gonzalez-Moles, M.; Solano-Reina, E. The use of gene therapy vs. corticotomy surgery in accelerating orthodontic tooth movement. Orthod. Craniofac. Res. 2011, 14, 138–148. [Google Scholar] [CrossRef]
  30. Li, Y.; Jacox, L.A.; Little, S.H.; Ko, C.C. Orthodontic tooth movement: The biology and clinical implications. Kaohsiung J. Med. Sci. 2018, 34, 207–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Odagaki, N.; Ishihara, Y.; Wang, Z.; Ei Hsu Hlaing, E.; Nakamura, M.; Hoshijima, M.; Hayano, S.; Kawanabe, N.; Kamioka, H. Role of Osteocyte-PDL Crosstalk in Tooth Movement via SOST/Sclerostin. J. Dent. Res. 2018, 97, 1374–1382. [Google Scholar] [CrossRef] [PubMed]
  32. Morse, A.; McDonald, M.M.; Kelly, N.H.; Melville, K.M.; Schindeler, A.; Kramer, I.; Kneissel, M.; van der Meulen, M.C.; Little, D.G. Mechanical load increases in bone formation via a sclerostin-independent pathway. J. Bone Miner. Res. 2014, 29, 2456–2467. [Google Scholar] [CrossRef]
  33. Wijenayaka, A.R.; Kogawa, M.; Lim, H.P.; Bonewald, L.F.; Findlay, D.M.; Atkins, G.J. Sclerostin stimulates osteocyte support of osteoclast activity by a RANKL-dependent pathway. PLoS ONE 2011, 6, e25900. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Galea, G.L.; Lanyon, L.E.; Price, J.S. Sclerostin’s role in bone’s adaptive response to mechanical loading. Bone 2017, 96, 38–44. [Google Scholar] [CrossRef] [Green Version]
  35. Shu, R.; Bai, D.; Sheu, T.; He, Y.; Yang, X.; Xue, C.; He, Y.; Zhao, M.; Han, X. Sclerostin Promotes Bone Remodeling in the Process of Tooth Movement. PLoS ONE 2017, 12, e0167312. [Google Scholar] [CrossRef]
  36. Lu, W.; Zhang, X.; Firth, F.; Mei, L.; Yi, J.; Gong, C.; Li, H.; Zheng, W.; Li, Y. Sclerostin injection enhances orthodontic tooth movement in rats. Arch. Oral Biol. 2019, 99, 43–50. [Google Scholar] [CrossRef] [PubMed]
  37. Robling, A.G.; Niziolek, P.J.; Baldridge, L.A.; Condon, K.W.; Allen, M.R.; Alam, I.; Mantila, S.M.; Gluhak-Heinrich, J.; Bellido, T.M.; Harris, S.E.; et al. Mechanical stimulation of bone in vivo reduces osteocyte expression of Sost/sclerostin. J. Biol. Chem. 2008, 283, 5866–5875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Kamiya, N.; Mishina, Y. New insights on the roles of BMP signaling in bone-A review of recent mouse genetic studies. Biofactors 2011, 37, 75–82. [Google Scholar] [CrossRef] [Green Version]
  39. Gao, Y.; Zhang, M.; Tian, X.; Wang, M.; Zhang, F. Experimental animal study on BMP-3 expression in periodontal tissues in the process of orthodontic tooth movement. Exp. Ther. Med. 2019, 17, 193–198. [Google Scholar] [CrossRef]
  40. Iglesias-Linares, A.; Yanez-Vico, R.M.; Moreno-Fernandez, A.M.; Mendoza-Mendoza, A.; Solano-Reina, E. Corticotomy-assisted orthodontic enhancement by bone morphogenetic protein-2 administration. J. Oral Maxillofac. Surg. 2012, 70, e124–e132. [Google Scholar] [CrossRef] [PubMed]
  41. Manokawinchoke, J.; Pavasant, P.; Sawangmake, C.; Limjeerajarus, N.; Limjeerajarus, C.N.; Egusa, H.; Osathanon, T. Intermittent compressive force promotes osteogenic differentiation in human periodontal ligament cells by regulating the transforming growth factor-beta pathway. Cell Death Dis. 2019, 10, 761. [Google Scholar] [CrossRef]
