Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 2610 word(s) 2610 2021-06-09 10:01:36 |
2 update references and layout + 24 word(s) 2634 2021-06-10 03:13:40 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Steinberg, T. Oral Periodontal Tissues. Encyclopedia. Available online: https://encyclopedia.pub/entry/10683 (accessed on 29 March 2024).
Steinberg T. Oral Periodontal Tissues. Encyclopedia. Available at: https://encyclopedia.pub/entry/10683. Accessed March 29, 2024.
Steinberg, Thorsten. "Oral Periodontal Tissues" Encyclopedia, https://encyclopedia.pub/entry/10683 (accessed March 29, 2024).
Steinberg, T. (2021, June 09). Oral Periodontal Tissues. In Encyclopedia. https://encyclopedia.pub/entry/10683
Steinberg, Thorsten. "Oral Periodontal Tissues." Encyclopedia. Web. 09 June, 2021.
Oral Periodontal Tissues
Edit

The periodontium is a complex composite tissue consisting of the gingiva, the periodontal ligament (PDL), cementum, and the alveolar bone. Beneath harboring various cell types, amongst others, gingival keratinocytes (GKs), gingival fibroblasts (GFs), periodontal ligament fibroblasts (PDLFs), cementoblasts and osteoblasts, the ECMs of the different periodontal cell and tissue entities are considerably different .

mechanotransduction (MT) nuclear mechanotransduction (NMT) YAP/TAZ extracellular matrix (ECM) gingipain proteases periodontitis oral squamous cell carcinoma (OSCC) regeneration

1. The ECM, Focal Adhesions and Adherens Junctions in Periodontal Health and Disease

The periodontium is a complex composite tissue consisting of the gingiva, the periodontal ligament (PDL), cementum, and the alveolar bone. Beneath harboring various cell types, amongst others, gingival keratinocytes (GKs), gingival fibroblasts (GFs), periodontal ligament fibroblasts (PDLFs), cementoblasts and osteoblasts, the ECMs of the different periodontal cell and tissue entities are considerably different [1].
The cementum and the alveolar bone consist of an inorganic hydroxyapatite matrix and collagen type I as the major organic compound [2]. Additionally, the glycosaminoglycans (GAGs) decorin and biglycan as well as the glycoproteins (GPs) osteonectin, osteopontin, fibronectin, and osteocalcin can be found in the ECM of both tissues [3][4]. The PDL harbors the fibrous collagen types I and III, and small amounts of collagen V and VI [5]. Above, the basement membrane collagen types IV and VII and collagen type XII, which is important for fibrillar organization, are expressed in the PDL [6]. Elastin, fibronectin (FN) and chondroitin-/dermatan-/keratin-sulfate, containing GAGs, also support PDL function [7].
The gingiva contributes to periodontal integrity via epithelium and connective tissue. While the GKs form a stratified squamous epithelial layer, GFs are embedded in a lamina propria with collagen type I, III, elastin and many other macromolecules [8].
Physical ECM properties are key determinants of cell behavior in vitro and in vivo. The stiffness of the ECM, quantified by the Young’s modulus, as well as viscoelastic properties, spatial arrangement of adhesion points, and other geometric constraints, influence cellular responses through MT and other signaling hubs [9][10][11]. This means that ECM composition, homeostasis and MT are tightly coupled and are, therefore, highly interdependent [12]. Cell morphology, migration, proliferation, differentiation, and apoptosis are consequently not only influenced by biochemical signals, but also by the direct mechanical properties of the respective ECM environment [13]. Interestingly, cellular responses, namely the actomyosin-derived cell-inherent contraction forces, as discussed below, seem to directly reflect ECM stiffness, meaning that the Young’s modulus of the ECM is encoded within the cell’s response to that specific microenvironment [14].
Mechanistically, the above-mentioned ECM constituents directly or indirectly, i.e., mediated through adaptor proteins, interact with neighboring cells via surface receptors. Regarding MT, the family of integrin proteins is especially important as they are the core linking hub between the ECM and the cytosol. Within the plasma membrane, integrins form heterodimers, which are composed of an α- and a β-subunit. Various combinations of heterodimers have been described in different experimental systems and they proved to have different ligand specificity. Table 1 summarizes the most important integrin heterodimers and their corresponding ligands relevant to periodontal MT [15][16][17][18][19][20][21]. As can be seen, one heterodimer can sometimes bind more than one ligand (e.g., α1β1-integrin) and one ligand, such as fibronectin, is recognized by multiple heterodimers. Therefore, the tissue- or cell-specific expression pattern of integrins determines its interaction with the ECM. Periodontal cell populations foremostly harbor α2β1, α3β1 and α5β1 integrins. Of interest, the expression pattern of periodontal integrins changes in response to damage or during wound healing or carcinogenesis [22].
Table 1. Selected integrin heterodimers and their corresponding ligands.
Integrin Heterodimers α1β1 α2β1 α3β1 α5β1 αVβ1 αVβ3 αVβ5
Ligand(s) Different collagens
(e.g. type I)
Different collagens (e.g. type I) laminins fibronectin fibronectin fibronectin vitronectin
Upon mechanical loading, such as during mastication, occlusion or orthodontic treatment, the ECM of the periodontal tissues is deformed. Exemplarily, occlusion forces that exert compressive load onto a tooth are transmitted to the PDL, which serves as a push-pull transducer [23]. This means that the ECM components within the PDL, such as the collagen fibers, are stretched. As they are either directly or indirectly connected to the integrin receptors and these proteins undergo a conformational change, which is herewith the consequence of the initial mechanical stimulus. Therefore, these surface receptors are considered mechanoreceptors, as they transmit the ECM’s physical state into the cell’s interior.
