Next Article in Journal
Functional Assessment of a New PBX1 Variant in a 46,XY Fetus with Severe Syndromic Difference of Sexual Development through CRISPR-Cas9 Gene Editing
Previous Article in Journal
Expression of Long Noncoding RNAs in Fibroblasts from Mucopolysaccharidosis Patients
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Effect of the Rho-Kinase/ROCK Signaling Pathway on Cytoskeleton Components

1
Sir John Walsh Research Institute, Faculty of Dentistry, University of Otago, Dunedin 9016, New Zealand
2
Department of Oral Diagnostic and Surgical Sciences, Faculty of Dentistry, University of Otago, 310 Great King Street, Dunedin 9016, New Zealand
3
Department of Oral Sciences, Faculty of Dentistry, University of Otago, 310 Great King Street, Dunedin 9016, New Zealand
*
Authors to whom correspondence should be addressed.
Genes 2023, 14(2), 272; https://doi.org/10.3390/genes14020272
Submission received: 20 December 2022 / Revised: 10 January 2023 / Accepted: 16 January 2023 / Published: 20 January 2023
(This article belongs to the Section Molecular Genetics and Genomics)

Abstract

:
The mechanical properties of cells are important in tissue homeostasis and enable cell growth, division, migration and the epithelial-mesenchymal transition. Mechanical properties are determined to a large extent by the cytoskeleton. The cytoskeleton is a complex and dynamic network composed of microfilaments, intermediate filaments and microtubules. These cellular structures confer both cell shape and mechanical properties. The architecture of the networks formed by the cytoskeleton is regulated by several pathways, a key one being the Rho-kinase/ROCK signaling pathway. This review describes the role of ROCK (Rho-associated coiled-coil forming kinase) and how it mediates effects on the key components of the cytoskeleton that are critical for cell behaviour.

1. Introduction

The biomechanical properties of cells are inextricably linked to their intracellular structural components, such as the cytoskeleton [1]. Internal and external mechanical forces both act through the cytoskeleton to affect cellular mechanical properties, behaviour, cell spreading, movement, polarity and cytokinesis [2,3,4,5,6,7]. The cytoskeleton also plays essential biomechanical roles in connecting the plasma membrane of the cell, as well as internal membranes such as the endoplasmic reticulum, to the rest of the cell. The cytoskeletal network also regulates the diffusion of intracellular polymers that are larger than the network mesh size and possibly controls the permeation of water and small solutes through the cell [8]. Rho-associated coiled-coil forming kinase (ROCK) is considered to be a key regulator of the cytoskeleton and affects various important cellular functions such as cell shape, motility, secretion, proliferation and gene expression [9,10,11,12]. For example, various neural processes, including cell migration, axonal guidance, dendritic spine architecture, axonal regeneration and cell survival, are dependent on the ROCK signaling pathway [13]. The ROCK signaling pathway plays an important role in the reorganization of both microtubules and actin during the process formation of podocytes [14]. ROCK is associated with cancer progression and ROCK protein expression is higher in a number of cancer types such as primary stomach carcinoma, colon and bladder cancers and malignant melanoma [15]. A high level of ROCK protein expression correlates with poor overall survival in osteosarcoma and more aggressive behaviour in hepatocellular carcinomas [16,17,18]. This review is focused on the current knowledge of the regulation of the Rho-kinase/ROCK signaling pathway and the effects of ROCK on cytoskeletal components.

2. The Cytoskeleton

The main components of the cytoskeleton include microfilaments (specifically actin filaments), intermediate filaments and microtubules, which connect to each other, cellular organelles and the cell/nuclear membrane. They are biopolymers and organized into a three-dimensional network that resists deformation but can rearrange in response to internal or external forces and are thus responsible for maintaining the integrity of the cellular components. In general, there are four main functions of the cytoskeleton: organization of cellular contents; connection of cellular constituents to the external environment; mechanical resistance to deformation and generation of forces to move the cell or change its shape (Table 1) [2,19].
The three types of cytoskeletal components differ from each other, both chemically and physically. Generally, the most significant difference is their rigidity, which can be represented by the persistence length (lP) [7]. The persistence length is a mechanical property quantifying the bending stiffness of a polymer, which is defined as the distance over which the filament is bent by thermal forces and is proportional to the stiffness of the polymer [66]. If a polymer is considered to be a uniform cylinder, lP = BS/kBT, where BS is bending stiffness, kB is the Boltzmann constant and T is the absolute temperature [67]. Filament stiffness, length, and geometry of cross-linking together determine the mechanical properties of cytoskeletal networks. The most common difference among the cytoskeletal filament types is their bending stiffness. The lP of microfilaments is 10–17 μm, the lP of intermediate filaments is less than 0.3–1.0 μm, and the lP of microtubules is more than 1 mm [68]. This shows that microtubules are not only larger but also stiffer than microfilaments and intermediate filament.

3. Microfilaments

Microfilaments, or more specifically actin filaments, are semiflexible (lP ~ 10–17 μm) polymers of actin (~375 amino acids with a molecular mass of ~42-kDa) folded into a U-shaped double helix structure. Each actin monomer has four subdomains and a bound adenine nucleotide, either ATP, ADP-Pi, or ADP, and a bound divalent cation, Mg2+ [67,69,70]. Microfilaments are ~7 nm in diameter and up to several micrometres in length [71]. Bending, torsional, and twist-bend coupling elasticities influence the microfilament’s mechanical properties [70]. The properties of microfilaments are determined by the strength and distribution of inter-subunit contacts [72]. Microfilaments can assemble into different structures, such as networks and bundles that are closely packed and crosslinked by trans-membrane adhesion proteins.
The mechanical properties of microfilament networks are determined by the physical properties of the individual actin molecules, the connections/interactions between actin monomers, and the three-dimensional geometry of the actin arrangement [8]. Microfilament networks appear to be the primary mechanical component of the cytoskeleton and are particularly abundant and dense near the inner surface of the plasma membrane, where they form a network with a variety of capping, binding, branching, and severing proteins. G-actin (individual globular actin) is capable of polymerisation and can bind either ATP or ADP. The pool of unpolymerized actin in normal cells is nearly exclusively made up of ATP-G-actin, with little to no ADP-G-actin in the pool [73]. Polymerization is not an energy-consuming process [74]. ATP is hydrolysed a long time after polymerization occurs [75]. F-actin is formed by polymerization and the addition of G-actin at the growing (+) end of the microfilament. The process of adding a G-actin unit at one end and removing it at the other (-) end is called treadmilling [76,77]. The actin cytoskeleton network is constantly assembling and disassembling by treadmilling, which can drive the formation of protrusive structures such as filopodia and lamellipodia. As a result, the network not only provides mechanical support, determines cell shape, but also drives cell locomotion by the extension of pseudopods, thereby enabling cellular migration, division, and particle engulfment [20,21,22]. The connection between the microfilaments and the trans-membrane adhesion proteins also facilitates communication between the intracellular and extracellular mechanical signals, which allows cells to detect and respond to both chemical and mechanical signals from their extracellular environment [78]. Microfilament networks are viscoelastic as they both store and dissipate mechanical energy [79]. The elasticity of the networks allows them to resist deformation like a simple spring. This resistance to deformation and ability to recover was found to be dependent on rho-kinase mediated contractility [80].

4. Intermediate Filaments

Intermediate filaments play an important role in cell mechanics, signaling and homeostasis. They belong to a superfamily of highly conserved, α-helix-rich fibrous proteins that are encoded by approximately 65–70 different genes [81,82,83,84]. They are more flexible biopolymers (lP = 0.3–1.0 μm) than microfilaments and microtubules. Their diameter is from 8 to 12 nm, intermediate in size between microfilaments and microtubules [7]. Thus far, six major types of intermediate filament have been described [85], and their abundance varies between cell types. Examples of the different intermediate filaments are: acidic keratins (type I), basic keratins (type II), vimentin, glial fibrillary acidic protein (GFAP), peripherin and desmin (type III), neurofilaments (type IV), lamin (type V), and filensin and phakinin (type VI) [86]. Type I and II keratins are highly expressed in epithelial cells [29,32,87,88]. The type III protein vimentin is present mainly in cells of mesenchymal origin [41], whereas desmin, GFAP and peripherin are mainly found in muscle, glial cells, and neurons of the peripheral system, respectively [36,89]. Type IV proteins are found in many types of mature neurons [42,85]. Type V proteins are the nuclear lamins which form a filamentous support inside the inner nuclear membrane [36,47]. The type VI proteins, phakinin and filensin, are lens fibre cell specific intermediate filaments [52,56,90].
Intermediate filaments are vital determinants of intracellular organelle organization and form a delicate network in the cytoplasm, encapsulating the nucleus and radiating toward the plasma membrane [19]. They are the main proteins providing mechanical resistance to the cells, because of their flexibility, elasticity and extensibility [85]. Both in vitro and in vivo studies have shown that intermediate filaments are important contributors to the elasticity and tensile strength of cells [91,92,93]. Indeed, intermediate filaments have been shown to determine cell stiffness in keratinocytes [93]. Indirect perturbation of cytoplasmic intermediate filaments has detrimental effects on cell stiffness. Research has shown that lipids such as sphingosylphosphorylcholine, induce perinuclear rearrangement of intermediate filaments leading to a significant increase in cellular elasticity, which could facilitate metastasis [94].

5. Microtubules

Microtubules are cylindrical polymers of α and β-tubulin dimers. There are multiple genes encoding tubulin protein isotypes [95,96]. According to The HUGO Gene Nomenclature Committee, there are 24 tubulin genes (10 alpha tubulins, 10 beta tubulins, 1 delta, 1 epsilon and 2 gamma tubulins) and 3 pseudogenes [97]. These tubulin proteins polymerize into protofilaments (α–β)n, and arrange in a spiral to form a hollow tube with a 25 nm diameter [67,98]. In addition, γ-tubulin which is located at the centrosome, has an important role in initiating microtubule assembly. Microtubules have an lP > 1 mm and are stiffer than microfilaments and intermediate filaments. They are relatively brittle and can be fractured more easily than intermediate filaments [68]. The published flexural rigidity ranges from 1 × 10−24 to 32 × 10−24 Nm2 [99]. Microtubules form a dynamic scaffold for the cells and are involved in a variety of functions, including the transportation of intracellular vesicles and organelles throughout the cell during normal physiological processes, force generation (by polymerization, depolymerization, or interactions with motor proteins), formation of intracellular transport platforms (for anchoring, signaling and force-coupling roles), segregation of chromosomes, assembly of the mitotic spindle in dividing cells, axon extension in neurons, directional cell migration and differentiation, maintaining cell stiffness/shape, and regulation of cell morphology and cell mechanics [57,100,101]. Microtubules undergo constant cycles of polymerization and depolymerization utilizing a guanosine triphosphate (GTP) dependent process (dynamic instability). The polymerization of microtubules begins with the formation of 13 linear protofilaments which are composed of head to tail arrays of tubulin dimers. Similar to actin, microtubules have two distinct ends (plus and minus ends). The unprotected minus ends are unstable and often require stabilizer proteins (calmodulin-regulated spectrin-associated proteins or Partonins) to prevent depolymerization [102]. On the other hand, microtubules elongate by the addition of tubulin subunits to the plus end [103]. They grow by αβ-tubulin heterodimers bound to GTP binding to the growing end of the microtubule. The GTP bound to the αβ-tubulin heterodimers can be hydrolyzed to GDP during or immediately after polymerization [104]. This hydrolysis decreases the binding affinity of tubulin polymers for adjacent molecules, thereby favouring depolymerization and resulting in the dynamic behaviour of microtubules (alternation between cycles of growth and shrinkage) [105]. Microtubules can therefore undergo periods of growth and disassembly, which means that microtubules have half-lives of only several minutes within the cell [106]. This is particularly important for the remodelling of the cytoskeleton that occurs during mitosis, and for force generation and signaling [107]. In order to control this dynamic instability, many organisms have developed proteins that perturb microtubule dynamics, several of which are in use as cancer chemotherapeutics and anti-inflammatory drugs [105]. For example, the anti-cancer drug paclitaxel (Taxol) stabilizes microtubules and prevents their disassembly, promoting mitotic arrest and cell death [108]. In epithelial cells, microtubules are responsible for the spatial organization of the secretory and endocytic apparatuses, the facilitation of exocytic and post-endocytic protein transportation, participation in protein sorting and apico-basolateral polarity [109].
In addition to these three main cytoskeleton components, many other proteins are associated with the cytoskeleton, such as spectrin, kinesin, plectin, emerin, Sad1 and UNC-84 proteins, Klarsicht, ANC-1, and Syne homology proteins, and they are important for cellular mechanics. [19]. Together, the cellular mechanical properties are determined by complex interactions involving cytoskeletal proteins and polymers. There are several pathways that lead to major rearrangements of the cytoskeletal structures, namely: Rho-kinase/ROCK; PI3K/AKT; β-Catenin-independent Wnt (Wnt ligands stabilize β-catenin; β-catenin helps link cadherin adhesion molecules to cytoskeleton); glycogen synthase kinase-3 beta (Gsk-3β); Mitogen-activated protein kinase (MAPK); and cyclic GMP-dependent protein kinase [110,111,112,113,114,115,116,117,118,119] (Table 2).
Other kinases, such as PI4P and protein kinase C, are also important in regulating interactions of the cell membrane proteins and protein scaffolds involved in vesicle budding, and cytoskeletal organization [197,198]. In particular, ROCK has been considered as a key regulator of the cytoskeleton as it mediates various important cellular functions such as cell shape, motility, secretion, proliferation, and gene expression [199].

