Back to Journals » Cancer Management and Research » Volume 14

Cofilin Acts as a Booster for Progression of Malignant Tumors Represented by Glioma

Authors Lv S, Chen Z , Mi H, Yu X

Received 14 September 2022

Accepted for publication 10 November 2022

Published 24 November 2022 Volume 2022:14 Pages 3245—3269

DOI https://doi.org/10.2147/CMAR.S389825

Checked for plagiarism Yes

Review by Single anonymous peer review

Peer reviewer comments 2

Editor who approved publication: Dr Ahmet Emre Eşkazan



Shihong Lv,1,* Zhiye Chen,2,3,* Hailong Mi,3 Xingjiang Yu3

1Department of Gastroenterology, The Second Affiliated Hospital of Mudanjiang Medical College, Mudanjiang Medical College, Mudanjiang, 157011, People’s Republic of China; 2Department of Neurosurgery, Tongji Hospital, Tongji Medical College, Huazhong University of Science and Technology, Wuhan, 430030, People’s Republic of China; 3Department of Histology and Embryology, Tongji Medical College, Huazhong University of Science and Technology, Wuhan, 430030, People’s Republic of China

*These authors contributed equally to this work

Correspondence: Xingjiang Yu, Department of Histology and Embryology, College of Basic Medicine, Tongji Medical College, Huazhong University of Science and Technology, #13 Hangkong Road, Baofeng Street, Qiaokou District, Wuhan, Hubei Province, 430000, People’s Republic of China, Tel/Fax +86-13146677984, Email [email protected]

Abstract: Cofilin, as a depolymerization factor of actin filaments, has been widely studied. Evidences show that cofilin has a role in actin structural reorganization and dynamic regulation. In recent years, several studies have demonstrated a regulatory role for cofilin in the migration and invasion mediated by cell dynamics and epithelial to mesenchymal transition (EMT)/EMT-like process, apoptosis, radiotherapy resistance, immune escape, and transcriptional dysregulation of malignant tumor cells, particularly glioma cells. On this basis, it is practical to evaluate cofilin as a biomarker for predicting tumor metastasis and prognosis. Targeting cofilin regulating kinases, Lin11, Isl-1 and Mec-3 kinases (LIM kinases/LIMKs) and their major upstream molecules inhibits tumor cell migration and invasion and targeting cofilin-mediated mitochondrial pathway induces apoptosis of tumor cells represent effective options for the development of novel anti-malignant tumor drug, especially anti-glioma drugs. This review explores the structure, general biological function, and regulation of cofilin, with an emphasis on the critical functions and prospects for clinical therapeutic applications of cofilin in malignant tumors represented by glioma.

Keywords: cofilin, glioma, malignant tumor, migration, invasion, apoptosis

Graphical Abstract:

Introduction

Cofilin belongs to the actin-depolymerizing factor family, attached to the actin-binding protein in eukaryotic cells.1 Researcher first isolated cofilin from the chicken embryo brain.2 Emerging reports show that, except for red blood cells and sperm cells, cofilin is expressed in nearly all eukaryotic cells.3–5 Cofilin is preferentially distributed in cell regions characterized by a high turnover rate of actin filaments, including the edge of dividing cells and the front part of migrating cells.6,7

Compelling pieces of evidence have demonstrated an indispensable role for cofilin in regulating the actin-cytoskeleton dynamics. Cofilin mainly promotes depolymerization and polymerization of actin filaments by cleaving actin filaments, accelerating the depolymerization of actin monomers from the pointed end of actin filaments and stimulating actin nucleation.8–10 Evidence indicates that the above process is contingent on the pH of the microenvironment and the concentration ratio of cofilin relative to actin and other actin-binding proteins.11 The phosphorylation/dephosphorylation of the N-terminal Serine3 (Ser3) site of cofilin functions as a switch in the assembly and disassembly of actin. Cofilin can be inactivated through phosphorylation driven by LIMKs and testicular protein kinases (TESKs). However, slingshot (SSH) and chronophin (CIN) can dephosphorylate cofilin to restore its activity.12–14

For decades, since the discovery of cofilin, there is increasing interest in finding the potential application of cofilin-driven regulation of cell movement to malignant tumor cells. Recent reports have demonstrated high cofilin expression in various malignant tumor cells, including glioma.15–17

Existing studies provide concrete evidence that cofilin is associated with regulating the migration, invasion, EMT/EMT-like, and apoptosis of malignant tumor cells and other important physiological functions both in vivo and in vitro. As the “driver” of tumor cell migration and invasion, the formation and maintenance of cell pseudopodia (including filopodia, lamellipodia, and invadopodia) are contingent on actin polymerization and depolymerization. Based on the precise adjustment of actin homeostasis, cofilin potentially regulates the direction of tumor cell migration by inducing the formation of lamellar pseudopods.18 Additionally, studies indicate that the special spatial structure of the pseudopodia and enzymes secreted by pseudopodia which degrade the extracellular matrix can effectively promote the invasion and metastasis of tumor cells.19–21All in all, existing researches suggest activated cofilin has four different functions in gliomas: (I) promote tumor migration and invasion by regulating cell dynamics and EMT/EMT-like process; (II) promote tumor cell apoptosis via the mitochondrial pathway; (III) enhance tumor radioresistance through activating Ras-related C3 botulinum toxin substrate 1 (Rac1)/Wiskott-Aldrich syndrome protein (WASP)-family verprolin-homologous protein2 (WAVE2)/Actin-related protein2/3 (Arp2/3) signaling pathway. (IV) is Involved in chemoresistance, immune escape and transcriptional dysregulation. Based on existing research results and reasonable inductive reasoning, it is practical to evaluate cofilin as a biomarker for predicting tumor metastasis and prognosis. Targeting cofilin regulating kinases, LIMKs and their major upstream molecules inhibits tumor cell migration and invasion and targeting cofilin-mediated mitochondrial pathway induces apoptosis of tumor cells represent effective options for the development of novel anti-tumor drugs.

Here, to explore the precise functions of cofilin in malignant tumors, including glioma, this review clarifies the structure, general biological function, and regulation of cofilin, with an emphasis on the critical functions and prospects for clinical therapeutic applications of cofilin in malignant tumors represented by glioma.

Methods

Relevant articles from PubMed/Medline and Embase (1980–2022) were searched and collected using the phrases

Cofilin, Glioma, Malignant tumor, Migration, Invasion, EMT/epithelial to mesenchymal transformation, Apoptosis, Radiotherapy resistance, Immune escape, Transcriptional dysregulation, Prognosis biomarker.

The Gene Characteristics and Protein Structure of Cofilin

The exon-intron junction sequence of the cofilin gene is quite conservative among different species. Most mammals, including human, are characterized by two cofilin variants, non-muscular type (cofilin-1) and muscle type (cofilin-2). At the genetic level, cofilin-1 gene and cofilin-2 gene are located on chromosome 11q13 and chromosome 14, respectively.22,23 Although cofilin-2 can transcribe two different messenger RNAs (mRNAs), the same polypeptides are produced via selective splicing. It is speculated that the location and/or stability are different. Moreover, the introns may confer a change in the tissue specificity of cofilin gene expression, thereby increasing its expression level.24

The structure of cofilin protein is highly conserved in both non-functional and functional region, with a common fold comprising five internal β-chains surrounded by four α-helices. It is a sequence in which cofilin binds to actin, promoting the depolymerization of filamentous actin, referred to as actin depolymerization factor homology (ADF-H) domain.25,26 Studies have demonstrated the existence of this homologous domain in other actin-binding protein families, including Abp1p, drebrins,27 twinfilin,28 and coactosin,29 with a size range between 113 amino acid (AA) and 168AA.27 The differences in chemical properties and small structural changes of the surface exposure residues are ascribed to the diversity of ADF-H domains of different actin-binding proteins in binding to globular actin (G-actin) and filamentous actin (F-actin).28 Different from actin depolymerization factor (ADF), two types of cofilin have a short β chain which binds to the β34 chains at the C-terminal, changes the environment adjacent to Lysine96 (Lys96) and exerts a direct impact on the binding of cofilin to F-actin. Also, the presence of a short sequence of Tryptophan100 (Trp100) ~ Methionine115 in cofilin protein, which binds to phosphatidylinositol 4-diphosphate, contributes to the interaction of cofilin with actin. In vivo and in vitro experiments have shown that human cofilin is a pH-sensitive actin-depolymerizing protein. Structural analysis reveals that cofilin may be related to a unique salt bridge comprising Aspartic acid98 and Histidine133 at positions 98 and 133, respectively.30,31

Basic Functions of Cofilin

The most critical role of cofilin is to mediate the structural recombination and enhance the dynamic balance of actin. Actin filaments maintain a certain length in a steady-state through “Treadmilling”, an adenosine triphosphate (ATP)-dependent process32,33(Figure 1).

Figure 1 Cofilin participates in the “treadmilling” model of actin filaments. The combination of the abundant G-actin monomers provided by the G-actin pool with ATP can continuously bind to the barbed end (plus end) of F-actin to maintain the continuous growth of actin filaments. On the one hand, activated cofilin binds to ADP-actin on F-actin, cleaving and disaggregating actin filaments. Then, the dissociated ADP-G-actin quickly exchanges ATP and ADP to supplement G-actin monomer pool. On the other hand, cofilin can also bind to the dissociated ADP-G-actin, thereby preventing the exchange of nucleotides in the ADP-G-actin complex and inhibiting the polymerization of actin. LIMK1 can inhibit this process by phosphorylating cofilin. Conversely, SSH1 can dephosphorylate cofilin and activate it. Created with BioRender.com.

In previous research, the microfilaments were artificially divided into two ports (barbed end and pointed end) according to the arrow-shaped decorative morphology of myosin combined with F-actin. The assembly speed of the barbed end is faster than the pointed end, which is why it is called the plus end, whereas the pointed end is called the minus end.34 Under physiological conditions, the polymerization of actin monomers, including G-actin occurs via four processes: actin activation, nucleation, elongation, and steady-state.35 The nucleation period is notably a key step in actin polymerization.36 Actin can be polymerized into dimer, trimer, or tetramer. However, actin dimer is extremely unstable. An actin dimer can form an actin polymer core, and continue to polymerize which is mediated specifically by some nucleating proteins such as ARP2/3 complex, formins, spirel, etc.37,38 Actin monomers binding to ATP exhibit a high affinity to the plus end of F-actin. In this view, actin monomers consecutively bind to the plus end to allow the continued growth of F-actin.39 As the F-actin growing, the bound ATP is potentially hydrolyzed into adenosine diphosphate (ADP) and phosphoric acid. In consequence, the affinity decreases between ADP-actin and the minus end of F-actin. Driven by energy from hydrolysis of ATP, ADP-actin can readily fall off from the minus end.40 Subsequently, the dissociated ADP-G-actin rapidly substitutes ADP with ATP in which profilin promotes the exchange of nucleotides.41 In a steady-state, the growth rate of the plus end of the actin filament is equal to the dissociation rate of the minus end. As a result, the polymerization and depolymerization of actin attain a balanced state, and the actin filament maintains a relatively constant length.42,43 Cofilin and capping protein independently act on actin to realize the precise regulation of “Treadmilling” of actin filaments.44

The kinetic model of Roland et al demonstrates that in order to exert the depolymerization activity, cofilin cleaves actin filaments and increases the rate of that actin monomers leave the ends of actin filaments.45 Cofilin binds to ADP-actin on F-actin in a 1:1 molar ratio, destroys and cuts actin filaments by rotating each actin subunit to approximately 4–5° and twisting the actin filaments to reduce the crossover length. The cleaving of actin filaments results in several short actin fragments, which avails more ADP-actin to bind to cofilin, and accelerate actin depolymerization under sufficient cofilin level.34,46 Cofilin also effectively increases the depolymerization rate of actin monomers from the end of actin filaments. Fluorescence-labeled ADP-binding actin subunits have been utilized in vitro experiments. Reports indicate that, according to fluorescence intensity changes before and after cofilin incubation, the nucleotide exchange rate increases in a concentration-dependent manner after cofilin incubation. This indirectly effectuates an increase in the actin monomer depolymerization rate.8,47 On the other hand, the binding of cofilin to dissociated ADP-G-actin is a potential hindrance to the exchange of nucleotides in the ADP-G-actin complex, delaying the transition to ATP-G-actin. Consequently, this blocks the reassembly of actin monomers into F-actin and indirectly contributes to depolymerization.48

The cofilin with actin cleaving activity can promote the polymerization of actin monomers. Cofilin depolymerization generates more free actin monomers which can be used for polymerization, supplement the monomer pool and increase the effective concentration of actin monomers. The resultant ATP-binding actin monomers can be added to the plus end of F-actin when needed, thereby maintaining the high elongation of actin filaments and enhancing the dynamics of the actin cytoskeleton. Cofilin also drives actin polymerization directly or indirectly by producing several free plus ends.35,49

In addition to the above, cofilin also can stimulate nucleation by stabilizing the “long-distance” actin dimer (the first intermediate in spontaneous assembly).50 Previous studies have demonstrated two mutants of cofilin that affect cofilin binding to F-actin: one involves two basic amino acid residues at the beginning of the β4 chains, whereas the other involves three amino acid residues (two basic and one acidic) in the C-terminal helix.51 Pope et al constructed two mutants K96Q and S3D via site-directed mutagenesis and confirmed the existence of two different actin-binding sites on cofilin. In the state of actin-binding, the interaction of the two sites with different subdomains of the two subunits in the double initiation helix promotes dimerization of actin monomers, consequently inducing their nucleation and assembly.52 In the co-cultured experiment of cofilin and actin, the nucleation activity of actin decreased at the cofilin to actin ratio exceeding 2:1. This observation could be ascribed to the excessive cofilin binding to two cofilin-binding sites on the actin monomer, which blocked the dimerization of the double-bound actin monomer. The experiment further revealed that appropriate cofilin concentration could promote the dimerization and polymerization of actin monomers, and accelerate nucleation.53

Collectively, in order to promote the dynamics of the actin cytoskeleton, cofilin induces the depolymerization and cleaving of actin filaments as well as the polymerization and nucleation of actin monomers.