  42. Quinn, J.M.; Itoh, K.; Udagawa, N.; Hausler, K.; Yasuda, H.; Shima, N.; Mizuno, A.; Higashio, K.; Takahashi, N.; Suda, T.; et al. Transforming growth factor beta affects osteoclast differentiation via direct and indirect actions. J. Bone Miner. Res. 2001, 16, 1787–1794. [Google Scholar] [CrossRef]
  43. Itonaga, I.; Sabokbar, A.; Sun, S.G.; Kudo, O.; Danks, L.; Ferguson, D.; Fujikawa, Y.; Athanasou, N.A. Transforming growth factor-beta induces osteoclast formation in the absence of RANKL. Bone 2004, 34, 57–64. [Google Scholar] [CrossRef]
  44. Wang, L.L.; Zhu, H.; Liang, T. Changes of transforming growth factor beta 1 in rat periodontal tissue during orthodontic tooth movement. Chin. J. Dent. Res. 2000, 3, 19–22. [Google Scholar]
  45. Uematsu, S.; Mogi, M.; Deguchi, T. Increase of transforming growth factor-beta 1 in gingival crevicular fluid during human orthodontic tooth movement. Arch. Oral Biol. 1996, 41, 1091–1095. [Google Scholar] [CrossRef]
  46. Nagai, M.; Yoshida, A.; Sato, N.; Wong, D.T. Messenger RNA level and protein localization of transforming growth factor-beta1 in experimental tooth movement in rats. Eur. J. Oral Sci. 1999, 107, 475–481. [Google Scholar] [CrossRef] [PubMed]
  47. Van Schepdael, A.; Vander Sloten, J.; Geris, L. A mechanobiological model of orthodontic tooth movement. Biomech. Model. Mechanobiol. 2013, 12, 249–265. [Google Scholar] [CrossRef] [PubMed]
  48. Masella, R.S.; Meister, M. Current concepts in the biology of orthodontic tooth movement. Am. J. Orthod. Dentofac. Orthop. 2006, 129, 458–468. [Google Scholar] [CrossRef]
  49. Qin, L.; Liu, W.; Cao, H.; Xiao, G. Molecular mechanosensors in osteocytes. Bone Res. 2020, 8, 23. [Google Scholar] [CrossRef]
  50. Huang, H.; Yang, R.; Zhou, Y.H. Mechanobiology of Periodontal Ligament Stem Cells in Orthodontic Tooth Movement. Stem Cells Int. 2018, 2018, 6531216. [Google Scholar] [CrossRef] [Green Version]
  51. Klein-Nulend, J.; Bakker, A.D.; Bacabac, R.G.; Vatsa, A.; Weinbaum, S. Mechanosensation and transduction in osteocytes. Bone 2013, 54, 182–190. [Google Scholar] [CrossRef]
  52. Mabuchi, R.; Matsuzaka, K.; Shimono, M. Cell proliferation and cell death in periodontal ligaments during orthodontic tooth movement. J. Periodontal Res. 2002, 37, 118–124. [Google Scholar] [CrossRef] [PubMed]
  53. Seo, B.M.; Miura, M.; Gronthos, S.; Bartold, P.M.; Batouli, S.; Brahim, J.; Young, M.; Robey, P.G.; Wang, C.Y.; Shi, S. Investigation of multipotent postnatal stem cells from human periodontal ligament. Lancet 2004, 364, 149–155. [Google Scholar] [CrossRef]
  54. Zheng, J.; Chen, S.; Albiero, M.L.; Vieira, G.H.A.; Wang, J.; Feng, J.Q.; Graves, D.T. Diabetes Activates Periodontal Ligament Fibroblasts via NF-kappaB In Vivo. J. Dent. Res. 2018, 97, 580–588. [Google Scholar] [CrossRef]
  55. Dacquin, R.; Starbuck, M.; Schinke, T.; Karsenty, G. Mouse alpha1(I)-collagen promoter is the best known promoter to drive efficient Cre recombinase expression in osteoblast. Dev. Dyn. 2002, 224, 245–251. [Google Scholar] [CrossRef]
  56. Giannopoulou, C.; Cimasoni, G. Functional characteristics of gingival and periodontal ligament fibroblasts. J. Dent. Res. 1996, 75, 895–902. [Google Scholar] [CrossRef]
  57. Basdra, E.K.; Komposch, G. Osteoblast-like properties of human periodontal ligament cells: An in vitro analysis. Eur. J. Orthod. 1997, 19, 615–621. [Google Scholar] [CrossRef]