Intracellularly, integrins are linked to various signaling proteins, which function as a molecular clutch that couples integrin–ECM interaction to intracellular biochemical signaling, the centerpiece of MT. Histologically, these integrin-dependent cell-to-matrix connection structures are called FAs [24].
Next to the plasma membrane, a plethora of proteins form the “integrin signaling layer”, where focal adhesion kinase (FAK), the head domain of talin and paxillin interact with the cytoplasmic integrin domain. Vinculin and the tail of talin are designated as the “force transduction layer” [25]. Finally, these proteins are linked to the actin cytoskeleton through an “actin binding layer”, which consists of α-actinin, vasodilator-simulated phosphoprotein (VASP), and zyxin [26]. Beneath the listed proteins, many other cellular key players, such as sarcoma (Src)-family kinases or the extracellular signal-regulated kinases 1/2 (ERK1/2) have been shown to directly interact with FAs and regulate the activity and assembly status of its components (Figure 1A) [24][27][28].
Figure 1. The role of focal adhesions and adherens junctions in mechanotransduction. (A): Focal adhesions (FAs) are adhesion structures that bind extracellular matrix (ECM) ligands via integrin receptors. The latter are composed of varying combinations of an α- and a β-subunit. Each heterodimer has specific ECM ligands (see Table 1). Intracellularly, integrins are linked to various signaling molecules that constitute a molecular clutch, which transmits mechanical information from the ECM into the cell’s interior and vice versa. Focal adhesion kinase (FAK), paxillin, talin, zyxin, vinculin, vasodilator-simulated phosphoprotein (VASP) and α-actinin are examples of important FAs proteins, connecting integrin receptors to the actin cytoskeleton (yellow). The small GTP-binding proteins Ras-related C3 botulinum toxin substrate 1 (Rac1), cell division control protein homologue 42 (Cdc42), and Ras homologue A (RhoA), together with Rho-associated, coiled-coil-containing protein kinase (ROCK) modulate the dynamic de- and repolymerization of globular (G)-actin (yellow dots) and filamentous (F)-actin. FAK activity and subcellular localization of yes-associated protein (YAP) and its cellular homologue transcriptional co-activator with PDZ motif (TAZ) are strongly interconnected. The linker of nucleoskeleton and cytoskeleton (LINC) complex couples the cytoplasmic cytoskeleton to the nucleus. Both mechanisms are important to regulate gene expression in response to mechanical signals. Details are described in the main text. (B): Cell-to-cell adhesion depends on adherens junctions (AJs). Cadherins, as exemplified by E-Cadherin, are transmembrane proteins that bind other cadherins on neighboring cells in a Ca2+-dependent manner (red dots). Intracellularly, cadherins are linked to various proteins, such as p120, α-catenin, β-catenin, and vinculin, which indirectly connect cadherins to the actin cytoskeleton (yellow). YAP/TAZ regulation is also dependent on AJs integrity. β-catenin can also serve as a transcription factor in the nucleus and its subcellular localization contributes to determining cell behavior. Further details are described in the main text.
The actin cytoskeleton is the common downstream target of all MT pathways [29][30]. It not only consists of filamentous (F) actin, but also of actin-regulating proteins (ARPs), which regulate the dynamic building and destruction of the filaments from globular (G) actin monomers. G-actin binds adenosine-triphosphate (ATP) and hydrolyses ATP to adenosine-diphosphate (ADP) within F-actin. This reaction goes along with conformational changes of actin monomers and contributes to the dynamic turnover of actin-related cytoskeletal structures [31][32][33]. Actin polymerization is regulated by a class of guanosine–triphosphate (GTP) binding proteins, known as Ras homologue A (RhoA), cell division control protein homolog 42 (Cdc42), Ras-related C3 botulinum toxin substrate 1 (Rac1), as well as the Rho-associated, coiled-coil-containing protein kinase 1 (ROCK1) (Figure 1A) [34][35][36][37]. These small GTPases are addressed by the mechanotransducing proteins of FAs and their activity state determines the polymerization of G-actin as well as the formation of lamellipodia or filopodia. ROCK1 is even more directly involved in periodontal differentiation, homeostasis, and regeneration, as the inhibition of ROCK1 prevents proper the differentiation of PDL cells into osteoblasts and reduces ECM regeneration via the downregulation of collagen I and fibronectin [38][39]. These findings underscore the essential role of actin cytoskeleton regulation in the MT of the periodontium.
Actin filaments are additionally stabilized by ARPs, such as Arp2/3, tropomyosin or profilin. Contrarily, severing proteins, such as gelsolin, support F-actin depolymerization or destruction [40][41]. Altogether, the complex interplay of actin regulatory proteins governs actin’s dynamic de- and repolymerization and enables complex cellular processes, such as cell division or migration [42]. Further details of these complex cytoskeletal regulatory principles are beyond the scope of this review and interested readers are referred to other comprehensive discussions on this subject [43][44].