Regulation of Cofilin

The activity of cofilin protein is primarily tuned by phosphorylation and dephosphorylation. In particularly, the phosphorylation/dephosphorylation of the most conserved N-terminal Ser3 is the key regulatory factor of the assembly and disassembly of cofilin-actin.3,54 Phosphorylation of cofilin on the Ser3 site inhibits the activity of cleaving F-actin to produce actin monomer and decreases the intracellular concentration of G-actin. This consequently increases the number and length of F-actin and slightly improves the F-actin stability. Contrarily, dephosphorylation of the Ser3 site activates F-actin depolymerization.48,55,56

There are two main types of phosphorylases involved in cofilin Ser 3 phosphorylation, include LIMKs and TESKs.57–59 LIMKs, including LIMK1 and LIMK2, are serine/threonine kinases in eukaryotes. LIMK1 is mainly expressed in brain tissue, especially in the cerebral cortex, and moderately expressed in embryonic and mature tissues, heart, and skeletal muscle. LIMK2 is widely expressed in nearly all tissues.60,61 Structural analysis revealed two N-terminal LIM domains of LIM kinase, including a C-terminal kinase domain and a PDZ (Postsynaptic density 95, PSD-85; Discs large, Dlg; Zonula occludens-1, ZO-1) domain. The LIM domain comprises a pair of zinc finger domains containing abundant cysteine/histidine sequences. The LIM domain allows for LIMK interaction with many macromolecules, therefore contributes largely in regulating kinase activity.62

Various proteins and miRNA (microRNA)s are involved in regulating the activity of LIMK. LIMKs could be activated by effector kinases of Ras homology (Rho),63 Rac,57,58 and cell division control protein 42 (CDC42) proteins,64 including Rho-associated kinase1/2 (ROCK1/2),65,66 p21-activated protein kinase1/2/4 (PAK1/2/4),67 and myotonic dystrophin-related CDC42-binding kinase-α.68 To achieve this, these effector kinases phosphorylate the conserved threonine residues in the LIMK domain (Threonine508(Thr508) in LIMK1 or Thr505 in LIMK2), which are the primary factors tuning the activity of LIMKs.66,69 Bone morphogenetic protein (BMP) inhibits both the activity of LIMK1 and its response to ROCK driven by the interaction of type II BMP receptor (BMPR) tail region with LIMK1.70 However, this effect seizes with the formation of BMP-BMPRII-BMPRIA complex and LIMK detachment from the tail of BMPRII.70 Vascular endothelial growth factor (VEGF) induces the activation of mitochondrial activated protein kinase (MAPK)-activated protein kinase-2 (MAPKAPK-2/MK2) by activating MAPK.71 On the other hand, MK2 mediates LIMK1 phosphorylation at Ser323 site directly, and regulates cell migration via the LIMK1/cofilin signal pathway.71 Moreover, MK2 activation induces the phosphorylation of heat shock protein27 (Hsp27). Studies show that phosphorylated Hsp27 potentially promotes actin polymerization by relieving the capping effect on actin filaments, and this drives the formation of pseudopodia and enhances cell migration ability.71,72 Elsewhere, Pandey et al found that defatted invertase lysophosphatidic acid could regulate platelet aggregation and secretion by inducing phosphorylation/dephosphorylation of cofilin in platelets enriched with fibrinogen.73 Particularly, the regulatory mechanism occurs in two steps: Ca2+-mediated rapid dephosphorylation of cofilin and ROCK/LIMK1-dependent cofilin re-phosphorylation. MiRNA-134 is suggested to bind to the mRNA of LIMK, blocking LIMK transcription. Related studies have revealed an excellent clinical application prospect of miRNA-134 in epileptic seizure inhibition and antidepressant.74,75 However, TESK1 is regulated via integrin-mediated signaling pathways, which is independent of PAK1/ROCK activation. This mechanism significantly promotes integrin-mediated cofilin phosphorylation and actin recombination.59,76

Current research demonstrates that dephosphorylation of cofilin on Ser 3 site is potentially mediated by two specific protein phosphatases, SSH and CIN.48,77,78 A previous exploration revealed changes in the morphology of Drosophila epithelial cells lacking the SSH gene, suggesting that SSH may influence cell extension by regulating the polymerization and depolymerization of actin filaments.78 Through in vitro experiments, researchers have shown that SSH and its two homologs (hSSH-1/2L) promote the dissociation of F-actin via dephosphorylation and activation of cofilin. SSH also performs upstanding dephosphorylation activity of LIMK. The direct binding of the N-terminal 1–533 of SSH1 to the LIMK1 kinase domain 310–633 has an effect on the Thr508 site of LIMK1, thereby blocking the activity of LIMK1.79,80 PAK4 and protein kinase D1 phosphorylated hSSH have been demonstrated to inhibit the phosphatase activity of SSH on LIMK1 and cofilin. Existing reports indicate that SSH can enhance phosphatase activity by binding to F-actin via multiple F-actin binding sites. In vitro co-immunoprecipitation assay demonstrates that at least three sites or domains in SSH1 are involved in binding F-actin, including Trp458 site, N-terminal leucin-His-Lys motif, and C-terminal Lys-ketoglutarate reductase motif.81–83 Simultaneously, the N-terminal or C-terminal domain of SSH can stabilize F-actin assembly, independent of the phosphatase activity of SSH.84

Mounting evidence shows that cofilin-mediated depolymerization of actin filaments and the polymerization of actin monomers are contingent on the concentration of active cofilin, the relative concentration of actin monomers, and some actin-binding proteins.10,39,85,86 Among these factors, the local concentration of active cofilin in cells may be a direct determinant of its action mechanism on actin filaments. At low cofilin concentration, some cofilin binds to actin filaments, and the number of torsional strain interfaces between twisted regions (long-range effects of cofilin binding) and non-twisted regions of actin filaments is the largest. At this time, the cleaving activity of cofilin to F-actin filament is at the maximum.87 Within a certain concentration threshold, an increase of cofilin concentration promotes the cooperative binding of cofilin to F-actin. As a result, cofilin cuts F-actin briefly and stabilize it in a twisted state.88 Conversely, in vitro experiments indicate that high cofilin concentrations promote actin nucleation and polymerization, with no cut-off. However, in cases where the intracellular cofilin is abnormally overexpressed or is subjected to oxidative pressure, the ratio of cofilin to actin exceeds a certain critical value. In consequence, cofilin-ADP-actin aggregates into bundles, forming cofilin–actin rods, and blocks cofilin-driven depolymerization and cleavage of F-actin.89

Cofilin in Malignant Tumors Represented by Glioma

Compared to normal tissues, cofilin expression is significantly increased in glioblastoma90 and a variety of other tumors including hepatoblastoma,91 non-small cell lung cancer,92 prostate cancer,93–95 breast cancer.94,96 The analysis of clinical-pathological samples has shown the positive correlation of high expression of cofilin with the malignant degree of the tumor, metastasis risk, tumor cell dedifferentiation, and the short survival time of patients94,95,97,98 (Table 1).

Table 1 Changes in Expression Levels of Cofilin in Tumor Cells and Their Correlated Phenotypes

In vitro experimental evidence suggests that after cofilin knockout, the proliferation, migration, and invasion of tumor cells are inhibited in varying degrees. On the contrary, overexpression of cofilin enhanced these abilities, demonstrating a close correlation of cofilin with tumor proliferation, migration, and invasion. These facts suggest an indispensable role for cofilin in tumorigenesis and development.

Cofilin Regulates Tumor Migration and Invasion

The migration and invasion of tumor cells in the peripheral tissue promote the spread of tumor and are the main causes of death of patients with malignant tumors.

Cofilin Regulates Tumor Cell Pseudopodia Formation by Affecting the Recombination of the Actin Cytoskeleton and Regulating Cell Dynamics

Many inducers of the invasion and metastasis of tumor cells, such as WASP family proteins, Arp2/3 complexes, LIMKs, and cofilin, bind to cell surface receptors to stimulate intracellular signal transduction pathways, regulate actin cytoskeleton recombination, and induce tumor cell invasion and metastasis.110,111 Several pieces of proof have confirmed that enhancing or inhibiting cofilin expression is potentially associated with significant differences in cell dynamics of metastatic tumors, thereby influencing the metastasis and invasion of cancer. Through transient and stable overexpression of cofilin in human glioblastoma cell line U373 MG, Yap et al found that cofilin overexpression below the threshold concentration of cofilin could ameliorate the movement rate of tumor cells and enhance their invasion and migration ability in a concentration-dependent manner.112 In particular, the migration rate of cells in the experimental group with overexpressed cofilin levels was 4.5 times higher than that in the untreated control group. On the other hand, inhibition of cofilin expression via short hairpin RNA-mediated silencing of cofilin could reduce cell movement.112 A variety of glioblastoma patient-derived tumor cell lines, including U87, U251, LN229 and D54 demonstrated accelerated motility to invade the normal brain tissue under hypoxic conditions.102 Inhibition of Src family kinases by Dasatinib eliminates this phenotype. The correlation analysis identified multiple molecules including cofilin and other three molecules associated with motility in U251 cells. The findings support the idea that cofilin plays a key role in hypoxia-induced accelerated glioma cell invasion.102

Cytoskeleton, especially the microfilament cytoskeleton composed of actin, has a crucial role in the movement of cell pseudopodia. Cell pseudopodia make an indispensable contribution to the migration and invasion of tumor cells, including filopodia, lamellipodia, invadopodia, podosomes, etc. Actin is the primary component of motor cell pseudopodia and the main player in the invasion and metastasis of tumor cells.113,114 Related studies employing the immunofluorescence labeling method have demonstrated that actin microfilaments composed of actin are widely distributed in the pseudopodia of motor cells, whereas the cytoskeleton is composed of tubulin is mainly distributed in the interior of the cells. Studies using in vivo fluorescence labeling method of actin have confirmed that the microfilament cytoskeleton plays a vital role in the extension and morphological construction of the leading edge of metastatic tumor cells.115 In this view, actin-binding proteins, including cofilin, play a critical role in maintaining the dynamic changes of actin, lamellipodia formation, and cell movement, and are indispensable regulators of cellular pseudopodia formation.116 Ghosh et al using the cage mimic of cofilin performed experiments in tumor cells to avoid the trapped mimic being phosphorylated.117 In the same experiment, caged cofilin increased the plus end of actin in the body and extended pseudopodia in the uncaged cofilin.117 The findings demonstrate that dephosphorylated cofilin promotes the recombination of the actin cytoskeleton through depolymerization and re-polymerization, induces the formation and extension of pseudopodia on the leading edge of cells, and regulates the “Step forward” of cells, to ameliorate tumor metastasis. Simultaneously, the highly localized activity of cofilin is directed to produce cellular pseudopodia, which determines the direction of cell movement and is the “steering wheel” of cells.118 Studies have also revealed that cofilin can competitively bind to actin with Arp2/3 complex.119 Such binding greatly weakens the affinity between Arp2/3 complex and actin filament, accelerates the removal of actin filament branches, and regulates the movement of tumor cells.119

Previous outcomes indicate that through phosphorylation of multiple amino acid sites, the cofilin activity is regulated via the LIMK-Cofilin pathway. This consequently regulates cytoskeleton remodeling. Similarly, in tumor cells, through the Cofilin pathway, a series of biological behaviors such as invasion and migration of tumor cells can be tuned120 (Figure 2).

Figure 2 Cofilin regulates tumor migration and invasion represented by glioma. Dephosphorylating cofilin on Ser3 mediates its activation. Activated cofilin promotes tumor cell migration and invasion by regulating cell dynamics and EMT/EMT-like process. Cofilin can be inactivated through phosphorylation driven by LIMKs on Ser3 site, while SSH and CIN can dephosphorylate cofilin to restore its’ activity. LIMK is activated by the Rho-GTPases including Rac, CDC42 and Rho through their effector kinases, PAK1 and ROCK1/2 mediating the phosphorylation of LIMK on Thr508 resulting in its activation. Multiply proteins known to regulate the activity of cofilin via Rho-GTPases/ROCK&PAK/LIMK/cofilin signaling pathway resulting in regulating the actin-cytoskeleton dynamics, are SEP17, P2Y2R, NKCC1, CIRP8, Nogo-A, CD73, PRP4, AKT2, PKCζ, Cathepsin B, uPAR, P4, WDR1, Arp2/3. Some non-coding RNAs including miR451, miR29-A/b/c and ARST are also considered to participate in the regulation of cofilin through similar process. Non-SMAD signaling of TGF-β can induce rapid activation of RhoA, followed by phosphorylation and activation of LIMK2/cofilin1 pathway, resulting in regulating EMT/EMT-like process. Furthermore, TWIST1, KITENIN, VSIG4 and β-elemene are considered to be involved in the direct regulation of EMT/EMT-like process. Created with BioRender.com.