  58. Hyeran Helen Jeon, C.-Y.Y.; Shin, M.K.; Wang, J.; Patel, J.H.; Chung, C.; Graves, D.T. Osteoblast lineage cells and periodontal ligament fibroblasts regulate orthodontic tooth movement that is dependent on Nuclear Factor-kappa B (NF-kB) activation. Angle Orthod. 2021, in press. [Google Scholar]
  59. Men, Y.; Wang, Y.; Yi, Y.; Jing, D.; Luo, W.; Shen, B.; Stenberg, W.; Chai, Y.; Ge, W.P.; Feng, J.Q.; et al. Gli1+ Periodontium Stem Cells Are Regulated by Osteocytes and Occlusal Force. Dev. Cell 2020, 54, 639–654 e636. [Google Scholar] [CrossRef] [PubMed]
  60. Liu, A.Q.; Zhang, L.S.; Chen, J.; Sui, B.D.; Liu, J.; Zhai, Q.M.; Li, Y.J.; Bai, M.; Chen, K.; Jin, Y.; et al. Mechanosensing by Gli1(+) cells contributes to the orthodontic force-induced bone remodelling. Cell Prolif. 2020, 53, e12810. [Google Scholar] [CrossRef] [Green Version]
  61. Dupont, S.; Morsut, L.; Aragona, M.; Enzo, E.; Giulitti, S.; Cordenonsi, M.; Zanconato, F.; Le Digabel, J.; Forcato, M.; Bicciato, S.; et al. Role of YAP/TAZ in mechanotransduction. Nature 2011, 474, 179–183. [Google Scholar] [CrossRef] [PubMed]
  62. Yang, Y.; Wang, B.K.; Chang, M.L.; Wan, Z.Q.; Han, G.L. Cyclic Stretch Enhances Osteogenic Differentiation of Human Periodontal Ligament Cells via YAP Activation. Biomed. Res. Int. 2018, 2018, 2174824. [Google Scholar] [CrossRef]
  63. Sun, B.; Wen, Y.; Wu, X.; Zhang, Y.; Qiao, X.; Xu, X. Expression pattern of YAP and TAZ during orthodontic tooth movement in rats. J. Mol. Histol. 2018, 49, 123–131. [Google Scholar] [CrossRef] [PubMed]
  64. Zhang, L.; Liu, W.; Zhao, J.; Ma, X.; Shen, L.; Zhang, Y.; Jin, F.; Jin, Y. Mechanical stress regulates osteogenic differentiation and RANKL/OPG ratio in periodontal ligament stem cells by the Wnt/beta-catenin pathway. Biochim. Biophys. Acta 2016, 1860, 2211–2219. [Google Scholar] [CrossRef]
  65. Pelaez, D.; Acosta Torres, Z.; Ng, T.K.; Choy, K.W.; Pang, C.P.; Cheung, H.S. Cardiomyogenesis of periodontal ligament-derived stem cells by dynamic tensile strain. Cell Tissue Res. 2017, 367, 229–241. [Google Scholar] [CrossRef]
  66. Chen, J.; Zhang, W.; Backman, L.J.; Kelk, P.; Danielson, P. Mechanical stress potentiates the differentiation of periodontal ligament stem cells into keratocytes. Br. J. Ophthalmol. 2018, 102, 562–569. [Google Scholar] [CrossRef]
  67. Liu, F.; Wen, F.; He, D.; Liu, D.; Yang, R.; Wang, X.; Yan, Y.; Liu, Y.; Kou, X.; Zhou, Y. Force-Induced H2S by PDLSCs Modifies Osteoclastic Activity during Tooth Movement. J. Dent. Res. 2017, 96, 694–702. [Google Scholar] [CrossRef] [PubMed]
  68. He, D.; Liu, F.; Cui, S.; Jiang, N.; Yu, H.; Zhou, Y.; Liu, Y.; Kou, X. Mechanical load-induced H2S production by periodontal ligament stem cells activates M1 macrophages to promote bone remodeling and tooth movement via STAT1. Stem Cell Res. Ther. 2020, 11, 112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Huang, H.; Williams, R.C.; Kyrkanides, S. Accelerated orthodontic tooth movement: Molecular mechanisms. Am. J. Orthod. Dentofac. Orthop. 2014, 146, 620–632. [Google Scholar] [CrossRef] [PubMed]
  70. Florencio-Silva, R.; Sasso, G.R.; Sasso-Cerri, E.; Simoes, M.J.; Cerri, P.S. Biology of Bone Tissue: Structure, Function, and Factors That Influence Bone Cells. Biomed. Res. Int. 2015, 2015, 421746. [Google Scholar] [CrossRef] [Green Version]