6. Rho-Associated Kinase/ROCK Pathway and Associated Genes

6.1. Rho GTPases

Rho GTPases are key regulators of several cellular processes, including cytoskeletal shape, gene expression, cell cycle progression, cell polarity and cell migration [200,201,202]. Rho GTPases are a sub-family within the Ras superfamily of G-proteins and comprise at least 20 members, including Rho (A, B, C isoforms), Rac (1, 2, 3 isoforms), Cdc42 (Cdc42Hs, G25K isoforms), Rnd1/Rho6, Rnd2/Rho7, Rnd3/RhoE, RhoD, RhoG, TC10 and TTF [203,204,205]. In particular, RhoA, RhoB and RhoC, have similar effectors and modes of action, which impact many cellular processes [206]. They serve as gate control molecules by cycling between a GTP-bound active and GDP-bound inactive states-facilitated by Rho GTPase-specific guanine nucleotide exchange factors (GEFs) [207,208,209]. Subsequently, Rho proteins regulate downstream pathways which affect cell migration, adhesion, proliferation, apoptosis and the cell cycle via induction of contractile fibre bundles such as microfilaments, modulating microtubules and regulating intermediate filament turnover [203,204,210,211]. Rho GTPases interact with at least 30 effector proteins and initiate downstream signaling [212,213]. The effectors of three members of the family (RhoA, Rac1 and Cdc42) have been well studied. RhoA regulates the contraction of moving cells, whereas Rac1 and Cdc42 mediate the formation of lamellipodia and filopodia, respectively [214]. Rho GTPase effectors can be classified into different groups according to their functions, such as scaffold proteins, kinases, actin-binding proteins, formin-like molecules, and phospholipases [215]. For example, ROCK1,2 (serine/threonine kinases) and Dia1,2 (scaffold proteins) are important RhoA effectors that regulate cytoskeleton nucleation and polymerization [206]. Wiskott–Aldrich syndrome protein (WASP; a scaffold protein) and neural-WASP (scaffold protein N-Wasp) are Cdc42 effectors. N-Wasp mediates actin polymerization via the Arp2/3 complex [216]. Formin-homology-domain-containing protein (formin-like molecule Fhod1) is one of the Rac1 effectors that regulates actin cytoskeleton organization and gene transcription [217]. p21-activated (serine/threonine) kinase is the effector for both Rac1 and Cdc42 [218], it regulates cell shape and polarity through phosphorylation of multiple cytoskeletal proteins.

6.2. ROCK

Rho-associated kinase ROCK is one of the most important effectors downstream of Rho GTPase [219] (Figure 1). Human ROCKs consist of two isoforms, ROCK 1 and ROCK 2. ROCK 1 is located on chromosome 18 (18q11.1, human), and its mRNA is highly expressed in bone marrow and adipose cells (https://www.ncbi.nlm.nih.gov/gene/6093#gene-expression, accessed on 21 December 2022) whereas ROCK2 is located on chromosome 2 (2p25.1, human) and its mRNA is expressed abundantly in adipose tissue and the colon (https://www.ncbi.nlm.nih.gov/gene/9475, accessed on 21 December 2022) [220,221]. ROCK proteins are activated by binding Rho-GTPase which then, directly and indirectly, regulate the cytoskeleton. ROCK proteins induce a wide range of cellular responses that involve microfilaments, intermediate filaments and microtubules (Figure 2). Their downstream targets are membrane distal, including LIM kinases, myosin phosphatase target subunit of myosin light chain phosphatase, myosin light chain (MLC), collapsing response mediator protein, and ezrin-radixin-moesin (ERM) proteins. For example, ROCK controls the organization and stabilization of actin filaments by phosphorylating a number of proteins, such as MLC [222], LIM1/2 [223], myosin phosphatase target subunit 1 (MYPT1) [224], ERM [225], adducin [226], calponin [227], myristoylated alanine-rich C kinase substrate (MARCKS) [228], elongation factor-1 alpha (EF1α) [229], troponin I/T [230] and profilin [231]. ROCK phosphorylates MLC either directly or inactivates MLC phosphatase, resulting in the induction of actin-myosin contractility [232]. ROCK also activates LIM-Kinase by phosphorylation and LIM-Kinase in turn phosphorylates cofilin, which inhibits cofilin’s actin depolymerization activity, resulting in the stabilization of the actin cytoskeleton [223] (Figure 2). Both the ROCK/MYPT1/MLC and ROCK/LIM kinases/cofilin pathways are key elements in stress fibre assembly and cell adhesion [233]. Moreover, although ROCK1 and 2 have high sequence identity, they differ functionally, especially in relation to actin regulation (Figure 1). ROCK 1 is thought to be involved in destabilizing actin via regulating MLC and actin-myosin contraction, whereas ROCK 2 acts via cofilin to stabilize the actin cytoskeleton [233]. However, Wang et al., found that ROCK 1 phosphorylates LIM-kinase to inactivate cofilin, so regulating cell adhesion and invasion, but not ROCK 2 does not [234]. Also, Rochelle et al., demonstrated that ROCK 1 mediated amoeboid motility via destrin but not via cofilin [235].
ROCK proteins regulate different types of intermediate filaments via phosphorylation. For example, ROCK proteins can breakdown and separate the intermediate filaments via phosphorylation. It also upregulates the cell stiffness by interacting with keratin intermediate filaments [130,236,237,238,239]. Interestingly, activation of ROCK induces the deformation of intermediate filaments (Figure 2) with an instant release of inactive ROCK. The released ROCK is transported to the periphery of the cell and reactivated by Rho-GTP, it can then accelerate the phosphorylation and disassembly of intermediate filaments [240]. In addition, ROCK regulates microtubules by phosphorylation of several microtubule-associated proteins such as MAP2/Tau [241], collapsin response mediator protein 2 (CRMP2) [242] and Doublecortin [243] (Figure 2). ROCK-induced phosphorylation of MAP2/Tau leads to destabilization of the microtubules [12,244]. CRMPs stabilize microtubules however this is inhibited by phosphorylation via ROCK [245]. Doublecortin is a microtubule-associated protein. Phosphorylation of Doublecortin via ROCK inhibits microtubule bundling [243]. In addition, ROCK regulates microtubule acetylation via phosphorylation of the tubulin polymerization promoting protein 1 (TPPP1/p25), resulting in a decreased cellular level of acetylated tubulin and increased cell migration [246].
There are six genes that either enhance (disabled 2 (Dab2), synaptopodin 2 (Synpo2) and thymus cell antigen 1, theta (Thy1)) or inhibit (cysteine and histidine-rich domain (CHORD)-containing, zinc-binding protein 1 (Chordc1), heart of glass (Heg1), and Ras interacting protein 1 (Rasip1)) ROCK activity. For example, Dab2 (on chromosome 15), which plays an important role in cell proliferation and differentiation, is capable of upregulating RhoA/ROCK signaling [247]. SYNPO2 (known as myopodin), encodes an actin-binding protein, and has been characterized as a tumour suppressor which positively regulates ROCK [248]. It may promote cell migration through activation of the ROCK signaling pathway via increasing levels of Rho-GTP [249]. Thy1 (on chromosome 9) triggers the actin cytoskeleton remodelling via the Thy-1-CBP-Csk-Src-RhoA-ROCK axis [250]. However, overexpressed Morgana/chp-1, encoded by the CHORDC1 gene (on chromosome 9), negatively regulates Rho, and binds then inhibits ROCK 1 and 2 [251]. Rasip1 (on chromosome 7) increases the activity of Cdc42 and suppresses actomyosin contractility via inhibition of RhoA [252]. Silencing HEG1 (on chromosome 16) expression by siRNA or shRNA increases phosphorylation of MLC and formation of stress fibres indicating increased ROCK signaling [253]. Together with these regulating genes, the ROCK signaling pathway plays an important role in cytoskeleton and cellular mechanics.
ROCK inhibitors act to suppress the ROCK pathway through multiple mechanisms. They can relax the trabecular meshwork through inhibiting the actin cytoskeleton contractile tone. They can induce beta-catenin nuclear translocation and inhibit cell migration or block ATP-dependent phosphorylation therefore inhibiting ROCK1 and ROCK2 and can inhibit Rho GTPases from binding to ROCK [233,254,255]. More than 170 chemicals are capable of inhibiting ROCK, either selectively or non-selectively [256]. Most of them are ATP-competitive kinase inhibitors [257,258]. Structurally, they can be classified into isoquinoline/isoquinolinone, indazole, pyridine, pyrimidine, pyrrolopyridine, pyrazole, benzimidazole, benzothiazole, benzathiophene, benzamide, phthalazinone, aminofurazan, quinazoline, and boron derivatives [256]. Several ROCK inhibitors have been used in clinical trials for many therapeutic indications, such as ophthalmology, cardiovascular disease, anti-erectile dysfunction, and cancers. For example, both Ripasudil and Netarsudil have been proven to be of benefit in the treatment of glaucoma [259]. Fasudil, which is a potent vasodilator, has been approved for the treatment of cerebral vasospasm in Japan and China since 1995 [260]. Y-27632 improved ventricular hypertrophy, fibrosis, and function in a rat model [261,262]. Daily administration of Y-27632 also improved erectile responses in a rat model [263,264]. The use of ROCK inhibitors in cancer, such as gastric, bone, lung, breast, renal, liver and prostate cancers, has been widely characterised [265,266,267,268,269,270,271]. For example, the selective ROCK inhibitor RKI-1447 was shown to have significant anti-invasive and anti-tumour effects [272]. AT13148, an inhibitor of ROCK and AKT kinases, has been found to have antimetastatic and antiproliferative activity [273]. Despite the potential of ROCK inhibitors, the number of clinical trials for human cancer is still limited.
Although ROCK 1 and ROCK 2 are highly homologous kinases and are both involved in the Rho/ROCK signaling pathway, their target substrates and physiological/pathological activities differ. Unfortunately, most ROCK inhibitors cannot selectively target one or other of the two isoforms; such inhibitors would be very useful. Dimerization is required for the RhoE activation of ROCK 1 [274], and although the homology between the kinase domains of ROCK 1 and ROCK 2 is ~90%, the homology between the N-terminal dimerization domains [275] of the two proteins is only ~60% [276]. Therefore dimerization-disrupting peptides could be developed and used to selectively target the dimerization domains of the two isoforms [276]. Autophosphorylation of ROCK 1 at Ser1333 and ROCK 2 at Ser1366 is required for activation of the kinases [277]. It has been shown that the interaction between Thr405 in the hydrophobic motif of ROCK 2 and Asp39 in the N-terminal extension was essential for both kinase activation and dimerization [278]. This provides further support for the development of peptides that prevent dimerization and kinase activity of ROCK isoforms.

7. Conclusions

The cytoskeleton gives cells their shape, structure and mechanical support and plays an essential role in cellular mechanics. ROCK is an effector of the small Rho GTPase and when activated influences many cellular functions and processes by regulating different elements of the cytoskeleton. Many cellular and physiological functions are mediated by ROCK and its activity is often elevated in some disorders, making them good targets for the development of new drugs. ROCK inhibitions have been found to effectively manage several diseases in humans and animal models. Progress has been made towards understanding how non-selective ROCK inhibitors work for several diseases and conditions and efforts should be made to develop ROCK inhibitors with improved specificity and sensitivity.