Multiple studies have demonstrated high expression of LIMIK1 in malignant gliomas, prostate tumors, breast cancer, lung cancer, etc. In these cases, LIMK1-induced regulation of the balance between cofilin phosphorylation and dephosphorylation can promote the invasion and migration of tumor cells.121–123 Cofilin is the only known enzyme substrate of LIMK1. LIMKs inactivate cofilin by phosphorylating Ser3 at the amino-terminal of cofilin, thus blocking cofilin-driven depolymerization of F-actin. These events consequently, accelerate F-actin polymerization to a lamellipodia structure, which changes the adhesion state of tumor cells to the extracellular matrix and promotes the movement and migration of tumor cells. Previous work by Guo et al indicated that in LN-229 cells, the down-regulation of protein kinase Czeta (PKCζ) expression resulted in dynamic remodeling faults of cytoskeleton in response to the epidermal growth factor (EGF) by dysregulation of LIMK/cofilin phosphorylation.124 Similarly, down-regulation of Intersectin1-s (ITSN1-s) inhibited EGF-induced activation of PAK1/LIMK/cofilin, corresponding to their faults in G-actin polymerization.125 LIMK is a major downstream mediator of PAK1. Active PAK1 phosphorylates LIMK by binding it at Thr508 sites in helix C and the activation loop, which results in an absolute increase in the activity of LIMK toward cofilin. Conversely, inhibiting the activity of PAK1 could suppress the LIMK-mediated F/G-actin structural remodeling.64 The results verified that the down-regulation of ITSN1-s resulted in cytoskeleton dynamic remodeling faults in response to EGF by dysregulation of PAK1/LIMK/cofilin phosphorylation in LN229 cells.125 The discovery by Zhu et al holds that abnormal TSG101 expression in human glioma patients potentially induces the activation of the member C of Rho (RhoC)/LIMK1/Cofilin pathway.15 Rho GTPases activate ROCK, which in turn activates LIMK to mediate cofilin inactivation through phosphorylation of Ser3 sites. As a consequence, actin filament cleavage is inhibited, and this ultimately promotes the migration and invasion of human gliomas. Contrarily, the ability of tumor cell migration and invasion can be inhibited significantly by the interference of LIMIK1 with small hairpin RNA. In a 3D culture system, LIMK inhibition potentially reduces the level of cofilin phosphorylation, inhibits the extension of F-actin, damaging its stability. In addition, by blocking the activity of LIMK, the transcriptional activity of serum reactive factors and the contractility of collagen decreased significantly, consequently inhibiting both the degradation of the extracellular matrix and the formation of invasive pseudopodia.126 Enzymes secreted by invasive pseudopodia can degrade and reconstruct the extracellular matrix. The action of these enzymes reduces the resistance of tumor cell migration and greatly assists tumor cells to invade through the basement membrane and blood vessel wall.20,127,128 In addition, the Rho/ROCK signaling pathway involved in significantly greater contractile force generation required for the assembly of stress fibers and tail retraction during cell movement. Using specific Rho kinase inhibitors targeting this pathway reduces the level of phosphorylated myosin II and cofilin, subsequently prevents tumor cell migration by inhibiting the dynamic assembly of f-actin structures.101 P2Y2 nucleotide receptor (P2Y2R),129 reticulon-4 isoform A (Nogo-A),130 electroneutral Na+-K+-2Cl-co-transporter1 (NKCC1),131 miR-451,132 C1q tumor necrosis factor-related peptide8 (CTRP8),133 miR-29a/b/c134 etc. are considered regulating phosphorylated LIMK1/2 and cofilin through targeting different members of Rho family of GTPases, including Rac1, the member A of Rho (RhoA) and CDC42, subsequently regulating glioma cell migration and invasion. Hou et al reported that Septin 7 (SEPT7) is considered to be a cytoskeletal regulatory protein with GTPase activity, which is involved in the progression of glioma.134 It was found that SEPT7 overexpression reduced the ratio of F-actin/G-actin and the level of total cofilin in glioma cells, while increased the level of p-cofilin, suggesting that SEPT7 inhibited the migration of glioma cells by promoting cofilin phosphorylation and accelerating F-actin depolymerization.135 Numerous kinases have been identified as crucial participants in regulating the migration and invasion of tumor cells by directly interfering with and regulating the phosphorylation and dephosphorylation of cofilin in glioma. Zhang et al showed that Akt serine/threonine kinase 2 (Akt2) mediated the EGF-induced activation of cofilin, suggesting that Akt2 directly involved in cytoskeleton dynamic remodeling of glioma cells by regulating actin polymerization.136 Schulze et al observed a negative correlation between CIN expression and level of cofilin phosphorylation in glioma tissue samples.137 In vitro verification of patient-derived glioblastoma cell lines indicated that chronophin-depleted cells showed increased phosphorylation of cofilin and polymerized actin, consistent with the cell phenotypes of enhanced cell migration orientation, invasiveness in vitro.137 In vitro experiments further showed CIN knockdown increased the relative expression level of p-cofilin, which was consistent with decreased efficiency of colony formation in glioblastoma cell lines. Furthermore, treatment with Y-27632 (ROCK-inhibitor) resulted in the opposite phenotype compared to colony formation, supporting the hypothesis that phosphorylation regulation of cofilin leads to this phenotype.138 These findings proved that cofilin is one of the substrates of CIN, which has been seen as a major regulator of cell migration and invasion in glioma. Cathepsin B,139 urokinase-type plasminogen activator receptor (uPAR),139 Progesterone (P4),140 WD-repeat containing protein1 (WDR1)100 have been also thought to be important activators of cofilin dephosphorylation, play key roles in glioma cell migration and invasion. In contrast, ARST (aldolase A/ALDOA-related specific transcript, a novel long non-coding RNA) can directly bind to the glycolytic enzyme ALDOA, which binds to cofilin to maintain an orderly dynamic balance of the polymerization and depolymerization of actin filaments, thereby exerting a tumor-suppressing effect141 (Table 2).

Table 2 Targets of Tumor Migration and Invasion Mediated by Cofilin

As one of the key regulators of cofilin activity, activated SSH phosphatase in tumor cells activates cofilin. To achieve this, activated SSH phosphatase induces cofilin dephosphorylation, increases the concentration and stability of F-actin, accelerates F-actin polymerization and nucleation, induces lamellipodia formation, and promotes tumor cell metastasis and invasion. On the contrary, SSH inactivation blocks the above process83,142,143 Furthermore, the C-terminal Trp458 site and N-terminal pleckstrin homologous domain of SSH phosphatase domain has been demonstrated to regulate cofilin phosphatase activity by binding F-actin.84,144

It is noteworthy that tension of actin filaments might affect the activity of cofilin. Mounting evidences explore that actin filaments can directly sense the change of tension by regulating their sensitivity to cofilin-mediated cleaving. Briefly, the tension in actin filaments reduces cofilin binding, which effectuates a decrease in cleavage activity. Cofilin preferentially binds to soft and twisted F-actin. As stretching increasing the tension in the filament, the twisted rotation of the filament decreases accordingly, consequently inhibiting the interaction of cofilin with F-actin, and promoting the movement and migration of tumor cells by enhancing F-actin stability and promoting lamellipodia formation.145,146

Cofilin Regulates Tumor EMT/EMT-Like Process

EMT is the process by which epithelial cells acquire mesenchymal characteristics, and is also an important biological pathway via which various epithelial-derived malignant tumor cells acquire the capabilities of migration and invasion.147 In the EMT process, the apex-basal polarity and the adhesion which directly affects the cytoskeleton and cell shape gradually disappears, leading to an increase in cell foot processes and motility, and enhanced capabilities of migration and invasion.148 Cofilin 1 is a crucial regulator of tumor cell invasion, metastasis and EMT, which is highly correlated with tumor occurrence and development. Cofilin 1-mediated regulation of tumor cell EMT is mainly manifested in cytoskeleton changes and transformation into mesenchymal cells to reduce the dependence on the growth of external matrix adhesion.17

At the molecular level, activation of the EMT program is accompanied by changes in various epithelial and mesenchymal molecular markers. The expression of epithelial molecular markers E-cadherin is down-regulated, while the interstitial molecular markers N-cadherin, vimentin, fibronectin, and β-catenin are up-regulated. Among them, the E-to N-cadherin switch, also called the “cadherin switch” is considered as the most critical landmark event during EMT.149–151

Several studies have revealed EMT or EMT-like changes in malignant glioma cells.152 For example, Tso152 et al reported that a series of gene overexpression related to mature mesenchymal tissue in primary glioblastoma biopsy specimens and cell lines, including cartilage/ligament/tendon (Chitinase-3-like protein 1, collagen type IV and VI, and lysyl oxidase), endothelial/epithelial cells (Interleukin-8, Matrix metallopeptidase 9, Epidermal growth factor receptor, and VEGF and myelo-supportive stromal tissue (Pre-B cell colony enhancing factor 1 and Tenascin-C), etc. These observations provide support that primary glioblastoma has a similar stromal cell phenotype and invasiveness.153 Lser et al found that, different from epithelial tumors, including breast cancer, thyroid cancer, and lung cancer, E-cadherin and N-cadherin are not expressed or are expressed in extremely low levels in most gliomas by analyzing the transcriptome data from the The Cancer Genome Atlas (TCGA) database.154 Glioblastoma multiforme (GBM) might be in an intermediate state between epithelial and mesenchymal phenotypes (tend to mesenchymal state). As such, the “cadherin switch” in EMT of glioma cells may not be essential, though other pathways that mediate tumor invasion to antagonize this effect potentially exist. In this view, the occurrence of glioma EMT cannot be judged only by “cadherin conversion”.154,155 Mikheeva also reported that GBM cells with overexpression of Twist1 which has been confirmed to inhibit E-cadherin and activate N-cadherin in previous research do not undergo the typical conversion of E-cadherin to N-cadherin.156 These pieces of evidence demonstrate that cadherin conversion is not associated with the phenotype of GBM aggressive mesenchymal cells.

In addition to morphological changes, EMT is closely associated with stem cell-like characteristics of cancer cells. Studies exploring various human tumors have revealed that the activation of the EMT induces the initial state of the tumor, also called the cancer stem cells (CSCs) state. CSCs are a subgroup of cancer cells similar to normal stem cells, characterized by the ability to automatically regenerate, proliferate and differentiate into various cancer cell lineages through symmetric and asymmetric cell division. This consequently can induce tumor heterogeneity, drug resistance, and the emergence of specific surface markers.157,158 A previous investigation by Mani159 et al found that the induction of EMT in human breast epithelial cells (HMLEs) could induce the expression of mesenchymal characteristics and stem cell markers. Cells induced by Twists have been shown to exhibit an enhanced ability to form mammary glands. On the other hand, stem cells isolated from HMLEs cells also have the potential to express EMT markers.159 Elsewhere, Gupta et al found that EMT program activation by blocking E-cadherin expression in HMLEs cells can increase the number of CSCs significantly.160 These studies demonstrate a close interaction between EMT and CSCs.159,160 Mounting pieces of evidence have further confirmed the existence of similar effects in glioma cells. For instance, Lee et al investigated that the expression of the cancer-promoting factor, KAI1 C-terminal interacting tetraspanin (KITENIN), in GL261 cells significantly increased the mRNA and protein expression levels of various EMT markers (N-cadherin, Zinc finger E-box binding homeobox1/2 (ZEB1/2), Snail, and Slug), and promoted the expression of glioma stem markers (Aldehyde dehydrogenase 1, CD133, erythropoietin-producing hepatocellular receptor B1, and CD44).161 Another investigation by Zhang et al revealed that overexpression of V-set and immunoglobulin domain-containing 4 (VSIG4) potentially induced EMT, significantly promoted the migration and invasion of glioblastoma U87MG cells, and supported the formation of glioma stem cell phenotypes in U87MG cells, simultaneously.162 On the contrary, let-7g-5p can inhibit EMT by down-regulating VSIG4 expression in glioblastoma. This is matched with the reduced migration and invasion ability of glioma stem cell.162 Zhu et al also demonstrated that the expression of mesenchymal markers, N-cadherin and β-catenin was down-regulated significantly after treatment with β-Elemene, a natural plant drug extracted from Curcuma wenyujin.163 On the other hand, the expression of epithelial marker, E-calcium significantly increased in glioma U87 cells, while the expression of CD133, a stem marker of glioma, was significantly reduced. These data prove that there is a synergistic effect existing between the EMT activation and the increase of CSCs in gliomas, which jointly targets the migration and invasion of tumor cells163 (Figure 2).

A variety of mechanisms have been revealed to play an indispensable role in EMT regulation. EMT-inducing transcription factors, including Snail, Slug, ZEB1/2, Twist1/2, Goosecoid, and Forkhead Box C, etc., can inhibit the expression of E-cadherin which plays an indispensable role in activating EMT.164

Evidence shows that the transforming growth factor-beta (TGF-β) is the most critical factor driving the EMT state of glioblastoma. TGF-β signaling can mediate EMT activation via the classic small mothers against decapentaplegic (SMAD) pathway or non-SMAD pathway. In the classic SMAD pathway, the TGF-β type II receptor is activated by binding to TGF-β. The activated RII recruits and binds to the TGF-β type I receptor, forming an RII-ligand-RI heterodimerization body. Subsequently, the trimer binds to intracellular SMAD family proteins and transfers to the nucleus in the form of a complex, which directly stimulates EMT-induced transcription factors. These events block the expression of epithelial cell marker proteins and promote the expression of a specific protein of mesenchymal cells, thereby promoting the activation of the EMT state. In contrast, in non-SMAD signaling pathways, TGF-β signals trigger the EMT process by activating phosphatidylinositol-3-kinase (PI3K)-AKT- mechanistic target of rapamycin or MAPK signals.149,165 non-SMAD signaling of TGF-β induces rapid activation of RhoA, followed by phosphorylation and activation of LIMK/cofilin1 pathway.17 Zhijun et al revealed that ecto-5′-nucleotidase/NT5E (CD73) can activate the RhoGAP involved in the β-catenin-N-cadherin and NMDA receptor signaling protein by transducing adenosinergic signals in the microenvironment, thereby inhibiting the phosphorylation of the RhoA-LIMK-cofilin pathway. They further showed that CD73 can bind to β-catenin directly and activate the EMT process.166 Elsewhere, UlIslam et al found that pre-mRNA processing factor 4B (PRP4) overexpression potentially induce the expression of phosphoprotein phosphatase 1A, thereby blocking the Rho-ROCK-LIMK-cofilin pathway and inducing cofilin dephosphorylation. This results in actin cytoskeleton rearrangement, down-regulation of E-cadherin, and induction of EMT, which successively promote the progression of human colon cancer167 (Figure 2).

Cofilin Regulates Tumor Cell Apoptosis

Cell death and renewal is an indispensable part of the whole life process of multicellular organisms capable of timely eliminating mechanical and damaged cells and critically contributed to the development of various tissues and organs and the immune system. Apoptosis is a crucial pathway of cell death in vivo. It is a programmed death process precisely regulated by a series of signal molecules, characterized by cell shrinkage, chromatin pyknosis, and apoptotic body shape, evident under light and electron microscope.168 Apoptosis occurs in various stages of embryonic development, tissue remodeling, immune regulation, and tumor degeneration. Of note, unwanted, severely damaged and potentially dangerous cells can be eliminated via apoptosis.169,170

The molecular mechanism of apoptosis is very complex, involving multiple cellular signal pathways and a variety of cytokines. Current evidence shows that eukaryotic cells trigger apoptosis mainly via the mitochondrial pathway or internal pathway and death receptor-mediated pathway or external pathway.171 Under normal circumstances, the apoptosis process is strictly regulated by the body to ensure the stability of various tissues and organs in the whole life process. However, when apoptosis regulation is out of balance, excessive cell proliferation or excessive apoptosis transpires, leading to liver cancer, ovarian cancer, Alzheimer’s disease, etc.172,173

Previous studies have shown that cofilin plays an important role during the initiation phase of apoptosis. After induction of apoptosis, cofilin, activated by dephosphorylation at Ser3 are translocated to the outer mitochondrial membrane before release of cytochrome c into the cytosol.174 Moreover, the actin-binding activity of cofilin is crucial for its apoptosis-inducing activity. Dephosphorylated cofilin translocates to the outer mitochondrial membrane while binding to actin, leading to cytoskeletal remodeling, which in turn affects mitochondrial function. The result is mitochondrial dysfunction, cytochrome c release and apoptosis174 (Figure 3).

Figure 3 Cofilin regulates tumor cell apoptosis. Activated cofilin translocated to the mitochondria can directly bind to the potential cleavage sites in the outer mitochondrial membrane, which interacts with DRP1 to promote cytochrome C release and accelerate mitochondria breakage, resulting in tumor cell apoptosis. Created with BioRender.com.