  71. Dallas, S.L.; Prideaux, M.; Bonewald, L.F. The osteocyte: An endocrine cell … and more. Endocr. Rev. 2013, 34, 658–690. [Google Scholar] [CrossRef] [Green Version]
  72. Gluhak-Heinrich, J.; Ye, L.; Bonewald, L.F.; Feng, J.Q.; MacDougall, M.; Harris, S.E.; Pavlin, D. Mechanical loading stimulates dentin matrix protein 1 (DMP1) expression in osteocytes in vivo. J. Bone Miner. Res. 2003, 18, 807–817. [Google Scholar] [CrossRef]
  73. Moin, S.; Kalajzic, Z.; Utreja, A.; Nihara, J.; Wadhwa, S.; Uribe, F.; Nanda, R. Osteocyte death during orthodontic tooth movement in mice. Angle Orthod. 2014, 84, 1086–1092. [Google Scholar] [CrossRef]
  74. Bivi, N.; Condon, K.W.; Allen, M.R.; Farlow, N.; Passeri, G.; Brun, L.R.; Rhee, Y.; Bellido, T.; Plotkin, L.I. Cell autonomous requirement of connexin 43 for osteocyte survival: Consequences for endocortical resorption and periosteal bone formation. J. Bone Miner. Res. 2012, 27, 374–389. [Google Scholar] [CrossRef]
  75. Kogianni, G.; Mann, V.; Noble, B.S. Apoptotic bodies convey activity capable of initiating osteoclastogenesis and localized bone destruction. J. Bone Miner. Res. 2008, 23, 915–927. [Google Scholar] [CrossRef]
  76. Xiong, J.; Onal, M.; Jilka, R.L.; Weinstein, R.S.; Manolagas, S.C.; O’Brien, C.A. Matrix-embedded cells control osteoclast formation. Nat. Med. 2011, 17, 1235–1241. [Google Scholar] [CrossRef] [Green Version]
  77. Nakashima, T.; Hayashi, M.; Fukunaga, T.; Kurata, K.; Oh-Hora, M.; Feng, J.Q.; Bonewald, L.F.; Kodama, T.; Wutz, A.; Wagner, E.F.; et al. Evidence for osteocyte regulation of bone homeostasis through RANKL expression. Nat. Med. 2011, 17, 1231–1234. [Google Scholar] [CrossRef]
  78. Matsumoto, T.; Iimura, T.; Ogura, K.; Moriyama, K.; Yamaguchi, A. The role of osteocytes in bone resorption during orthodontic tooth movement. J. Dent. Res. 2013, 92, 340–345. [Google Scholar] [CrossRef]
  79. Tatsumi, S.; Ishii, K.; Amizuka, N.; Li, M.; Kobayashi, T.; Kohno, K.; Ito, M.; Takeshita, S.; Ikeda, K. Targeted ablation of osteocytes induces osteoporosis with defective mechanotransduction. Cell Metab. 2007, 5, 464–475. [Google Scholar] [CrossRef] [Green Version]
  80. Kassem, M.; Abdallah, B.M.; Saeed, H. Osteoblastic cells: Differentiation and trans-differentiation. Arch. Biochem. Biophys. 2008, 473, 183–187. [Google Scholar] [CrossRef]
  81. Matic, I.; Matthews, B.G.; Wang, X.; Dyment, N.A.; Worthley, D.L.; Rowe, D.W.; Grcevic, D.; Kalajzic, I. Quiescent Bone Lining Cells Are a Major Source of Osteoblasts During Adulthood. Stem Cells 2016, 34, 2930–2942. [Google Scholar] [CrossRef] [Green Version]
  82. Kohli, S.S.; Kohli, V.S. Role of RANKL-RANK/osteoprotegerin molecular complex in bone remodeling and its immunopathologic implications. Indian J. Endocrinol. Metab. 2011, 15, 175–181. [Google Scholar] [CrossRef] [PubMed]
  83. Wang, L.; Li, J.Y.; Zhang, X.Z.; Liu, L.; Wan, Z.M.; Li, R.X.; Guo, Y. Involvement of p38MAPK/NF-kappaB signaling pathways in osteoblasts differentiation in response to mechanical stretch. Ann. Biomed. Eng. 2012, 40, 1884–1894. [Google Scholar] [CrossRef]
  84. Matsuo, K.; Irie, N. Osteoclast-osteoblast communication. Arch. Biochem. Biophys. 2008, 473, 201–209. [Google Scholar] [CrossRef]
  85. Zhang, S.; Wang, X.; Li, G.; Chong, Y.; Zhang, J.; Guo, X.; Li, B.; Bi, Z. Osteoclast regulation of osteoblasts via RANKRANKL reverse signal transduction in vitro. Mol. Med. Rep. 2017, 16, 3994–4000. [Google Scholar] [CrossRef] [PubMed]