Thus far, the description of the actin cytoskeleton and its dynamics does not explain how cells maintain their shape during mechanical stimulation and how cellular integrity can be achieved by mechanisms of FAs-related signaling. To this end, cells need the ability to actively generate forces to withstand external deformation or to exert mechanical stimuli on their environment. This is possible through the action of cytoplasmic motor proteins, known as myosins, which are coupled to the actin cytoskeleton. Via the hydrolysis of ATP, these motor proteins can actively move along actin filaments [45][46][47]. Besides functions in cargo transport, myosins can, therefore, generate tension and traction forces through the relative displacement of actin filaments [48][49][50][51]. This mechanism immediately explains that FAs are not only outside-in signaling platforms that transmit mechanical ECM signals into the cell, but that actomyosin-generated cytoskeletal forces can also be transmitted to the ECM with the help of integrins and their neighboring adaptor proteins. Therefore, FAs are bidirectional mechanosensitive signaling hubs.
Of interest, this is of enormous importance for ECM homeostasis and regeneration in the PDL, as FN and collagen fibrillogenesis depends on intracellularly derived contractile forces. Actin stress fibers serve as a guide trail for the centripetal movement of α5β1-bound FN, which supports FN-FN interactions [52][53]. Collagen molecules can then be deposited on the pre-existing FN fibrils. As this process is force-dependent, characteristic ECM structures, such as the parallel arrays of collagen fibers within the PDL, can be explained via this mechanism and are, therefore, a result of bidirectional MT related to FAs. Interestingly, a recent study gave new insights into the actual nonuniformity of the PDL and revealed the mechanical properties of different subregions within the gomphosis. The so-called collar region is characterized by a high proportion of collagen type I, making it resistant to tensile forces due to high mechanical stiffness. Contrary to that, the furcation region is less stiff and contains less type I collagen, which seems to be associated with a dual function in resisting compressive loads [54].
In periodontitis, collagen and other ECM components are degraded by proteases, such as MMPs or bacteria-derived gingipains (see Section 4) [55]. This changes integrin-dependent MT, leading to a decrease in intracellular actomyosin contractility (outside-in-signaling). Consequently, inside-out signaling is also impaired, which leads to incorrect collagen fibril deposition, worsening the catabolic destruction of the periodontium [56][22].
FAs related signaling also comprise an important step in (alveolar) osteocyte differentiation and the cell’s response to fluid shear stress. The reduction in the protein sclerostin by mechanical loading is mediated by FAK-dependent phosphorylation of the histone deacetylase 5 (HDAC5), which is translocated into the nucleus in response to this post-translational modification. Sclerostin suppression leads to an increase in bone formation and thus mediates an adaptation process by which shear stress leads to mechanical strengthening of the exposed tissue [57][58].
Besides cell-to-matrix adhesion, cell-to-cell adhesion is of great importance, especially in epithelial tissues, such as parts of the gingiva. The main constituents of the so-called adherens junctions (AJs) are cadherin (Cad)-family members (Cads), which are Ca2+-dependent, membrane-embedded proteins that connect cells via homophilic interaction [59]. In the cytoplasm, Cads are connected to various proteins, amongst others α-catenin, p120, vinculin or β-catenin [60][61][62][63]. These adaptor proteins are comparable to the integrin-linked intracellular mechanotransducers, as they connect Cads to the actin cytoskeleton. The same mechanisms and principles as discussed above also apply for AJs and qualify the actin cytoskeleton not only as the common final pathway of AJs and FAs signaling, but also as a crosstalk platform that integrates mechanical cues transmitted through various MT pathways [64][65]. ERK1/2, YAP, and its cellular homologue transcriptional co-activator with PDZ motif (TAZ), vinculin and FAK, are both addressed by AJs and FAs, underscoring the complex mutuality of cell-to-cell and cell-to-matrix adhesion (Figure 1B) [66][67][68][69]. This is the reason why current systemic approaches try to elucidate the tissue- and cell-type specific interplay and fine-regulation of signaling crosstalk related to MT [70]. It is of great interest, to shed light into these principles in periodontal tissues.
In the periodontal context, Cads fulfill different functions, ranging from maintenance of cellular differentiation to epithelial barrier function, tumor suppression, and MT-related tissue homeostasis. Specifically, β-catenin, in its function as a transcriptional regulator, is important in PDLF differentiation and simultaneously inhibits the cementoblastic phenotype [71].
The epithelial E-Cad, which is expressed in GKs, plays a significant role both in periodontitis and oral carcinogenesis. Patients suffering from periodontitis show decreased protein levels of E-Cad, which is indirect evidence for a dysfunctional epithelial barrier function. This further promotes the inflammatory process [72]. Downregulation of E-Cad has also been reported in many carcinomas, where it represents a key step during epithelial-to-mesenchymal transition (EMT). Thereby, epithelial cells detach from their surrounding cells and develop a migratory, fibroblastoid phenotype, which is a prerequisite for tissue invasion and metastasis. Regarding oral carcinogenesis, reduced expression levels of E-Cad and β-catenin are indicators of the progression from dysplasia to cancer and an aggressive OSCC phenotype [73]. As chronic inflammation is a risk factor for cancer development, the loss of E-Cad-related barrier function and MT offers valuable insights into the link between periodontitis and OSCC [74][75].