Author Contributions

Conceptualization, G.G., L.M., R.D.C. and D.E.C.; writing—original draft preparation, G.G.; writing—review and editing, R.D.C., D.E.C. and L.M.; visualization, G.G., L.M. and R.D.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by 2021 SJWRI PhD Research Grant, Grant number 118940.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Guan, G.; He, Y.; Mei, L. Atomic force microscopy: A nanobiotechnology for cellular research. Nano TransMed. 2022, 1, e9130004. [Google Scholar] [CrossRef]
  2. Fletcher, D.A.; Mullins, R.D. Cell mechanics and the cytoskeleton. Nature 2010, 463, 485–492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Zhu, C.; Bao, G.; Wang, N. Cell mechanics: Mechanical response, cell adhesion, and molecular deformation. Annu. Rev. Biomed. Eng. 2000, 2, 189–226. [Google Scholar] [CrossRef]
  4. Walker, M.; Rizzuto, P.; Godin, M.; Pelling, A.E. Structural and mechanical remodeling of the cytoskeleton maintains tensional homeostasis in 3D microtissues under acute dynamic stretch. Sci. Rep. 2020, 10, 7696. [Google Scholar] [CrossRef] [PubMed]
  5. Reichl, E.M.; Effler, J.C.; Robinson, D.N. The stress and strain of cytokinesis. Trends. Cell Biol. 2005, 15, 200–206. [Google Scholar] [CrossRef] [PubMed]
  6. Dalous, J.; Burghardt, E.; Müller-Taubenberger, A.; Bruckert, F.; Gerisch, G.; Bretschneider, T. Reversal of cell polarity and actin-myosin cytoskeleton reorganization under mechanical and chemical stimulation. Biophys. J. 2008, 94, 1063–1074. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Ananthakrishnan, R.; Ehrlicher, A. The forces behind cell movement. Int. J. Biol. Sci. 2007, 3, 303–317. [Google Scholar] [CrossRef] [PubMed]
  8. Pegoraro, A.F.; Janmey, P.; Weitz, D.A. Mechanical Properties of the Cytoskeleton and Cells. Cold Spring Harb. Perspect. Biol. 2017, 9, a022038. [Google Scholar] [CrossRef] [Green Version]
  9. Riento, K.; Ridley, A.J. ROCKs: Multifunctional kinases in cell behaviour. Nat. Rev. Mol. Cell Biol. 2003, 4, 446–456. [Google Scholar] [CrossRef]
  10. 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]
  11. Maekawa, M.; Ishizaki, T.; Boku, S.; Watanabe, N.; Fujita, A.; Iwamatsu, A.; Obinata, T.; Ohashi, K.; Mizuno, K.; Narumiya, S. Signaling from Rho to the actin cytoskeleton through protein kinases ROCK and LIM-kinase. Science 1999, 285, 895–898. [Google Scholar] [CrossRef]
  12. Amano, M.; Nakayama, M.; Kaibuchi, K. Rho-kinase/ROCK: A key regulator of the cytoskeleton and cell polarity. Cytoskeleton 2010, 67, 545–554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Schmandke, A.; Schmandke, A.; Strittmatter, S.M. ROCK and Rho: Biochemistry and neuronal functions of Rho-associated protein kinases. Neuroscientist 2007, 13, 454–469. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Gao, S.Y.; Li, C.Y.; Chen, J.; Pan, L.; Saito, S.; Terashita, T.; Saito, K.; Miyawaki, K.; Shigemoto, K.; Mominoki, K.; et al. Rho-ROCK signal pathway regulates microtubule-based process formation of cultured podocytes - Inhibition of ROCK promoted process elongation. Nephron Exp. Nephrol. 2004, 97, e49–e61. [Google Scholar] [CrossRef]
  15. Morgan-Fisher, M.; Wewer, U.M.; Yoneda, A. Regulation of ROCK activity in cancer. J. Histochem. Cytochem. 2013, 61, 185–198. [Google Scholar] [CrossRef] [Green Version]
  16. Lane, J.; Martin, T.A.; Watkins, G.; Mansel, R.E.; Jiang, W.G. The expression and prognostic value of ROCK I and ROCK II and their role in human breast cancer. Int. J. Oncol. 2008, 33, 585–593. [Google Scholar] [CrossRef] [Green Version]
  17. Liu, X.; Choy, E.; Hornicek, F.J.; Yang, S.; Yang, C.; Harmon, D.; Mankin, H.; Duan, Z. ROCK1 as a potential therapeutic target in osteosarcoma. J. Orthop. Res. 2011, 29, 1259–1266. [Google Scholar] [CrossRef] [PubMed]
  18. Wong, C.C.; Wong, C.M.; Tung, E.K.; Man, K.; Ng, I.O. Rho-kinase 2 is frequently overexpressed in hepatocellular carcinoma and involved in tumor invasion. Hepatology 2009, 49, 1583–1594. [Google Scholar] [CrossRef]
  19. Darling, E.M.; Di Carlo, D. High-Throughput Assessment of Cellular Mechanical Properties. Annu. Rev. Biomed. Eng. 2015, 17, 35–62. [Google Scholar] [CrossRef]
  20. Jiang, X.; Qin, Y.; Kun, L.; Zhou, Y. The Significant Role of the Microfilament System in Tumors. Front. Oncol. 2021, 11, 620390. [Google Scholar] [CrossRef]
  21. Falahzadeh, K.; Banaei-Esfahani, A.; Shahhoseini, M. The potential roles of actin in the nucleus. Cell Journal 2015, 17, 7–14. [Google Scholar] [PubMed]
  22. Gordón-Alonso, M.; Veiga, E.; Sánchez-Madrid, F. Actin dynamics at the immunological synapse. Cell Health Cytoskelet. 2010, 2, 33–47. [Google Scholar]
  23. Blessing, C.A.; Ugrinova, G.T.; Goodson, H.V. Actin and ARPs: Action in the nucleus. Trends Cell Biol. 2004, 14, 435–442. [Google Scholar] [CrossRef]
  24. Halpain, S. Actin in a supporting role. Nat. Neurosci. 2003, 6, 101–102. [Google Scholar] [CrossRef]
  25. Lanier, L.M.; Gertler, F.B. Actin cytoskeleton: Thinking globally, actin’ locally. Curr. Biol. 2000, 10, R655–R657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Pollard, T.D.; Blanchoin, L.; Mullins, R.D. Molecular mechanisms controlling actin filament dynamics in nonmuscle cells. Annu. Rev. Biophys. Biomol. Struct. 2000, 29, 545–576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Schmidt, A.; Hall, M.N. Signaling to the actin cytoskeleton. Annu. Rev. Cell Dev. Biol. 1998, 14, 305–338. [Google Scholar] [CrossRef] [PubMed]
  28. Bretscher, A. Microfilaments and membranes. Curr. Opin. Cell Biol. 1993, 5, 653–660. [Google Scholar] [CrossRef]
  29. Gül, D.; Habtemichael, N.; Dietrich, D.; Dietrich, J.; Gößwein, D.; Khamis, A.; Deuss, E.; Künzel, J.; Schneider, G.; Strieth, S.; et al. Identification of cytokeratin24 as a tumor suppressor for the management of head and neck cancer. Biol. Chem. 2022, 403, 869–890. [Google Scholar] [CrossRef]
  30. Jacob, J.T.; Coulombe, P.A.; Kwan, R.; Omary, M.B. Types I and II Keratin Intermediate Filaments. Cold Spring Harb. Perspect. Biol. 2018, 10, a018275. [Google Scholar] [CrossRef] [Green Version]
  31. Saitoh, T.; Sato, K.; Tonogi, M.; Tanaka, Y.; Yamane, G.Y. Expression of Cytokeratin 13, 14, 17, and 19 in 4-nitroquinoline-1-oxide-induced Oral Carcinogenesis in Rat. Bull. Tokyo Dent. Coll. 2016, 57, 241–251. [Google Scholar] [CrossRef] [PubMed]
  32. Schweizer, J.; Bowden, P.E.; Coulombe, P.A.; Langbein, L.; Lane, E.B.; Magin, T.M.; Maltais, L.; Omary, M.B.; Parry, D.A.; Rogers, M.A.; et al. New consensus nomenclature for mammalian keratins. J. Cell Biol. 2006, 174, 169–174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Honda, Y.; Koike, K.; Kubo, Y.; Masuko, S.; Arakawa, Y.; Ando, S. In vitro assembly properties of human type I and II hair keratins. Cell Struct. Funct. 2014, 39, 31–43. [Google Scholar] [CrossRef] [Green Version]
  34. Infante, C.; Ponce, M.; Asensio, E.; Zerolo, R.; Manchado, M. Molecular characterization of a novel type II keratin gene (sseKer3) in the Senegalese sole (Solea senegalensis): Differential expression of keratin genes by salinity. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 2011, 160, 15–23. [Google Scholar] [CrossRef]
  35. Moll, R.; Divo, M.; Langbein, L. The human keratins: Biology and pathology. Histochem. Cell Biol. 2008, 129, 705–733. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Bowden, P.E.; Quinlan, R.A.; Breitkreutz, D.; Fusenig, N.E. Proteolytic modification of acidic and basic keratins during terminal differentiation of mouse and human epidermis. Eur. J. Biochem. 1984, 142, 29–36. [Google Scholar] [CrossRef] [PubMed]
  37. Ridge, K.M.; Eriksson, J.E.; Pekny, M.; Goldman, R.D. Roles of vimentin in health and disease. Genes Dev. 2022, 36, 391–407. [Google Scholar] [CrossRef]
  38. Yue, Q.; Feng, L.; Cao, B.; Liu, M.; Zhang, D.; Wu, W.; Jiang, B.; Yang, M.; Liu, X.; Guo, D. Proteomic analysis revealed the important role of vimentin in human cervical carcinoma HeLa cells treated with gambogic acid. Mol. Cell. Proteom. 2016, 15, 26–44. [Google Scholar] [CrossRef] [Green Version]
  39. Kidd, M.E.; Shumaker, D.K.; Ridge, K.M. The role of Vimentin intermediate filaments in the progression of lung cancer. Am. J. Respir. Cell Mol. Biol. 2014, 50, 1–6. [Google Scholar] [CrossRef] [Green Version]
  40. Menko, A.S.; Bleaken, B.M.; Libowitz, A.A.; Zhang, L.; Stepp, M.A.; Walker, J.L. A central role for vimentin in regulating repair function during healing of the lens epithelium. Mol. Biol. Cell 2014, 25, 776–790. [Google Scholar] [CrossRef]
  41. Herrmann, H.; Bär, H.; Kreplak, L.; Strelkov, S.V.; Aebi, U. Intermediate filaments: From cell architecture to nanomechanics. Nat. Rev. Mol. Cell Biol. 2007, 8, 562–573. [Google Scholar] [CrossRef] [PubMed]
  42. Wang, R.; Li, Q.; Tang, D.D. Role of vimentin in smooth muscle force development. Am. J. Physiol. Cell Physiol. 2006, 291, C483–C489. [Google Scholar] [CrossRef] [PubMed]
  43. Didonna, A.; Opal, P. The role of neurofilament aggregation in neurodegeneration: Lessons from rare inherited neurological disorders. Mol. Neurodegener. 2019, 14, 19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Yuan, A.; Nixon, R.A. Specialized roles of neurofilament proteins in synapses: Relevance to neuropsychiatric disorders. Brain Res. Bull. 2016, 126, 334–346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Yuan, A.; Rao, M.V.; Veeranna; Nixon, R.A. A. Neurofilaments at a glance. J. Cell Sci. 2012, 125, 3257–3263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Lin, H.; Schlaepfer, W.W. Role of neurofilament aggregation in motor neuron disease. Ann. Neurol. 2006, 60, 399–406. [Google Scholar] [CrossRef]
  47. Hoffman, P.N. Distinct roles of neurofilament and tubulin gene expression in axonal growth. Ciba Found. Symp. 1988, 138, 192–204. [Google Scholar]
  48. Khadija, S.G.; Chen, F.; Hadden, T.; Commissaris, R.L.; Kowluru, A. Biology and regulatory roles of nuclear lamins in cellular function and dysfunction. Recent Pat. Endocr. Metab. Immune Drug Discov. 2015, 9, 111–120. [Google Scholar] [CrossRef]
  49. Carmosino, M.; Torretta, S.; Procino, G.; Gerbino, A.; Forleo, C.; Favale, S.; Svelto, M. Role of nuclear Lamin A/C in cardiomyocyte functions. Biol. Cell 2014, 106, 346–358. [Google Scholar] [CrossRef]
  50. Burke, B.; Stewart, C.L. The nuclear lamins: Flexibility in function. Nat. Rev. Mol. Cell Biol. 2013, 14, 13–24. [Google Scholar] [CrossRef]
  51. Lopez-Soler, R.I.; Moir, R.D.; Spann, T.P.; Stick, R.; Goldman, R.D. A role for nuclear lamins in nuclear envelope assembly. J. Cell Biol. 2001, 154, 61–70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Parnaik, V.K. Role of Nuclear Lamins in Nuclear Organization, Cellular Signaling, and Inherited Diseases. Int. Rev. Cell Mol. Biol. 2008, 266, 157–206. [Google Scholar] [CrossRef]
  53. Oka, M.; Kudo, H.; Sugama, N.; Asami, Y.; Takehana, M. The function of filensin and phakinin in lens transparency. Mol. Vis. 2008, 14, 815–822. [Google Scholar] [PubMed]
  54. Pittenger, J.T.; Hess, J.F.; FitzGerald, P.G. Identifying the role of specific motifs in the lens fiber cell-specific intermediate filament phakosin. Invest. Ophthalmol. Vis. Sci. 2007, 48, 5132–5141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Blankenship, T.N.; Hess, J.F.; FitzGerald, P.G. Development- and differentiation-dependent reorganization of intermediate filaments in fiber cells. Investig. Ophthalmol. Vis. Sci. 2001, 42, 735–742. [Google Scholar] [PubMed]
  56. Goulielmos, G.; Remington, S.; Schwesinger, F.; Georgatos, S.D.; Gounari, F. Contributions of the structural domains of filensin in polymer formation and filament distribution. J. Cell Sci. 1996, 109, 447–456. [Google Scholar] [CrossRef] [PubMed]
  57. Georgatos, S.D.; Gounari, F.; Remington, S. The beaded intermediate filaments and their potential functions in eye lens. Bioessays 1994, 16, 413–418. [Google Scholar] [CrossRef]
  58. Hlavaty, D.; Lechler, T. Roles for microtubules in the proliferative and differentiated cells of stratified epithelia. Curr. Opin. Cell Biol. 2021, 68, 98–104. [Google Scholar] [CrossRef]
  59. Chang, W.; Gu, J.G. Role of microtubules in Piezo2 mechanotransduction of mouse Merkel cells. J. Neurophysiol. 2020, 124, 1824–1831. [Google Scholar] [CrossRef]