  86. Boyce, B.F.; Xing, L. The RANKL/RANK/OPG pathway. Curr. Osteoporos. Rep. 2007, 5, 98–104. [Google Scholar] [CrossRef]
  87. Holland, R.; Bain, C.; Utreja, A. Osteoblast differentiation during orthodontic tooth movement. Orthod. Craniofac. Res. 2019, 22, 177–182. [Google Scholar] [CrossRef]
  88. Uribe, F.; Kalajzic, Z.; Bibko, J.; Nanda, R.; Olson, C.; Rowe, D.; Wadhwa, S. Early effects of orthodontic forces on osteoblast differentiation in a novel mouse organ culture model. Angle Orthod. 2011, 81, 284–291. [Google Scholar] [CrossRef] [PubMed]
  89. Ibrahimi Disha, S.; Furlani, B.; Drevensek, G.; Plut, A.; Yanagisawa, M.; Hudoklin, S.; Prodan Zitnik, I.; Marc, J.; Drevensek, M. The role of endothelin B receptor in bone modelling during orthodontic tooth movement: A study on ETB knockout rats. Sci. Rep. 2020, 10, 14226. [Google Scholar] [CrossRef] [PubMed]
  90. Premaraj, S.; Souza, I.; Premaraj, T. Mechanical loading activates beta-catenin signaling in periodontal ligament cells. Angle Orthod. 2011, 81, 592–599. [Google Scholar] [CrossRef] [PubMed]
  91. Kramer, I.; Halleux, C.; Keller, H.; Pegurri, M.; Gooi, J.H.; Weber, P.B.; Feng, J.Q.; Bonewald, L.F.; Kneissel, M. Osteocyte Wnt/beta-catenin signaling is required for normal bone homeostasis. Mol. Cell. Biol. 2010, 30, 3071–3085. [Google Scholar] [CrossRef] [Green Version]
  92. Chen, B.; Li, X.D.; Liu, D.X.; Wang, H.; Xie, P.; Liu, Z.Y.; Hou, G.Q.; Chang, B.; Du, S.X. Canonical Wnt signaling is required for Panax notoginseng saponin-mediated attenuation of the RANKL/OPG ratio in bone marrow stromal cells during osteogenic differentiation. Phytomedicine 2012, 19, 1029–1034. [Google Scholar] [CrossRef]
  93. Fu, H.D.; Wang, B.K.; Wan, Z.Q.; Lin, H.; Chang, M.L.; Han, G.L. Wnt5a mediated canonical Wnt signaling pathway activation in orthodontic tooth movement: Possible role in the tension force-induced bone formation. J. Mol. Histol. 2016, 47, 455–466. [Google Scholar] [CrossRef] [PubMed]
  94. Sawakami, K.; Robling, A.G.; Ai, M.; Pitner, N.D.; Liu, D.; Warden, S.J.; Li, J.; Maye, P.; Rowe, D.W.; Duncan, R.L.; et al. The Wnt co-receptor LRP5 is essential for skeletal mechanotransduction but not for the anabolic bone response to parathyroid hormone treatment. J. Biol. Chem. 2006, 281, 23698–23711. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Cui, Y.; Niziolek, P.J.; MacDonald, B.T.; Zylstra, C.R.; Alenina, N.; Robinson, D.R.; Zhong, Z.; Matthes, S.; Jacobsen, C.M.; Conlon, R.A.; et al. Lrp5 functions in bone to regulate bone mass. Nat. Med. 2011, 17, 684–691. [Google Scholar] [CrossRef]
  96. Niziolek, P.J.; Warman, M.L.; Robling, A.G. Mechanotransduction in bone tissue: The A214V and G171V mutations in Lrp5 enhance load-induced osteogenesis in a surface-selective manner. Bone 2012, 51, 459–465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Holland, R.; Bain, C.; Alrasheed, R.S.; Robling, A.G.; Utreja, A. The effect of overexpression of Lrp5 on orthodontic tooth movement. Orthod. Craniofac. Res. 2020. [Google Scholar] [CrossRef] [PubMed]
  98. Lim, W.H.; Liu, B.; Mah, S.J.; Yin, X.; Helms, J.A. Alveolar bone turnover and periodontal ligament width are controlled by Wnt. J. Periodontol. 2015, 86, 319–326. [Google Scholar] [CrossRef] [PubMed]