Beneath E-Cad, Cadherin 11 (Cad11) is another member of the cadherin family expressed in periodontal tissues [76][77]. Upon mechanical loading, expression of Cad11 and β-catenin decreases in PDLFs, which consequently leads to a reduction in collagen 1 synthesis and changes in cellular morphology [78]. These findings clearly show that not only FAs but also AJs are involved in ECM homeostasis and regeneration and that cell-to-cell adhesion is not limited to intercellular information exchange. This hypothesis is supported by other experimental results, where knock-down of Cad11 impairs elastin and collagen synthesis [79]. In mice, Cad11 deficiency reduces cell contractility, which again represents the involvement of AJs in both ECM structure and MT [80].
Taken together, the findings presented in this chapter show that the ECM and its molecular composition are important determinants of periodontal cell behavior in the context of MT. As the periodontal tissues are affected by diseases, such as periodontitis or OSCC development, the relevance of MT in these pathophysiological processes is of clinical relevance. So far, the convergence of FAs and AJs-mediated signal transduction on the cytoskeleton has been discussed. However, the mechanisms by which the information encoded within the contractility and tension of the actin-cytoskeleton is translated into cellular adaptation, which depend on further cellular key players, such as the mechanoresponsive co-transcriptional activators YAP/TAZ and the nuclear cytoskeleton, will be discussed in the subsequent sections.

2. Conclusions

MT-immanent mechanobiological signals are key determinants of cell behavior and tissue adaptation to the external cell and tissue environment. The interplay of molecular mechanosensors and mechanotransducers is complex and represents the tight interrelationship between different cellular compartments and signaling hubs. FAs, AJs, YAP/TAZ, the cytoskeleton, LINC, and the nucleoskeleton only represent a small subset of cellular players involved in MT and NMT but reveal important principles of how the cell percepts and integrates biophysical mechanical information. Periodontal tissues are of special interest to MT research and are a paradigm of how the above-discussed signaling networks, cascades, and molecules govern tissue development, homeostasis and regeneration, cell proliferation, differentiation, and pathologic processes, such as periodontitis or OSCC. By further elucidating the cell-type specific and spatiotemporal fine-tuning of mechanobiological processes, the future translation of these principles into clinical applications will prove to be a strong tool in the field of oral regenerative medicine.

References

  1. Batool, H.; Nadeem, A.; Kashif, M.; Shahzad, F.; Tahir, R.; Afzal, N. Salivary levels of IL-6 and IL-17 could be an indicator of disease severity in patients with calculus associated chronic periodontitis. BioMed Res. Int. 2018, 2018, 8531961.
  2. Arroyo, R.; López, S.; Romo, E.; Montoya, G.; Hoz, L.; Pedraza, C.; Garfias, Y.; Arzate, H. Carboxy-Terminal Cementum Protein 1-Derived Peptide 4 (cemp1-p4) Promotes Mineralization through wnt/β-catenin Signaling in Human Oral Mucosa Stem Cells. Int. J. Mol. Sci. 2020, 21, 1307.
  3. Martins, L.; Amorim, B.R.; Salmon, C.R.; Leme, A.F.P.; Kantovitz, K.R.; Nociti Jr, F.H. Novel LRAP-binding partner revealing the plasminogen activation system as a regulator of cementoblast differentiation and mineral nodule formation in vitro. J. Cell. Physiol. 2020, 235, 4545–4558.
  4. Zhang, Y.; Yang, Y.; Xu, M.; Zheng, J.; Xu, Y.; Chen, G.; Guo, Q.; Tian, W.; Guo, W. The Dual Effects of Reactive Oxygen Species on the Mandibular Alveolar Bone Formation in SOD1 Knockout Mice: Promotion or Inhibition. Oxidative Med. Cell. Longev. 2021, 2021, 8847140.
  5. Denes, B.J.; Ait-Lounis, A.; Wehrle-Haller, B.; Kiliaridis, S. Core matrisome protein signature during periodontal ligament maturation from pre-occlusal eruption to occlusal function. Front. Physiol. 2020, 11, 174.
  6. Zvackova, I.; Matalova, E.; Lesot, H. Regulators of collagen fibrillogenesis during molar development in the mouse. Front. Physiol. 2017, 8, 554.
  7. Kurylo, M.P.; Grandfield, K.; Marshall, G.W.; Altoe, V.; Aloni, S.; Ho, S.P. Effect of proteoglycans at interfaces as related to location, architecture, and mechanical cues. Arch. Oral Biol. 2016, 63, 82–92.
  8. Wei, L.; Chen, Q.; Zheng, Y.; Nan, L.; Liao, N.; Mo, S. Potential Role of Integrin α5β1/Focal Adhesion Kinase (FAK) and Actin Cytoskeleton in the Mechanotransduction and Response of Human Gingival Fibroblasts Cultured on a 3-Dimension Lactide-Co-Glycolide (3D PLGA) Scaffold. Med Sci. Monit. Int. Med J. Exp. Clin. Res. 2020, 26, e921621–e921626.
  9. Naqvi, S.M.; McNamara, L.M. Stem Cell Mechanobiology and the Role of Biomaterials in Governing Mechanotransduction and Matrix Production for Tissue Regeneration. Front. Bioeng. Biotechnol. 2020, 8, 1375.
  10. Hetmanski, J.H.; Jones, M.C.; Chunara, F.; Schwartz, J.-M.; Caswell, P.T. Combinatorial mathematical modelling approaches to interrogate rear retraction dynamics in 3D cell migration. PLoS Comput. Biol. 2021, 17, e1008213.