  60. Matis, M. The Mechanical Role of Microtubules in Tissue Remodeling. Bioessays 2020, 42, 1900244. [Google Scholar] [CrossRef] [Green Version]
  61. Logan, C.M.; Menko, A.S. Microtubules: Evolving roles and critical cellular interactions. Exp. Biol. Med. 2019, 244, 1240–1254. [Google Scholar] [CrossRef]
  62. Cirillo, L.; Gotta, M.; Meraldi, P. The elephant in the room: The role of microtubules in cancer. Adv. Exp. Med. Biol. 2017, 1002, 93–124. [Google Scholar]
  63. Ganguly, A.; Yang, H.; Sharma, R.; Patel, K.D.; Cabral, F. The role of microtubules and their dynamics in cell migration. J. Biol. Chem. 2012, 287, 43359–43369. [Google Scholar] [CrossRef] [PubMed]
  64. van der Vaart, B.; Akhmanova, A.; Straube, A. Regulation of microtubule dynamic instability. Biochem. Soc. Trans. 2009, 37, 1007–1013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Kulić, I.M.; Brown, A.E.X.; Kim, H.; Kural, C.; Blehm, B.; Selvin, P.R.; Nelson, P.C.; Gelfand, V.I. The role of microtubule movement in bidirectional organelle transport. Proc. Natl. Acad. Sci. USA 2008, 105, 10011–10016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Morse, D.C. Viscoelasticity of Concentrated Isotropic Solutions of Semiflexible Polymers. 1. Model and Stress Tensor. Macromolecules 1998, 31, 7030–7043. [Google Scholar] [CrossRef]
  67. Wen, Q.; Janmey, P.A. Polymer physics of the cytoskeleton. Curr. Opin. Solid. State Mater. Sci. 2011, 15, 177–182. [Google Scholar] [CrossRef] [Green Version]
  68. Wen, Q.; Janmey, P.A. Effects of non-linearity on cell-ECM interactions. Exp. Cell Res. 2013, 319, 2481–2489. [Google Scholar] [CrossRef] [Green Version]
  69. Gittes, F.; Mickey, B.; Nettleton, J.; Howard, J. Flexural rigidity of microtubules and actin filaments measured from thermal fluctuations in shape. J. Cell. Biol. 1993, 120, 923–934. [Google Scholar] [CrossRef]
  70. De La Cruz, E.M.; Gardel, M.L. Actin Mechanics and Fragmentation. J. Biol. Chem. 2015, 290, 17137–17144. [Google Scholar] [CrossRef] [Green Version]
  71. Cooper, G.M. The Cell: A Molecular Approach, 6th ed.; Sinauer Associates: Sunderland, MA, USA, 2013. [Google Scholar]
  72. De La Cruz, E.M.; Roland, J.; McCullough, B.R.; Blanchoin, L.; Martiel, J.L. Origin of twist-bend coupling in actin filaments. Biophys. J. 2010, 99, 1852–1860. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Atkinson, S.J.; Hosford, M.A.; Molitoris, B.A. Mechanism of Actin Polymerization in Cellular ATP Depletion*. J. Biol. Chem. 2004, 279, 5194–5199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Kudryashov, D.S.; Reisler, E. ATP and ADP actin states. Biopolymers 2013, 99, 245–256. [Google Scholar] [CrossRef]
  75. Janmey, P.A.; Hvidt, S.; Oster, G.F.; Lamb, J.; Stossel, T.P.; Hartwig, J.H. Effect of ATP on actin filament stiffness. Nature 1990, 347, 95–99. [Google Scholar] [CrossRef] [PubMed]
  76. Brieher, W. Mechanisms of actin disassembly. Mol. Biol. Cell 2013, 24, 2299–2302. [Google Scholar] [CrossRef] [PubMed]
  77. Wegner, A. Treadmilling of actin at physiological salt concentrations: An analysis of the critical concentrations of actin filaments. J. Mol. Biol. 1982, 161, 607–615. [Google Scholar] [CrossRef] [PubMed]
  78. Janmey, P.A.; Miller, R.T. Mechanisms of mechanical signaling in development and disease. J. Cell Sci. 2011, 124, 9–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Xu, J.; Schwarz, W.H.; Käs, J.A.; Stossel, T.P.; Janmey, P.A.; Pollard, T.D. Mechanical Properties of Actin Filament Networks Depend on Preparation, Polymerization Conditions, and Storage of Actin Monomers. Biophys. J. 1998, 74, 2731–2740. [Google Scholar] [CrossRef] [Green Version]
  80. Haase, K.; Pelling, A.E. The role of the actin cortex in maintaining cell shape. Commun. Integr. Biol. 2013, 6, e26714. [Google Scholar] [CrossRef] [Green Version]
  81. Qin, Z.; Kreplak, L.; Buehler, M.J. Hierarchical structure controls nanomechanical properties of vimentin intermediate filaments. PLoS ONE 2009, 4, e7294. [Google Scholar] [CrossRef]
  82. Qin, Z.; Buehler, M.J. Flaw tolerance of nuclear intermediate filament lamina under extreme mechanical deformation. ACS Nano 2011, 5, 3034–3042. [Google Scholar] [CrossRef] [Green Version]
  83. Hesse, M.; Magin, T.M.; Weber, K. Genes for intermediate filament proteins and the draft sequence of the human genome: Novel keratin genes and a surprisingly high number of pseudogenes related to keratin genes 8 and 18. J. Cell Sci. 2001, 114, 2569–2575. [Google Scholar] [CrossRef] [PubMed]
  84. Strnad, P.; Stumptner, C.; Zatloukal, K.; Denk, H. Intermediate filament cytoskeleton of the liver in health and disease. Histochem. Cell Biol. 2008, 129, 735–749. [Google Scholar] [CrossRef]
  85. Sanghvi-Shah, R.; Weber, G.F. Intermediate Filaments at the Junction of Mechanotransduction, Migration, and Development. Front. Cell Dev. Biol. 2017, 5, 81. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Oshima, R.G. Intermediate filaments: A historical perspective. Exp. Cell Res. 2007, 313, 1981–1994. [Google Scholar] [CrossRef] [Green Version]
  87. Karantza, V. Keratins in health and cancer: More than mere epithelial cell markers. Oncogene 2011, 30, 127–138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Huang, T.L.; Chou, C.C. Effect of mutations on the hydrophobic interactions of the hierarchical molecular structure and mechanical properties of epithelial keratin 1/10. Int. J. Biol. Macromol. 2022, 212, 442–450. [Google Scholar] [CrossRef]
  89. Lowery, J.; Kuczmarski, E.R.; Herrmann, H.; Goldman, R.D. Intermediate Filaments Play a Pivotal Role in Regulating Cell Architecture and Function. J. Biol. Chem. 2015, 290, 17145–17153. [Google Scholar] [CrossRef] [Green Version]
  90. FitzGerald, P.; Sun, N.; Shibata, B.; Hess, J.F. Expression of the type VI intermediate filament proteins CP49 and filensin in the mouse lens epithelium. Mol. Vis. 2016, 22, 970–989. [Google Scholar]
  91. Fudge, D.; Russell, D.; Beriault, D.; Moore, W.; Lane, E.B.; Vogl, A.W. The intermediate filament network in cultured human keratinocytes is remarkably extensible and resilient. PLoS ONE 2008, 3, e2327. [Google Scholar] [CrossRef] [Green Version]
  92. Nolting, J.-F.; Möbius, W.; Köster, S. Mechanics of individual keratin bundles in living cells. Biophys. J. 2014, 107, 2693–2699. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Ma, L.; Xu, J.; Coulombe, P.A.; Wirtz, D. Keratin Filament Suspensions Show Unique Micromechanical Properties *. J. Biol. Chem. 1999, 274, 19145–19151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Beil, M.; Micoulet, A.; von Wichert, G.; Paschke, S.; Walther, P.; Omary, M.B.; Van Veldhoven, P.P.; Gern, U.; Wolff-Hieber, E.; Eggermann, J.; et al. Sphingosylphosphorylcholine regulates keratin network architecture and visco-elastic properties of human cancer cells. Nat. Cell Biol. 2003, 5, 803–811. [Google Scholar] [CrossRef]
  95. Janke, C.; Magiera, M.M. The tubulin code and its role in controlling microtubule properties and functions. Nat. Rev. Mol. Cell Biol. 2020, 21, 307–326. [Google Scholar] [CrossRef]
  96. Roll-Mecak, A. The Tubulin Code in Microtubule Dynamics and Information Encoding. Dev. Cell 2020, 54, 7–20. [Google Scholar] [CrossRef]
  97. Tubulins. HGNC Database, HUGO Gene Nomenclature Committee (HGNC), European Molecular Biology Laboratory, European Bi-oinformatics Institute (EMBL-EBI), Wellcome Genome Campus, Hinxton, Cambridge CB10 1SD, United Kingdom. 2022. Available online: https://www.genenames.org/data/genegroup/#!/group/778. (accessed on 21 December 2022).
  98. Sept, D.; MacKintosh, F.C. Microtubule elasticity: Connecting all-atom simulations with continuum mechanics. Phys. Rev. Lett. 2010, 104, 018101. [Google Scholar] [CrossRef] [Green Version]
  99. Schaap, I.A.T.; Carrasco, C.; de Pablo, P.J.; MacKintosh, F.C.; Schmidt, C.F. Elastic Response, Buckling, and Instability of Microtubules under Radial Indentation. Biophys. J. 2006, 91, 1521–1531. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Avila, J. Microtubule functions. Life Sci. 1992, 50, 327–334. [Google Scholar] [CrossRef] [PubMed]
  101. Kent, I.A.; Lele, T.P. Microtubule-based force generation. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2017, 9, e1428. [Google Scholar] [CrossRef] [Green Version]
  102. Akhmanova, A.; Steinmetz, M.O. Control of microtubule organization and dynamics: Two ends in the limelight. Nat. Rev. Mol. Cell Biol. 2015, 16, 711–726. [Google Scholar] [CrossRef]
  103. van Haren, J.; Wittmann, T. Microtubule Plus End Dynamics—Do We Know How Microtubules Grow?: Cells boost microtubule growth by promoting distinct structural transitions at growing microtubule ends. Bioessays 2019, 41, e1800194. [Google Scholar] [CrossRef]
  104. MacRae, T.H. Towards an understanding of microtubule function and cell organization: An overview. Biochem. Cell Biol. 1992, 70, 835–841. [Google Scholar] [CrossRef] [PubMed]
  105. Zwetsloot, A.J.; Tut, G.; Straube, A. Measuring microtubule dynamics. Essays Biochem. 2018, 62, 725–735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Sept, D.; Baker, N.A.; McCammon, J.A. The physical basis of microtubule structure and stability. Protein Sci. 2003, 12, 2257–2261. [Google Scholar] [CrossRef]
  107. McIntosh, J.R.; Grishchuk, E.L.; West, R.R. Chromosome-microtubule interactions during mitosis. Annu. Rev. Cell Dev. Biol. 2002, 18, 193–219. [Google Scholar] [CrossRef]
  108. Xiao, H.; Verdier-Pinard, P.; Fernandez-Fuentes, N.; Burd, B.; Angeletti, R.; Fiser, A.; Horwitz, S.B.; Orr, G.A. Insights into the mechanism of microtubule stabilization by Taxol. Proc. Natl. Acad. Sci. USA 2006, 103, 10166–10173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Müsch, A. Microtubule Organization and Function in Epithelial Cells. Traffic 2004, 5, 896–903. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Lai, S.-L.; Chien, A.J.; Moon, R.T. Wnt/Fz signaling and the cytoskeleton: Potential roles in tumorigenesis. Cell Res. 2009, 19, 532–545. [Google Scholar] [CrossRef]
  111. Hoffman, L.; Jensen, C.C.; Yoshigi, M.; Beckerle, M. Mechanical signals activate p38 MAPK pathway-dependent reinforcement of actin via mechanosensitive HspB1. Mol. Biol. Cell 2017, 28, 2661–2675. [Google Scholar] [CrossRef]
  112. Sauzeau, V.; Le Jeune, H.; Cario-Toumaniantz, C.; Smolenski, A.; Lohmann, S.M.; Bertoglio, J.; Chardin, P.; Pacaud, P.; Loirand, G. Cyclic GMP-dependent Protein Kinase Signaling Pathway Inhibits RhoA-induced Ca2+ Sensitization of Contraction in Vascular Smooth Muscle*. J. Biol. Chem. 2000, 275, 21722–21729. [Google Scholar] [CrossRef] [Green Version]
  113. Zou, L.; Zhang, J.; Han, J.; Li, W.; Su, F.; Xu, X.; Zhai, Z.; Xiao, F. cGMP interacts with tropomyosin and downregulates actin-tropomyosin-myosin complex interaction. Respir. Res. 2018, 19, 201. [Google Scholar] [CrossRef] [Green Version]
  114. James, R.G.; Conrad, W.H.; Moon, R.T. Beta-catenin-independent Wnt pathways: Signals, core proteins, and effectors. Methods Mol. Biol. 2008, 468, 131–144. [Google Scholar] [CrossRef]
  115. Roarty, K.; Pfefferle, A.D.; Creighton, C.J.; Perou, C.M.; Rosen, J.M. Ror2-mediated alternative Wnt signaling regulates cell fate and adhesion during mammary tumor progression. Oncogene 2017, 36, 5958–5968. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Beurel, E.; Grieco, S.F.; Jope, R.S. Glycogen synthase kinase-3 (GSK3): Regulation, actions, and diseases. Pharmacol. Ther. 2015, 148, 114–131. [Google Scholar] [CrossRef] [PubMed]
  117. Deng, S.; Leong, H.C.; Datta, A.; Gopal, V.; Kumar, A.P.; Yap, C.T. PI3K/AKT Signaling Tips the Balance of Cytoskeletal Forces for Cancer Progression. Cancers 2022, 14, 1652. [Google Scholar] [CrossRef]
  118. May-Simera, H.L.; Kelley, M.W. Cilia, Wnt signaling, and the cytoskeleton. Cilia 2012, 1, 7. [Google Scholar] [CrossRef] [Green Version]
  119. Shapiro, L. The multi-talented beta-catenin makes its first appearance. Structure 1997, 5, 1265–1268. [Google Scholar] [CrossRef] [Green Version]
  120. Shi, J.; Wei, L. Rho Kinases in Embryonic Development and Stem Cell Research. Arch. Immunol. Ther. Exp. 2022, 70, 4. [Google Scholar] [CrossRef]