  99. Lu, J.; Duan, Y.; Zhang, M.; Wu, M.; Wang, Y. Expression of Wnt3a, Wnt10b, beta-catenin and DKK1 in periodontium during orthodontic tooth movement in rats. Acta Odontol. Scand. 2016, 74, 217–223. [Google Scholar] [CrossRef] [PubMed]
  100. Pocaterra, A.; Romani, P.; Dupont, S. YAP/TAZ functions and their regulation at a glance. J. Cell Sci. 2020, 133, jcs230425. [Google Scholar] [CrossRef]
  101. Huelter-Hassler, D.; Tomakidi, P.; Steinberg, T.; Jung, B.A. Orthodontic strain affects the Hippo-pathway effector YAP concomitant with proliferation in human periodontal ligament fibroblasts. Eur. J. Orthod. 2017, 39, 251–257. [Google Scholar] [CrossRef]
  102. Klein-Nulend, J.; Bacabac, R.G.; Bakker, A.D. Mechanical loading and how it affects bone cells: The role of the osteocyte cytoskeleton in maintaining our skeleton. Eur. Cells Mater. 2012, 24, 278–291. [Google Scholar] [CrossRef]
  103. Feller, L.; Khammissa, R.A.; Schechter, I.; Thomadakis, G.; Fourie, J.; Lemmer, J. Biological Events in Periodontal Ligament and Alveolar Bone Associated with Application of Orthodontic Forces. Sci. World J. 2015, 2015, 876509. [Google Scholar] [CrossRef] [Green Version]
  104. Feller, L.; Khammissa, R.A.; Schechter, I.; Moodley, A.; Thomadakis, G.; Lemmer, J. Periodontal Biological Events Associated with Orthodontic Tooth Movement: The Biomechanics of the Cytoskeleton and the Extracellular Matrix. Sci. World J. 2015, 2015, 894123. [Google Scholar] [CrossRef] [Green Version]
  105. McBeath, R.; Pirone, D.M.; Nelson, C.M.; Bhadriraju, K.; Chen, C.S. Cell shape, cytoskeletal tension, and RhoA regulate stem cell lineage commitment. Dev. Cell 2004, 6, 483–495. [Google Scholar] [CrossRef] [Green Version]
  106. Pan, J.; Wang, T.; Wang, L.; Chen, W.; Song, M. Cyclic strain-induced cytoskeletal rearrangement of human periodontal ligament cells via the Rho signaling pathway. PLoS ONE 2014, 9, e91580. [Google Scholar] [CrossRef] [PubMed]
  107. Geiger, B.; Bershadsky, A.; Pankov, R.; Yamada, K.M. Transmembrane crosstalk between the extracellular matrix--cytoskeleton crosstalk. Nat. Rev. Mol. Cell Biol. 2001, 2, 793–805. [Google Scholar] [CrossRef]
  108. Young, S.R.; Gerard-O’Riley, R.; Kim, J.B.; Pavalko, F.M. Focal adhesion kinase is important for fluid shear stress-induced mechanotransduction in osteoblasts. J. Bone Miner. Res. 2009, 24, 411–424. [Google Scholar] [CrossRef] [PubMed]
  109. Santos, A.; Bakker, A.D.; Zandieh-Doulabi, B.; de Blieck-Hogervorst, J.M.; Klein-Nulend, J. Early activation of the beta-catenin pathway in osteocytes is mediated by nitric oxide, phosphatidyl inositol-3 kinase/Akt, and focal adhesion kinase. Biochem. Biophys. Res. Commun. 2010, 391, 364–369. [Google Scholar] [CrossRef]
  110. Chen, C.S.; Tan, J.; Tien, J. Mechanotransduction at cell-matrix and cell-cell contacts. Annu. Rev. Biomed. Eng. 2004, 6, 275–302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Yuan, X.; Yang, S. Cilia/Ift protein and motor -related bone diseases and mouse models. Front. Biosci. 2015, 20, 515–555. [Google Scholar] [CrossRef] [Green Version]
  112. Song, J.; Wang, L.; Fan, F.; Wei, J.; Zhang, J.; Lu, Y.; Fu, Y.; Wang, S.; Juncos, L.A.; Liu, R. Role of the Primary Cilia on the Macula Densa and Thick Ascending Limbs in Regulation of Sodium Excretion and Hemodynamics. Hypertension 2017, 70, 324–333. [Google Scholar] [CrossRef] [PubMed]