  11. Kao, T.-W.; Chiou, A.; Lin, K.-H.; Liu, Y.-S.; Lee, O.K.-S. Alteration of 3D Matrix Stiffness Regulates Viscoelasticity of Human Mesenchymal Stem Cells. Int. J. Mol. Sci. 2021, 22, 2441.
  12. Humphrey, J.D.; Dufresne, E.R.; Schwartz, M.A. Mechanotransduction and extracellular matrix homeostasis. Nat. Rev. Mol. Cell Biol. 2014, 15, 802–812.
  13. Hegedűs, O.; Juriga, D.; Sipos, E.; Voniatis, C.; Juhász, Á.; Idrissi, A.; Zrínyi, M.; Varga, G.; Jedlovszky-Hajdú, A.; Nagy, K.S. Free thiol groups on poly (aspartamide) based hydrogels facilitate tooth-derived progenitor cell proliferation and differentiation. PLoS ONE 2019, 14, e0226363.
  14. Feld, L.; Kellerman, L.; Mukherjee, A.; Livne, A.; Bouchbinder, E.; Wolfenson, H. Cellular contractile forces are nonmechanosensitive. Sci. Adv. 2020, 6, eaaz6997.
  15. Reyes-Ramos, A.M.; Álvarez-García, Y.R.; Solodin, N.; Almodovar, J.; Alarid, E.T.; Torres-Garcia, W.; Domenech, M. Collagen I Fibrous Substrates Modulate the Proliferation and Secretome of Estrogen Receptor-Positive Breast Tumor Cells in a Hormone-Restricted Microenvironment. ACS Biomater. Sci. Eng. 2021.
  16. Wei, D.; Li, C.; Ye, J.; Xiang, F.; Liu, J. Extracellular Collagen Mediates Osteosarcoma Progression Through an Integrin α2β1/JAK/STAT3 Signaling Pathway. Cancer Manag. Res. 2020, 12, 12067.
  17. Xing, Q.; Parvizi, M.; Higuita, M.L.; Griffiths, L.G. Basement membrane proteins modulate cell migration on bovine pericardium extracellular matrix scaffold. Sci. Rep. 2021, 11, 4607.
  18. Al-Yafeai, Z.; Orr, A.W. Quantification of integrin activation and ligation in adherent cells. In The Integrin Interactome; Springer: New York, NY, USA, 2021; pp. 17–25.
  19. Ye, Y.; Zhang, R.; Feng, H. Fibronectin promotes tumor cells growth and drugs resistance through a CDC42-YAP-dependent signaling pathway in colorectal cancer. Cell Biol. Int. 2020, 44, 1840–1849.
  20. Roy, S.; Spinali, K.; Schmuck, E.G.; Kink, J.A.; Hematti, P.; Raval, A.N. Cardiac fibroblast derived matrix-educated macrophages express VEGF and IL-6, and recruit mesenchymal stromal cells. J. Immunol. Regen. Med. 2020, 10, 100033.
  21. Sugahara, M.; Nakaoki, Y.; Yamaguchi, A.; Hashimoto, K.; Miyamoto, Y. Vitronectin is involved in the morphological transition of neurites in retinoic acid-induced neurogenesis of neuroblastoma cell line neuro2a. Neurochem. Res. 2019, 44, 1621–1635.
  22. Jakhu, H.; Gill, G.; Singh, A. Role of integrins in wound repair and its periodontal implications. J. Oral Biol. Craniofac. Res. 2018, 8, 122–125.
  23. Jang, A.; Wang, B.; Ustriyana, P.; Gansky, S.A.; Maslenikov, I.; Useinov, A.; Prevost, R.; Ho, S.P. Functional adaptation of interradicular alveolar bone to reduced chewing loads on dentoalveolar joints in rats. Dent. Mater. 2021, 37, 486–495.
  24. Case, L.B.; Waterman, C.M. Integration of actin dynamics and cell adhesion by a three-dimensional, mechanosensitive molecular clutch. Nat. Cell Biol. 2015, 17, 955–963.
  25. Husari, A.; Steinberg, T.; Dieterle, M.P.; Prucker, O.; Rühe, J.; Jung, B.; Tomakidi, P. On the relationship of YAP and FAK in hMSCs and osteosarcoma cells: Discrimination of FAK modulation by nuclear YAP depletion or YAP silencing. Cell. Signal. 2019, 63, 109382.
  26. Belgardt, E.; Steinberg, T.; Husari, A.; Dieterle, M.P.; Hülter-Hassler, D.; Jung, B.; Tomakidi, P. Force-responsive Zyxin modulation in periodontal ligament cells is regulated by YAP rather than TAZ. Cell. Signal. 2020, 72, 109662.
  27. Gao, W.-J.; Liu, J.-X.; Xie, Y.; Luo, P.; Liu, Z.-Q.; Liu, L.; Zhou, H. Suppression of macrophage migration by down-regulating Src/FAK/P130Cas activation contributed to the anti-inflammatory activity of sinomenine. Pharmacol. Res. 2021, 167, 105513.
  28. Zhang, L.; Yan, H.; Tai, Y.; Xue, Y.; Wei, Y.; Wang, K.; Zhao, Q.; Wang, S.; Kong, D.; Midgley, A.C. Design and Evaluation of a Polypeptide that Mimics the Integrin Binding Site for EDA Fibronectin to Block Profibrotic Cell Activity. Int. J. Mol. Sci. 2021, 22, 1575.