It deserves a special mention that apoptosis induced by a mitochondrial pathway mediated by cofilin dephosphorylation, is closely associated with the dynamin-related protein1 (DRP1) and phosphatase and tensin homolog (PTEN)-induced kinase1/ Parkinson protein2 pathways, reported in other malignant tumors.175 Mitochondrial mitotic protein, DRP1[175−177], locates in the cytoplasm, exerts function in the form of polymers. Following the mitochondria stimulation by various physical and chemical factors, the mitochondrial outer membrane mitochondrial adaptor fission1 summons the DRP1 and transports it to the potential cleavage site at the outer mitochondrial membrane. DRP1 polymers form a ring structure around mitochondria, which generates energy through guanosine triphosphate hydrolysis, and gradually compresses the mitochondria until it breaks. This process promotes mitochondria division.176–178 In addition, the cofilin translocated to the mitochondria can directly bind to the potential cleavage sites in the outer mitochondrial membrane, forming a similar circular complex around the mitochondria, which interacts with DRP1 to accelerate mitochondria breakage98 (Figure 3). Hu179 et al transfected human breast cancer cell line MDA-MB-231 using activated plasmids cofilin (S3A) and DRP1 (S637A) that mimic cofilin and DRP1 dephosphorylation, and revealed that the two plasmids induced the translocation of cofilin and DRP1 from the cytoplasm to mitochondria, inducing polymer-mediated mitochondrial cleavage. Contrarily, tumor cells transfected with cofilin (S3E) and DRP1 (S637D), which mimic cofilin and DRP1 phosphorylation, could inhibit mitochondrial division and apoptosis.179 These results suggest that the synergistic effect of cofilin and DRP1 is critical in promoting mitochondrial pathway-mediated apoptosis. Recent data shows that mitochondrial translocation of DRP1 and cofilin is vital to this process. Silencing of DRP1 and cofilin to prevent mitochondrial translocation of DRP1 and cofilin respectively was demonstrated to significantly inhibit the mitochondrial division and fragmentation and significantly decrease apoptosis.180,181

Cofilin Regulates Tumor Radioresistance

Radiotherapy is currently one of the most effective non-surgical treatments for a variety of malignant tumors including glioma. Unfortunately, due to the high inherent radiation resistance of glioma cells, especially glioblastoma cells, the effectiveness of radiation therapy is greatly reduced.182,183 Therefore, there is no doubt that it is vital to explore the causes of glioma radiotherapy resistance and enhance the radiosensitivity of glioma.

Through two-dimensional high-performance liquid chromatography- mass spectrometry, Wenbo Sun et al found that both phosphoglycerate kinase1 and cofilin1 are significantly highly expressed in the tumor tissue from glioma patients who are tolerant to radiotherapy.99,103 The capacities of cell viability, migration and invasion were significantly down-regulated and the proportion of cells arrested in G2/M phase was increased in cofilin1‑silenced radioresistant-U251 and U251 cells, compared with that of the untreated cells. In contrast, overexpressing cofilin1 showed progressive malignant behavior of tumor in U251 cells.184 Similarly, when the Rac1-WAVE2-Arp2/3 signaling transduction was blocked, cell proliferation, migration and invasion abilities were decreased with the down-regulation of cofilin1.185 These results indicated that cofilin1 plays an important role in enhancing radioresistance through activating Rac1-WAVE2-Arp2/3-Cofilin signaling pathway in U251 human glioma cells (Figure 4).

Figure 4 Cofilin regulates tumor radioresistance and other malignant behaviors. (A) Cofilin can enhance tumor radioresistance through activating Rac1-WAVE2-Arp2/3-Cofilin signaling pathway. (B) Low-dose EMAP-II promotes increased the activation of MLC and cofilin via RhoA/ROCK/MLC/Cofilin pathway, thereby triggering actin cytoskeleton dynamic remodeling and ultimately increasing the permeability of BTB. (C) HIF1α induced by TNFα impulses cofilin-mediated changes in actin filament dynamics to stabilize MHC-I clusters through affecting mitochondrial localization of HKII, eventually affecting tumor immune-related outcomes. (D) Activation of c-Myc promotes nuclear translocation of cofilin and F-actin cytoskeleton dynamic remodeling, which may in turn affect tumor cell division and metastasis. Created with BioRender.com.

Cofilin Regulates Other Malignant Behaviors of Tumors

The blood-brain barrier (BBB) maintains the homeostasis of the central nervous system by forming neurovascular units, which are consisted of endothelial cells, neurons, glia, smooth muscle cells, and pericytes.186 However, the BBB also blocks effectiveness of treatments for brain tumors. The BBB is disrupted during the progression of brain tumor, resulting in a vasculature with high heterogeneity known as the blood–tumor barrier (BTB).187 BTB exhibits different structural and functional characteristics than normal cerebral capillaries that form BBB. Although BTB is more permeable than BBB, its heterogeneous permeability to small and macromolecules can still hinder the entry of antitumor drugs into brain tumor tissue.187,188 Therefore, it is crucial to explore the methods to selectively open the BTB to improve the effectiveness of chemotherapy for glioma. Low-dose endothelial monocyte-activating polypeptide-II (EMAP-II) promotes increased phosphorylation of myosin light chain (MLC) and cofilin by inducing activation of RhoA/ROCK, thereby triggering actin cytoskeleton dynamic remodeling and ultimately increasing the permeability of BTB189,190 (Figure 4) Furthermore, inhibition of PI3K significantly suppress ed EMAP-II-induced phosphorylation of both MLC and cofilin, suggesting that the activation of PI3K was involved in regulation of MLC and cofilin phosphorylation mediated by RhoA/ROCK signaling pathway.191

Cofilin is involved in the immune escape process of glioma. The expression of Hexokinase-II (HKII) is regulated by hypoxic induced factor-1α (HIF-1α)192,193 which is believed to be the basis for maintaining high metabolic requirements in malignant tumors under hypoxia. This is also conducive to promote various malignant behaviors of tumors, such as immune evasion in gliomas. Considering that HIF-1α regulates gene transcription of Major Histocompatibility Complex class I (MHC-I) induced by inflammation,194 and as actin cytoskeleton is actively involved in enhancing the formation of stable MHC-I clusters, changes of cofilin activity closely correlated with glioma hypermetabolism and immune escape. Ghosh et al demonstrate that HIF-1α induced by tumor necrosis factor alpha both affects mitochondrial localization of HKII and impulses cofilin-mediated changes in actin filament dynamics to stabilize MHC-I clusters in glioma cells, eventually affecting immune-related outcomes195 (Figure 4) In addition, cofilin1 has been identified as one of five key T-cell target immunogenic tumor-associated antigens (TAA) for immunity therapy in mutations in isocitrate dehydrogenase glioma patients.

Notably, one study in medulloblastoma showed cofilin is associated with RNA polymerase II -mediated transcriptional activity.196 The protein sequence of cofilin contains nuclear localization signals, which allows the transport of depolymerized actin to the nucleus,197 thereby regulating transcription and chromatin structure. Activation of c-myc, a transcription factor encoded by the proto-oncogene c-myc could promote nuclear translocation of cofilin and F-actin cytoskeleton dynamic remodeling in medulloblastoma cells which may in turn affect tumor cell division and metastasis196 (Figure 4).

Prospects of Cofilin in Clinical Therapeutic Applications of Malignant Tumors Represented by Glioma

Evaluate Cofilin as a Biomarker for Predicting Tumor Metastasis and Prognosis

Previous studies in various types of tumors have been observed the positive correlation of high expression of cofilin with the malignant degree of the tumor, metastasis risk, tumor cell dedifferentiation, and the short survival time of patients.94,95,97,98 In a retrospective cohort of non-small cell lung cancer patients, survival analysis showed that patient with high cofilin-1 had a lower overall survival rate (P<0.05), which be used to distinguish between good and poor prognosis. However, it is important to note that no correlation was found between age, sex or histological type and patient outcome or level of cofilin.198 Li WP et al reported that the presence of cofilin in Juvenile nasopharyngeal angiofibroma (JNA) was correlated with tumor stage (p=0.012) and volume of intraoperative hemorrhage (p<0.001). The recurrence rate was higher in JNA patients with high cofilin expression than in those with low cofilin expression (p = 0.012). Similarly, high levels of cofilin indicated tumor progression and poor prognosis of patients in glioma,99 pancreatic ductal adenocarcinoma98 and other lethal malignancies.

In conclusion, all of these studies indicate that the expression of cofilin and its activation status are closely associated with the malignancy of tumors, which is expected to be a biomarker for predicting tumor metastasis and prognosis.

Targeting Cofilin Regulating Kinases, LIMKs and Their Major Upstream Molecules Inhibits Tumor Cell Migration and Invasion

LIMKs promote polymerization of actin via the LIMK-Cofilin pathway, which contributes to the formation and extension of actin filaments, resulting in driving membrane forward at the front part of migrating tumor cells. Thus, targeting cofilin regulating kinases, LIMKs and their major upstream molecules is significant for inhibiting tumor cell migration and invasion.

Several antineoplastic drugs have been shown to inhibit cell migration and invasion by interfering with and regularing the kinase activity of LIMKs in glioma. For instance, BMS-5 and Cucurbitacin I have been demonstrated directly blocking the cofilin regulating kinases, LIMK1 and LIMK2. The cell viability of glioma cells treated with BMS-5 and Cucurbitacin I was significant decreases, while no cytotoxicity was observed in normal astrocytes lacking LIMKs. BMS-5 and Cucurbitacin I promoted enhancement of GBM cell adhesion, suppressed cell migration and invasion.90 T56-LIMKi has high specificity in inhibiting LIMK2, without almost any cross-reactivity with LIMK1. It decreases level of phosphorylated cofilin, subsequently inhibits growth of a variety of tumor cell lines, including glioma, schwannoma and pancreatic cancer.199 Wang X et al reported that Alantolactone (ATL), a natural small molecule inhibitor, activated cofilin by specifically inhibiting the activity of LIMKs enzyme, which thus upregulated the ratio of G/F actin through actin polymerization inhibition, and blocked the migration and invasion of GBM cell.200 Treatment with recombinant bone morphogenetic protein 9 protein triggers the activation of SMADs in patient-derived GBM cells, and strongly suppresses cell proliferation and invasion by inhibiting the activation of RhoA/Cofilin and PI3K/AKT/MAPK pathways.201 In addition, it was reported that Eucalyptal A, a natural phloroglucinol-terpene adducts could downregulate expression of Serine/Arginine splicing factor1 (SRSF1) and rescue SRSF1-mediated unexpected alternative splicing of myosin IB mRNA, eventually exert anti-GBM effect by the PAK/Cofilin and 3-phosphoinositide-dependent kinase 1 (PDK1)/AKT/c-Myc signaling pathways.202 As mentioned above, miRNAs, such as miR-451132 and miR-29a/b/c,134 inhibit glioma cell migration and invasion through regulating the cofilin activity via LIMK/cofilin pathway, acting as promising inhibitors of tumor migration and invasion.

Targeting Cofilin-Mediated Mitochondrial Pathway Induces Apoptosis of Tumor Cells

Activated cofilin translocates to the external mitochondrial membrane, inducing Mitochondrial outer membrane permeabilization (MOMP) and lysis of mitochondria, while interacting with DRP1 to cause excessive mitochondrial division and mitochondrial damage, ultimately leading to cytochrome C release and driving the apoptotic process. Therefore, targeting cofilin to induce apoptosis of glioma cells represent effective option for the development of novel anti-tumor drugs.

Various stimulants, including iso-lanolin lactone,203 allyl isothiocyanate (AITC),204 uric acid,205 4- (methylthio) butyl isothiocyanate,206 and ATL200 induce apoptosis of glioma cells via the cofilin-mediated mitochondrial pathway. Isoalantolactone (Iso) is a lactone compound extracted from the roots of Eucalyptus plants, which has been revealed to exert upstanding insect repellent and antibacterial effects.207 Previous in vitro studies show that Iso has an anti-tumor effect.208–210 A recent investigation of glioblastoma cell lines, U87MG, U251, and U118, explored that Iso inhibits cofilin phosphorylation by targeting LIMK kinase and down-regulating LIMK activity. This, in consequence, induces the transfer of cofilin and G-actin to the mitochondrial inner membrane and promotes the release of cytochrome C from the mitochondria to the cytoplasm. AITC induces cofilin dephosphorylation by activating serine/threonine protein phosphatase1/2A and activates mitochondrial pathway-mediated apoptosis by promoting G-actin translocation to mitochondria. Evidence suggests that the blockade of the ROCK1/PTEN/PI3K signal pathway plays an important role in the activation of cofilin.211 Similarly, ATL activated cofilin by specifically inhibiting the activity of LIMKs enzyme and upregulated the proportion of G/F-actin. Activated cofilin can be jointly transferred to the mitochondria for initiating release of cytochrome c across the mitochondrial membrane, which eventually induce glioma cell apoptosis.200

Conclusion

Compelling evidence demonstrates that cofilin, as a depolymerization factor of actin filaments, potentially plays a crucial role in mediating actin structural reorganization and dynamic regulation via different mechanisms, including depolymerization, repolymerization, severing, and nucleation. Several active enzymes such as LIMK and SSH can achieve precise regulation of cofilin through phosphorylation or dephosphorylation. In recent years, emerging pieces of evidence have shown a regulatory role for cofilin in the migration, invasion, EMT, apoptosis, radiotherapy and chemotherapy resistance, immune escape and transcriptional dysregulation of malignant tumors represented by glioma. Given findings, it is practical to evaluate cofilin as a biomarker for predicting tumor metastasis and prognosis. Targeting cofilin regulating kinases, LIMKs and their major upstream molecules inhibits tumor cell migration and invasion and targeting cofilin-mediated mitochondrial pathway induces apoptosis of tumor cells represent effective options for the development of novel anti-malignant tumor drug, especially anti-glioma drugs.

Existing studies generally support that LIMK1/2 are significantly upregulated in glioma (especially glioblastoma) and many other infiltrative tumors, and strongly regulate tumor invasive motility and progression, which support further investigation of LIMK1/2 as options for the development of novel anti-tumor drugs.121–123,212 However, some investigations on the final effect of LIMK reveal the opposite results, in which, LIMK overexpression inhibits tumor metastasis and invasion, whereas LIMK silencing promotes tumor cell metastasis and invasion. Zebda et al found that promotion of LIMK-regulated cofilin phosphorylation in metastatic rat breast cancer MTLn3 cells almost completely inactivated cofilin, significantly decreased actin turnover under sufficient G-actin levels, and almost completely inhibited lamellipodia formation, making the cell protrude, lengthen and move slowly, unidirectionally in a straight line.213 These phenomena are related to the loss of consistent chemosensitivity to EGF stimulators on the surfaces of MTLn3 cells, indicating that LIMK-induced cofilin phosphorylation can not only inhibit the dynamic renewal of actin and lamellipodia formation but also impede the response of whole tumor cells to a chemotactic stimulator in any area of its surface. This eventually influences the chemotaxis and movement of tumor cells.213 Wang et al suggested that it is the overall activity of the LIMK1/cofilin pathway and not that of LIMK1 expression alone that determines the migration and invasion status of malignant tumor cells.214 Furthermore, different tumor cell types used in the different studies might have contributed to the inconsistency of results. Future studies are required to focus on elucidating the differential regulatory mechanisms of the effect of the LIMK/cofilin pathway in gliomas and other malignancies. Besides that, it is important to study whether cofilin has a similar promoting effect on tumor angiogenesis and tumor cell proliferation in gliomas compared to other malignancies, which will be critical for future drug development targeting cofilin.