  121. Tang, L.; Dai, F.; Liu, Y.; Yu, X.; Huang, C.; Wang, Y.; Yao, W. RhoA/ROCK signaling regulates smooth muscle phenotypic modulation and vascular remodeling via the JNK pathway and vimentin cytoskeleton. Pharmacol. Res. 2018, 133, 201–212. [Google Scholar] [CrossRef]
  122. Schofield, A.V.; Bernard, O. Rho-associated coiled-coil kinase (ROCK) signaling and disease. Crit. Rev. Biochem. Mol. Biol. 2013, 48, 301–316. [Google Scholar] [CrossRef]
  123. Sit, S.-T.; Manser, E. Rho GTPases and their role in organizing the actin cytoskeleton. J. Cell Sci. 2011, 124, 679–683. [Google Scholar] [CrossRef] [Green Version]
  124. Sarasa-Renedo, A.; Tunç-Civelek, V.; Chiquet, M. Role of RhoA/ROCK-dependent actin contractility in the induction of tenascin-C by cyclic tensile strain. Exp. Cell Res. 2006, 312, 1361–1370. [Google Scholar] [CrossRef]
  125. Woods, A.; Wang, G.; Beier, F. RhoA/ROCK signaling regulates Sox9 expression and actin organization during chondrogenesis. J. Biol. Chem. 2005, 280, 11626–11634. [Google Scholar] [CrossRef] [Green Version]
  126. Da Silva, J.S.; Medina, M.; Zuliani, C.; Di Nardo, A.; Witke, W.; Dotti, C.G. RhoA/ROCK regulation of neuritogenesis via profilin IIa-mediated control of actin stability. J. Cell Biol. 2003, 162, 1267–1279. [Google Scholar] [CrossRef]
  127. Amano, M.; Fukata, Y.; Kaibuchi, K. Regulation and functions of Rho-associated kinase. Exp. Cell Res. 2000, 261, 44–51. [Google Scholar] [CrossRef]
  128. Yang, L.; Tang, L.; Dai, F.; Meng, G.; Yin, R.; Xu, X.; Yao, W. Raf-1/CK2 and RhoA/ROCK signaling promote TNF-α-mediated endothelial apoptosis via regulating vimentin cytoskeleton. Toxicology 2017, 389, 74–84. [Google Scholar] [CrossRef]
  129. Lei, S.; Tian, Y.P.; Xiao, W.D.; Li, S.; Rao, X.C.; Zhang, J.L.; Yang, J.; Hu, X.M.; Chen, W. ROCK is Involved in Vimentin Phosphorylation and Rearrangement Induced by Dengue Virus. Cell Biochem. Biophys. 2013, 67, 1333–1342. [Google Scholar] [CrossRef] [Green Version]
  130. Hirose, M.; Ishizaki, T.; Watanabe, N.; Uehata, M.; Kranenburg, O.; Moolenaar, W.H.; Matsumura, F.; Maekawa, M.; Bito, H.; Narumiya, S. Molecular dissection of the Rho-associated protein kinase (p160ROCK)- regulated neurite remodeling in neuroblastoma N1E-115 cells. J. Cell Biol. 1998, 141, 1625–1636. [Google Scholar] [CrossRef] [Green Version]
  131. Becker, K.N.; Pettee, K.M.; Sugrue, A.; Reinard, K.A.; Schroeder, J.L.; Eisenmann, K.M. The Cytoskeleton Effectors Rho-Kinase (ROCK) and Mammalian Diaphanous-Related (mDia) Formin Have Dynamic Roles in Tumor Microtube Formation in Invasive Glioblastoma Cells. Cells 2022, 11, 1559. [Google Scholar] [CrossRef]
  132. Heng, Y.W.; Lim, H.H.; Mina, T.; Utomo, P.; Zhong, S.; Lim, C.T.; Koh, C.G. TPPP acts downstream of RhoA-ROCK-LIMK2 to regulate astral microtubule organization and spindle orientation. J. Cell Sci. 2012, 125, 1579–1590. [Google Scholar] [CrossRef] [Green Version]
  133. Fonseca, A.V.; Freund, D.; Bornhäuser, M.; Corbeil, D. Polarization and migration of hematopoietic stem and progenitor cells rely on the RhoA/ROCK I pathway and an active reorganization of the microtubule network. J. Biol. Chem. 2010, 285, 31661–31671. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Takesono, A.; Heasman, S.J.; Wojciak-Stothard, B.; Garg, R.; Ridley, A.J. Microtubules regulate migratory polarity through Rho/ ROCK signaling in T cells. PLoS ONE 2010, 5, e8774. [Google Scholar] [CrossRef]
  135. Han, D.; Sun, J.; Fan, D.; Zhang, C.; Du, S.; Zhang, W. Simvastatin ameliorates oxygen glucose deprivation/reoxygenation-induced pulmonary endothelial barrier dysfunction by restoring cell-cell junctions and actin cytoskeleton dynamics via the PI3K/Akt signaling pathway. Am. J. Transl. Res. 2020, 12, 5586–5596. [Google Scholar] [PubMed]
  136. Lien, E.C.; Dibble, C.C.; Toker, A. PI3K signaling in cancer: Beyond AKT. Curr. Opin. Cell Biol. 2017, 45, 62–71. [Google Scholar] [CrossRef] [PubMed]
  137. Kakinuma, N.; Roy, B.C.; Zhu, Y.; Wang, Y.; Kiyama, R. Kank regulates RhoA-dependent formation of actin stress fibers and cell migration via 14-3-3 in PI3K-Akt signaling. J. Cell Biol. 2008, 181, 537–549. [Google Scholar] [CrossRef]
  138. Qian, Y.; Zhong, X.; Flynn, D.C.; Zheng, J.Z.; Qiao, M.; Wu, C.; Dedhar, S.; Shi, X.; Jiang, B.H. ILK mediates actin filament rearrangements and cell migration and invasion through PI3K/Akt/Rac1 signaling. Oncogene 2005, 24, 3154–3165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Qian, Y.; Corum, L.; Meng, Q.; Blenis, J.; Zheng, J.Z.; Shi, X.; Flynn, D.C.; Jiang, B.H. PI3K induced actin filament remodeling through Akt and p70S6K1: Implication of essential role in cell migration. Am. J. Physiol. Cell Physiol. 2004, 286, C153–C163. [Google Scholar] [CrossRef]
  140. Krasilnikov, M.A. Phosphatidylinositol-3 kinase dependent pathways: The role in control of cell growth, survival, and malignant transformation. Biochemistry 2000, 65, 59–67. [Google Scholar]
  141. Roux, A.; Loranger, A.; Lavoie, J.N.; Marceau, N. Keratin 8/18 regulation of insulin receptor signaling and trafficking in hepatocytes through a concerted phosphoinositide-dependent Akt and Rab5 modulation. FASEB J. 2017, 31, 3555–3573. [Google Scholar] [CrossRef] [Green Version]
  142. Wang, R.C.; Wei, Y.; An, Z.; Zou, Z.; Xiao, G.; Bhagat, G.; White, M.; Reichelt, J.; Levine, B. Akt-mediated regulation of autophagy and tumorigenesis through Beclin 1 phosphorylation. Science 2012, 338, 956–959. [Google Scholar] [CrossRef] [Green Version]
  143. Kong, L.; Schäfer, G.; Bu, H.; Zhang, Y.; Zhang, Y.; Klocker, H. Lamin A/C protein is overexpressed in tissue-invading prostate cancer and promotes prostate cancer cell growth, migration and invasion through the PI3K/AKT/PTEN pathway. Carcinogenesis 2012, 33, 751–759. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Tseng, Y.H.; Yang, C.C.; Lin, S.C.; Cheng, C.C.; Lin, S.H.; Liu, C.J.; Chang, K.W. Areca nut extract upregulates vimentin by activating PI3K/AKT signaling in oral carcinoma. J. Oral Pathol. Med. 2011, 40, 160–166. [Google Scholar] [CrossRef] [PubMed]
  145. Chakrabarty, S.; Nag, D.; Ganguli, A.; Das, A.; Ghosh Dastidar, D.; Chakrabarti, G. Theaflavin and epigallocatechin-3-gallate synergistically induce apoptosis through inhibition of PI3K/Akt signaling upon depolymerizing microtubules in HeLa cells. J. Cell. Biochem. 2019, 120, 5987–6003. [Google Scholar] [CrossRef] [PubMed]
  146. Fu, Y.F.; Liu, X.; Gao, M.; Zhang, Y.N.; Liu, J. Endoplasmic reticulum stress induces autophagy and apoptosis while inhibiting proliferation and drug resistance in multiple myeloma through the PI3K/Akt/mTOR signaling pathway. Oncotarget 2017, 8, 61093–61106. [Google Scholar] [CrossRef] [Green Version]
  147. Kitagishi, Y.; Nakanishi, A.; Ogura, Y.; Matsuda, S. Dietary regulation of PI3K/AKT/GSK-3β pathway in Alzheimer’s disease. Alzheimers Res. Ther. 2014, 6, 35. [Google Scholar] [CrossRef]
  148. Onishi, K.; Higuchi, M.; Asakura, T.; Masuyama, N.; Gotoh, Y. The PI3K-Akt pathway promotes microtubule stabilization in migrating fibroblasts. Genes Cells 2007, 12, 535–546. [Google Scholar] [CrossRef]
  149. Fujiwara, Y.; Hosokawa, Y.; Watanabe, K.; Tanimura, S.; Ozaki, K.I.; Kohno, M. Blockade of the phosphatidylinositol-3-kinase-Akt signaling pathway enhances the induction of apoptosis by microtubule-destabilizing agents in tumor cells in which the pathway is constitutively activated. Mol. Cancer Ther. 2007, 6, 1133–1142. [Google Scholar] [CrossRef] [Green Version]
  150. Zhang, J.; Hu, Q.; Jiang, X.; Wang, S.; Zhou, X.; Lu, Y.; Huang, X.; Duan, H.; Zhang, T.; Ge, H.; et al. Actin Alpha 2 Downregulation Inhibits Neural Stem Cell Proliferation and Differentiation into Neurons through Canonical Wnt/β-Catenin Signaling Pathway. Oxid. Med. Cell. Longev. 2022, 2022, 7486726. [Google Scholar] [CrossRef]
  151. Galli, C.; Piemontese, M.; Lumetti, S.; Ravanetti, F.; MacAluso, G.M.; Passeri, G. Actin cytoskeleton controls activation of Wnt/β-catenin signaling in mesenchymal cells on implant surfaces with different topographies. Acta Biomater. 2012, 8, 2963–2968. [Google Scholar] [CrossRef]
  152. Lehmann, M.; Hu, Q.; Hu, Y.; Hafner, K.; Costa, R.; van den Berg, A.; Königshoff, M. Chronic WNT/β-catenin signaling induces cellular senescence in lung epithelial cells. Cell. Signal. 2020, 70, 109588. [Google Scholar] [CrossRef]
  153. Tian, Z.; Zhang, X.; Zhao, Z.; Zhang, F.; Deng, T. The Wnt/β-catenin signaling pathway affects the distribution of cytoskeletal proteins in Aβ treated PC12 cells. J. Integr. Neurosci. 2019, 18, 309–312. [Google Scholar] [CrossRef] [PubMed]
  154. Bermeo, S.; Vidal, C.; Zhou, H.; Duque, G. Lamin A/C Acts as an Essential Factor in Mesenchymal Stem Cell Differentiation Through the Regulation of the Dynamics of the Wnt/β-Catenin Pathway. J. Cell. Biochem. 2015, 116, 2344–2353. [Google Scholar] [CrossRef] [PubMed]
  155. Prasad, C.P.; Mirza, S.; Sharma, G.; Prashad, R.; DattaGupta, S.; Rath, G.; Ralhan, R. Epigenetic alterations of CDH1 and APC genes: Relationship with activation of Wnt/β-catenin Pathway in invasive ductal carcinoma of breast. Life Sci. 2008, 83, 318–325. [Google Scholar] [CrossRef]
  156. Bierie, B.; Nozawa, M.; Renou, J.P.; Shillingford, J.M.; Morgan, F.; Oka, T.; Taketo, M.M.; Cardiff, R.D.; Miyoshi, K.; Wagner, K.U.; et al. Activation of β-catenin in prostate epithelium induces hyperplasias and squamous transdifferentiation. Oncogene 2003, 22, 3875–3887. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Puri, D.; Ponniah, K.; Biswas, K.; Basu, A.; Dey, S.; Lundquist, E.A.; Ghosh-Roy, A. Wnt signaling establishes the microtubule polarity in neurons through regulation of kinesin-13. J. Cell Biol. 2021, 220, e202005080. [Google Scholar] [CrossRef] [PubMed]
  158. Ou, D.; Chen, L.; He, J.; Rong, Z.; Gao, J.; Li, Z.; Liu, L.; Tang, F.; Li, J.; Deng, Y.; et al. CDK11 negatively regulates Wnt/β-catenin signaling in the endosomal compartment by affecting microtubule stability. Cancer Biol. Med. 2020, 17, 328–342. [Google Scholar] [CrossRef]
  159. Huang, P.; Senga, T.; Hamaguchi, M. A novel role of phospho-β-catenin in microtubule regrowth at centrosome. Oncogene 2007, 26, 4357–4371. [Google Scholar] [CrossRef] [Green Version]
  160. Ciani, L.; Krylova, O.; Smalley, M.J.; Dale, T.C.; Salinas, P.C. A divergent canonical WNT-signaling pathway regulates microtubule dynamics: Dishevelled signals locally to stabilize microtubules. J. Cell Biol. 2004, 164, 243–253. [Google Scholar] [CrossRef] [Green Version]
  161. Peifer, M.; Polakis, P. Wnt signaling in oncogenesis and embryogenesis—A look outside the nucleus. Science 2000, 287, 1606–1609. [Google Scholar] [CrossRef]
  162. Hajka, D.; Budziak, B.; Pietras, Ł.; Duda, P.; McCubrey, J.A.; Gizak, A. GSK3 as a Regulator of Cytoskeleton Architecture: Consequences for Health and Disease. Cells 2021, 10, 2092. [Google Scholar] [CrossRef]
  163. Yoshino, Y.; Suzuki, M.; Takahashi, H.; Ishioka, C. Inhibition of invasion by glycogen synthase kinase-3 beta inhibitors through dysregulation of actin re-organisation via down-regulation of WAVE2. Biochem. Biophys. Res. Commun. 2015, 464, 275–280. [Google Scholar] [CrossRef] [PubMed]
  164. Watanabe, T.; Noritake, J.; Kakeno, M.; Matsui, T.; Harada, T.; Wang, S.; Itoh, N.; Sato, K.; Matsuzawa, K.; Iwamatsu, A.; et al. Phosphorylation of CLASP2 by GSK-3β regulates its interaction with IQGAP1, EB1 and microtubules. J. Cell Sci. 2009, 122, 2969–2979. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Sun, T.; Rodriguez, M.; Kim, L. Glycogen synthase kinase 3 in the world of cell migration. Dev. Growth Differ. 2009, 51, 735–742. [Google Scholar] [CrossRef]
  166. Vaidya, R.J.; Ray, R.M.; Johnson, L.R. Akt-mediated GSK-3β inhibition prevents migration of polyamine-depleted intestinal epithelial cells via Rac1. Cell. Mol. Life Sci. 2006, 63, 2871–2879. [Google Scholar] [CrossRef] [PubMed]
  167. Sen, S.; Lagas, S.; Roy, A.; Kumar, H. Cytoskeleton saga: Its regulation in normal physiology and modulation in neurodegenerative disorders. Eur. J. Pharmacol. 2022, 925, 175001. [Google Scholar] [CrossRef] [PubMed]