  29. Boujemaa-Paterski, R.; Martins, B.; Eibauer, M.; Beales, C.T.; Geiger, B.; Medalia, O. Talin-activated vinculin interacts with branched actin networks to initiate bundles. eLife 2020, 9, e53990.
  30. Damiano-Guercio, J.; Kurzawa, L.; Mueller, J.; Dimchev, G.; Schaks, M.; Nemethova, M.; Pokrant, T.; Brühmann, S.; Linkner, J.; Blanchoin, L. Loss of Ena/VASP interferes with lamellipodium architecture, motility and integrin-dependent adhesion. eLife 2020, 9, e55351.
  31. Hsiao, B.-Y.; Chen, C.-H.; Chi, H.-Y.; Yen, P.-R.; Yu, Y.-Z.; Lin, C.-H.; Pang, T.-L.; Lin, W.-C.; Li, M.-L.; Yeh, Y.-C. Human Costars Family Protein ABRACL Modulates Actin Dynamics and Cell Migration and Associates with Tumorigenic Growth. Int. J. Mol. Sci. 2021, 22, 2037.
  32. Roopnarine, O.; Thomas, D.D. Mechanistic analysis of actin-binding compounds that affect the kinetics of cardiac myosin-actin interaction. J. Biol. Chem. 2021, 196, 100471.
  33. Mani, S.; Katkar, H.H.; Voth, G.A. Compressive and Tensile Deformations Alter ATP Hydrolysis and Phosphate Release Rates in Actin Filaments. J. Chem. Theory Comput. 2021, 17, 1900–1913.
  34. Zeng, Y.; Cao, Y.; Liu, L.; Zhao, J.; Zhang, T.; Xiao, L.; Jia, M.; Tian, Q.; Yu, H.; Chen, S. SEPT9_i1 regulates human breast cancer cell motility through cytoskeletal and RhoA/FAK signaling pathway regulation. Cell Death Dis. 2019, 10, 720.
  35. Kurotsu, S.; Sadahiro, T.; Fujita, R.; Tani, H.; Yamakawa, H.; Tamura, F.; Isomi, M.; Kojima, H.; Yamada, Y.; Abe, Y. Soft Matrix Promotes Cardiac Reprogramming via Inhibition of YAP/TAZ and Suppression of Fibroblast Signatures. Stem Cell Rep. 2020, 15, 612–628.
  36. Nikoloudaki, G.; Snider, P.; Simmons, O.; Conway, S.J.; Hamilton, D.W. Periostin and matrix stiffness combine to regulate myofibroblast differentiation and fibronectin synthesis during palatal healing. Matrix Biol. 2020, 94, 31–56.
  37. Liu, Y.; Rens, E.G.; Edelstein-Keshet, L. Spots, stripes, and spiral waves in models for static and motile cells. J. Math. Biol. 2021, 82, 1–38.
  38. Ugawa, Y.; Yamamoto, T.; Kawamura, M.; Yamashiro, K.; Shimoe, M.; Tomikawa, K.; Hongo, S.; Maeda, H.; Takashiba, S. Rho-kinase regulates extracellular matrix-mediated osteogenic differentiation of periodontal ligament cells. Cell Biol. Int. 2017, 41, 651–658.
  39. Yamamoto, T.; Ugawa, Y.; Yamashiro, K.; Shimoe, M.; Tomikawa, K.; Hongo, S.; Kochi, S.; Ideguchi, H.; Maeda, H.; Takashiba, S. Osteogenic differentiation regulated by Rho-kinase in periodontal ligament cells. Differentiation 2014, 88, 33–41.
  40. Liu, C.; Zhang, Y.; Ren, H. Profilin promotes formin-mediated actin filament assembly and vesicle transport during polarity formation in pollen. Plant Cell 2021.
  41. Matarrese, P.; Vona, R.; Ascione, B.; Paggi, M.G.; Mileo, A.M. Physical Interaction between HPV16E7 and the Actin-Binding Protein Gelsolin Regulates Epithelial-Mesenchymal Transition via HIPPO-YAP Axis. Cancers 2021, 13, 353.
  42. Choi, C.K.; Vicente-Manzanares, M.; Zareno, J.; Whitmore, L.A.; Mogilner, A.; Horwitz, A.R. Actin and α-actinin orchestrate the assembly and maturation of nascent adhesions in a myosin II motor-independent manner. Nat. Cell Biol. 2008, 10, 1039–1050.
  43. Yamamoto, T.; Ugawa, Y.; Kawamura, M.; Yamashiro, K.; Kochi, S.; Ideguchi, H.; Takashiba, S. Modulation of microenvironment for controlling the fate of periodontal ligament cells: The role of Rho/ROCK signaling and cytoskeletal dynamics. J. Cell Commun. Signal. 2018, 12, 369–378.
  44. Romero, S.; Le Clainche, C.; Gautreau, A.M. Actin polymerization downstream of integrins: Signaling pathways and mechanotransduction. Biochem. J. 2020, 477, 1–21.
  45. Ramirez, I.; Gholkar, A.A.; Velasquez, E.F.; Guo, X.; Tofig, B.; Damoiseaux, R.; Torres, J.Z. The myosin regulatory light chain Myl5 localizes to mitotic spindle poles and is required for proper cell division. Cytoskeleton 2021, 78, 23–35.
  46. Jarvis, K.J.; Bell, K.M.; Loya, A.K.; Swank, D.M.; Walcott, S. Force-velocity and tension transient measurements from Drosophila jump muscle reveal the necessity of both weakly-bound cross-bridges and series elasticity in models of muscle contraction. Arch. Biochem. Biophys. 2021, 701, 108809.
  47. Salomon, J.; Gaston, C.; Magescas, J.; Duvauchelle, B.; Canioni, D.; Sengmanivong, L.; Mayeux, A.; Michaux, G.; Campeotto, F.; Lemale, J. Contractile forces at tricellular contacts modulate epithelial organization and monolayer integrity. Nat. Commun. 2017, 8, 13998.
  48. Schreiber, C.; Amiri, B.; Heyn, J.C.; Rädler, J.O.; Falcke, M. On the adhesion–velocity relation and length adaptation of motile cells on stepped fibronectin lanes. Proc. Natl. Acad. Sci. USA 2021, 118.
  49. Santa-Cruz Mateos, C.; Valencia-Expósito, A.; Palacios, I.M.; Martín-Bermudo, M.D. Integrins regulate epithelial cell shape by controlling the architecture and mechanical properties of basal actomyosin networks. PLoS Genet. 2020, 16, e1008717.
  50. Uçar, M.C.; Lipowsky, R. Collective force generation by molecular motors is determined by strain-induced unbinding. Nano Lett. 2019, 20, 669–676.
  51. Cong, J.; Fang, B.; Wang, Q.; Su, Y.; Gu, T.; Luo, T. The mechanobiology of actin cytoskeletal proteins during cell–cell fusion. J. R. Soc. Interface 2019, 16, 20190022.
  52. Van Helvert, S.; Friedl, P. Strain stiffening of fibrillar collagen during individual and collective cell migration identified by AFM nanoindentation. ACS Appl. Mater. Interfaces 2016, 8, 21946–21955.
  53. Wu, C.; Bauer, J.; Juliano, R.; McDonald, J. The alpha 5 beta 1 integrin fibronectin receptor, but not the alpha 5 cytoplasmic domain, functions in an early and essential step in fibronectin matrix assembly. J. Biol. Chem. 1993, 268, 21883–21888.
  54. Connizzo, B.; Sun, L.; Lacin, N.; Gendelman, A.; Solomonov, I.; Sagi, I.; Grodzinsky, A.; Naveh, G. Nonuniformity in Periodontal Ligament: Mechanics and Matrix Composition. J. Dent. Res. 2020, 100, 179–186.
  55. Attia, M.S.; Alblowi, J.A. Effect of Subantimicrobial Dose Doxycycline Treatment on Gingival Crevicular Fluid Levels of MMP-9 and MMP-13 in Periodontitis Stage 2, Grade B in Subjects with Type 2 Diabetes Mellitus. J. Immunol. Res. 2020, 2020, 2807259.
  56. Behm, C.; Nemec, M.; Blufstein, A.; Schubert, M.; Rausch-Fan, X.; Andrukhov, O.; Jonke, E. Interleukin-1β Induced Matrix Metalloproteinase Expression in Human Periodontal Ligament-Derived Mesenchymal Stromal Cells under In Vitro Simulated Static Orthodontic Forces. Int. J. Mol. Sci. 2021, 22, 1027.
  57. Sato, T.; Verma, S.; Andrade, C.D.C.; Omeara, M.; Campbell, N.; Wang, J.S.; Cetinbas, M.; Lang, A.; Ausk, B.J.; Brooks, D.J. A FAK/HDAC5 signaling axis controls osteocyte mechanotransduction. Nat. Commun. 2020, 11, 3282.
  58. Li, X.; Ominsky, M.S.; Niu, Q.T.; Sun, N.; Daugherty, B.; D’Agostin, D.; Kurahara, C.; Gao, Y.; Cao, J.; Gong, J. Targeted deletion of the sclerostin gene in mice results in increased bone formation and bone strength. J. Bone Miner. Res. 2008, 23, 860–869.
  59. Muhamed, I.; Wu, J.; Sehgal, P.; Kong, X.; Tajik, A.; Wang, N.; Leckband, D.E. E-cadherin-mediated force transduction signals regulate global cell mechanics. J. Cell Sci. 2016, 129, 1843–1854.
  60. Pence, L.J.; Kourtidis, A.; Feathers, R.W.; Haddad, M.T.; Sotiriou, S.; Decker, P.A.; Nassar, A.; Ocal, I.T.; Shah, S.S.; Anastasiadis, P.Z. PLEKHA7, an Apical Adherens Junction Protein, Suppresses Inflammatory Breast Cancer in the Context of High E-Cadherin and p120-Catenin Expression. Int. J. Mol. Sci. 2021, 22, 1275.
  61. Krischak, A.; Kowaliuk, J.; Sarsarshahi, S.; Dörr, W.; Kleiter, M. Effect of irradiation on the expression of E-cadherin and β-catenin in early and late radiation sequelae of the urinary bladder and its modulation by NF-κB inhibitor thalidomide. Strahlenther. Onkol. 2021, 197, 537–546.
  62. Kluger, C.; Braun, L.; Sedlak, S.M.; Pippig, D.A.; Bauer, M.S.; Miller, K.; Milles, L.F.; Gaub, H.E.; Vogel, V. Different vinculin binding sites use the same mechanism to regulate directional force transduction. Biophys. J. 2020, 118, 1344–1356.