Abbreviations

EMT, epithelial to mesenchymal transition; LIM kinases/LIMKs, Lin11, Isl-1 and Mec-3 kinases; Ser, Serine; TESKs, testicular protein kinases; SSH, slingshot; CIN, chronophin; Rac1, Ras-related C3 botulinum toxin substrate1; WASP, Wiskott-Aldrich syndrome protein; WAVE2, WASP family verprolin-homologous protein2; Arp2/3, Actin-related protein 2/3; mRNA, messenger RNA; ADF-H, actin depolymerization factor homology; AA, amino acid; G-actin, globular actin; F-actin, filamentous actin; ADF, actin depolymerization factor; Lys, Lysine; Trp, Tryptophan; ATP, adenosine triphosphate; PDZ, Postsynaptic density 95, PSD-85; Discs large, Dlg; Zonula occludens-1, ZO-1; miRNA, microRNA; Rho, Ras homology; RhoC, the member C of Rho; RhoA, the member A of Rho; CDC42, cell division control protein 42; ROCK, Rho-associated kinase; PAK, p21-activated protein kinase; Thr, Threonine; BMP, bone morphogenetic protein; BMPRII/IA, BMPR receptor II/IA; VEGF, vascular endothelial growth factor; MAPK, mitochondrial activated protein kinase; MAPKAPK-2/MK2, MAPK-activated protein kinase-2; Hsp27, heat shock protein27; PKCζ, protein kinase Czeta; EGF, epidermal growth factor; ITSN1-s, Intersectin1-s; P2Y2R, P2Y2 nucleotide receptor; Nogo-A, reticulon-4 isoform A; NKCC1, electroneutral Na+-K+-2Cl-co-transporter1; CTRP8, C1q tumor necrosis factor-related peptide8; SEPT7, Septin 7; Akt2, Akt serine/threonine kinase 2; uPAR, urokinase-type plasminogen activator receptor; P4, Progesterone; WDR1, WD-repeat containing protein1; ARST, aldolase A/ALDOA-related specific transcript; TCGA, The Cancer Genome Atlas; GBM, glioblastoma multiforme; CSCs, cancer stem cells; HMLEs, human breast epithelial cells; KITENIN, KAI1 C-terminal interacting tetraspanin; ZEB1/2, Zinc finger E-box binding homeobox1/2; VSIG4, V-set and immunoglobulin domain-containing 4; TGF-β, transforming growth factor-beta; SMAD, small mothers against decapentaplegic; PI3K, phosphatidylinositol-3-kinase; CD73, ecto-5′-nucleotidase, NT5E; PRP4, pre-mRNA processing factor 4B; DRP1, dynamin-related protein1; PTEN, phosphatase and tensin homolog; BBB, blood-brain barrier; BTB, blood–tumor barrier; EMAP-II, endothelial monocyte-activating polypeptide-II; MLC, myosin light chain; HKII, Hexokinase-II; HIF-1α, hypoxic induced factor-1α; MHC-I, Major Histocompatibility Complex class I; JNA, Juvenile nasopharyngeal angiofibroma; ATL, Alantolactone; SRSF1, Serine/Arginine splicing factor 1; AITC, allyl isothiocyanate; Iso, Isoalantolactone.

Acknowledgments

Graphical abstract and figures presented in this manuscript were created with BioRender.

Author Contributions

S.L. and Z.C. mainly wrote the manuscript and drew all figures and graphical abstract. H.M. revised the spelling and grammar of the manuscript. X.Y. designed the overall review and proofed the manuscript. All authors gave final approval of the version to be published, have agreed on the journal to which the article has been submitted and agree to be accountable for all aspects of the work.

Funding

This work was supported by the National Natural Science Foundation of China (Grant No. 81974452 to X.Y.) and Heilongjiang Provincial Health Commission (Grant No. 2019-393 to S.L.).

Disclosure

The authors declare no conflict of interest.

References

1. Nishida E, Maekawa S, Sakai H. Cofilin, a protein in porcine brain that binds to actin filaments and inhibits their interactions with myosin and tropomyosin. Biochemistry. 1984;23(22):5307–5313. doi:10.1021/bi00317a032

2. Bamburg JR, Harris HE, Weeds AG. Partial purification and characterization of an actin depolymerizing factor from brain. FEBS Lett. 1980;121(1):178–182. doi:10.1016/0014-5793(80)81292-0

3. Bamburg JR. Proteins of the ADF/cofilin family: essential regulators of actin dynamics. Annu Rev Cell Dev Biol. 1999;15:185–230. doi:10.1146/annurev.cellbio.15.1.185

4. Maciver SK, Hussey PJ. The ADF/cofilin family: actin-remodeling proteins. Genome Biol. 2002;3(5):3007. doi:10.1186/gb-2002-3-5-reviews3007

5. Pollard TD, Blanchoin L, Mullins RD. Molecular mechanisms controlling actin filament dynamics in nonmuscle cells. Annu Rev Biophys Biomol Struct. 2000;29:545–576. doi:10.1146/annurev.biophys.29.1.545

6. Bamburg JR, Bray D. Distribution and cellular localization of actin depolymerizing factor. J Cell Biol. 1987;105(6 Pt 1):2817–2825. doi:10.1083/jcb.105.6.2817

7. Svitkina TM, Borisy GG. Arp2/3 complex and actin depolymerizing factor/cofilin in dendritic organization and treadmilling of actin filament array in lamellipodia. J Cell Biol. 1999;145(5):1009–1026. doi:10.1083/jcb.145.5.1009

8. Carlier MF, Laurent V, Santolini J, et al. Actin depolymerizing factor (ADF/cofilin) enhances the rate of filament turnover: implication in actin-based motility. J Cell Biol. 1997;136(6):1307–1322. doi:10.1083/jcb.136.6.1307

9. Maciver SK. How ADF/cofilin depolymerizes actin filaments. Curr Opin Cell Biol. 1998;10(1):140–144. doi:10.1016/s0955-0674(98)80097-5

10. Wioland H, Guichard B, Senju Y, et al. ADF/Cofilin accelerates actin dynamics by severing filaments and promoting their depolymerization at both ends. Curr Biol. 2017;27(13):1956–67.e7. doi:10.1016/j.cub.2017.05.048

11. Bernstein BW, Painter WB, Chen H, Minamide LS, Abe H, Bamburg JR. Intracellular pH modulation of ADF/cofilin proteins. Cell Motil Cytoskeleton. 2000;47(4):319–336. doi:10.1002/1097-0169(200012)47:4<319::Aid-cm6>3.0.Co;2-i

12. Yoo Y, Ho HJ, Wang C, Guan JL. Tyrosine phosphorylation of cofilin at Y68 by v-Src leads to its degradation through ubiquitin-proteasome pathway. Oncogene. 2010;29(2):263–272. doi:10.1038/onc.2009.319

13. Frantz C, Barreiro G, Dominguez L, et al. Cofilin is a pH sensor for actin free barbed end formation: role of phosphoinositide binding. J Cell Biol. 2008;183(5):865–879. doi:10.1083/jcb.200804161

14. Ishikawa-Ankerhold HC, Daszkiewicz W, Schleicher M, Müller-Taubenberger A. Actin-interacting protein 1 contributes to intranuclear rod assembly in dictyostelium discoideum. Sci Rep. 2017;7(1):40310. doi:10.1038/srep40310

15. Zhu Y, Xu Y, Chen T, et al. TSG101 promotes the proliferation, migration, and invasion of human glioma cells by regulating the AKT/GSK3β/β-catenin and RhoC/cofilin pathways. Mol Neurobiol. 2021;58(5):2118–2132. doi:10.1007/s12035-020-02231-7

16. Lee MH, Kundu JK, Chae JI, Shim JH. Targeting ROCK/LIMK/cofilin signaling pathway in cancer. Arch Pharm Res. 2019;42(6):481–491. doi:10.1007/s12272-019-01153-w

17. Sousa-Squiavinato ACM, Rocha MR, Barcellos-de-souza P, de Souza WF, Morgado-Diaz JA. Cofilin-1 signaling mediates epithelial-mesenchymal transition by promoting actin cytoskeleton reorganization and cell-cell adhesion regulation in colorectal cancer cells. Biochim Biophys Acta Mol Cell Res. 2019;1866(3):418–429. doi:10.1016/j.bbamcr.2018.10.003

18. Popow-Woźniak A, Mazur AJ, Mannherz HG, Malicka-Błaszkiewicz M, Nowak D. Cofilin overexpression affects actin cytoskeleton organization and migration of human colon adenocarcinoma cells. Histochem Cell Biol. 2012;138(5):725–736. doi:10.1007/s00418-012-0988-2

19. Wicki A, Lehembre F, Wick N, Hantusch B, Kerjaschki D, Christofori G. Tumor invasion in the absence of epithelial-mesenchymal transition: podoplanin-mediated remodeling of the actin cytoskeleton. Cancer Cell. 2006;9(4):261–272. doi:10.1016/j.ccr.2006.03.010

20. Yu X, Zech T, McDonald L, et al. N-WASP coordinates the delivery and F-actin-mediated capture of MT1-MMP at invasive pseudopods. J Cell Biol. 2012;199(3):527–544. doi:10.1083/jcb.201203025

21. Dalaka E, Kronenberg NM, Liehm P, Segall JE, Prystowsky MB, Gather MC. Direct measurement of vertical forces shows correlation between mechanical activity and proteolytic ability of invadopodia. Sci Adv. 2020;6(11):eaax6912. doi:10.1126/sciadv.aax6912

22. Ono S, Minami N, Abe H, Obinata T. Characterization of a novel cofilin isoform that is predominantly expressed in mammalian skeletal muscle. J Biol Chem. 1994;269(21):15280–15286.

23. Gillett GT, Fox MF, Rowe PS, Casimir CM, Povey S. Mapping of human non-muscle type cofilin (CFL1) to chromosome 11q13 and muscle-type cofilin (CFL2) to chromosome 14. Ann Hum Genet. 1996;60(3):201–211. doi:10.1111/j.1469-1809.1996.tb00423.x

24. Thirion C, Stucka R, Mendel B, et al. Characterization of human muscle type cofilin (CFL2) in normal and regenerating muscle. Eur J Biochem. 2001;268(12):3473–3482. doi:10.1046/j.1432-1327.2001.02247.x

25. Goroncy AK, Koshiba S, Tochio N, et al. NMR solution structures of actin depolymerizing factor homology domains. Protein Sci. 2009;18(11):2384–2392. doi:10.1002/pro.248

26. Singh BK, Sattler JM, Chatterjee M, Huttu J, Schüler H, Kursula I. Crystal structures explain functional differences in the two actin depolymerization factors of the malaria parasite. J Biol Chem. 2011;286(32):28256–28264. doi:10.1074/jbc.M111.211730

27. Peitsch WK, Grund C, Kuhn C, et al. Drebrin is a widespread actin-associating protein enriched at junctional plaques, defining a specific microfilament Anchorage system in polar epithelial cells. Eur J Cell Biol. 1999;78(11):767–778. doi:10.1016/s0171-9335(99)80027-2

28. Lappalainen P, Kessels MM, Cope MJ, Drubin DG. The ADF homology (ADF-H) domain: a highly exploited actin-binding module. Mol Biol Cell. 1998;9(8):1951–1959. doi:10.1091/mbc.9.8.1951

29. Dos Remedios CG, Chhabra D, Kekic M, et al. Actin binding proteins: regulation of cytoskeletal microfilaments. Physiol Rev. 2003;83(2):433–473. doi:10.1152/physrev.00026.2002

30. Pope BJ, Zierler-Gould KM, Kühne R, Weeds AG, Ball LJ. Solution structure of human cofilin: actin binding, pH sensitivity, and relationship to actin-depolymerizing factor. J Biol Chem. 2004;279(6):4840–4848. doi:10.1074/jbc.M310148200

31. Van Troys M, Dewitte D, Verschelde JL, Goethals M, Vandekerckhove J, Ampe C. The competitive interaction of actin and PIP2 with actophorin is based on overlapping target sites: design of a gain-of-function mutant. Biochemistry. 2000;39(40):12181–12189. doi:10.1021/bi000816c

32. Wegner A. Treadmilling of actin at physiological salt concentrations. An analysis of the critical concentrations of actin filaments. J Mol Biol. 1982;161(4):607–615. doi:10.1016/0022-2836(82)90411-9

33. Neuhaus JM, Wanger M, Keiser T, Wegner A. Treadmilling of actin. J Muscle Res Cell Motil. 1983;4(5):507–527. doi:10.1007/bf00712112

34. Pollard TD, Borisy GG. Cellular motility driven by assembly and disassembly of actin filaments. Cell. 2003;112(4):453–465. doi:10.1016/s0092-8674(03)00120-x

35. Plastino J, Blanchoin L. Dynamic stability of the actin ecosystem. J Cell Sci. 2018;132(4). doi:10.1242/jcs.219832

36. Pollard TD, Cooper JA. Quantitative analysis of the effect of Acanthamoeba profilin on actin filament nucleation and elongation. Biochemistry. 1984;23(26):6631–6641. doi:10.1021/bi00321a054

37. Buracco S, Claydon S, Insall R. Control of actin dynamics during cell motility. F1000Res. 2019;8. doi:10.12688/f1000research.18669.1.