  113. Whitfield, J.F. Primary cilium--is it an osteocyte’s strain-sensing flowmeter? J. Cell Biochem. 2003, 89, 233–237. [Google Scholar] [CrossRef]
  114. Yuan, X.; Yang, S. Primary Cilia and Intraflagellar Transport Proteins in Bone and Cartilage. J. Dent. Res. 2016, 95, 1341–1349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Yang, S.; Wang, C. The intraflagellar transport protein IFT80 is required for cilia formation and osteogenesis. Bone 2012, 51, 407–417. [Google Scholar] [CrossRef] [Green Version]
  116. Qiu, N.; Xiao, Z.; Cao, L.; Buechel, M.M.; David, V.; Roan, E.; Quarles, L.D. Disruption of Kif3a in osteoblasts results in defective bone formation and osteopenia. J. Cell Sci. 2012, 125, 1945–1957. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Nauli, S.M.; Alenghat, F.J.; Luo, Y.; Williams, E.; Vassilev, P.; Li, X.; Elia, A.E.; Lu, W.; Brown, E.M.; Quinn, S.J.; et al. Polycystins 1 and 2 mediate mechanosensation in the primary cilium of kidney cells. Nat. Genet. 2003, 33, 129–137. [Google Scholar] [CrossRef]
  118. Wang, P.; Tang, C.; Wu, J.; Yang, Y.; Yan, Z.; Liu, X.; Shao, X.; Zhai, M.; Gao, J.; Liang, S.; et al. Pulsed electromagnetic fields regulate osteocyte apoptosis, RANKL/OPG expression, and its control of osteoclastogenesis depending on the presence of primary cilia. J. Cell Physiol. 2019, 234, 10588–10601. [Google Scholar] [CrossRef] [PubMed]
  119. Delaine-Smith, R.M.; Sittichokechaiwut, A.; Reilly, G.C. Primary cilia respond to fluid shear stress and mediate flow-induced calcium deposition in osteoblasts. FASEB J. 2014, 28, 430–439. [Google Scholar] [CrossRef] [Green Version]
  120. Malone, A.M.; Anderson, C.T.; Tummala, P.; Kwon, R.Y.; Johnston, T.R.; Stearns, T.; Jacobs, C.R. Primary cilia mediate mechanosensing in bone cells by a calcium-independent mechanism. Proc. Natl. Acad. Sci. USA 2007, 104, 13325–13330. [Google Scholar] [CrossRef] [Green Version]
  121. Xiao, Z.; Zhang, S.; Mahlios, J.; Zhou, G.; Magenheimer, B.S.; Guo, D.; Dallas, S.L.; Maser, R.; Calvet, J.P.; Bonewald, L.; et al. Cilia-like structures and polycystin-1 in osteoblasts/osteocytes and associated abnormalities in skeletogenesis and Runx2 expression. J. Biol. Chem. 2006, 281, 30884–30895. [Google Scholar] [CrossRef] [Green Version]
  122. Shalish, M.; Will, L.A.; Fukai, N.; Hou, B.; Olsen, B.R. Role of polycystin-1 in bone remodeling: Orthodontic tooth movement study in mutant mice. Angle Orthod. 2014, 84, 885–890. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Riquelme, M.A.; Cardenas, E.R.; Xu, H.; Jiang, J.X. The Role of Connexin Channels in the Response of Mechanical Loading and Unloading of Bone. Int. J. Mol. Sci. 2020, 21, 1146. [Google Scholar] [CrossRef] [Green Version]
  124. Su, M.; Borke, J.L.; Donahue, H.J.; Li, Z.; Warshawsky, N.M.; Russell, C.M.; Lewis, J.E. Expression of connexin 43 in rat mandibular bone and periodontal ligament (PDL) cells during experimental tooth movement. J. Dent. Res. 1997, 76, 1357–1366. [Google Scholar] [CrossRef] [PubMed]
  125. Plotkin, L.I.; Stains, J.P. Connexins and pannexins in the skeleton: Gap junctions, hemichannels and more. Cell. Mol. Life Sci. 2015, 72, 2853–2867. [Google Scholar] [CrossRef] [Green Version]