  63. Chandran, R.; Kale, G.; Philippe, J.-M.; Lecuit, T.; Mayor, S. Distinct actin-dependent nanoscale assemblies underlie the dynamic and hierarchical organization of E-cadherin. Curr. Biol. 2021, 31, 1726–1736.e4.
  64. Huang, S.-C.; Liang, J.Y.; Vu, L.V.; Faye, H.Y.; Ou, A.C.; Ou, J.P.; Zhang, H.S.; Burnett, K.M.; Benz, E.J., Jr. Epithelial-specific isoforms of protein 4.1 R promote adherens junction assembly in maturing epithelia. J. Biol. Chem. 2020, 295, 191–211.
  65. Ishiyama, N.; Sarpal, R.; Wood, M.N.; Barrick, S.K.; Nishikawa, T.; Hayashi, H.; Kobb, A.B.; Flozak, A.S.; Yemelyanov, A.; Fernandez-Gonzalez, R. Force-dependent allostery of the α-catenin actin-binding domain controls adherens junction dynamics and functions. Nat. Commun. 2018, 9, 5121.
  66. Jasuja, H.; Kar, S.; Katti, D.R.; Katti, K. Perfusion bioreactor enabled fluid-derived shear stress conditions for novel bone metastatic prostate cancer testbed. Biofabrication 2021, 13, 035004.
  67. Liu, Y.-H.; Chen, C.-C.; Hsueh, Y.-J.; Hung, L.-M.; Ma, D.H.-K.; Chen, H.-C.; Len, W.-B.; Meir, Y.-J.J. Extraneous E-Cadherin Engages the Deterministic Process of Somatic Reprogramming through Modulating STAT3 and Erk1/2 Activity. Cells 2021, 10, 284.
  68. Ma, Y.-C.; Yang, Z.-S.; Ma, L.-Q.; Shu, R.; Zou, C.-G.; Zhang, K.-Q. YAP in epithelium senses gut barrier loss to deploy defenses against pathogens. PLoS Pathog. 2020, 16, e1008766.
  69. Monster, J.L.; Donker, L.; Vliem, M.J.; Win, Z.; Matthews, H.K.; Cheah, J.S.; Yamada, S.; de Rooij, J.; Baum, B.; Gloerich, M. An asymmetric junctional mechanoresponse coordinates mitotic rounding with epithelial integrity. J. Cell Biol. 2021, 220.
  70. Zuidema, A.; Wang, W.; Sonnenberg, A. Crosstalk between Cell Adhesion Complexes in Regulation of Mechanotransduction. BioEssays 2020, 42, 2000119.
  71. Lim, J.C.; Bae, S.H.; Lee, G.; Ryu, C.J.; Jang, Y.J. Activation of β-catenin by TGF-β1 promotes ligament-fibroblastic differentiation and inhibits cementoblastic differentiation of human periodontal ligament cells. STEM CELLS 2020, 38, 1612–1623.
  72. Arun, R.; Hemalatha, R.; Arun, K.; Kumar, T. E-cadherin and CD1a expression in gingival epithelium in periodontal health, disease and post-treatment. Indian J. Dent. Res. 2010, 21, 396.
  73. Gupta, A.; Sharma, S.; Batra, M.; Abidullah, M.; Bhuvinder, S.; Katragadda, P. Role of E-cadherin in Progression of Oral Squamous Cell Carcinoma: A Retrospective Immunohistochemical Study. J. Contemp. Dent. Pract. 2018, 19, 1105–1110.
  74. Abe-Yutori, M.; Chikazawa, T.; Shibasaki, K.; Murakami, S. Decreased expression of E-cadherin by Porphyromonas gingivalis-lipopolysaccharide attenuates epithelial barrier function. J. Periodontal Res. 2017, 52, 42–50.
  75. Sowmya, S.; Rao, R.S.; Prasad, K. Development of clinico-histopathological predictive model for the assessment of metastatic risk of oral squamous cell carcinoma. J. Carcinog. 2020, 19, 2.
  76. Johnson, C.L.; Merryman, W.D. Side-specific valvular endothelial-interstitial cell mechano-communication via cadherin-11. J. Biomech. 2021, 119, 110253.
  77. Piao, S.; Inglehart, R.C.; Scanlon, C.S.; Russo, N.; Banerjee, R.; D’Silva, N.J. CDH 11 inhibits proliferation and invasion in head and neck cancer. J. Oral Pathol. Med. 2017, 46, 89–97.
  78. Feng, L.; Zhang, Y.; Kou, X.; Yang, R.; Liu, D.; Wang, X.; Song, Y.; Cao, H.; He, D.; Gan, Y. Cadherin-11 modulates cell morphology and collagen synthesis in periodontal ligament cells under mechanical stress. Angle Orthod. 2017, 87, 193–199.
  79. Row, S.; Liu, Y.; Alimperti, S.; Agarwal, S.K.; Andreadis, S.T. Cadherin-11 is a novel regulator of extracellular matrix synthesis and tissue mechanics. J. Cell Sci. 2016, 129, 2950–2961.
  80. Alimperti, S.; You, H.; George, T.; Agarwal, S.K.; Andreadis, S.T. Cadherin-11 regulates both mesenchymal stem cell differentiation into smooth muscle cells and the development of contractile function in vivo. J. Cell Sci. 2014, 127, 2627–2638.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 494
Revisions: 2 times (View History)
Update Date: 10 Jun 2021
1000/1000