38. Pollard TD. Regulation of actin filament assembly by Arp2/3 complex and formins. Annu Rev Biophys Biomol Struct. 2007;36:451–477. doi:10.1146/annurev.biophys.35.040405.101936

39. Van Troys M, Huyck L, Leyman S, Dhaese S, Vandekerkhove J, Ampe C. Ins and outs of ADF/cofilin activity and regulation. Eur J Cell Biol. 2008;87(8–9):649–667. doi:10.1016/j.ejcb.2008.04.001

40. Kueh HY, Brieher WM, Mitchison TJ. Dynamic stabilization of actin filaments. Proc Natl Acad Sci USA. 2008;105(43):16531–16536. doi:10.1073/pnas.0807394105

41. Rotty JD, Wu C, Haynes EM, et al. Profilin-1 serves as a gatekeeper for actin assembly by Arp2/3-dependent and -independent pathways. Dev Cell. 2015;32(1):54–67. doi:10.1016/j.devcel.2014.10.026

42. Pollard TD. Rate constants for the reactions of ATP- and ADP-actin with the ends of actin filaments. J Cell Biol. 1986;103(6 Pt 2):2747–2754. doi:10.1083/jcb.103.6.2747

43. Fujiwara I, Takahashi S, Tadakuma H, Funatsu T, Ishiwata S. Microscopic analysis of polymerization dynamics with individual actin filaments. Nat Cell Biol. 2002;4(9):666–673. doi:10.1038/ncb841

44. Wang R, Carlsson AE. How capping protein enhances actin filament growth and nucleation on biomimetic beads. Phys Biol. 2015;12(6):066008. doi:10.1088/1478-3975/12/6/066008

45. Roland J, Berro J, Michelot A, Blanchoin L, Martiel JL. Stochastic severing of actin filaments by actin depolymerizing factor/cofilin controls the emergence of a steady dynamical regime. Biophys J. 2008;94(6):2082–2094. doi:10.1529/biophysj.107.121988

46. McGough A, Pope B, Chiu W, Weeds A. Cofilin changes the twist of F-actin: implications for actin filament dynamics and cellular function. J Cell Biol. 1997;138(4):771–781. doi:10.1083/jcb.138.4.771

47. Blanchoin L, Pollard TD, Mullins RD. Interactions of ADF/cofilin, Arp2/3 complex, capping protein and profilin in remodeling of branched actin filament networks. Curr Biol. 2000;10(20):1273–1282. doi:10.1016/s0960-9822(00)00749-1

48. Bernstein BW, Bamburg JR. ADF/cofilin: a functional node in cell biology. Trends Cell Biol. 2010;20(4):187–195. doi:10.1016/j.tcb.2010.01.001

49. Rosenblatt J, Peluso P, Mitchison TJ. The bulk of unpolymerized actin in Xenopus egg extracts is ATP-bound. Mol Biol Cell. 1995;6(2):227–236. doi:10.1091/mbc.6.2.227

50. Sept D, McCammon JA. Thermodynamics and kinetics of actin filament nucleation. Biophys J. 2001;81(2):667–674. doi:10.1016/s0006-3495(01)75731-1

51. Fedorov AA, Lappalainen P, Fedorov EV, Drubin DG, Almo SC. Structure determination of yeast cofilin. Nat Struct Biol. 1997;4(5):366–369. doi:10.1038/nsb0597-366

52. Pope BJ, Gonsior SM, Yeoh S, McGough A, Weeds AG. Uncoupling actin filament fragmentation by cofilin from increased subunit turnover. J Mol Biol. 2000;298(4):649–661. doi:10.1006/jmbi.2000.3688

53. Andrianantoandro E, Pollard TD. Mechanism of actin filament turnover by severing and nucleation at different concentrations of ADF/cofilin. Mol Cell. 2006;24(1):13–23. doi:10.1016/j.molcel.2006.08.006

54. Bamburg JR, Bernstein BW. ADF/cofilin. Curr Biol. 2008;18(7):R273–5. doi:10.1016/j.cub.2008.02.002

55. Agnew BJ, Minamide LS, Bamburg JR. Reactivation of phosphorylated actin depolymerizing factor and identification of the regulatory site. J Biol Chem. 1995;270(29):17582–17587. doi:10.1074/jbc.270.29.17582

56. Kanellos G, Frame MC. Cellular functions of the ADF/cofilin family at a glance. J Cell Sci. 2016;129(17):3211–3218. doi:10.1242/jcs.187849

57. Arber S, Barbayannis FA, Hanser H, et al. Regulation of actin dynamics through phosphorylation of cofilin by LIM-kinase. Nature. 1998;393(6687):805–809. doi:10.1038/31729

58. Yang N, Higuchi O, Ohashi K, et al. Cofilin phosphorylation by LIM-kinase 1 and its role in Rac-mediated actin reorganization. Nature. 1998;393(6687):809–812. doi:10.1038/31735

59. Toshima J, Toshima JY, Takeuchi K, Mori R, Mizuno K. Cofilin phosphorylation and actin reorganization activities of testicular protein kinase 2 and its predominant expression in testicular Sertoli cells. J Biol Chem. 2001;276(33):31449–31458. doi:10.1074/jbc.M102988200

60. Scott RW, Olson MF. LIM kinases: function, regulation and association with human disease. J Mol Med (Berl). 2007;85(6):555–568. doi:10.1007/s00109-007-0165-6

61. Po’uha ST, Shum MS, Goebel A, Bernard O, Kavallaris M. LIM-kinase 2, a regulator of actin dynamics, is involved in mitotic spindle integrity and sensitivity to microtubule-destabilizing drugs. Oncogene. 2010;29(4):597–607. doi:10.1038/onc.2009.367

62. Hamill S, Lou HJ, Turk BE, Boggon TJ. Structural basis for noncanonical substrate recognition of Cofilin/ADF proteins by LIM kinases. Mol Cell. 2016;62(3):397–408. doi:10.1016/j.molcel.2016.04.001

63. Maekawa M, Ishizaki T, Boku S, et al. Signaling from Rho to the actin cytoskeleton through protein kinases ROCK and LIM-kinase. Science. 1999;285(5429):895–898. doi:10.1126/science.285.5429.895

64. Edwards DC, Sanders LC, Bokoch GM, Gill GN. Activation of LIM-kinase by Pak1 couples Rac/Cdc42 GTPase signalling to actin cytoskeletal dynamics. Nat Cell Biol. 1999;1(5):253–259. doi:10.1038/12963

65. Amano T, Tanabe K, Eto T, Narumiya S, Mizuno K. LIM-kinase 2 induces formation of stress fibres, focal adhesions and membrane blebs, dependent on its activation by Rho-associated kinase-catalysed phosphorylation at threonine-505. Biochem J. 2001;354(Pt 1):149–159. doi:10.1042/0264-6021:

66. Sumi T, Matsumoto K, Nakamura T. Specific activation of LIM kinase 2 via phosphorylation of threonine 505 by ROCK, a Rho-dependent protein kinase. J Biol Chem. 2001;276(1):670–676. doi:10.1074/jbc.M007074200

67. Ahmed T, Shea K, Masters JR, Jones GE, Wells CM. A PAK4-LIMK1 pathway drives prostate cancer cell migration downstream of HGF. Cell Signal. 2008;20(7):1320–1328. doi:10.1016/j.cellsig.2008.02.021

68. Sumi T, Matsumoto K, Shibuya A, Nakamura T. Activation of LIM kinases by myotonic dystrophy kinase-related Cdc42-binding kinase alpha. J Biol Chem. 2001;276(25):23092–23096. doi:10.1074/jbc.C100196200

69. 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(5):3577–3582. doi:10.1074/jbc.275.5.3577

70. Foletta VC, Lim MA, Soosairajah J, et al. Direct signaling by the BMP type II receptor via the cytoskeletal regulator LIMK1. J Cell Biol. 2003;162(6):1089–1098. doi:10.1083/jcb.200212060

71. Schlau M, Terheyden-Keighley D, Theis V, Mannherz HG, Theiss C. VEGF triggers the activation of cofilin and the Arp2/3 complex within the growth cone. Int J Mol Sci. 2018;19(2). doi:10.3390/ijms19020384

72. Song IS, Kang SS, Kim ES, et al. Heat shock protein 27 phosphorylation is involved in epithelial cell apoptosis as well as epithelial migration during corneal epithelial wound healing. Exp Eye Res. 2014;118:36–41. doi:10.1016/j.exer.2013.11.002

73. Pandey D, Goyal P, Siess W. Lysophosphatidic acid stimulation of platelets rapidly induces Ca2+-dependent dephosphorylation of cofilin that is independent of dense granule secretion and aggregation. Blood Cells Mol Dis. 2007;38(3):269–279. doi:10.1016/j.bcmd.2007.01.002

74. Reschke CR, Silva LFA, Vangoor VR, et al. Systemic delivery of antagomirs during blood-brain barrier disruption is disease-modifying in experimental epilepsy. Mol Ther. 2021;29(6):2041–2052. doi:10.1016/j.ymthe.2021.02.021

75. Castañeda P, Muñoz M, García-Rojo G, et al. Association of N-cadherin levels and downstream effectors of Rho GTPases with dendritic spine loss induced by chronic stress in rat hippocampal neurons. J Neurosci Res. 2015;93(10):1476–1491. doi:10.1002/jnr.23602

76. Toshima J, Toshima JY, Amano T, Yang N, Narumiya S, Mizuno K. Cofilin phosphorylation by protein kinase testicular protein kinase 1 and its role in integrin-mediated actin reorganization and focal adhesion formation. Mol Biol Cell. 2001;12(4):1131–1145. doi:10.1091/mbc.12.4.1131

77. Huang TY, DerMardirossian C, Bokoch GM. Cofilin phosphatases and regulation of actin dynamics. Curr Opin Cell Biol. 2006;18(1):26–31. doi:10.1016/j.ceb.2005.11.005

78. Niwa R, Nagata-Ohashi K, Takeichi M, Mizuno K, Uemura T. Control of actin reorganization by Slingshot, a family of phosphatases that dephosphorylate ADF/cofilin. Cell. 2002;108(2):233–246. doi:10.1016/s0092-8674(01)00638-9

79. Ohta Y, Kousaka K, Nagata-Ohashi K, et al. Differential activities, subcellular distribution and tissue expression patterns of three members of Slingshot family phosphatases that dephosphorylate cofilin. Genes Cells. 2003;8(10):811–824. doi:10.1046/j.1365-2443.2003.00678.x

80. Soosairajah J, Maiti S, Wiggan O, et al. Interplay between components of a novel LIM kinase-slingshot phosphatase complex regulates cofilin. EMBO j. 2005;24(3):473–486. doi:10.1038/sj.emboj.7600543

81. Kurita S, Gunji E, Ohashi K, Mizuno K. Actin filaments-stabilizing and -bundling activities of cofilin-phosphatase Slingshot-1. Genes Cells. 2007;12(5):663–676. doi:10.1111/j.1365-2443.2007.01078.x

82. Yamamoto M, Nagata-Ohashi K, Ohta Y, Ohashi K, Mizuno K. Identification of multiple actin-binding sites in cofilin-phosphatase Slingshot-1L. FEBS Lett. 2006;580(7):1789–1794. doi:10.1016/j.febslet.2006.02.034

83. Kurita S, Watanabe Y, Gunji E, Ohashi K, Mizuno K. Molecular dissection of the mechanisms of substrate recognition and F-actin-mediated activation of cofilin-phosphatase Slingshot-1. J Biol Chem. 2008;283(47):32542–32552. doi:10.1074/jbc.M804627200

84. Mizuno K. Signaling mechanisms and functional roles of cofilin phosphorylation and dephosphorylation. Cell Signal. 2013;25(2):457–469. doi:10.1016/j.cellsig.2012.11.001

85. Ichetovkin I, Grant W, Condeelis J. Cofilin produces newly polymerized actin filaments that are preferred for dendritic nucleation by the Arp2/3 complex. Curr Biol. 2002;12(1):79–84. doi:10.1016/s0960-9822(01)00629-7

86. Orlova A, Shvetsov A, Galkin VE, et al. Actin-destabilizing factors disrupt filaments by means of a time reversal of polymerization. Proc Natl Acad Sci U S A. 2004;101(51):17664–17668. doi:10.1073/pnas.0407525102

87. Bobkov AA, Muhlrad A, Pavlov DA, Kokabi K, Yilmaz A, Reisler E. Cooperative effects of cofilin (ADF) on actin structure suggest allosteric mechanism of cofilin function. J Mol Biol. 2006;356(2):325–334. doi:10.1016/j.jmb.2005.11.072

88. Chan C, Beltzner CC, Pollard TD. Cofilin dissociates Arp2/3 complex and branches from actin filaments. Curr Biol. 2009;19(7):537–545. doi:10.1016/j.cub.2009.02.060

89. Minamide LS, Striegl AM, Boyle JA, Meberg PJ, Bamburg JR. Neurodegenerative stimuli induce persistent ADF/cofilin-actin rods that disrupt distal neurite function. Nat Cell Biol. 2000;2(9):628–636. doi:10.1038/35023579

90. Park JB, Agnihotri S, Golbourn B, et al. Transcriptional profiling of GBM invasion genes identifies effective inhibitors of the LIM kinase-Cofilin pathway. Oncotarget. 2014;5(19):9382–9395. doi:10.18632/oncotarget.2412

91. Liu W, Zhang Q, Tang Q, et al. Lycorine inhibits cell proliferation and migration by inhibiting ROCK1/cofilin‑induced actin dynamics in HepG2 hepatoblastoma cells. Oncol Rep. 2018;40(4):2298–2306. doi:10.3892/or.2018.6609

92. Wei R, Zhang Y, Shen L, et al. Comparative proteomic and radiobiological analyses in human lung adenocarcinoma cells. Mol Cell Biochem. 2012;359(1–2):151–159. doi:10.1007/s11010-011-1008-x

93. Collazo J, Zhu B, Larkin S, et al. Cofilin drives cell-invasive and metastatic responses to TGF-β in prostate cancer. Cancer Res. 2014;74(8):2362–2373. doi:10.1158/0008-5472.Can-13-3058

94. Maimaiti Y, Tan J, Liu Z, et al. Overexpression of cofilin correlates with poor survival in breast cancer: a tissue microarray analysis. Oncol Lett. 2017;14(2):2288–2294. doi:10.3892/ol.2017.6413

95. Lu LI, Fu NI, Luo XU, Li XY, Li XP. Overexpression of cofilin 1 in prostate cancer and the corresponding clinical implications. Oncol Lett. 2015;9(6):2757–2761. doi:10.3892/ol.2015.3133

96. Zhang Y, Tong X. Expression of the actin-binding proteins indicates that cofilin and fascin are related to breast tumour size. J Int Med Res. 2010;38(3):1042–1048. doi:10.1177/147323001003800331

97. Wu Q, Jiang Y, Cui S, Wang Y, Wu X. The role of cofilin-l in vulvar squamous cell carcinoma: a marker of carcinogenesis, progression and targeted therapy. Oncol Rep. 2016;35(5):2743–2754. doi:10.3892/or.2016.4625

98. Satoh M, Takano S, Sogawa K, et al. Immune-complex level of cofilin-1 in sera is associated with cancer progression and poor prognosis in pancreatic cancer. Cancer Sci. 2017;108(4):795–803. doi:10.1111/cas.13181

99. Sun W, Yan H, Qian C, et al. Cofilin-1 and phosphoglycerate kinase 1 as promising indicators for glioma radiosensibility and prognosis. Oncotarget. 2017;8(33):55073–55083. doi:10.18632/oncotarget.19025

100. Xu H, Chen Y, Tan C, et al. High expression of WDR1 in primary glioblastoma is associated with poor prognosis. Am J Transl Res. 2016;8(2):1253–1264.

101. Erkutlu I, Cigiloglu A, Kalender ME, et al. Correlation between Rho-kinase pathway gene expressions and development and progression of glioblastoma multiforme. Tumour Biol. 2013;34(2):1139–1144. doi:10.1007/s13277-013-0655-9

102. Tang Z, Araysi LM, Fathallah-Shaykh HM. c-Src and neural Wiskott-Aldrich syndrome protein (N-WASP) promote low oxygen-induced accelerated brain invasion by gliomas. PLoS One. 2013;8(9):e75436. doi:10.1371/journal.pone.0075436

103. Yan H, Yang K, Xiao H, Zou YJ, Zhang WB, Liu HY. Over-expression of cofilin-1 and phosphoglycerate kinase 1 in astrocytomas involved in pathogenesis of radioresistance. CNS Neurosci Ther. 2012;18(9):729–736. doi:10.1111/j.1755-5949.2012.00353.x

104. Yang ZL, Miao X, Xiong L, et al. CFL1 and Arp3 are biomarkers for metastasis and poor prognosis of squamous cell/adenosquamous carcinomas and adenocarcinomas of gallbladder. Cancer Invest. 2013;31(2):132–139. doi:10.3109/07357907.2012.756113

105. Tsai CH, Lin LT, Wang CY, et al. Over-expression of cofilin-1 suppressed growth and invasion of cancer cells is associated with up-regulation of let-7 microRNA. Biochim Biophys Acta. 2015;1852(5):851–861. doi:10.1016/j.bbadis.2015.01.007

106. Chung H, Kim B, Jung SH, et al. Does phosphorylation of cofilin affect the progression of human bladder cancer? BMC Cancer. 2013;13(1):45. doi:10.1186/1471-2407-13-45

107. Zheng Y, Fang Y, Li S, Zheng B. 血浆cofilin蛋白检测在肺癌诊断中的价值. 南方医科大学学报 [Detection of plasma cofilin protein for diagnosis of lung cancer]. Nan Fang Yi Ke Da Xue Xue Bao. 2013;33(10):1551–1553. Chinese

108. Becker M, De Bastiani MA, Müller CB, Markoski MM, Castro MA, Klamt F. High cofilin-1 levels correlate with cisplatin resistance in lung adenocarcinomas. Tumour Biol. 2014;35(2):1233–1238. doi:10.1007/s13277-013-1164-6

109. Ding SJ, Li Y, Shao XX, et al. Proteome analysis of hepatocellular carcinoma cell strains, MHCC97-H and MHCC97-L, with different metastasis potentials. Proteomics. 2004;4(4):982–994. doi:10.1002/pmic.200300653

110. Yamaguchi H, Condeelis J. Regulation of the actin cytoskeleton in cancer cell migration and invasion. Biochim Biophys Acta. 2007;1773(5):642–652. doi:10.1016/j.bbamcr.2006.07.001

111. Yamaguchi H, Lorenz M, Kempiak S, et al. Molecular mechanisms of invadopodium formation: the role of the N-WASP-Arp2/3 complex pathway and cofilin. J Cell Biol. 2005;168(3):441–452. doi:10.1083/jcb.200407076

112. Yap CT, Simpson TI, Pratt T, Price DJ, Maciver SK. The motility of glioblastoma tumour cells is modulated by intracellular cofilin expression in a concentration-dependent manner. Cell Motil Cytoskeleton. 2005;60(3):153–165. doi:10.1002/cm.20053

113. Bisi S, Disanza A, Malinverno C, Frittoli E, Palamidessi A, Scita G. Membrane and actin dynamics interplay at lamellipodia leading edge. Curr Opin Cell Biol. 2013;25(5):565–573. doi:10.1016/j.ceb.2013.04.001

114. Ideses Y, Brill-Karniely Y, Haviv L, Ben-Shaul A, Bernheim-Groswasser A. Arp2/3 branched actin network mediates filopodia-like bundles formation in vitro. PLoS One. 2008;3(9):e3297. doi:10.1371/journal.pone.0003297

115. Kiuchi T, Nagai T, Ohashi K, Mizuno K. Measurements of spatiotemporal changes in G-actin concentration reveal its effect on stimulus-induced actin assembly and lamellipodium extension. J Cell Biol. 2011;193(2):365–380. doi:10.1083/jcb.201101035

116. Shishkin S, Eremina L, Pashintseva N, Kovalev L, Kovaleva M. Cofilin-1 and other ADF/cofilin superfamily members in human malignant cells. Int J Mol Sci. 2016;18(1):10. doi:10.3390/ijms18010010

117. Ghosh M, Ichetovkin I, Song X, Condeelis JS, Lawrence DS. A new strategy for caging proteins regulated by kinases. J Am Chem Soc. 2002;124(11):2440–2441. doi:10.1021/ja017592l

118. Delorme V, Machacek M, DerMardirossian C, et al. Cofilin activity downstream of Pak1 regulates cell protrusion efficiency by organizing lamellipodium and lamella actin networks. Dev Cell. 2007;13(5):646–662. doi:10.1016/j.devcel.2007.08.011

119. Hoffmann L, Rust MB, Culmsee C. Actin(g) on mitochondria - A role for cofilin1 in neuronal cell death pathways. Biol Chem. 2019;400(9):1089–1097. doi:10.1515/hsz-2019-0120

120. Mardilovich K, Gabrielsen M, McGarry L, et al. Elevated LIM kinase 1 in nonmetastatic prostate cancer reflects its role in facilitating androgen receptor nuclear translocation. Mol Cancer Ther. 2015;14(1):246–258. doi:10.1158/1535-7163.Mct-14-0447

121. Yoshioka K, Foletta V, Bernard O, Itoh K. A role for LIM kinase in cancer invasion. Proc Natl Acad Sci U S A. 2003;100(12):7247–7252. doi:10.1073/pnas.1232344100

122. Manetti F. Recent findings confirm LIM domain kinases as emerging target candidates for cancer therapy. Curr Cancer Drug Targets. 2012;12(5):543–560. doi:10.2174/156800912800673266

123. Zhang W, Gan N, Zhou J. Immunohistochemical investigation of the correlation between LIM kinase 1 expression and development and progression of human ovarian carcinoma. J Int Med Res. 2012;40(3):1067–1073. doi:10.1177/147323001204000325

124. Guo H, Gu F, Li W, et al. Reduction of protein kinase C zeta inhibits migration and invasion of human glioblastoma cells. J Neurochem. 2009;109(1):203–213. doi:10.1111/j.1471-4159.2009.05946.x

125. Ma Y, Wang B, Li W, et al. Intersectin1-s is involved in migration and invasion of human glioma cells. J Neurosci Res. 2011;89(7):1079–1090. doi:10.1002/jnr.22616

126. Scott RW, Hooper S, Crighton D, et al. LIM kinases are required for invasive path generation by tumor and tumor-associated stromal cells. J Cell Biol. 2010;191(1):169–185. doi:10.1083/jcb.201002041

127. Choi JA, Lim IK. TIS21/BTG2 inhibits invadopodia formation by downregulating reactive oxygen species level in MDA-MB-231 cells. J Cancer Res Clin Oncol. 2013;139(10):1657–1665. doi:10.1007/s00432-013-1484-3

128. Aung A, Seo YN, Lu S, et al. 3D traction stresses activate protease-dependent invasion of cancer cells. Biophys J. 2014;107(11):2528–2537. doi:10.1016/j.bpj.2014.07.078

129. Kłopocka W, Korczyński J, Pomorski P. Cytoskeleton and nucleotide signaling in glioma C6 cells. Adv Exp Med Biol. 2020;1202:109–128. doi:10.1007/978-3-030-30651-9_6

130. Jin SG, Ryu HH, Li SY, et al. Nogo-A inhibits the migration and invasion of human malignant glioma U87MG cells. Oncol Rep. 2016;35(6):3395–3402. doi:10.3892/or.2016.4737

131. Schiapparelli P, Guerrero-Cazares H, Magaña-Maldonado R, et al. NKCC1 regulates migration ability of glioblastoma cells by modulation of actin dynamics and interacting with cofilin. EBioMedicine. 2017;21:94–103. doi:10.1016/j.ebiom.2017.06.020

132. Zhao K, Wang L, Li T, et al. The role of miR-451 in the switching between proliferation and migration in malignant glioma cells: AMPK signaling, mTOR modulation and Rac1 activation required. Int J Oncol. 2017;50(6):1989–1999. doi:10.3892/ijo.2017.3973

133. Glogowska A, Thanasupawat T, Beiko J, Pitz M, Hombach-Klonisch S, Klonisch T. Novel CTRP8-RXFP1-JAK3-STAT3 axis promotes Cdc42-dependent actin remodeling for enhanced filopodia formation and motility in human glioblastoma cells. Mol Oncol. 2022;16(2):368–387. doi:10.1002/1878-0261.12981

134. Shi C, Ren L, Sun C, et al. miR-29a/b/c function as invasion suppressors for gliomas by targeting CDC42 and predict the prognosis of patients. Br J Cancer. 2017;117(7):1036–1047. doi:10.1038/bjc.2017.255

135. Hou M, Liu X, Cao J, Chen B. SEPT7 overexpression inhibits glioma cell migration by targeting the actin cytoskeleton pathway. Oncol Rep. 2016;35(4):2003–2010. doi:10.3892/or.2016.4609

136. Zhang B, Gu F, She C, et al. Reduction of Akt2 inhibits migration and invasion of glioma cells. Int J Cancer. 2009;125(3):585–595. doi:10.1002/ijc.24314

137. Schulze M, Fedorchenko O, Zink TG, et al. Chronophin is a glial tumor modifier involved in the regulation of glioblastoma growth and invasiveness. Oncogene. 2016;35(24):3163–3177. doi:10.1038/onc.2015.376

138. Schulze M, Hutterer M, Sabo A, et al. Chronophin regulates active vitamin B6 levels and transcriptomic features of glioblastoma cell lines cultured under non-adherent, serum-free conditions. BMC Cancer. 2018;18(1):524. doi:10.1186/s12885-018-4440-4

139. Gondi CS, Kandhukuri N, Kondraganti S, et al. Down-regulation of uPAR and cathepsin B retards cofilin dephosphorylation. Int J Oncol. 2006;28(3):633–639.

140. Piña-Medina AG, Hansberg-Pastor V, González-Arenas A, Cerbón M, Camacho-Arroyo I. Progesterone promotes cell migration, invasion and cofilin activation in human astrocytoma cells. Steroids. 2016;105:19–25. doi:10.1016/j.steroids.2015.11.008

141. Sun J, He D, Fu Y, et al. A novel lncRNA ARST represses glioma progression by inhibiting ALDOA-mediated actin cytoskeleton integrity. J Exp Clin Cancer Res. 2021;40(1):187. doi:10.1186/s13046-021-01977-9

142. Peterburs P, Heering J, Link G, Pfizenmaier K, Olayioye MA, Hausser A. Protein kinase D regulates cell migration by direct phosphorylation of the cofilin phosphatase slingshot 1 like. Cancer Res. 2009;69(14):5634–5638. doi:10.1158/0008-5472.Can-09-0718

143. Zhuang H, Li Q, Zhang X, et al. Downregulation of glycine decarboxylase enhanced cofilin-mediated migration in hepatocellular carcinoma cells. Free Radic Biol Med. 2018;120:1–12. doi:10.1016/j.freeradbiomed.2018.03.003

144. Takahashi K, Okabe H, Kanno SI, Nagai T, Mizuno K. A pleckstrin homology-like domain is critical for F-actin binding and cofilin-phosphatase activity of Slingshot-1. Biochem Biophys Res Commun. 2017;482(4):686–692. doi:10.1016/j.bbrc.2016.11.095

145. Hayakawa K, Tatsumi H, Sokabe M. Actin filaments function as a tension sensor by tension-dependent binding of cofilin to the filament. J Cell Biol. 2011;195(5):721–727. doi:10.1083/jcb.201102039

146. Matsushita S, Inoue Y, Hojo M, Sokabe M, Adachi T. Effect of tensile force on the mechanical behavior of actin filaments. J Biomech. 2011;44(9):1776–1781. doi:10.1016/j.jbiomech.2011.04.012

147. Pastushenko I, Blanpain C. EMT transition states during tumor progression and metastasis. Trends Cell Biol. 2019;29(3):212–226. doi:10.1016/j.tcb.2018.12.001

148. Kalluri R, Weinberg RA. The basics of epithelial-mesenchymal transition. J Clin Invest. 2009;119(6):1420–1428. doi:10.1172/jci39104

149. Lamouille S, Xu J, Derynck R. Molecular mechanisms of epithelial-mesenchymal transition. Nat Rev Mol Cell Biol. 2014;15(3):178–196. doi:10.1038/nrm3758

150. Aiello NM, Maddipati R, Norgard RJ, et al. EMT subtype influences epithelial plasticity and mode of cell migration. Dev Cell. 2018;45(6):681–95.e4. doi:10.1016/j.devcel.2018.05.027

151. Cervantes-Arias A, Pang LY, Argyle DJ. Epithelial-mesenchymal transition as a fundamental mechanism underlying the cancer phenotype. Vet Comp Oncol. 2013;11(3):169–184. doi:10.1111/j.1476-5829.2011.00313.x

152. Kahlert UD, Nikkhah G, Maciaczyk J. Epithelial-to-mesenchymal(-like) transition as a relevant molecular event in malignant gliomas. Cancer Lett. 2013;331(2):131–138. doi:10.1016/j.canlet.2012.12.010

153. Tso CL, Shintaku P, Chen J, et al. Primary glioblastomas express mesenchymal stem-like properties. Mol Cancer Res. 2006;4(9):607–619. doi:10.1158/1541-7786.Mcr-06-0005

154. Iser IC, Pereira MB, Lenz G, Wink MR. The epithelial-to-mesenchymal transition-like process in glioblastoma: an updated systematic review and in silico investigation. Med Res Rev. 2017;37(2):271–313. doi:10.1002/med.21408

155. Lewis-Tuffin LJ, Rodriguez F, Giannini C, et al. Misregulated E-cadherin expression associated with an aggressive brain tumor phenotype. PLoS One. 2010;5(10):e13665. doi:10.1371/journal.pone.0013665

156. Mikheeva SA, Mikheev AM, Petit A, et al. TWIST1 promotes invasion through mesenchymal change in human glioblastoma. Mol Cancer. 2010;9:194. doi:10.1186/1476-4598-9-194

157. Yadav AK, Desai NS. Cancer stem cells: acquisition, characteristics, therapeutic implications, targeting strategies and future prospects. Stem Cell Rev Rep. 2019;15(3):331–355. doi:10.1007/s12015-019-09887-2

158. Lau EY, Ho NP, Lee TK. Cancer stem cells and their microenvironment: biology and therapeutic implications. Stem Cells Int. 2017;2017:3714190. doi:10.1155/2017/3714190

159. Mani SA, Guo W, Liao MJ, et al. The epithelial-mesenchymal transition generates cells with properties of stem cells. Cell. 2008;133(4):704–715. doi:10.1016/j.cell.2008.03.027

160. Gupta PB, Onder TT, Jiang G, et al. Identification of selective inhibitors of cancer stem cells by high-throughput screening. Cell. 2009;138(4):645–659. doi:10.1016/j.cell.2009.06.034

161. Lee KH, Ahn EJ, Oh SJ, et al. KITENIN promotes glioma invasiveness and progression, associated with the induction of EMT and stemness markers. Oncotarget. 2015;6(5):3240–3253. doi:10.18632/oncotarget.3087

162. Zhang XH, Qian Y, Li Z, Zhang NN, Xie YJ. Let-7g-5p inhibits epithelial-mesenchymal transition consistent with reduction of glioma stem cell phenotypes by targeting VSIG4 in glioblastoma. Oncol Rep. 2016;36(5):2967–2975. doi:10.3892/or.2016.5098

163. Zhu T, Li X, Luo L, et al. Reversion of malignant phenotypes of human glioblastoma cells by β-elemene through β-catenin-mediated regulation of stemness-, differentiation- and epithelial-to-mesenchymal transition-related molecules. J Transl Med. 2015;13(1):356. doi:10.1186/s12967-015-0727-2

164. Tania M, Khan MA, Fu J. Epithelial to mesenchymal transition inducing transcription factors and metastatic cancer. Tumour Biol. 2014;35(8):7335–7342. doi:10.1007/s13277-014-2163-y

165. Derynck R, Muthusamy BP, Saeteurn KY. Signaling pathway cooperation in TGF-β-induced epithelial-mesenchymal transition. Curr Opin Cell Biol. 2014;31:56–66. doi:10.1016/j.ceb.2014.09.001

166. Xu Z, Gu C, Yao X, et al. CD73 promotes tumor metastasis by modulating RICS/RhoA signaling and EMT in gastric cancer. Cell Death Dis. 2020;11(3):202. doi:10.1038/s41419-020-2403-6

167. Islam SU, Ahmed MB, Lee SJ, et al. PRP4 kinase induces actin rearrangement and epithelial-mesenchymal transition through modulation of the actin-binding protein cofilin. Exp Cell Res. 2018;369(1):158–165. doi:10.1016/j.yexcr.2018.05.018

168. Häcker G. The morphology of apoptosis. Cell Tissue Res. 2000;301(1):5–17. doi:10.1007/s004410000193

169. Nijhawan D, Honarpour N, Wang X. Apoptosis in neural development and disease. Annu Rev Neurosci. 2000;23(1):73–87. doi:10.1146/annurev.neuro.23.1.73

170. Pistritto G, Trisciuoglio D, Ceci C, Garufi A, D’Orazi G. Apoptosis as anticancer mechanism: function and dysfunction of its modulators and targeted therapeutic strategies. Aging. 2016;8(4):603–619. doi:10.18632/aging.100934

171. Igney FH, Krammer PH. Death and anti-death: tumour resistance to apoptosis. Nat Rev Cancer. 2002;2(4):277–288. doi:10.1038/nrc776

172. Nezhat F, Cohen C, Rahaman J, Gretz H, Cole P, Kalir T. Comparative immunohistochemical studies of bcl-2 and p53 proteins in benign and malignant ovarian endometriotic cysts. Cancer. 2002;94(11):2935–2940. doi:10.1002/cncr.10566

173. Nasu K, Nishida M, Kawano Y, et al. Aberrant expression of apoptosis-related molecules in endometriosis: a possible mechanism underlying the pathogenesis of endometriosis. Reprod Sci. 2011;18(3):206–218. doi:10.1177/1933719110392059

174. Chua BT, Volbracht C, Tan KO, Li R, Yu VC, Li P. Mitochondrial translocation of cofilin is an early step in apoptosis induction. Nat Cell Biol. 2003;5(12):1083–1089. doi:10.1038/ncb1070

175. Hatch AL, Gurel PS, Higgs HN. Novel roles for actin in mitochondrial fission. J Cell Sci. 2014;127(Pt 21):4549–4560. doi:10.1242/jcs.153791

176. Hu C, Huang Y, Li L. Drp1-Dependent Mitochondrial Fission Plays Critical Roles in Physiological and Pathological Progresses in Mammals. Int J Mol Sci. 2017;18(1):144. doi:10.3390/ijms18010144

177. Guo L, Cui C, Wang J, et al. PINCH-1 regulates mitochondrial dynamics to promote proline synthesis and tumor growth. Nat Commun. 2020;11(1):4913. doi:10.1038/s41467-020-18753-6

178. Taguchi N, Ishihara N, Jofuku A, Oka T, Mihara K. Mitotic phosphorylation of dynamin-related GTPase Drp1 participates in mitochondrial fission. J Biol Chem. 2007;282(15):11521–11529. doi:10.1074/jbc.M607279200

179. Hu J, Zhang H, Li J, et al. ROCK1 activation-mediated mitochondrial translocation of Drp1 and cofilin are required for arnidiol-induced mitochondrial fission and apoptosis. J Exp Clin Cancer Res. 2020;39(1):37. doi:10.1186/s13046-020-01545-7

180. Zhang L, Cao H, Tao H, Yang J, Gong W, Hu Q. Effect of the interference with DRP1 expression on the biological characteristics of glioma stem cells. Exp Ther Med. 2021;22(1):696. doi:10.3892/etm.2021.10128

181. Otera H, Miyata N, Kuge O, Mihara K. Drp1-dependent mitochondrial fission via MiD49/51 is essential for apoptotic cristae remodeling. J Cell Biol. 2016;212(5):531–544. doi:10.1083/jcb.201508099

182. Stupp R, Hegi ME, Mason WP, et al. Effects of radiotherapy with concomitant and adjuvant temozolomide versus radiotherapy alone on survival in glioblastoma in a randomised Phase III study: 5-year analysis of the EORTC-NCIC trial. Lancet Oncol. 2009;10(5):459–466. doi:10.1016/s1470-2045(09)70025-7

183. Stupp R, Mason WP, van den Bent MJ, et al. Radiotherapy plus concomitant and adjuvant temozolomide for glioblastoma. N Engl J Med. 2005;352(10):987–996. doi:10.1056/NEJMoa043330

184. Du HQ, Chen L, Wang Y, et al. Increasing radiosensitivity with the downregulation of cofilin-1 in U251 human glioma cells. Mol Med Rep. 2015;11(5):3354–3360. doi:10.3892/mmr.2014.3125

185. Zhou T, Wang CH, Yan H, et al. Inhibition of the Rac1-WAVE2-Arp2/3 signaling pathway promotes radiosensitivity via downregulation of cofilin-1 in U251 human glioma cells. Mol Med Rep. 2016;13(5):4414–4420. doi:10.3892/mmr.2016.5088

186. Abbott NJ, Rönnbäck L, Hansson E. Astrocyte-endothelial interactions at the blood-brain barrier. Nat Rev Neurosci. 2006;7(1):41–53. doi:10.1038/nrn1824

187. Arvanitis CD, Ferraro GB, Jain RK. The blood-brain barrier and blood-tumour barrier in brain tumours and metastases. Nat Rev Cancer. 2020;20(1):26–41. doi:10.1038/s41568-019-0205-x

188. Idbaih A, Ducray F, Sierra Del Rio M, Hoang-Xuan K, Delattre JY. Therapeutic application of noncytotoxic molecular targeted therapy in gliomas: growth factor receptors and angiogenesis inhibitors. Oncologist. 2008;13(9):978–992. doi:10.1634/theoncologist.2008-0056

189. Xie H, Xue YX, Liu LB, Liu YH, Wang P. Role of RhoA/ROCK signaling in endothelial-monocyte-activating polypeptide II opening of the blood-tumor barrier: role of RhoA/ROCK signaling in EMAP II opening of the BTB. J Mol Neurosci. 2012;46(3):666–676. doi:10.1007/s12031-011-9564-9

190. Li Z, Liu YH, Xue YX, Liu LB, Wang P. Signal mechanisms underlying low-dose endothelial monocyte-activating polypeptide-II-induced opening of the blood-tumor barrier. J Mol Neurosci. 2012;48(1):291–301. doi:10.1007/s12031-012-9776-7

191. Li Z, Liu XB, Liu YH, et al. Low-dose endothelial monocyte-activating polypeptide-II increases blood-tumor barrier permeability by activating the RhoA/ROCK/PI3K signaling pathway. J Mol Neurosci. 2016;59(2):193–202. doi:10.1007/s12031-015-0668-5

192. Kim JW, Gao P, Liu YC, Semenza GL, Dang CV. Hypoxia-inducible factor 1 and dysregulated c-Myc cooperatively induce vascular endothelial growth factor and metabolic switches hexokinase 2 and pyruvate dehydrogenase kinase 1. Mol Cell Biol. 2007;27(21):7381–7393. doi:10.1128/mcb.00440-07

193. Wolf A, Agnihotri S, Micallef J, et al. Hexokinase 2 is a key mediator of aerobic glycolysis and promotes tumor growth in human glioblastoma multiforme. J Exp Med. 2011;208(2):313–326. doi:10.1084/jem.20101470

194. Ghosh S, Paul A, Sen E. Tumor necrosis factor α-induced hypoxia-inducible factor 1α-β-catenin axis regulates major histocompatibility complex class I gene activation through chromatin remodeling. Mol Cell Biol. 2013;33(14):2718–2731. doi:10.1128/mcb.01254-12

195. Ghosh S, Gupta P, Sen E. TNFα driven HIF-1α-hexokinase II axis regulates MHC-I cluster stability through actin cytoskeleton. Exp Cell Res. 2016;340(1):116–124. doi:10.1016/j.yexcr.2015.11.016

196. Lewinska A, Klukowska-Rötzler J, Deregowska A, Adamczyk-Grochala J, Wnuk M. c-Myc activation promotes cofilin-mediated F-actin cytoskeleton remodeling and telomere homeostasis as a response to oxidant-based DNA damage in medulloblastoma cells. Redox Biol. 2019;24:101163. doi:10.1016/j.redox.2019.101163

197. Obrdlik A, The PP. F-actin severing protein cofilin-1 is required for RNA polymerase II transcription elongation. Nucleus. 2011;2(1):72–79. doi:10.4161/nucl.2.1.14508

198. Müller CB, de Barros RL, Castro MA, et al. Validation of cofilin-1 as a biomarker in non-small cell lung cancer: application of quantitative method in a retrospective cohort. J Cancer Res Clin Oncol. 2011;137(9):1309–1316. doi:10.1007/s00432-011-1001-5

199. Rak R, Haklai R, Elad-Tzfadia G, Wolfson HJ, Carmeli S, Novel KY. LIMK2 Inhibitor blocks Panc-1 tumor growth in a mouse xenograft model. Oncoscience. 2014;1(1):39–48. doi:10.18632/oncoscience.7

200. Wang X, Zou S, Ren T, et al. Alantolactone suppresses the metastatic phenotype and induces the apoptosis of glioblastoma cells by targeting LIMK kinase activity and activating the cofilin/G‑actin signaling cascade. Int J Mol Med. 2021;47(5). doi:10.3892/ijmm.2021.4901

201. Porcù E, Maule F, Boso D, et al. BMP9 counteracts the tumorigenic and pro-angiogenic potential of glioblastoma. Cell Death Differ. 2018;25(10):1808–1822. doi:10.1038/s41418-018-0149-9

202. Hua D, Zhao Q, Yu Y, et al. Eucalyptal A inhibits glioma by rectifying oncogenic splicing of MYO1B mRNA via suppressing SRSF1 expression. Eur J Pharmacol. 2021;890:173669. doi:10.1016/j.ejphar.2020.173669

203. Xing JS, Wang X, Lan YL, et al. Isoalantolactone inhibits IKKβ kinase activity to interrupt the NF-κB/COX-2-mediated signaling cascade and induces apoptosis regulated by the mitochondrial translocation of cofilin in glioblastoma. Cancer Med. 2019;8(4):1655–1670. doi:10.1002/cam4.2013

204. Tang Q, Ji Q, Tang Y, et al. Mitochondrial translocation of cofilin-1 promotes apoptosis of gastric cancer BGC-823 cells induced by ursolic acid. Tumour Biol. 2014;35(3):2451–2459. doi:10.1007/s13277-013-1325-7

205. Li R, Wang X, Zhang XH, Chen HH, Liu YD. Ursolic acid promotes apoptosis of SGC-7901 gastric cancer cells through ROCK/PTEN mediated mitochondrial translocation of cofilin-1. Asian Pac J Cancer Prev. 2014;15(22):9593–9597. doi:10.7314/apjcp.2014.15.22.9593

206. Prełowska M, Kaczyńska A, Herman-Antosiewicz A. 4-(Methylthio)butyl isothiocyanate inhibits the proliferation of breast cancer cells with different receptor status. Pharmacol Rep. 2017;69(5):1059–1066. doi:10.1016/j.pharep.2017.04.014

207. Wang Q, Gao S, Wu GZ, et al. Total sesquiterpene lactones isolated from Inula helenium L. attenuates 2,4-dinitrochlorobenzene-induced atopic dermatitis-like skin lesions in mice. Phytomedicine. 2018;46:78–84. doi:10.1016/j.phymed.2018.04.036

208. Wang J, Cui L, Feng L, et al. Isoalantolactone inhibits the migration and invasion of human breast cancer MDA-MB-231 cells via suppression of the p38 MAPK/NF-κB signaling pathway. Oncol Rep. 2016;36(3):1269–1276. doi:10.3892/or.2016.4954

209. Li Z, Qin B, Qi X, Mao J, Wu D. Isoalantolactone induces apoptosis in human breast cancer cells via ROS-mediated mitochondrial pathway and downregulation of SIRT1. Arch Pharm Res. 2016;39(10):1441–1453. doi:10.1007/s12272-016-0815-8

210. Fan Y, Weng Z, Gao H, et al. Isoalantolactone enhances the radiosensitivity of UMSCC-10A Cells via specific inhibition of Erk1/2 phosphorylation. PLoS One. 2015;10(12):e0145790. doi:10.1371/journal.pone.0145790

211. Li GB, Cheng Q, Liu L, et al. Mitochondrial translocation of cofilin is required for allyl isothiocyanate-mediated cell death via ROCK1/PTEN/PI3K signaling pathway. Cell Commun Signal. 2013;11:50. doi:10.1186/1478-811x-11-50

212. Chen J, Ananthanarayanan B, Springer KS, et al. Suppression of LIM kinase 1 and LIM kinase 2 limits glioblastoma invasion. Cancer Res. 2020;80(1):69–78. doi:10.1158/0008-5472.Can-19-1237

213. Zebda N, Bernard O, Bailly M, Welti S, Lawrence DS, Condeelis JS. Phosphorylation of ADF/cofilin abolishes EGF-induced actin nucleation at the leading edge and subsequent lamellipod extension. J Cell Biol. 2000;151(5):1119–1128. doi:10.1083/jcb.151.5.1119

214. Wang W, Mouneimne G, Sidani M, et al. The activity status of cofilin is directly related to invasion, intravasation, and metastasis of mammary tumors. J Cell Biol. 2006;173(3):395–404. doi:10.1083/jcb.200510115

Creative Commons License © 2022 The Author(s). This work is published and licensed by Dove Medical Press Limited. The full terms of this license are available at https://www.dovepress.com/terms.php and incorporate the Creative Commons Attribution - Non Commercial (unported, v3.0) License. By accessing the work you hereby accept the Terms. Non-commercial uses of the work are permitted without any further permission from Dove Medical Press Limited, provided the work is properly attributed. For permission for commercial use of this work, please see paragraphs 4.2 and 5 of our Terms.