  168. Lee, W.C.; Kan, D.; Chen, Y.Y.; Han, S.K.; Lu, K.S.; Chien, C.L. Suppression of extensive neurofilament phosphorylation rescues α-internexin/peripherin-overexpressing PC12 cells from neuronal cell death. PLoS ONE 2012, 7, e43883. [Google Scholar] [CrossRef] [PubMed]
  169. Kim, J.Y.; Kim, Y.M.; Yang, C.H.; Cho, S.K.; Lee, J.W.; Cho, M. Functional regulation of Slug/Snail2 is dependent on GSK-3β-mediated phosphorylation. FEBS J. 2012, 279, 2929–2939. [Google Scholar] [CrossRef]
  170. Sasaki, T.; Taoka, M.; Ishiguro, K.; Uchida, A.; Saito, T.; Toshiaki Isobe, D.; Hisanaga, S.I. In vivo and in vitro phosphorylation at Ser-493 in the glutamate (E)-segment of neurofilament-H subunit by glycogen synthase kinase 3β. J. Biol. Chem. 2002, 277, 36032–36039. [Google Scholar] [CrossRef] [Green Version]
  171. Guidato, S.; Tsai, L.H.; Woodgett, J.; Miller, C.C.J. Differential cellular phosphorylation of neurofilament heavy side-arms by glycogen synthase kinase-3 and cydin-dependent kinase-5. J. Neurochem. 1996, 66, 1698–1706. [Google Scholar] [CrossRef]
  172. Fumoto, K.; Hoogenraad, C.C.; Kikuchi, A. GSK-3β-regulated interaction of BICD with dynein is involved in microtubule anchorage at centrosome. EMBO J. 2006, 25, 5670–5682. [Google Scholar] [CrossRef] [Green Version]
  173. Grimes, C.A.; Jope, R.S. The multifaceted roles of glycogen synthase kinase 3β in cellular signaling. Prog. Neurobiol. 2001, 65, 391–426. [Google Scholar] [CrossRef]
  174. Joe, S.Y.; Yang, S.G.; Lee, J.H.; Park, H.J.; Koo, D.B. Stabilization of F-Actin Cytoskeleton by Paclitaxel Improves the Blastocyst Developmental Competence through P38 MAPK Activity in Porcine Embryos. Biomedicines 2022, 10, 1867. [Google Scholar] [CrossRef] [PubMed]
  175. Yang, C.; Patel, K.; Harding, P.; Sorokin, A.; Glass Ii, W.F. Regulation of TGF-β1/MAPK-mediated PAI-1 gene expression by the actin cytoskeleton in human mesangial cells. Exp. Cell Res. 2007, 313, 1240–1250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Fujiwara, M.; Jin, E.; Ghazizadeh, M.; Kawanami, O. Activation of PAR4 induces a distinct actin fiber formation via p38 MAPK in human lung endothelial cells. J. Histochem. Cytochem. 2005, 53, 1121–1129. [Google Scholar] [CrossRef] [Green Version]
  177. Paliga, A.J.M.; Natale, D.R.; Watson, A.J. p38 mitogen-activated protein kinase (MAPK) first regulates filamentous actin at the 8-16-cell stage during preimplantation development. Biol. Cell 2005, 97, 629–640. [Google Scholar] [CrossRef]
  178. Tania, M.; Khan, M.A.; Fu, J. Epithelial to mesenchymal transition inducing transcription factors and metastatic cancer. Tumour Biol. 2014, 35, 7335–7342. [Google Scholar] [CrossRef]
  179. Wang, P.B.; Chen, Y.; Ding, G.R.; Du, H.W.; Fan, H.Y. Keratin 18 induces proliferation, migration, and invasion in gastric cancer via the MAPK signalling pathway. Clin. Exp. Pharmacol. Physiol. 2021, 48, 147–156. [Google Scholar] [CrossRef] [PubMed]
  180. Wöll, S.; Windoffer, R.; Leube, R.E. p38 MAPK-dependent shaping of the keratin cytoskeleton in cultured cells. J. Cell Biol. 2007, 177, 795–807. [Google Scholar] [CrossRef] [PubMed]
  181. Schechter, R.; Yanovitch, T.; Abboud, M.; Johnson III, G.; Gaskins, J. Effects of brain endogenous insulin on neurofilament and MAPK in fetal rat neuron cell cultures. Brain Res. 1998, 808, 270–278. [Google Scholar] [CrossRef]
  182. Cheng, T.J.; Lai, Y.K. Identification of mitogen-activated protein kinase-activated protein kinase-2 as a vimentin kinase activated by okadaic acid in 9L rat brain tumor cells. J. Cell. Biochem. 1998, 71, 169–181. [Google Scholar] [CrossRef]
  183. Li, L.; Hu, J.; He, T.; Zhang, Q.; Yang, X.; Lan, X.; Zhang, D.; Mei, H.; Chen, B.; Huang, Y. P38/MAPK contributes to endothelial barrier dysfunction via MAP4 phosphorylation-dependent microtubule disassembly in inflammation-induced acute lung injury. Sci. Rep. 2015, 5, 8895. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Hu, J.Y.; Chu, Z.G.; Han, J.; Dang, Y.M.; Yan, H.; Zhang, Q.; Liang, G.P.; Huang, Y.S. The p38/MAPK pathway regulates microtubule polymerization through phosphorylation of MAP4 and Op18 in hypoxic cells. Cell. Mol. Life Sci. 2010, 67, 321–333. [Google Scholar] [CrossRef] [PubMed]
  185. Lee, S.E.; Kim, J.H.; Kim, N.H. Inactivation of MAPK affects centrosome assembly, but not actin filament assembly, in mouse oocytes maturing in vitro. Mol. Reprod. Dev. 2007, 74, 904–911. [Google Scholar] [CrossRef] [PubMed]
  186. Fan, M.; Chambers, T.C. Role of mitogen-activated protein kinases in the response of tumor cells to chemotherapy. Drug Resist. Updates 2001, 4, 253–267. [Google Scholar] [CrossRef] [PubMed]
  187. Reszka, A.A.; Seger, R.; Diltz, C.D.; Krebs, E.G.; Fischer, E.H. Association of mitogen-activated protein kinase with the microtubule cytoskeleton. Proc. Natl. Acad. Sci. USA 1995, 92, 8881–8885. [Google Scholar] [CrossRef] [PubMed]
  188. Butt, E.; Gambaryan, S.; Göttfert, N.; Galler, A.; Marcus, K.; Meyer, H.E. Actin binding of human LIM and SH3 protein is regulated by cGMP- and cAMP-dependent protein kinase phosphorylation on serine 146. J. Biol. Chem. 2003, 278, 15601–15607. [Google Scholar] [CrossRef] [Green Version]
  189. Butt, E.; Immler, D.; Meyer, H.E.; Kotlyarov, A.; Laaß, K.; Gaestel, M. Heat shock protein 27 is a substrate of cGMP-dependent protein kinase in intact human platelets. Phosphorylation-induced actin polymerization caused by HSP27 mutants. J. Biol. Chem. 2001, 276, 7108–7113. [Google Scholar] [CrossRef]
  190. Sandau, K.B.; Gantner, F.; Brüne, B. Nitric oxide-induced F-actin disassembly is mediated via cGMP, cAMP, and protein kinase A activation in rat mesangial cells. Exp. Cell Res. 2001, 271, 329–336. [Google Scholar] [CrossRef]
  191. Pryzwansky, K.B.; Wyatt, T.A.; Lincoln, T.M. Cyclic guanosine monophosphate-dependent protein kinase is targeted to intermediate filaments and phosphorylates vimentin in A23187-stimulated human neutrophils. Blood 1995, 85, 222–230. [Google Scholar] [CrossRef] [Green Version]
  192. MacMillan-Crow, L.A.; Lincoln, T.M. High-affinity binding and localization of the cyclic GMP-dependent protein kinase with the intermediate filament protein vimentin. Biochemistry 1994, 33, 8035–8043. [Google Scholar] [CrossRef]
  193. Wyatt, T.A.; Lincoln, T.M.; Pryzwansky, K.B. Regulation of human neutrophil degranulation by LY-83583 and L-arginine: Role of cGMP-dependent protein kinase. Am. J. Physiol. Cell Physiol. 1993, 265, C201–C211. [Google Scholar] [CrossRef] [PubMed]
  194. Wyatt, T.A.; Lincoln, T.M.; Pryzwansky, K.B. Vimentin is transiently co-localized with and phosphorylated by cyclic GMP-dependent protein kinase in formyl-peptide-stimulated neutrophils. J. Biol. Chem. 1991, 266, 21274–21280. [Google Scholar] [CrossRef] [PubMed]
  195. Xia, C.; Nguyen, M.; Garrison, A.K.; Zhao, Z.; Wang, Z.; Sutherland, C.; Ma, L. CNP/cGMP signaling regulates axon branching and growth by modulating microtubule polymerization. Dev. Neurobiol. 2013, 73, 673–687. [Google Scholar] [CrossRef] [PubMed]
  196. Gong, K.; Xing, D.; Li, P.; Hilgers, R.H.; Hage, F.G.; Oparil, S.; Chen, Y.F. cGMP inhibits TGF-beta signaling by sequestering Smad3 with cytosolic beta2-tubulin in pulmonary artery smooth muscle cells. Mol. Endocrinol. 2011, 25, 1794–1803. [Google Scholar] [CrossRef] [Green Version]
  197. Guo, J.; Wenk, M.R.; Pellegrini, L.; Onofri, F.; Benfenati, F.; De Camilli, P. Phosphatidylinositol 4-kinase type IIα is responsible for the phosphatidylinositol 4-kinase activity associated with synaptic vesicles. Proc. Natl. Acad. Sci. USA 2003, 100, 3995–4000. [Google Scholar] [CrossRef]
  198. Michalczyk, I.; Sikorski, A.F.; Kotula, L.; Junghans, R.P.; Dubielecka, P.M. The emerging role of protein kinase Cθ in cytoskeletal signaling. J. Leukoc. Biol. 2013, 93, 319–327. [Google Scholar] [CrossRef] [Green Version]
  199. Liao, J.K.; Seto, M.; Noma, K. Rho kinase (ROCK) inhibitors. J. Cardiovasc. Pharmacol. 2007, 50, 17–24. [Google Scholar] [CrossRef] [Green Version]
  200. Apodaca, G. Endocytic traffic in polarized epithelial cells: Role of the actin and microtubule cytoskeleton. Traffic 2001, 2, 149–159. [Google Scholar] [CrossRef]
  201. Ridley, A.J. Rho proteins: Linking signaling with membrane trafficking. Traffic 2001, 2, 303–310. [Google Scholar] [CrossRef]
  202. Etienne-Manneville, S.; Hall, A. Rho GTPases in cell biology. Nature 2002, 420, 629–635. [Google Scholar] [CrossRef]
  203. Jaffe, A.B.; Hall, A. Rho GTPases: Biochemistry and biology. Annu. Rev. Cell Dev. Biol. 2005, 21, 247–269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Burridge, K.; Wennerberg, K. Rho and Rac take center stage. Cell 2004, 116, 167–179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Bishop, A.L.; Hall, A. Rho GTPases and their effector proteins. Biochem. J. 2000, 348 Pt 2, 241–255. [Google Scholar] [CrossRef]
  206. Narumiya, S.; Tanji, M.; Ishizaki, T. Rho signaling, ROCK and mDia1, in transformation, metastasis and invasion. Cancer Metastasis Rev. 2009, 28, 65–76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Wittinghofer, A.; Vetter, I.R. Structure-function relationships of the G domain, a canonical switch motif. Annu. Rev. Biochem. 2011, 80, 943–971. [Google Scholar] [CrossRef] [PubMed]
  208. Zheng, Y. Dbl family guanine nucleotide exchange factors. Trends Biochem. Sci. 2001, 26, 724–732. [Google Scholar] [CrossRef]
  209. Hodge, R.G.; Ridley, A.J. Regulating Rho GTPases and their regulators. Nat. Rev. Mol. Cell Biol. 2016, 17, 496–510. [Google Scholar] [CrossRef]
  210. Jiu, Y.; Peränen, J.; Schaible, N.; Cheng, F.; Eriksson, J.E.; Krishnan, R.; Lappalainen, P. Vimentin intermediate filaments control actin stress fiber assembly through GEF-H1 and RhoA. J. Cell Sci. 2017, 130, 892–902. [Google Scholar] [CrossRef] [Green Version]
  211. Inaba, H.; Yamakawa, D.; Tomono, Y.; Enomoto, A.; Mii, S.; Kasahara, K.; Goto, H.; Inagaki, M. Regulation of keratin 5/14 intermediate filaments by CDK1, Aurora-B, and Rho-kinase. Biochem. Biophys. Res. Commun. 2018, 498, 544–550. [Google Scholar] [CrossRef]
  212. Dharmawardhane, S. Rho family GTPases in cancer. Cancers 2021, 13, 1271. [Google Scholar] [CrossRef]
  213. Ellenbroek, S.I.J.; Collard, J.G. Rho GTPases: Functions and association with cancer. Clin. Exp. Metastasis 2007, 24, 657–672. [Google Scholar] [CrossRef]
  214. El-Sibai, M.; Pertz, O.; Pang, H.; Yip, S.C.; Lorenz, M.; Symons, M.; Condeelis, J.S.; Hahn, K.M.; Backer, J.M. RhoA/ROCK-mediated switching between Cdc42- and Rac1-dependent protrusion in MTLn3 carcinoma cells. Exp. Cell Res. 2008, 314, 1540–1552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Bustelo, X.R.; Sauzeau, V.; Berenjeno, I.M. GTP-binding proteins of the Rho/Rac family: Regulation, effectors and functions in vivo. Bioessays 2007, 29, 356–370. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Pinyol, R.; Haeckel, A.; Ritter, A.; Qualmann, B.; Kessels, M.M. Regulation of N-WASP and the Arp2/3 complex by Abp1 controls neuronal morphology. PLoS ONE 2007, 2, e400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Koka, S.; Neudauer, C.L.; Li, X.; Lewis, R.E.; McCarthy, J.B.; Westendorf, J.J. The formin-homology-domain-containing protein FHOD1 enhances cell migration. J. Cell Sci. 2003, 116, 1745–1755. [Google Scholar] [CrossRef] [PubMed]
  218. Royal, I.; Lamarche-Vane, N.; Lamorte, L.; Kaibuchi, K.; Park, M. Activation of cdc42, rac, PAK, and rho-kinase in response to hepatocyte growth factor differentially regulates epithelial cell colony spreading and dissociation. Mol. Biol. Cell 2000, 11, 1709–1725. [Google Scholar] [CrossRef] [Green Version]
  219. Watanabe, N.; Madaule, P.; Reid, T.; Ishizaki, T.; Watanabe, G.; Kakizuka, A.; Saito, Y.; Nakao, K.; Jockusch, B.M.; Narumiya, S. p140mDia, a mammalian homolog of Drosophila diaphanous, is a target protein for Rho small GTPase and is a ligand for profilin. EMBO J. 1997, 16, 3044–3056. [Google Scholar] [CrossRef]
  220. ROCK1, Rho Associated Coiled-Coil Containing Protein Kinase 1 [Homo Sapiens (Human)]. Bethesda (MD): National Library of Medicine (US), National Center for Biotechnology Information; 2022. Available online: https://www.ncbi.nlm.nih.gov/gene/60932022 (accessed on 20 December 2022).
  221. ROCK2, Rho Associated Coiled-Coil Containing Protein Kinase 1 [Homo Sapiens (Human)]. Bethesda (MD): National Library of Medicine (US), National Center for Biotechnology Information; 2022. Available online: https://www.ncbi.nlm.nih.gov/gene/9475 (accessed on 20 December 2022).
  222. Totsukawa, G.; Yamakita, Y.; Yamashiro, S.; Hartshorne, D.J.; Sasaki, Y.; Matsumura, F. Distinct roles of ROCK (Rho-kinase) and MLCK in spatial regulation of MLC phosphorylation for assembly of stress fibers and focal adhesions in 3T3 fibroblasts. J. Cell Biol. 2000, 150, 797–806. [Google Scholar] [CrossRef]
  223. Ohashi, K.; Nagata, K.; Maekawa, M.; Ishizaki, T.; Narumiya, S.; Mizuno, K. Rho-associated kinase ROCK activates LIM-kinase 1 by phosphorylation at threonine 508 within the activation loop. J. Biol. Chem. 2000, 275, 3577–3582. [Google Scholar] [CrossRef] [Green Version]
  224. Qiao, Y.N.; He, W.Q.; Chen, C.P.; Zhang, C.H.; Zhao, W.; Wang, P.; Zhang, L.; Wu, Y.Z.; Yang, X.; Peng, Y.J.; et al. Myosin phosphatase target subunit 1 (MYPT1) regulates the contraction and relaxation of vascular smooth muscle and maintains blood pressure. J. Biol. Chem. 2014, 289, 22512–22523. [Google Scholar] [CrossRef] [Green Version]
  225. Lee, J.H.; Katakai, T.; Hara, T.; Gonda, H.; Sugai, M.; Shimizu, A. Roles of p-ERM and Rho-ROCK signaling in lymphocyte polarity and uropod formation. J. Cell Biol. 2004, 167, 327–337. [Google Scholar] [CrossRef] [PubMed]
  226. Fukata, Y.; Oshiro, N.; Kinoshita, N.; Kawano, Y.; Matsuoka, Y.; Bennett, V.; Matsuura, Y.; Kaibuchi, K. Phosphorylation of adducin by Rho-kinase plays a crucial role in cell motility. J. Cell Biol. 1999, 145, 347–361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Shibukawa, Y.; Yamazaki, N.; Daimon, E.; Wada, Y. Rock-dependent calponin 3 phosphorylation regulates myoblast fusion. Exp. Cell Res. 2013, 319, 633–648. [Google Scholar] [CrossRef] [PubMed]
  228. Ikenoya, M.; Hidaka, H.; Hosoya, T.; Suzuki, M.; Yamamoto, N.; Sasaki, Y. Inhibition of rho-kinase-induced myristoylated alanine-rich C kinase substrate (MARCKS) phosphorylation in human neuronal cells by H-1152, a novel and specific Rho-kinase inhibitor. J. Neurochem. 2002, 81, 9–16. [Google Scholar] [CrossRef] [Green Version]
  229. Izawa, T.; Fukata, Y.; Kimura, T.; Iwamatsu, A.; Dohi, K.; Kaibuchi, K. Elongation factor-1 alpha is a novel substrate of rho-associated kinase. Biochem. Biophys. Res. Commun. 2000, 278, 72–78. [Google Scholar] [CrossRef]
  230. Vahebi, S.; Kobayashi, T.; Warren, C.M.; de Tombe, P.P.; Solaro, R.J. Functional effects of rho-kinase-dependent phosphorylation of specific sites on cardiac troponin. Circ. Res. 2005, 96, 740–747. [Google Scholar] [CrossRef] [Green Version]
  231. Shao, J.; Welch, W.J.; Diprospero, N.A.; Diamond, M.I. Phosphorylation of profilin by ROCK1 regulates polyglutamine aggregation. Mol. Cell. Biol. 2008, 28, 5196–5208. [Google Scholar] [CrossRef] [Green Version]
  232. Amano, M.; Ito, M.; Kimura, K.; Fukata, Y.; Chihara, K.; Nakano, T.; Matsuura, Y.; Kaibuchi, K. Phosphorylation and activation of myosin by Rho-associated kinase (Rho-kinase). J. Biol. Chem. 1996, 271, 20246–20249. [Google Scholar] [CrossRef] [Green Version]
  233. Shi, J.; Wu, X.; Surma, M.; Vemula, S.; Zhang, L.; Yang, Y.; Kapur, R.; Wei, L. Distinct roles for ROCK1 and ROCK2 in the regulation of cell detachment. Cell Death Dis. 2013, 4, e483. [Google Scholar] [CrossRef] [Green Version]
  234. Wang, J.; Liu, X.-H.; Yang, Z.-J.; Xie, B.; Zhong, Y.-S. The effect of ROCK-1 activity change on the adhesive and invasive ability of Y79 retinoblastoma cells. BMC Cancer 2014, 14, 89. [Google Scholar] [CrossRef] [Green Version]
  235. Rochelle, T.; Daubon, T.; Van Troys, M.; Harnois, T.; Waterschoot, D.; Ampe, C.; Roy, L.; Bourmeyster, N.; Constantin, B. p210bcr-abl induces amoeboid motility by recruiting ADF/destrin through RhoA/ROCK1. FASEB J. 2013, 27, 123–134. [Google Scholar] [CrossRef] [PubMed]
  236. Yasui, Y.; Amano, M.; Nagata, K.; Inagaki, N.; Nakamura, H.; Saya, H.; Kaibuchi, K.; Inagaki, M. Roles of Rho-associated kinase in cytokinesis; mutations in Rho-associated kinase phosphorylation sites impair cytokinetic segregation of glial filaments. J. Cell Biol. 1998, 143, 1249–1258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Hashimoto, R.; Nakamura, Y.; Goto, H.; Wada, Y.; Sakoda, S.; Kaibuchi, K.; Inagaki, M.; Takeda, M. Domain- and site-specific phosphorylation of bovine NF-L by Rho-associated kinase. Biochem. Biophys. Res. Commun. 1998, 245, 407–411. [Google Scholar] [CrossRef] [PubMed]
  238. Goto, H.; Kosako, H.; Inagaki, M. Regulation of intermediate filament organization during cytokinesis: Possible roles of Rho-associated kinase. Microsc. Res. Tech. 2000, 49, 173–182. [Google Scholar] [CrossRef]
  239. Bordeleau, F.; Myrand Lapierre, M.E.; Sheng, Y.; Marceau, N. Keratin 8/18 regulation of cell stiffness-extracellular matrix interplay through modulation of Rho-mediated actin cytoskeleton dynamics. PLoS ONE 2012, 7, e38780. [Google Scholar] [CrossRef] [PubMed]
  240. Sin, W.C.; Chen, X.Q.; Leung, T.; Lim, L. RhoA-binding kinase alpha translocation is facilitated by the collapse of the vimentin intermediate filament network. Mol. Cell. Biol. 1998, 18, 6325–6339. [Google Scholar] [CrossRef] [Green Version]
  241. Amano, M.; Kaneko, T.; Maeda, A.; Nakayama, M.; Ito, M.; Yamauchi, T.; Goto, H.; Fukata, Y.; Oshiro, N.; Shinohara, A.; et al. Identification of Tau and MAP2 as novel substrates of Rho-kinase and myosin phosphatase. J. Neurochem. 2003, 87, 780–790. [Google Scholar] [CrossRef]
  242. Arimura, N.; Ménager, C.; Kawano, Y.; Yoshimura, T.; Kawabata, S.; Hattori, A.; Fukata, Y.; Amano, M.; Goshima, Y.; Inagaki, M.; et al. Phosphorylation by Rho kinase regulates CRMP-2 activity in growth cones. Mol. Cell. Biol. 2005, 25, 9973–9984. [Google Scholar] [CrossRef] [Green Version]
  243. Amano, M.; Tsumura, Y.; Taki, K.; Harada, H.; Mori, K.; Nishioka, T.; Kato, K.; Suzuki, T.; Nishioka, Y.; Iwamatsu, A.; et al. A proteomic approach for comprehensively screening substrates of protein kinases such as Rho-kinase. PLoS ONE 2010, 5, e8704. [Google Scholar] [CrossRef]
  244. Zhang, J.; Dong, X.P. Dysfunction of microtubule-associated proteins of MAP2/tau family in Prion disease. Prion 2012, 6, 334–338. [Google Scholar] [CrossRef] [Green Version]
  245. Lin, P.C.; Chan, P.M.; Hall, C.; Manser, E. Collapsin response mediator proteins (CRMPs) are a new class of microtubule-associated protein (MAP) that selectively interacts with assembled microtubules via a taxol-sensitive binding interaction. J. Biol. Chem. 2011, 286, 41466–41478. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Schofield, A.V.; Steel, R.; Bernard, O. Rho-associated coiled-coil kinase (ROCK) protein controls microtubule dynamics in a novel signaling pathway that regulates cell migration. J. Biol. Chem. 2012, 287, 43620–43629. [Google Scholar] [CrossRef] [Green Version]
  247. Huang, C.H.; Cheng, J.C.; Chen, J.C.; Tseng, C.P. Evaluation of the role of Disabled-2 in nerve growth factor-mediated neurite outgrowth and cellular signalling. Cell. Signal. 2007, 19, 1339–1347. [Google Scholar] [CrossRef] [PubMed]
  248. OuYang, C.; Xie, Y.; Fu, Q.; Xu, G. SYNPO2 suppresses hypoxia-induced proliferation and migration of colorectal cancer cells by regulating YAP-KLF5 axis. Tissue Cell 2021, 73, 101598. [Google Scholar] [CrossRef] [PubMed]
  249. Kai, F.; Tanner, K.; King, C.; Duncan, R. Myopodin isoforms alter the chemokinetic response of PC3 cells in response to different migration stimuli via differential effects on Rho-ROCK signaling pathways. Carcinogenesis 2012, 33, 2100–2107. [Google Scholar] [CrossRef]
  250. Maldonado, H.; Calderon, C.; Burgos-Bravo, F.; Kobler, O.; Zuschratter, W.; Ramirez, O.; Härtel, S.; Schneider, P.; Quest, A.F.; Herrera-Molina, R.; et al. Astrocyte-to-neuron communication through integrin-engaged Thy-1/CBP/Csk/Src complex triggers neurite retraction via the RhoA/ROCK pathway. Biochim. Biophys. Acta Mol. Cell Res. 2017, 1864, 243–254. [Google Scholar] [CrossRef]
  251. Fusella, F.; Seclì, L.; Busso, E.; Krepelova, A.; Moiso, E.; Rocca, S.; Conti, L.; Annaratone, L.; Rubinetto, C.; Mello-Grand, M.; et al. The IKK/NF-κB signaling pathway requires Morgana to drive breast cancer metastasis. Nat. Commun. 2017, 8, 1636. [Google Scholar] [CrossRef] [Green Version]
  252. Barry, D.M.; Koo, Y.; Norden, P.R.; Wylie, L.A.; Xu, K.; Wichaidit, C.; Azizoglu, D.B.; Zheng, Y.; Cobb, M.H.; Davis, G.E.; et al. Rasip1-Mediated Rho GTPase Signaling Regulates Blood Vessel Tubulogenesis via Nonmuscle Myosin II. Circ. Res. 2016, 119, 810–826. [Google Scholar] [CrossRef] [Green Version]
  253. de Kreuk, B.J.; Gingras, A.R.; Knight, J.D.; Liu, J.J.; Gingras, A.C.; Ginsberg, M.H. Heart of glass anchors Rasip1 at endothelial cell-cell junctions to support vascular integrity. Elife 2016, 5, e11394. [Google Scholar] [CrossRef]
  254. Yamaguchi, H.; Kasa, M.; Amano, M.; Kaibuchi, K.; Hakoshima, T. Molecular Mechanism for the Regulation of Rho-Kinase by Dimerization and Its Inhibition by Fasudil. Structure 2006, 14, 589–600. [Google Scholar] [CrossRef] [Green Version]
  255. Wang, J.; Liu, X.; Zhong, Y. Rho/Rho-associated kinase pathway in glaucoma (Review). Int. J. Oncol. 2013, 43, 1357–1367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Feng, Y.; LoGrasso, P.V.; Defert, O.; Li, R. Rho Kinase (ROCK) Inhibitors and Their Therapeutic Potential. J. Med. Chem. 2016, 59, 2269–2300. [Google Scholar] [CrossRef]
  257. Liu, Y.; Gray, N.S. Rational design of inhibitors that bind to inactive kinase conformations. Nat. Chem. Biol. 2006, 2, 358–364. [Google Scholar] [CrossRef]
  258. Zhang, T.; Inesta-Vaquera, F.; Niepel, M.; Zhang, J.; Ficarro, S.B.; Machleidt, T.; Xie, T.; Marto, J.A.; Kim, N.; Sim, T.; et al. Discovery of potent and selective covalent inhibitors of JNK. Chem. Biol. 2012, 19, 140–154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  259. Moshirfar, M.; Parker, L.; Birdsong, O.C.; Ronquillo, Y.C.; Hofstedt, D.; Shah, T.J.; Gomez, A.T.; Hoopes, P.C.S. Use of Rho kinase Inhibitors in Ophthalmology: A Review of the Literature. Med. Hypothesis Discov. Innov. Ophthalmol. 2018, 7, 101–111. [Google Scholar] [PubMed]
  260. Zhao, J.; Zhou, D.; Guo, J.; Ren, Z.; Zhou, L.; Wang, S.; Xu, B.; Wang, R. Effect of fasudil hydrochloride, a protein kinase inhibitor, on cerebral vasospasm and delayed cerebral ischemic symptoms after aneurysmal subarachnoid hemorrhage. Neurol. Med. Chir. 2006, 46, 421–428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  261. Satoh, S.; Ueda, Y.; Koyanagi, M.; Kadokami, T.; Sugano, M.; Yoshikawa, Y.; Makino, N. Chronic inhibition of Rho kinase blunts the process of left ventricular hypertrophy leading to cardiac contractile dysfunction in hypertension-induced heart failure. J. Mol. Cell. Cardiol. 2003, 35, 59–70. [Google Scholar] [CrossRef]
  262. Rikitake, Y.; Oyama, N.; Wang, C.Y.; Noma, K.; Satoh, M.; Kim, H.H.; Liao, J.K. Decreased perivascular fibrosis but not cardiac hypertrophy in ROCK1+/- haploinsufficient mice. Circulation 2005, 112, 2959–2965. [Google Scholar] [CrossRef] [Green Version]
  263. Hannan, J.L.; Albersen, M.; Kutlu, O.; Gratzke, C.; Stief, C.G.; Burnett, A.L.; Lysiak, J.J.; Hedlund, P.; Bivalacqua, T.J. Inhibition of Rho-kinase improves erectile function, increases nitric oxide signaling and decreases penile apoptosis in a rat model of cavernous nerve injury. J. Urol. 2013, 189, 1155–1161. [Google Scholar] [CrossRef] [Green Version]
  264. Wingard, C.J.; Johnson, J.A.; Holmes, A.; Prikosh, A. Improved erectile function after Rho-kinase inhibition in a rat castrate model of erectile dysfunction. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2003, 284, R1572–R1579. [Google Scholar] [CrossRef] [Green Version]
  265. Shin, J.Y.; Kim, Y.I.; Cho, S.J.; Lee, M.K.; Kook, M.C.; Lee, J.H.; Lee, S.S.; Ashktorab, H.; Smoot, D.T.; Ryu, K.W.; et al. MicroRNA 135a suppresses lymph node metastasis through down-regulation of ROCK1 in early gastric cancer. PLoS ONE 2014, 9, e85205. [Google Scholar] [CrossRef] [PubMed]
  266. Li, E.; Zhang, J.; Yuan, T.; Ma, B. MiR-145 inhibits osteosarcoma cells proliferation and invasion by targeting ROCK1. Tumour Biol. 2014, 35, 7645–7650. [Google Scholar] [CrossRef] [PubMed]
  267. Li, J.; Song, Y.; Wang, Y.; Luo, J.; Yu, W. MicroRNA-148a suppresses epithelial-to-mesenchymal transition by targeting ROCK1 in non-small cell lung cancer cells. Mol. Cell. Biochem. 2013, 380, 277–282. [Google Scholar] [CrossRef]
  268. Cascione, M.; De Matteis, V.; Toma, C.C.; Pellegrino, P.; Leporatti, S.; Rinaldi, R. Morphomechanical and structural changes induced by ROCK inhibitor in breast cancer cells. Exp. Cell Res. 2017, 360, 303–309. [Google Scholar] [CrossRef]
  269. Ueno, K.; Hirata, H.; Shahryari, V.; Chen, Y.; Zaman, M.S.; Singh, K.; Tabatabai, Z.L.; Hinoda, Y.; Dahiya, R. Tumour suppressor microRNA-584 directly targets oncogene Rock-1 and decreases invasion ability in human clear cell renal cell carcinoma. Br. J. Cancer 2011, 104, 308–315. [Google Scholar] [CrossRef] [PubMed]
  270. Ogawa, T.; Tashiro, H.; Miyata, Y.; Ushitora, Y.; Fudaba, Y.; Kobayashi, T.; Arihiro, K.; Okajima, M.; Asahara, T. Rho-associated kinase inhibitor reduces tumor recurrence after liver transplantation in a rat hepatoma model. Am. J. Transplant. 2007, 7, 347–355. [Google Scholar] [CrossRef] [PubMed]
  271. Kroiss, A.; Vincent, S.; Decaussin-Petrucci, M.; Meugnier, E.; Viallet, J.; Ruffion, A.; Chalmel, F.; Samarut, J.; Allioli, N. Androgen-regulated microRNA-135a decreases prostate cancer cell migration and invasion through downregulating ROCK1 and ROCK2. Oncogene 2015, 34, 2846–2855. [Google Scholar] [CrossRef]
  272. Patel, R.A.; Forinash, K.D.; Pireddu, R.; Sun, Y.; Sun, N.; Martin, M.P.; Schönbrunn, E.; Lawrence, N.J.; Sebti, S.M. RKI-1447 is a potent inhibitor of the Rho-associated ROCK kinases with anti-invasive and antitumor activities in breast cancer. Cancer Res. 2012, 72, 5025–5034. [Google Scholar] [CrossRef] [Green Version]
  273. McLeod, R.; Kumar, R.; Papadatos-Pastos, D.; Mateo, J.; Brown, J.S.; Garces, A.H.I.; Ruddle, R.; Decordova, S.; Jueliger, S.; Ferraldeschi, R.; et al. First-in-Human Study of AT13148, a Dual ROCK-AKT Inhibitor in Patients with Solid Tumors. Clin. Cancer Res. 2020, 26, 4777–4784. [Google Scholar] [CrossRef] [PubMed]
  274. Garg, R.; Riento, K.; Keep, N.; Morris, J.D.; Ridley, A.J. N-terminus-mediated dimerization of ROCK-I is required for RhoE binding and actin reorganization. Biochem. J. 2008, 411, 407–414. [Google Scholar] [CrossRef] [Green Version]
  275. Jacobs, M.; Hayakawa, K.; Swenson, L.; Bellon, S.; Fleming, M.; Taslimi, P.; Doran, J. The structure of dimeric ROCK I reveals the mechanism for ligand selectivity. J. Biol. Chem. 2006, 281, 260–268. [Google Scholar] [CrossRef] [Green Version]
  276. Gu, Z.; Yan, T.; Yan, F. Rational design and improvement of the dimerization-disrupting peptide selectivity between ROCK-I and ROCK-II kinase isoforms in cerebrovascular diseases. J. Mol. Recognit. 2020, 33, e2835. [Google Scholar] [CrossRef] [PubMed]
  277. Hartmann, S.; Ridley, A.J.; Lutz, S. The Function of Rho-Associated Kinases ROCK1 and ROCK2 in the Pathogenesis of Cardiovascular Disease. Front. Pharmacol. 2015, 6, 276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  278. Couzens, A.L.; Saridakis, V.; Scheid, M.P. The hydrophobic motif of ROCK2 requires association with the N-terminal extension for kinase activity. Biochem. J. 2009, 419, 141–148. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Molecular structure of ROCK 1 and ROCK 2. They are similar and both contain a kinase domain, coiled-coil region, rho-binding domain, putative pleckstrin homology domain (PHD) and cysteine-rich domain (CRD). Rho GEF = Rho guanine nucleotide exchange factor. Created with BioRender.com.
Figure 1. Molecular structure of ROCK 1 and ROCK 2. They are similar and both contain a kinase domain, coiled-coil region, rho-binding domain, putative pleckstrin homology domain (PHD) and cysteine-rich domain (CRD). Rho GEF = Rho guanine nucleotide exchange factor. Created with BioRender.com.
Genes 14 00272 g001
Figure 2. The functions of ROCK. Rho is activated by Rho guanine nucleotide exchange factor (GEF). Activated GTP-bound Rho stimulates ROCK. Activated ROCK activates or inhibits downstream targets by phosphorylation. Activated ROCK regulates microfilaments via the ROCK/MYPT1/MLC and ROCK/LIM kinases/cofilin pathways. ROCK also regulates intermediate filaments by phosphorylation. ROCK regulates microtubules by phosphorylation of several microtubule-associated proteins such as MAP2/Tau, collapsin response mediator protein 2 (CRMP2) and Doublecortin. Created with BioRender.com.
Figure 2. The functions of ROCK. Rho is activated by Rho guanine nucleotide exchange factor (GEF). Activated GTP-bound Rho stimulates ROCK. Activated ROCK activates or inhibits downstream targets by phosphorylation. Activated ROCK regulates microfilaments via the ROCK/MYPT1/MLC and ROCK/LIM kinases/cofilin pathways. ROCK also regulates intermediate filaments by phosphorylation. ROCK regulates microtubules by phosphorylation of several microtubule-associated proteins such as MAP2/Tau, collapsin response mediator protein 2 (CRMP2) and Doublecortin. Created with BioRender.com.
Genes 14 00272 g002
Table 1. Cytoskeleton components and their functions.
Table 1. Cytoskeleton components and their functions.
Cytoskeleton ComponentFunctionsReferences
Microfilament
ActinGeneration of forces in cellular contraction, endo- and exocytosis, secretion, vesicle transfer, cell division and cell integrity Jiang et al., 2021 [20]; Falahzadeh et al., 2015 [21]; Gordon-Alonso et al., 2010 [22]; Blessing et al., 2004 [23]; Halpain 2003 [24]; Lanier and Gertler, 2000 [25]; Pollard et al., 2000 [26]; Schmidt and Hall, 1998 [27]; Bretscher, 1993 [28].
Intermediate filament
Type I (Acidic Keratins)Expressed in epithelial cells, form the structural framework of cells and contribute to mechanical resilience Gül et al., 2022 [29]; Jacob et al., 2018 [30]; Saitoh et al., 2016 [31]; Schweizer et al., 2006 [32].
Type II (Basic Keratins)Expressed in epithelial cells, they also maintain the stability of the cell nucleus and mechanical stability of the whole cell Honda et al., 2014 [33]; Infante et al., 2011 [34]; Moll et al., 2008 [35]; Bowden et al., 1984 [36].
Type III (Vimentin)Regulate cell mechanics and coordinate mechanosensing, transduction, signaling pathways, motility and inflammatory responses Ridge et al., 2022 [37]; Yue et al., 2016 [38]; Kidd et al., 2014 [39]; Menko et al., 2014 [40]; Herrmann et al., 2007 [41]; Wang et al., 2006 [42]
Type IV (Neurofilaments)Provide structural support for axons and regulate axon radial growth Didonna and Opal, 2019 [43]; Yuan andNixon, 2016 [44]; Yuan et al., 2012 [45]; Lin and Schlaepfer, 2006 [46]; Hoffman1988 [47]
Type V (Nuclear Lamins)Attach chromatin domains to the nuclear periphery and localize some nuclear membrane proteins Khadija et al., 2015 [48]; Carmosino et al., 2014 [49]; Burke and Stewart, 2013 [50]; Lopez-Soler et al., 2001 [51]; Parnaik, 2008 [52]
Type VI (Phakinin and Filensin)Lens fiber and maintenance of lens transparency Oka et al., 2008 [53]; Pittenger et al., 2007 [54]; Blankenship et al., 2001 [55]; Goulielmos et al., 1996 [56]; Georgatos et al., 1994 [57]
Microtubule
TubulinMaintain the structure of the cell, transport secretory vesicles, organelles and macromolecular assemblies and assembly of mitotic spindle Hlavaty and Lechler, 2021 [58]; Chang and Gu, 2020 [59]; Matis 2020 [60]; Logan and Menko, 2019 [61]; Cirillo et al., 2017 [62]; Ganguly et al., 2012 [63]; van der Vaart et al., 2009 [64]; Kulić et al., 2008 [65]
Table 2. Pathways that regulate the cytoskeleton.
Table 2. Pathways that regulate the cytoskeleton.
PathwayFunctionCytoskeleton Components AffectedAuthors, Year
Rho-kinase/ROCKMajor signaling pathway that regulates the cytoskeleton and cell polarity Actin Shi and Wei, 2022 [120]; Tang et al., 2018 [121]; Schofield and Bernard, 2013 [122]; Sit and Manser, 2011 [123]; Sarasa-Renedo et al., 2006 [124]; Woods et al., 2005 [125]; McBeath et al., 2004 [10]; Da Silva et al., 2003 [126]; Amano et al., 2000 [127]; Maekawa et al., 1999 [11]
Intermediate filaments Tang et al., 2018 [121]; Yang et al., 2017 [128]; Lei et al., 2013 [129]; Schofield et al., 2013 [122]; Amano et al., 2010 [12]; Hirose et al., 1998 [130]
Microtubules Becker et al., 2022 [131]; Schofield et al., 2013 [122]; Heng et al., 2012 [132]; Fonseca et al., 2010 [133]; Takesono et al., 2010 [134]; Gao et al., 2004 [14]
PI3K/AKTRoles in the assembly of actin filaments, polymerization of microtubules and intermediate filaments Actin Han et al., 2020 [135]; Lien et al., 2017 [136]; Kakinuma et al., 2008 [137]; Qian et al., 2005 [138]; Qian et al., 2004 [139]; Krasilnikov 2000 [140]
Intermediate filaments Deng et al., 2022 [117]; Roux et al., 2017 [141]; Wang et al., 2012 [142]; Kong et al., 2012 [143]; Tseng et al., 2011 [144]
Microtubules Chakrabarty et al., 2019 [145]; Fu et al., 2017 [146]; Kitagishi et al., 2014 [147]; Onishi et al., 2007 [148]; Fujiwara et al., 2007 [149];
Wnt/β-cateninCell proliferation, differentiation, survival and adhesion (cytoskeleton reorganization) and regulation of microtubule stability. Wnt ligands stabilize β-catenin whereas β-catenin helps link cadherin adhesion molecules to cytoskeletonActin Zhang et al., 2022 [150]; Roarty et al., 2017 [115]; Galli et al., 2012 [151]; Lai et al., 2009 [110]; James et al., 2008 [114]
Intermediate filaments Lehmann et al., 2020 [152]; Tian et al., 2019 [153]; Bermeo et al., 2015 [154]; Prasad et al., 2008 [155]; Bierie et al., 2003 [156]
Microtubule Puri et al., 2021 [157]; Ou et al., 2020 [158]; Huang et al., 2007 [159]; Ciani et al., 2004 [160]; Peifer et al., 2000 [161]
Gsk-3βInhibition of Wnt signaling pathway. Responsible for actin branching, regulation of intermediate filaments and controlling microtubule dynamics Actin Hajka et al., 2021 [162]; Yoshino et al., 2015 [163]; Watanabe et al., 2009 [164]; Sun et al., 2009 [165]; Vaidya et al., 2006 [166]
Intermediate filaments Sen et al., 2022 [167]; Lee et al., 2012 [168]; Kim et al., 2012 [169]; Sasaki et al., 2002 [170]; Guidato et al., 1996 [171]
Microtubules Sen et al., 2022 [167]; Hajka et al., 2021 [162]; Beurel et al., 2015 [116]; Watanabe et al., 2009 [164]; Fumoto et al., 2006 [172]; Grimes and Jope, 2001 [173]
MAPKCytoskeleton remodelling, downregulation of vimentin and affects the polymerization and stability of microtubules Actin Joe et al., 2022 [174]; Hoffman et al., 2017 [111]; Yang et al., 2007 [175]; Fujiwara et al., 2005 [176]; Paliga et al., 2005 [177]
Intermediate filaments Tania et al., 2014 [178]; Wang et al., 2021 [179]; Wöll et al., 2007 [180]; Schechter et al., 1998 [181]; Cheng and Lai, 1998 [182]
Microtubules Li et al., 2015 [183]; Hu et al., 2010 [184]; Lee et al., 2007 [185]; Fan and Chambers, 2001 [186]; Reszka et al., 1995 [187];
cGMP kinaseInhibition of actin cytoskeleton organization, phosphorylation of vimentin and modulating microtubules and their associated proteins Actin Zou et al., 2018 [113]; Butt et al., 2003 [188]; Butt et al., 2001 [189]; Sandau et al., 2001 [190]; Sauzeau et al., 2000 [112]
Intermediate filaments Pryzwansky et al., 1995 [191]; MacMillan-Crow and Lincoln, 1994 [192]; Wyatt et al., 1993 [193]; Wyatt et al., 1991 [194]
Microtubules Xia et al., 2013 [195]; Gong et al., 2011 [196]
Glycogen synthase kinase-3 beta (Gsk-3β), Mitogen-activated protein kinase (MAPK), Cyclic GMP-dependent protein kinase (cGMP kinase).
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Guan, G.; Cannon, R.D.; Coates, D.E.; Mei, L. Effect of the Rho-Kinase/ROCK Signaling Pathway on Cytoskeleton Components. Genes 2023, 14, 272. https://doi.org/10.3390/genes14020272

AMA Style

Guan G, Cannon RD, Coates DE, Mei L. Effect of the Rho-Kinase/ROCK Signaling Pathway on Cytoskeleton Components. Genes. 2023; 14(2):272. https://doi.org/10.3390/genes14020272

Chicago/Turabian Style

Guan, Guangzhao, Richard D. Cannon, Dawn E. Coates, and Li Mei. 2023. "Effect of the Rho-Kinase/ROCK Signaling Pathway on Cytoskeleton Components" Genes 14, no. 2: 272. https://doi.org/10.3390/genes14020272

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