  126. Li, S.; Zhang, H.; Li, S.; Yang, Y.; Huo, B.; Zhang, D. Connexin 43 and ERK regulate tension-induced signal transduction in human periodontal ligament fibroblasts. J. Orthop. Res. 2015, 33, 1008–1014. [Google Scholar] [CrossRef] [Green Version]
  127. Xu, C.; Fan, Z.; Shan, W.; Hao, Y.; Ma, J.; Huang, Q.; Zhang, F. Cyclic stretch influenced expression of membrane connexin 43 in human periodontal ligament cell. Arch. Oral Biol. 2012, 57, 1602–1608. [Google Scholar] [CrossRef]
  128. Shen, Y.; Pan, Y.; Guo, S.; Sun, L.; Zhang, C.; Wang, L. The roles of mechanosensitive ion channels and associated downstream MAPK signaling pathways in PDLC mechanotransduction. Mol. Med. Rep. 2020, 21, 2113–2122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Jin, Y.; Li, J.; Wang, Y.; Ye, R.; Feng, X.; Jing, Z.; Zhao, Z. Functional role of mechanosensitive ion channel Piezo1 in human periodontal ligament cells. Angle Orthod. 2015, 85, 87–94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Li, X.; Han, L.; Nookaew, I.; Mannen, E.; Silva, M.J.; Almeida, M.; Xiong, J. Stimulation of Piezo1 by mechanical signals promotes bone anabolism. Elife 2019, 8. [Google Scholar] [CrossRef]
  131. Sasaki, F.; Hayashi, M.; Mouri, Y.; Nakamura, S.; Adachi, T.; Nakashima, T. Mechanotransduction via the Piezo1-Akt pathway underlies Sost suppression in osteocytes. Biochem. Biophys. Res. Commun. 2020, 521, 806–813. [Google Scholar] [CrossRef]
  132. Pei, F.; Liu, J.; Zhang, L.; Pan, X.; Huang, W.; Cen, X.; Huang, S.; Jin, Y.; Zhao, Z. The functions of mechanosensitive ion channels in tooth and bone tissues. Cell. Signal. 2021, 78, 109877. [Google Scholar] [CrossRef] [PubMed]
  133. Lyons, J.S.; Joca, H.C.; Law, R.A.; Williams, K.M.; Kerr, J.P.; Shi, G.; Khairallah, R.J.; Martin, S.S.; Konstantopoulos, K.; Ward, C.W.; et al. Microtubules tune mechanotransduction through NOX2 and TRPV4 to decrease sclerostin abundance in osteocytes. Sci. Signal. 2017, 10, eaan5748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Cytokines and Mechanosensory cells in OTM.
Figure 1. Cytokines and Mechanosensory cells in OTM.
Jcm 10 01733 g001
Figure 2. Possible Mechanosensors in Orthodontic Tooth Movement. (1) Cytoskeletons, (2) Focal adhesions: integrins, (3) Ion channels, (4) Primary cilia, and (5) Gap Junctions: connexins. TRPV4: transient receptor potential cation channel subfamily V member 4.
Figure 2. Possible Mechanosensors in Orthodontic Tooth Movement. (1) Cytoskeletons, (2) Focal adhesions: integrins, (3) Ion channels, (4) Primary cilia, and (5) Gap Junctions: connexins. TRPV4: transient receptor potential cation channel subfamily V member 4.
Jcm 10 01733 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jeon, H.H.; Teixeira, H.; Tsai, A. Mechanistic Insight into Orthodontic Tooth Movement Based on Animal Studies: A Critical Review. J. Clin. Med. 2021, 10, 1733. https://doi.org/10.3390/jcm10081733

AMA Style

Jeon HH, Teixeira H, Tsai A. Mechanistic Insight into Orthodontic Tooth Movement Based on Animal Studies: A Critical Review. Journal of Clinical Medicine. 2021; 10(8):1733. https://doi.org/10.3390/jcm10081733

Chicago/Turabian Style

Jeon, Hyeran Helen, Hellen Teixeira, and Andrew Tsai. 2021. "Mechanistic Insight into Orthodontic Tooth Movement Based on Animal Studies: A Critical Review" Journal of Clinical Medicine 10, no. 8: 1733. https://doi.org/10.3390/jcm10081733

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop