Next Article in Journal
Photodynamic Therapy Combined with Ferroptosis Is a Synergistic Antitumor Therapy Strategy
Previous Article in Journal
CD44 Intracellular Domain: A Long Tale of a Short Tail
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

PDZ and LIM Domain-Encoding Genes: Their Role in Cancer Development

Department of Respiratory and Critical Care Medicine, The Fourth Affiliated Hospital, Zhejiang University School of Medicine, Yiwu 322000, China
*
Authors to whom correspondence should be addressed.
These authors have contributed equally to this work.
Cancers 2023, 15(20), 5042; https://doi.org/10.3390/cancers15205042
Submission received: 10 September 2023 / Revised: 13 October 2023 / Accepted: 15 October 2023 / Published: 19 October 2023
(This article belongs to the Section Cancer Pathophysiology)

Abstract

:

Simple Summary

There are three subfamilies of human PDZ-LIM family proteins with a total of 10 protein molecules, and PDZ-LIM family proteins serve as a class of scaffolding proteins that assume the function of signal transduction. In this paper, we describe the signature structural domains and major regulatory signals of PDZ-LIM family proteins and provide an overview and discussion of their functions in various tumors and major diseases, aiming to provide directions for future disease (mainly tumor) prevention and drug development.

Abstract

PDZ-LIM family proteins (PDLIMs) are a kind of scaffolding proteins that contain PDZ and LIM interaction domains. As protein–protein interacting molecules, PDZ and LIM domains function as scaffolds to bind to a variety of proteins. The PDLIMs are composed of evolutionarily conserved proteins found throughout different species. They can participate in cell signal transduction by mediating the interaction of signal molecules. They are involved in many important physiological processes, such as cell differentiation, proliferation, migration, and the maintenance of cellular structural integrity. Studies have shown that dysregulation of the PDLIMs leads to tumor formation and development. In this paper, we review and integrate the current knowledge on PDLIMs. The structure and function of the PDZ and LIM structural domains and the role of the PDLIMs in tumor development are described.

1. Introduction

Cell signal transduction is a complex cascade process that requires precise spatiotemporal regulation to ensure the coordinated operation of signaling events. And scaffold proteins are indispensable for the coordinated and accurate transmission of signals. Although the term scaffold is widely used in academia and not strictly defined, it is currently widely believed that scaffold proteins are protein molecules that bind to two or more proteins for signal regulation [1]. Scaffold proteins mainly function in four ways: assembling signal components, regulating the localization of signaling molecules in the cell, assisting with positive and negative feedback regulation, and protecting activation signaling proteins from inactivation [2]. In these ways, scaffold proteins increase the flexibility of signaling while making it simpler and more efficient.
PDZ-LIM family proteins (PDLIMs) are kinds of scaffold proteins. Based on genetic structure and phylogenetics, they can be divided into four subtypes: ALP (PDLIM3), Elfin (PDLIM1, CLP36), Mystique (PDLIM2, SLIM), and RIL (PDLIM4) in the ALP subfamily (Figure 1); Enigma (LMP-1, PDLIM7), Enigma Homolog (ENH, PDLIM5), and ZASP (Cypher, Oracle, PDLIM6) in the Enigma subfamily (Figure 2); and LIMK1, LIMK2, and LIM Domain Only Protein 7 (LMO7, FBXO20) [3] in the LIM kinase subfamily (Figure 3). The current structural analysis shows that all other PDLIMs do not have enzymatically active structures except for the LIM kinase subfamily. All PDLIMs mainly act as scaffold proteins through the PDZ and LIM domains. Since their discovery, many PDLIM-binding proteins have been identified, including actin, kinases, and phosphatases [4,5,6,7,8]. Most of them are involved in cancer signaling. Therefore, the role of PDLIMs in cancer has received increasing academic attention. Abundant studies have demonstrated that PDLIMs are involved in the regulation of the tumor cell cycle, proliferation, metastasis, and epithelial–mesenchymal transition (EMT) by modulating signaling [9,10,11,12,13]. In this review, we focus on summarizing the role of PDLIMs in tumor biology while describing the major signaling pathways they are involved in regulating. Overall, we sought to highlight the importance of PDLIMs in tumorigenesis and development and, by reviewing the potential functions of these molecules in the regulation of signaling responses, stimulate and help future investigators develop new therapeutic strategies for tumors by considering their functions.

2. Structural Features of the PDLIMs

PDLIMs include at least one PDZ and one LIM domain, which contribute to the ability of PDLIMs to regulate the spatiotemporal transduction of signals through direct or indirect binding to other protein molecules and to perform a wide range of biological functions in the cell.

2.1. PDZ Domain

The PDZ domain is originally a 90-amino-acid tandem repeat found in the Drosophila postsynaptic density protein PSD-95, the tumor suppressor DGL, and the epithelial tight junction protein ZO-1 (hence the acronym PDZ) [14,15,16]. The classical PDZ domain is a “sandwich” structure consisting of two α-helices and six antiparallel β-helices [17]. As one of the most common protein–protein interaction domains, more than 200 PDZ domain structures have been resolved so far, including not only individual PDZ domains but also complexes of PDZ and its ligands and the dimerization of PDZ and PDZ. The N-terminal and C-terminal ends of the PDZ domain are in close proximity to each other, and the first α-helix contained therein forms a groove with the second β-fold, allowing the C-terminal peptide of the interacting protein to be embedded therein [18,19]. Based on the characteristics of the ligand C-terminal amino acid sequence, PDZ domains are traditionally classified into three categories: Class I PDZ proteins recognize the ligand C-terminal sequence X-S/T-X-Φ; Class II proteins recognize XΦ-X-Φ; and Class III proteins recognize X-D/E-X-Φ (where X indicates any amino acid and Φ indicates hydrophobic amino acid) [20,21]. However, more studies have shown that the recognition and binding of substrates by PDZ are mainly achieved through the seven amino acids at the end, and the PDZ domain can be classified into 16 different specific classes based on the characteristics of these binding motifs [22]. Moreover, both the GLGF sequence within the PDZ and the sequence within the ligand protein affect protein–ligand binding [23]. The classical example is that the PDZ domain of Syntrophin or PSD-95 proteins can interact with the β-hairpin finger structure inside nNOS proteins [24,25]. By analyzing the PDZ domain on a large scale, the researchers found that PDZ mutants can still explicitly select ligands and function after artificial mutations, which suggests that the PDZ domain has an intrinsic ability to bind substrates and is robust under mutational loading, which may be an ideal model to support the high speed of evolution of interaction networks. The authors then investigated the co-evolution of the PDZ with ligands, but due to the differences between in vitro evolution and natural selection in biology, the authors obtained an evolutionary model with a higher affinity for ligands but lower specificity. Future studies on the structural evolution of the PDZ need to be continued [26]. Tandemly repeated PDZ structural domains enable the spatially organized and coordinated binding of proteins to specific targets [27]. It is currently believed that PDZ binds mainly to membrane proteins, cytoplasmic signaling proteins, and cytoskeletal proteins. For example, the PDZ domain can bind to the palladin protein, α-actinin, and β-tropomyosin and then localize to F-actin to regulate the cytoskeleton and signaling [24,28]. The PDZ domain also binds to phospholipids and enhances their ability to mediate the aggregation of membrane proteins, playing an important role in maintaining cell polarization [29]. In addition, the PDZ domain not only binds to proteoglycan receptors and affects cell migration and adhesion but also binds to CNS glutamate receptors and affects synaptic plasticity, ion channel opening, and cell surface receptor expression [17,30,31]. As a central organizer of protein complexes on the plasma membrane and an important modular protein-interacting structural domain, the PDZ domain is involved in the formation of protein complexes that mediate cellular adhesion, ion transport, and signaling, thereby regulating cellular homeostasis. Several researchers have demonstrated the involvement of PDZ domains in the regulation of cell growth, differentiation, and morphogenetic movements during development using different animal models [23,32]. Continued in-depth studies of the specific binding properties of PDZ and the specific biological functions it exhibits will provide more insights into orderly intracellular signaling.

2.2. LIM Domain

Compared to the PDZ domain, the LIM domain is shorter, consisting of only 50–65 amino acids, and generally contains CX2CX16–23HX2CX2CX2CX16–21 CX2(CH/D) motifs [33,34]. LIM was first discovered in three transcription factors, LIN-11, ISL-1, and MEC-3, and was named from their initials. The LIM domain is rich in histidine and cysteine and can bind two zinc ions, forming a zinc finger structure [35]. Unlike GATA-like zinc finger domains, however, the LIM domain only mediates protein–protein interactions and does not recognize DNA. In contrast to PDZ domains, the recognition motif for the LIM domain has not been clearly defined to date [36]. According to the most common classification, proteins containing the LIM domain can be divided into four categories: class I LIM homomorphic domain proteins; class II LIM kinases; class III proteins with LIM domain at the C-terminus and containing other domains; and class IV LIM domain-only proteins. Similar to the PDZ domain, LIM domains act as a “scaffold”, playing a protein-binding function and taking part in a variety of intracellular signaling functions [37]. Tight turn structures containing important Tyr or Phe residues are the “sequence code” for endocytosis of receptors on the cell surface, and the LIM domain in proteins provides molecular recognition of Tyr-containing tight turn structures. As an example, Enigma binds to tyrosine tight loops in the insulin receptor via the LIM domain, thereby facilitating the activation of signaling pathways and maintaining normal physiological function [38,39]. The LIM domain, associated with cellular mechanosensing, is a major member of the integrin adhesion site, participates in the formation of cytoskeletal complexes, converts mechanical forces into biochemical signals, and plays an important role in regulating the balance between exogenous and endogenous forces [40,41,42]. In addition, the LIM structural domain can regulate gene expression by forming transcription factor complexes with other proteins or by nucleoplasmic shuttling [43]. In conclusion, the LIM domain plays indispensable roles in the cytoskeleton, cell differentiation, and organ development [44]. Several researchers have found that LIM domain mRNA is localized to the lymphatic, muscular, and circulatory systems, but whether the LIM domain plays a role in cardiac and hematopoietic morphogenesis remains to be further investigated [45]. Disruptions in signaling caused by LIM dysfunction are an important cause of disease, especially tumorigenesis. The large number of proteins contained in the LIM domain makes it difficult to understand its function in depth. However, the LIM domain is essential for the PDLIMs, and its presence is an important reason for the diversity of functions performed by this family of proteins.

2.3. Other Domains

In addition to the PDZ and LIM domains, some of the PDLIMs also contain other structural domains. For example, the ZM domain, first identified in ZASP proteins, mediates protein binding to α-actinin and is required for the Z-line recruitment of proteins in cultured myoblasts [46,47]. LIMK1 and 2 contain a kinase domain that regulates actin polymerization and microtubule disassembly [48]. LMO7 contains a calponin homolog (CH) domain, which is one of the most common modules of actin-binding proteins and is highly structurally conserved. The CH domain binds to actin, tubulin, and signaling proteins, regulates the actin cytoskeleton, activates signaling pathways, and mobilizes calcium ions [49,50]. The presence of these functional motifs allows the PDZ-LIM family to play a richer role in regulating signaling, which we will not repeat here.

3. PDLIMs and Signaling Pathways

The PDLIMs can be involved in various signaling pathways as upstream molecules or downstream targets. Among them, the most studied are integrin-related signaling, TGF-β signaling, MAPK signaling, and NF-κB signaling pathways.

3.1. Integrin Signaling Pathway

Integrins are cell surface receptors composed of α and β subunits that regulate cell–cell and cell–extracellular matrix interactions [51]. At least 18α and 8β subunits are assembled into 24 different integrins [52]. Although integrins lack kinase activity, they can recruit kinases and assemble signaling complexes on the cytoplasmic face of the plasma membrane, thereby inducing signal transduction [53]. Many PDLIMs have been found to interact with integrins and affect cytoskeletal and cell adhesion functions. PDLIM1 and PDLIM5 stabilize α5β1 integrin-α-actinin-actin junctions [54]. Moreover, α2β1 integrin can activate PDLIM1, but PDLIM1 may prevent α-actinin-1 from binding to integrin subunits and foci adhesion, directing the molecule to actin filaments and stress fibers [55]. ZASP can activate α5β1 integrins in mammals and αPS2βPS integrins in Drosophila with talin, and it can also activate integrins with vimentin [42,56]. PDLIM2 deficiency upregulates the expression of β1 integrin, leading to a loss of the normal phenotype in mammary epithelial cells [57]. Upon exposure to mechanical forces, PDLIM5 and PDLIM7 are recruited to the integrin adhesion complex, which allows tendon cells and muscles to adapt to and withstand mechanical stress [58]. LMO7 is negatively regulated by the integrin-dependent signaling p130Cas, and dysregulation of LMO7 has been associated with X-linked Emory-Dreyfus muscular dystrophy [59]. Integrins also affect the expression and activation of PDLIMs. The expression of integrin α6β4 upregulates PDLIM4 and promotes cell migration capacity [60], and integrins, through activation of LIMK/Cofilin, not only control the initial formation of filopodia (FLPs) and promote the maturation of adherence sites and cell proliferation [61] but also lead to the disruption of adhesion junctions and loss of spermatogenic epithelial cells [62]. PrPc deficiency triggers integrin aggregation and overactivation of the LIMK pathway, leading to impaired neuronal synapse formation [63] (Figure 4).

3.2. TGF-β Signaling Pathway

TGF-β family members include three TGF-β isoforms, bone morphogenetic proteins (BMPs), and other proteins [64]. TGF-β activates the LIMK signaling axis and affects actin-myosin contractility, which in turn triggers cytoskeletal rearrangements and alters cell migration and invasion [65]. Activation of the TGF-β/LIMK pathway in retinal pigment epithelial cells leads to proliferative retinopathy in vitro [66]. The SMAD3-SMAD4 complex translocates to the nucleus and promotes the transcription of miR-181b, a molecule that promotes LIMK/Cofilin degradation via Sema3A, leading to atrial fibrosis through actin remodeling, sodium cilia formation, increased SMA stability, and the EMT process [67]. PDLIM5 promotes the migratory invasive capacity of lung cancer by inhibiting SMAD3 degradation [68]. But in pulmonary artery smooth muscle cells, PDLIM5 negatively regulates the expression of SMC markers, and overexpression of PDLIM5 inhibits TGF-β/Smad signaling, which attenuates hypoxia-induced elevation of RVSP, RV hypertrophy, and arterial wall remodeling [69]. At a certain concentration and time, TGF-β-induced expression of LMO7 inhibits TGFβ self-induction and αvβ3 integrin transcription by regulating the stability of AP-1 protein, thus limiting extracellular matrix deposition and vascular fibrosis [70]. TGF-β upregulation of LMO7 stimulates the invasive ability of mouse ascites hepatocellular carcinoma cells [71]. Due to its osteoinductive properties, PDLIM7 recruits many bone-forming factors [72,73]. The overexpression of PDLIM7 upregulates TGF-β, BMP2, BMP-4, BMP-6, BMP-7, and BMP receptors [74,75,76,77,78,79]. For example, PDLIM7 is able to activate the BMP–2/Smad1/5 signaling pathway and induces the process of pulp/osteogenic differentiation of dental pulp stem cells. PDLIM7 mediates dendritic cell proliferation through the activation of LIMK1 upon binding to BMPRII [80] (Figure 5).

3.3. NF-κB Signaling Pathway

The term NF-κB is commonly used to describe the p50:p65 complex. In unstimulated cells, NF-κB is mainly located in the cytoplasm due to interactions with the NF-κB inhibitor (IkB) [81]. PDLIM1 inhibits the nuclear translocation of p65 and sequesters it in the cytoplasm, thereby inhibiting inflammatory signaling [82]. Nuclear PDLIM2 controls the overactivation of inflammatory signaling by promoting p65 degradation [83]. Similarly, PDLIM7 degrades p65 alone or synergistically with PDLIM2 [84]. PDLIM2 translocates to the cytoplasm to activate NF-κB and improve cell adhesion during cell differentiation [13]. The PDLIM2/NF-κB pathway inhibits high-fat-diet-induced hepatic lipogenesis and inflammation and protects articular chondrocytes from lipopolysaccharide-induced apoptosis, degeneration, and inflammatory damage in mice [85,86]. Not only that, PDLIM2 also promotes the degradation of NF-κB and STAT3 and inhibits the development of tumor resistance [87]. The Kaposi’s sarcoma herpesvirus (KSHV) inhibits PDLIM2, activates NF-κB and STAT3, and promotes tumorigenesis and maintenance [88].

3.4. MAPK Signaling Pathway

The mitogen-activated protein kinase (MAPK) family mainly consists of extracellularly regulated kinases (ERK), MAPK14, and c-jun N-terminal kinases or stress-activated protein kinases (JNK or SAPK) [89]. LMO7-deficient mice result in altered JNK and ERK pathway activation and ultimately cardiac dysfunction [90]. The PDLIM2 inhibition of MEK and ERK phosphorylation affects cell viability [91]. PDLIM3 and PDLIM5 increase p-p38 and thus regulate the proliferation and differentiation of chicken skeletal muscle satellite cells [92,93]. PDLIM4 promotes CCR7-driven dendritic cell migration through the JNK signaling pathway [94]. PDLIM6 deficiency increases p38 MAPK phosphorylation, which induces cardiomyocyte apoptosis and ultimately dilated cardiomyopathy [95]. In canine pulmonary artery SMCs, the inhibition of ERK MAPK or p38 MAPK suppresses LIMK2 expression and affects the reorganization of the inflammation-associated actin cytoskeleton [96] (Figure 6).

4. PDLIMs and Tumor

The PDLIMs serve as a platform for protein–protein interactions and are involved in the amplification and integration of signal transduction, and their indispensable role in intracellular signaling has prompted a series of studies aimed at understanding the mechanisms of the role of PDLIMs in tumorigenesis and development. These clinical and basic studies have shown that most members of PDLIMs (except for PDLIM6) are aberrantly expressed in a range of tumors, including lung cancer (LC), breast cancer (BC), and gastric cancer (GC), and are involved in the functional regulation of pathophysiology in a variety of tumors. However, the role of PDLIMs in tumor development is tissue-specific, and they play distinct roles in regulating the proliferation and migration of tumors of different tissue origins. Their specific functions need to be further investigated.

4.1. ALP Subfamily

4.1.1. PDLIM1

PDLIM1 functions mainly as a key regulatory protein for cell migration and invasion in various tumors, but its role in different tumors is inconsistent [97]. In glioma, PDLIM1 is a novel scaffolding protein for the neurotrophin receptor p75 (NTR) that promotes cell metastasis by facilitating PKA phosphorylation of the p75 S303 site to induce p75 activation [98]. Studies showed that PDLIM1 was highly expressed in BC patients’ tissue and plasma samples to promote BC cell migration and invasion. And mechanistic studies showed that PDLIM1 promotes cell polarization and migration by directly interacting with ACTN1 and ACTN4 to activate CDC42 [99,100]. However, in hepatocellular carcinoma (HCC), PDLIM1 was poorly expressed, and overexpression of PDLIM1 in HCC competed for binding with the ACTN4, resulting in the dissociation of ACTN4 from F-actin. This process effectively prevented the overgrowth of F-actin and downstream Hippo signaling activation, thus inhibiting HCC metastasis [101]. Similarly, PDLIM1 was lowly expressed in GC and promoted GC progression and cisplatin resistance [102]. PDLIM1 deletion or knockdown in highly metastatic colorectal cancer cells (CRCs) reduces the stability of the E-cadherin/β-catenin adhesion complex and inhibits the transcriptional activity of β-catenin on EMT-related genes [103]. Moreover, silencing of PDLIM1 in choriocarcinoma (CC) cells leads to cell expansion and loss of stress fibers and local adherent patches, a process that can be rescued by the overexpression of full-length PDLIM1 but not by mutants with deletion of the PDZ or LIM domain, confirming that PDLIM1 plays an important function in the pathogenesis of CC through its protein-binding domain [104]. In addition, PDLIM1 also acts as a tumor antigen to induce antibody responses in pancreatic adenocarcinoma (PAAD), but whether this immune response is tumor-specific needs further investigation [105].

4.1.2. PDLIM2

PDLIM2, located on chromosome 8p21, is frequently disrupted in various cancers [11,106]. In BC, conflicting findings regarding PDLIM2 expression and function have been reported. Ding et al. found that highly expressed miR-222 targets PDLIM2 and inhibits its expression, leading to malignant progression and lymphatic metastasis [107]. Moreover, the study also found hypermethylation in the PDLIM2 promoter region, and methylation inhibitors or vitamin D treatment resulted in demethylation of the PDLIM2 promoter, leading to upregulated PDLIM2 expression and BC progression arrest [108]. Further studies showed that PDLIM2 expression was associated with the EMT process. PDLIM2 is lowly expressed in non-EMT BC cells but highly expressed in infiltrating cells that do undergo EMT. Another study revealed that PDLIM2 promoted cell-directed migration, cytoskeletal polarization, and the EMT process by activating COP9 signaling vesicle activity and NF-κB and STAT3 signaling [109,110]. In addition, it was shown that high PDLIM2 expression in BC was associated with higher M2 macrophage infiltration.
In LC, PDLIM2 acts as a tumor suppressor gene. PDLIM2 represses multidrug resistance genes and cancer-associated genes, rendering cancer cells susceptible to immune attack and treatment [87]. Global or lung epithelial-specific deletion of PDLIM2 promotes LC development, chemoresistance, and resistance to anti-PD-1 therapies [87]. Moreover, the overexpression of PDLIM2 hinders the proliferation, colony formation, and invasive capacity of NCLCL cells by inhibiting NF-κB signaling [111].
The function of PDLIM2 has also been reported in other cancer types. In esophageal squamous cell carcinoma, PDLIM2 expression was significantly downregulated, and exon 7/8/9/10 of PDLIM2 was a novel valuable predictive characterization for esophageal cancer (EC) patients [112]. In HCC, PDLIM2 is also downregulated and associated with poorer prognosis. PDLIM2 exerts tumor-suppressive effects by inhibiting HCC cell proliferation, migration, invasion, EMT, and colony formation through the inhibition of β-catenin activity [113]. In addition, macrophage-derived exosomes from Laryngeal squamous cell carcinoma (LSCC) inhibited PDLIM2 expression by delivering miR-222-3p to LSCC cells, leading to elevated PFKL expression and enhanced glycolysis and thereby accelerating cell proliferation [114]. In ovarian cancer (OV), PDLIM2 is epigenetically suppressed, and PDLIM2 inhibition promotes OV growth in vivo and in vitro via NOS2-derived nitric oxide signaling, leading to M2-type macrophage recruitment [115]. In addition, some clinical analyses showed that PDLIM2 expression was also significantly downregulated in patients with metastatic CRC [116]. However, in prostate cancer (PC) and GC, PDLIM2 plays a pro-carcinogenic role, and specific deletion of PDLIM2 significantly suppressed cell proliferation and migration [91,117].

4.1.3. PDLIM3

PDLIM3 has been relatively poorly studied in tumors. In medulloblastoma, studies have shown that the abnormal expression of PDLIM3 and four other genes leads to the abnormal activation of Hedgehog (Hh) signaling, which is an important diagnostic marker for medulloblastoma patients [118]. A weighted gene co-expression network analysis of 401 patients with invasive uroepithelial carcinoma of the bladder revealed that high PDLIM3 expression was associated with poor patient prognosis [119]. However, Lu et al. found that PDLIM3 expression was significantly decreased in patients with bladder cancer (BCA) through a tumor cell differential expression gene screen, and the specific mechanism of action deserves further exploration [120]. PDLIM3 expression was also decreased in papillary thyroid carcinoma (PTC) samples from the Chernobyl region, but the correlation between PDLIM3 expression and clinical stage, metastasis, and prognosis of thyroid carcinoma (THCA) patients and its role in the pathogenesis of THCA is unclear [121].

4.1.4. PDLIM4

PDLIM4 is considered a tumor suppressor gene whose expression is downregulated in a variety of cancers. For example, it was shown that PDLIM4 expression is downregulated in OV; that its expression is negatively correlated with clinical stage, lymphatic metastasis, and patient survival; and that the overexpression of PDLIM4 inhibits OV cell proliferation, migration, invasion, and xenograft tumor growth by suppressing STAT3 signaling activation [122]. Hypermethylation of the PDLIM4 gene can be used as a diagnostic marker for PC. Numerous studies have shown that hypermethylation of the PDLIM4 gene leads to the development and malignant progression of PC, and functional experiments have confirmed that PDLIM4 can inhibit tumor growth by suppressing PC cell proliferation [123,124,125,126]. In addition, the PDLIM4 gene was also found to undergo hypermethylation in THCA and kidney cancer, but whether the change has a biological effect is still unknown [127,128]. For now, the role of PDLIM4 in BC is controversial. Feng et al. found that PDLIM4 was highly methylated in tumor tissue by analyzing 38 pairs of patients with primary BC lymphatic metastases [129], and other studies showed that PDLIM4 expression was negatively correlated with tumor size, differentiation status, and SPF (S-phase fraction) value [130]. However, it has also been shown that upregulation of PDLIM4 expression promotes BC cell migration [131]. Proper assessment of its role in the malignant progression of BC facilitates the implementation of individualized tumor treatment. These findings shake the perception of PDLIM4 as an oncogene, and the complexity of life signal transduction makes it possible for the same gene to perform completely opposite functions in different settings (Table 1).

4.2. Enigma Subfamily

4.2.1. PDLIM5

High expression of PDLIM5 in PC is significantly correlated with poor survival. And the rs17021918 mutation of PDLIM5 is negatively correlated with the progression of PC patients. Liu et al. found that PDLIM5 can inhibit AMPK ubiquitinated degradation by binding to it, thus promoting the malignant progression of PC [132,133]. Meanwhile, Yan et al. found that PDLIM5 is a novel substrate for AMPK, which can directly phosphorylate PDLIM5 Ser177, and mutations at this site significantly inhibit cell migration and attenuate lamellipodia formation [134]. Furthermore, PDLIM5 promotes THCA malignant progression by activating the Ras/ERK signaling pathway [135]. In LC, high expression of PDLIM5 accelerates metastasis by competitively binding Smad3 with the E3 ubiquitin ligase STUB1 and promoting TGFβ signaling activation and EMT [68]. Similarly, others have probed and found a trend toward better survival in NSCLC patients with low PDLIM5 expression [136]. High expression of PDLIM5 also promotes gefitinib resistance by inhibiting autophagy in PC9GR cells [137]. Knockdown of PDLIM5 through a novel mesoporous silica-loaded PDLIM5 siRNA (small interfering RNA) nanoplatform can reverse the gefitinib-resistant phenotype in PC9 [138].

4.2.2. PDLIM6

PDLIM6 is a myosin Z family protein expressed mainly in the transverse muscle, skeletal muscle, and heart. Its deletion or mutation causes mainly cardiac and skeletal muscle disorders. To date, there is no direct evidence of a role for PDLIM6 in tumor development [139,140,141].

4.2.3. PDLIM7

PDLIM7 plays a significant role in cancer development and progression. In acute myeloid leukemia (AML), high expression of PDLIM7 is an independent risk factor for poor event-free survival (EFS) and overall survival [142]. It is also associated with negative patient survival outcomes in BC, where it prevents the ubiquitination degradation of RETMEN2A mediated by cbl-c [143]. Furthermore, in liver metastasis BC cells, both PDLIM2 and PDLIM7 can bind to Claudin-2, a molecule known to promote BC liver metastasis, suggesting their involvement in BC metastasis [144].
In HCC and CRC, PDLIM7 enhances the growth of cancer cells. Mitogenic stimulating factors like serum, FGF, and HGF increase PDLIM7 transcription by inducing the serum response factor (SRF). PDLIM7 then inhibits the auto-ubiquitination of P53 by binding to MDM2, leading to P53 degradation and promoting cancer cell growth [145].
In THCA, PDLIM7 is aberrantly highly expressed in all subtypes and positively correlates with lymphatic metastasis and tumor malignancy. Subcellular localization analysis reveals differences in PDLIM7 distribution, with cytoplasmic localization in classical papillary thyroid carcinoma (PTC) and mixed cytoplasmic and nuclear staining in follicular papillary thyroid carcinoma (FVPTC), poorly differentiated thyroid carcinoma (PDTC), and undifferentiated thyroid carcinoma (ATC). This suggests that PDLIM7 localization may be associated with the malignant progression of THCA [146]. This suggests that PDLIM7 localization may be associated with the malignant progression of THCA. PDLIM7 also plays a role in regulating THCA progression by selectively binding to the THCA-specific RET/PTC2 mutant short heterodimer protein [147]. Knockdown of PDLIM7 in THCA cells leads to the downregulation of AKT and Survivin protein expression, inhibiting cell proliferation and promoting apoptosis [148].
Moreover, recent studies have identified PDLIM7 in cancer-associated fibroblasts binding to calponin 1 protein and inhibiting its degradation by E3 ubiquitin ligase NEDD4-1. This activation of ROCK1/MLC signaling leads to increased matrix stiffness and promotes the activation of YAP signaling in GC cells, making them resistant to 5-fluorouracil treatment [149]. However, a few studies have also shown that PDLIM7 exerts oncogenic functions, such as the downregulation of PDLIM7 expression in osteosarcoma (OS) tissues and the inhibition of cell proliferation and migration by PDLIM7 overexpression in OS cells [150].
Overall, PDLIM7 plays complex roles in various cancers, acting as both a promoter and inhibitor of tumor progression depending on the specific cancer type Table 2.

4.3. LMO7

LMO7 was reported to be a marker for BC diagnosis. LMO7 is highly expressed in invasive BC tissues and synergizes with Rho GTPase to activate myocardin-related transcription factors (MRTFs), which then upregulate SRF transcriptional activity to promote BC cell migration [151]. Analysis of the clinical correlation between LMO7 expression and lung adenocarcinoma (LUAD) patients showed that LMO7 expression was negatively associated with lymphatic metastasis and poor prognosis [152]. Animal experiments revealed that lacking LMO7 increased susceptibility to spontaneous LC in mice. In vitro experiments showed that LMO7 knockout resulting in numerical chromosome abnormalities was the major cause of LC induction [153]. Meanwhile, it has been shown that miR-96 promotes proliferation, migration, and drug resistance of LC cells through the downregulation of LMO7. These findings suggest that LMO7 acts as a tumor suppressor in LC pathogenesis [154]. LMO7 has been reported to be involved in tumor progression by fusing with other genes. Two case reports showed that LMO7 and ALK1 can form fusion mutations, which affect LC progression and targeted drug efficacy [155,156]. In addition. LMO7-BRAF rearrangement has been found in THCA, and the fusion protein formed stimulates ERK1/2 phosphorylation and promotes cell growth in a similar pattern to BRAF [90]. The overexpression of LMO7 in CRC cells and cervical cancer (CCA) cells interfered with spindle check site (SAC) action, prolonged mitosis, and inhibited cell proliferation [157]. In PAAD, LMO7 deficiency inhibited tumor cell proliferation and metastasis and the progression of subcutaneous hormonal tumors [158]. LMO7 expression was positively correlated with the metastatic ability of HCC, but it is puzzling that in vitro overexpression of LMO7 did not promote the migration of HCC cells [71]. LMO7 is also involved in tumor microenvironment regulation. DAPK3 induces K63-linked polyubiquitination of STING by phosphorylating LMO7, which in turn drives tumor-intrinsic innate immunity and tumor immune surveillance [159].

4.4. LIM Kinase

Of all PDLIMs, LIMK1/2 has been most extensively studied as a kinase molecule, and more studies continue to attempt to gain insight into the role of LIMK1/2 in tumor progression. The primary function of LIMK1/2 is to act as a signaling node that controls the dynamics of the actin cytoskeleton and regulates the phosphorylation of cofilin [160,161,162]. LIMK1/2 shares nearly 70% similarity in the kinase domain and binds to macromolecular ligands mainly through the LIM domain, which include Rho GTase, Rac1, Cdc42, RhoA, and their downstream effect molecules p21 activated protein kinase (PAK) 1–4 and ROCK1/2. ROCK1/2 and PAK activate LIMK through phosphorylation. Activated LIMK specifically binds to the downstream molecule cofilin and phosphorylates the Ser3 site of cofilin, leading to its inactivation. The occurrence of this event subsequently regulates cell morphology, motility, adhesion, and migration by modulating cellular actin backbone polymerization. As with all other kinase molecules, the development of inhibitors against LIMK1/2 is currently an ideal strategy for antitumor drug development [163,164]. The following sections compile the current understanding of the involvement of LIMK1/2 in tumor progression.

4.4.1. LIMK1

LIMK1 in GC

As shown, LIMK1 is highly expressed in GC patients and positively correlates with the degree of GC differentiation clinical stage, lymph node metastasis, and poor prognosis [165]. Treatment with PAK4 inhibitors (GL-1196, LC-0882, and LCH-7749944), diallyl disulfide (garlic extract, DADS), bitter ginseng, quercetin, and knockdown of RhoGDI2 or DGCR6L have been shown to inhibit GC malignant progression by inhibiting LIMK1/Cofilin signaling [166,167,168,169,170,171,172,173]. In addition, the clinical agent Dabrafenib has also been shown to inhibit GC metastasis by targeting LIMK [174].

LIMK1 in CRC

LIMK1 was highly expressed in CRC tissues and positively correlated with metastasis, overall survival, and the pathological grade of patients [175,176]. LIMK1 was found to promote CRC proliferation and metastasis by binding to STK25, MYH9, and ACTIN4 [177,178]. As in GC, DADS treatment also inhibited the malignant progression of CRC by suppressing LIMK1/Cofilin signaling [179]. In addition, miR-27b-3p and miR-145 have been reported to exert antitumor effects in CRC by inhibiting LIMK1/Cofilin signaling through targeting LIMK1 and PAK4, respectively [180,181]. However, the role of LIMK1/Cofilin signaling in CRC is controversial, as recent studies have shown that IRX5 and SSH3 promote CRC progression by inhibiting this signaling [182,183].

LIMK1 in BC

Several studies have confirmed that LIMK1 is significantly upregulated in BC patients. The specific Rock and Rho inhibitors Y-27632 and C3 inhibit LIMK1 activation to suppress BC cell migration and invasion [184,185]. LIMK1 promotes tumor growth and tumor angiogenesis by increasing uPA expression [186]. In addition, LIMK1 alters the distribution of cortical actin in MT1-MMP positive inclusions and then promotes BC cell migration [187]. Molecular biology studies showed that EBP50, SEMA3B, MEX3A, and SphK2/S1P affect BC cell migration by regulating LIMK1/cofilin signaling [188,189,190,191]. Pharmacological studies revealed that herbal monomer serralactone A inhibits BC cell migration by inhibiting LIMK1 activity [192]. Similarly, miR-200b-3p, miR-429-5p, miR-128-3p, miR-519d-3p, and miR-143-3p inhibit BC cell proliferation and migration by downregulating LIMK1 [193,194,195,196].

LIMK1 in PC

The current results show that ectopic expression of LIMK1 gives PC cells an aggressive phenotype [197]. MiR-143, miR-23a, and nuclear clusterin exert cancer-suppressive effects by blocking the activation of LIMK1/Cofilin [198,199,200]. In addition, CXCL12/CXCR4 antagonizes doxorubicin-induced cell cycle arrest by activating LIMK1/Cofilin signaling [201]. In androgen-dependent PC patients, the LIMK1 inhibitor inhibits PC cell proliferation and migration by altering microtubule-dynamics-impeding DHT-induced androgen receptor nuclear translocation, protein stability, and transcriptional activity [202].

LIMK1 in LC

LIMK1 expression was significantly upregulated in NSCLC patients. Bioinformatic analysis revealed that LIMK1 could be used as a biomarker for poor prognosis in LUAD and a potential target for immunotherapy [203,204]. Consistent with other tumors, in NSCLC, PAK4 can also regulate cell migration and invasion by regulating the phosphorylation of LIMK1 [205]. The herbal monomers Leuconostoc and Mucuna pruriens inhibit LIMK1/Cofilin signaling and thus LC cell migration [206,207]. miR-27b inhibits the proliferation and migration of NSCLC cells by suppressing LIMK1 expression [208]. In addition, knockdown of LIMK1 enhanced the sensitivity of LC cells to cisplatin [209].

LIMK1 in OS

Immunohistochemical analysis of cancer and paracancer tissues revealed that LIMK1 was highly expressed in OS tissues [210]. Insulin stimulation activates the LIMK1/Cofilin pathway and thus promotes OS cell proliferation [211]. 6-Hydroxythioglycine inhibits the migration of OS cells by reducing LIMK1 expression [212]. In addition, it was shown that LIMK1 expression was elevated in multidrug-resistant OS cells and that knockdown of LIMK1 enhances the sensitivity of multidrug-resistant cells to drugs [213].

LIMK1 in Cervical Cancer (CC)

The expression level of LIMK1 in CC was significantly higher than that in the control and heterogeneous hyperplasia groups. Further analysis revealed that LIMK1 expression was positively correlated with metastasis and negatively correlated with patients’ overall survival [214]. It was reported that miR-125a-5p enhanced the efficacy of cisplatin on CC cells by targeting LIMK1. In addition, FOXD3-AS1 promoted CC progression by competitively sponging miR-128p to upregulate LIMK1 expression [215,216].

LIMK1 in HCC

The role of LIMK1 in HCC has been discovered recently, and studies have shown that exo-miR-374c-5p inhibits EMT by targeting the LIMK1-Wnt/β-catenin signaling to suppress HCC metastasis [217], while lncRNA H19 upregulates LIMK1 by sponging miR-520a-3p to promote HCC progression [218]. Furthermore, studies on HCC revealed that EGF drives the nuclear translocation of LIMK1 by activating the interaction between p-ERK and LIMK1 and that nuclear LIMK1 directly binds to the promoter region of c-myc to stimulate c-myc transcription, thereby promoting HCC progression [219].

LIMK1 in Other Tumors

LIMK1 also plays an important role in other tumors. For example, overexpression of the integrin αVβ3 inhibits RhoA activation and thus LIMK1-mediated cofilin phosphorylation, which promotes melanoma cell migration and invasion [220]. It was found that miR-106a inhibits cell proliferation, migration, and EMT by suppressing LIMK1 expression in oral cancer cell lines [221]. In addition, miR-20a and miR-373 inhibited the progression of THCA/cutaneous squamous cell carcinoma and glioblastoma, respectively, by targeting LIMK1 [222,223,224]. LINC00941 acts as a sponge for miR-335-5p, upregulates ROCK1 to activate the LIMK1/Cofilin-1 pathway, and subsequently facilitates PAAD growth and metastasis [225]. Immunohistochemical staining showed that LIMK1 was highly expressed in OV tissues and correlated with the clinical stage of OV patients. In OV cells, ET-1/ETAR promotes LIMK1/Cofilin activation through activation of Rock signaling, thereby promoting cell migration [226]. In addition, miR-138 inhibits OV cell migration by targeting LIMK1, but LIMK1 is not involved in miR-138-regulated cell proliferation [227].

4.4.2. LIMK2

LIMK2 in CRC

Studies have shown that LIMK2 expression is downregulated in CRC patients and continues to be downregulated as the tumor deteriorates [228]. Filipe et al. further confirmed that LIMK2 is poorly expressed in the intestines of cancer-prone mice, as well as in human CRC cell lines and tumors. The reduced expression of LIMK2 is associated with shortened OS in patients. Further results revealed that LIMK2 inhibits the proliferation of stem cells, and LIMK2 deficiency promotes the growth of CRC in mice [229]. In addition, the overexpression of LIMK2 in CRC downregulates β-catenin and inhibits WNT signaling, thereby inhibiting cell proliferation and migration. However, LIMK2 has also been reported to be highly expressed in CRC tissues and positively correlated with clinical stage and lymph node metastasis [230]. Upregulated LIMK1/2 contributes to malignant progression and chemoresistance in CRC. In addition, miR-939-5p was shown to inhibit CRC metastasis by targeting LIMK2 [231].

LIMK2 in PC

LIMK2 is a specific target for the treatment of castration-resistant PC (CRPC), which is upregulated in androgen deprivation therapy. Inducible knockdown of LIMK2 completely reversed CRPC tumorigenesis in castrated mice. Mechanistic studies revealed that TWIST1 is a direct substrate of LIMK2. Under hypoxia, LIMK2 induces TWIST1 transcriptional activation and stabilizes TWIST1 by direct phosphorylation, thereby mediating CRPC development [232]. Recent studies have shown that SPOP, PTEN, and NKX3.1 are also substrates of LIMK1 and that LIMK1 affects PC progression and resistance by regulating the phosphorylation and degradation of these substrates [233,234,235].

LIMK2 in BC

Aurora A is a serine/threonine kinase that is overexpressed in most tumors. Emmanuel et al. identified LIMK2 as a substrate for Aurora A. Aurora A regulates LIMK2 enzyme activity, localization, and expression by promoting phosphorylation at S283, T494, and T505 sites of LIMK2, thereby inducing the occurrence of BC. In turn, LIMK2 promotes Aurora A expression through positive feedback regulation [236]. It has also been shown that LIMK2 promotes the metastatic progression of TNBC through the activation of SRPK1 [237]. In addition, consistent with LIMK1, SEMA3B also inhibits BC cell migration by suppressing LIMK2 activity [191].

LIMK2 in LC

Xu et al. found that MED12 knockdown activates LIMK2, causing abnormal remodeling of the actin cytoskeleton, disrupting the shedding of intercellular bridges, and leading to cytokinesis failure, thereby inhibiting NSCLC cell proliferation [238]. In contrast, lncRNA TUG1 upregulates LIMK2b (a splice variant of LIMK2), thereby promoting NSCLC proliferation and drug resistance [233]. Other studies on non-coding RNAs have also shown that PPVT1 and DHRS4-AS1 promote the expression of LIMK2 through miR-423-5p, thereby affecting the progression of lung squamous cell carcinoma (LUSC) [234].

LIMK2 in OS

When OS cells are stimulated by EGF, EGF activates RhoA and then phosphorylates LIMK2/Cofilin to promote cell migration. Knocking down LIMK2 in OS inhibits the formation of actin stress fibers and cell migration while also inhibiting the occurrence of EMT. In addition, BMPR2 and PD-L2 also regulate OS metastasis by regulating the RhoA-Rock-LIMK2 signal [235,239].

LIMK2 in Neuroblastoma

LIMK2 is highly expressed in vincristine- and colchicine-resistant neuroblastoma cell lines. And the inhibition of LIMK2 expression increased the sensitivity of neuroblastoma to vincristine and colchicine in [240]. Meanwhile, some findings suggest that LIMK2 expression is elevated in neuroblastoma cells resistant to microtubule-targeted drugs. Further, mechanistic findings suggest that LIMK2 is involved in the regulation of cellular drug resistance through the regulation of microtubule acetylation and microtubule protein polymerization protein 1 (TTP1) expression [241].

LIMK2 in Other Tumors

A zebrafish xenograft model showed that LIMK2 synergistically regulates PAAD development with LIMK1 [242]. In addition, the LIMK2 inhibitor T56-LIMKi inhibited the growth of subcutaneous tumor models of PAAD in mice by specifically inhibiting LIMK2 without cross-reacting with LIMK1 [243]. Wang et al. demonstrated for the first time that LIMK2 was highly expressed in BCA patients. Multifactorial logistic regression analysis showed that the SNP mutation rs2073859 (G-A mutation) in the UTR region of LIMK2 was significantly more prevalent in BCA patients than in controls, and the mutation was positively correlated with metastasis and the clinical stage of patients. Functional assays showed that LIMK2 overexpression promoted BCA cell proliferation, migration, and invasion [244]. Research has found that LIMK2b is downregulated in esophageal and THCA tissues. Overexpression of LIMK2b promotes Cofilin phosphorylation, leading to the arrest of these two types of tumor cells in the G2/M phase and a decrease in cell migration ability [245] (Table 3).

5. Conclusions and Future Directions

Cancer occurs when mutations or epigenetic alterations in genes that regulate signal transduction lead to the dysregulation of intracellular signaling homeostasis [249]. Many reports have highlighted the important role of scaffolding proteins as a “silver bullet” for signal transduction in maintaining these sophisticated signals. PDLIMs are several different groups of scaffolding proteins that play essential roles as mediators in various physiological processes such as cell proliferation, stemness, apoptosis, differentiation, and migration [97,250]. In this article, we highlight their role in tumor development. The level of PDLIM expression or protein dysfunction leads to tumor susceptibility and tumor progression, and the detection of PDLIM expression can effectively improve the identification of tumors. Although a large number of assays have clarified that the expression of PDLIMs is dysregulated in tumors, there is still a paucity of studies on the mechanism that triggers their dysregulation, with there being basically only miRNA-related studies in the literature, and continuing to search for upstream regulatory molecules has certain significance for inhibiting the signal at the source [251]. Considering the central role of PDLIMs in a wide range of tumors, targeting this class of molecules to synthesize corresponding inhibitors provides a new entry point for molecularly targeted therapies. To date, inhibitors targeting LIMK1 and LIMK2 have shown potent tumor-suppressive effects. Moreover, although there are no known active domains for the other eight proteins except LIMK1 and LIMK2, the emergence of the targeted protein degradation technology “PROTAC” offers the possibility of drug discovery for this class of non-druggable protein targets and is a worthy direction for future efforts [252].
However, as mentioned above, most PDLIMs have two-sided effects on tumor regulation, which may be related to the different protein signaling pathways interacting with each other in different environments; therefore, further exploration of the mechanism of action of PDLIMs in the process of tumorigenesis is of positive significance for the rational use and control of these molecules in tumor prevention and treatment. Meanwhile, PDLIMs of the same protein subfamily have great structural similarity, so, do they have synergistic effects among themselves in regulating tumor progression? At present, there are relatively blank studies on this point. A systematic and in-depth study of the relationship between PDLIMs and tumors will help people understand the status and role of PDLIMs in tumorigenesis, which is important for tumor diagnosis and treatment (Table 4).

Author Contributions

All authors contributed to writing and revising the manuscript content. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants from the National Natural Science Foundation of China (82203608 to Z.X.) and the Health Commission of Zhejiang Province of China (2021KY196 to Y.S.).

Data Availability Statement

The data can be shared up on request.

Acknowledgments

We are grateful to all of the participants who have made this research possible.

Conflicts of Interest

The authors declare that they have no competing interest.

References

  1. DiRusso, C.J.; Dashtiahangar, M.; Gilmore, T.D. Scaffold Proteins as Dynamic Integrators of Biological Processes. J. Biol. Chem. 2022, 298, 102628. [Google Scholar] [CrossRef]
  2. Shaw, A.S.; Filbert, E.L. Scaffold Proteins and Immune-Cell Signalling. Nat. Rev. Immunol. 2009, 9, 47–56. [Google Scholar] [CrossRef] [PubMed]
  3. Velthuis, A.J.W.T.; Bagowski, C.P. PDZ and LIM Domain-Encoding Genes: Molecular Interactions and Their Role in Development. Sci. World J. 2007, 7, 1470–1492. [Google Scholar] [CrossRef]
  4. Vallenius, T.; Luukko, K.; Mäkelä, T.P. CLP-36 PDZ-LIM Protein Associates with Nonmuscle Alpha-Actinin-1 and Alpha-Actinin-4. J. Biol. Chem. 2000, 275, 11100–11105. [Google Scholar] [CrossRef]
  5. Zhou, Q.; Ruiz-Lozano, P.; Martone, M.E.; Chen, J. Cypher, a Striated Muscle-Restricted PDZ and LIM Domain-Containing Protein, Binds to Alpha-Actinin-2 and Protein Kinase C. J. Biol. Chem. 1999, 274, 19807–19813. [Google Scholar] [CrossRef] [PubMed]
  6. Krcmery, J.; Camarata, T.; Kulisz, A.; Simon, H.-G. Nucleocytoplasmic Functions of the PDZ-LIM Protein Family: New Insights into Organ Development. Bioessays 2010, 32, 100–108. [Google Scholar] [CrossRef] [PubMed]
  7. Martinelli, V.C.; Kyle, W.B.; Kojic, S.; Vitulo, N.; Li, Z.; Belgrano, A.; Maiuri, P.; Banks, L.; Vatta, M.; Valle, G.; et al. ZASP Interacts with the Mechanosensing Protein Ankrd2 and P53 in the Signalling Network of Striated Muscle. PLoS ONE 2014, 9, e92259. [Google Scholar] [CrossRef]
  8. Lasorella, A.; Iavarone, A. The Protein ENH Is a Cytoplasmic Sequestration Factor for Id2 in Normal and Tumor Cells from the Nervous System. Proc. Natl. Acad. Sci. USA 2006, 103, 4976–4981. [Google Scholar] [CrossRef]
  9. Piao, S.; Zheng, L.; Zheng, H.; Zhou, M.; Feng, Q.; Zhou, S.; Ke, M.; Yang, H.; Wang, X. High Expression of PDLIM2 Predicts a Poor Prognosis in Prostate Cancer and Is Correlated with Epithelial-Mesenchymal Transition and Immune Cell Infiltration. J. Immunol. Res. 2022, 2022, 2922832. [Google Scholar] [CrossRef]
  10. Kundu, J.; Bakshi, S.; Joshi, H.; Bhadada, S.K.; Verma, I.; Sharma, S. Proteomic Profiling of Peripheral Blood Mononuclear Cells Isolated from Patients with Tuberculosis and Diabetes Copathogenesis—A Pilot Study. PLoS ONE 2020, 15, e0233326. [Google Scholar] [CrossRef]
  11. Loughran, G.; Healy, N.C.; Kiely, P.A.; Huigsloot, M.; Kedersha, N.L.; O’Connor, R. Mystique Is a New Insulin-like Growth Factor-I-Regulated PDZ-LIM Domain Protein That Promotes Cell Attachment and Migration and Suppresses Anchorage-Independent Growth. Mol. Biol. Cell 2005, 16, 1811–1822. [Google Scholar] [CrossRef] [PubMed]
  12. Ooshio, T.; Irie, K.; Morimoto, K.; Fukuhara, A.; Imai, T.; Takai, Y. Involvement of LMO7 in the Association of Two Cell-Cell Adhesion Molecules, Nectin and E-Cadherin, through Afadin and Alpha-Actinin in Epithelial Cells. J. Biol. Chem. 2004, 279, 31365–31373. [Google Scholar] [CrossRef] [PubMed]
  13. Healy, N.C.; O’connor, R. Sequestration of PDLIM2 in the Cytoplasm of Monocytic/Macrophage Cells Is Associated with Adhesion and Increased Nuclear Activity of NF-kappaB. J. Leukoc. Biol. 2009, 85, 481–490. [Google Scholar] [CrossRef] [PubMed]
  14. Itoh, M.; Nagafuchi, A.; Yonemura, S.; Kitani-Yasuda, T.; Tsukita, S. The 220-kD Protein Colocalizing with Cadherins in Non-Epithelial Cells Is Identical to ZO-1, a Tight Junction-Associated Protein in Epithelial Cells: cDNA Cloning and Immunoelectron Microscopy. J. Cell Biol. 1993, 121, 491–502. [Google Scholar] [CrossRef] [PubMed]
  15. Cho, K.-O.; Hunt, C.A.; Kennedy, M.B. The Rat Brain Postsynaptic Density Fraction Contains a Homolog of the Drosophila Discs-Large Tumor Suppressor Protein. Neuron 1992, 9, 929–942. [Google Scholar] [CrossRef] [PubMed]
  16. Woods, D.F.; Bryant, P.J. The Discs-Large Tumor Suppressor Gene of Drosophila Encodes a Guanylate Kinase Homolog Localized at Septate Junctions. Cell 1991, 66, 451–464. [Google Scholar] [CrossRef]
  17. Dev, K.K. Making Protein Interactions Druggable: Targeting PDZ Domains. Nat. Rev. Drug Discov. 2004, 3, 1047–1056. [Google Scholar] [CrossRef] [PubMed]
  18. Xue, L.; Zhang, F.; Chen, X.; Lin, J.; Shi, J. PDZ Protein Mediated Activity-Dependent LTP/LTD Developmental Switch at Rat Retinocollicular Synapses. Am. J. Physiol. Cell Physiol. 2010, 298, C1572–C1582. [Google Scholar] [CrossRef] [PubMed]
  19. Tao, Y.-X.; Johns, R.A. PDZ Domains at Excitatory Synapses: Potential Molecular Targets for Persistent Pain Treatment. Curr. Neuropharmacol. 2006, 4, 217–223. [Google Scholar] [CrossRef]
  20. Gallardo, R.; Ivarsson, Y.; Schymkowitz, J.; Rousseau, F.; Zimmermann, P. Structural Diversity of PDZ-Lipid Interactions. Chembiochem 2010, 11, 456–467. [Google Scholar] [CrossRef] [PubMed]
  21. Lenfant, N.; Polanowska, J.; Bamps, S.; Omi, S.; Borg, J.-P.; Reboul, J. A Genome-Wide Study of PDZ-Domain Interactions in C. Elegans Reveals a High Frequency of Non-Canonical Binding. BMC Genom. 2010, 11, 671. [Google Scholar] [CrossRef]
  22. Tonikian, R.; Zhang, Y.; Sazinsky, S.L.; Currell, B.; Yeh, J.-H.; Reva, B.; A Held, H.; A Appleton, B.; Evangelista, M.; Wu, Y.; et al. A Specificity Map for the PDZ Domain Family. PLoS Biol. 2008, 6, e239. [Google Scholar] [CrossRef] [PubMed]
  23. Fanning, A.S.; Anderson, J.M. PDZ Domains: Fundamental Building Blocks in the Organization of Protein Complexes at the Plasma Membrane. J. Clin. Investig. 1999, 103, 767–772. [Google Scholar] [CrossRef]
  24. Jeleń, F.; Oleksy, A.; Smietana, K.; Otlewski, J. PDZ Domains—Common Players in the Cell Signaling. Acta Biochim. Pol. 2003, 50, 985–1017. [Google Scholar] [CrossRef] [PubMed]
  25. E Brenman, J.; Chao, D.S.; Gee, S.H.; McGee, A.W.; E Craven, S.; Santillano, D.R.; Wu, Z.; Huang, F.; Xia, H.; Peters, M.F.; et al. Interaction of Nitric Oxide Synthase with the Postsynaptic Density Protein PSD-95 and Alpha1-Syntrophin Mediated by PDZ Domains. Cell 1996, 84, 757–767. [Google Scholar] [CrossRef] [PubMed]
  26. Ernst, A.; Gfeller, D.; Kan, Z.; Seshagiri, S.; Kim, P.M.; Bader, G.D.; Sidhu, S.S. Coevolution of PDZ Domain-Ligand Interactions Analyzed by High-Throughput Phage Display and Deep Sequencing. Mol. Biosyst. 2010, 6, 1782–1790. [Google Scholar] [CrossRef]
  27. Grootjans, J.J.; Reekmans, G.; Ceulemans, H.; David, G. Syntenin-Syndecan Binding Requires Syndecan-Synteny and the Co-Operation of Both PDZ Domains of Syntenin. J. Biol. Chem. 2000, 275, 19933–19941. [Google Scholar] [CrossRef]
  28. Kornau, H.-C.; Schenker, L.T.; Kennedy, M.B.; Seeburg, P.H. Domain Interaction between NMDA Receptor Subunits and the Postsynaptic Density Protein PSD-95. Science 1995, 269, 1737–1740. [Google Scholar] [CrossRef] [PubMed]
  29. Wu, H.; Feng, W.; Chen, J.; Chan, L.-N.; Huang, S.; Zhang, M. PDZ Domains of Par-3 as Potential Phosphoinositide Signaling Integrators. Mol. Cell 2007, 28, 886–898. [Google Scholar] [CrossRef]
  30. Perroy, J.; El Far, O.; Bertaso, F.; Pin, J.; Betz, H.; Bockaert, J.; Fagni, L. PICK1 Is Required for the Control of Synaptic Transmission by the Metabotropic Glutamate Receptor 7. EMBO J. 2002, 21, 2990–2999. [Google Scholar] [CrossRef]
  31. Gardiol, D. PDZ-Containing Proteins as Targets in Human Pathologies. FEBS J. 2012, 279, 3529. [Google Scholar] [CrossRef]
  32. Sheng, M.; Sala, C. PDZ Domains and the Organization of Supramolecular Complexes. Annu. Rev. Neurosci. 2001, 24, 1–29. [Google Scholar] [CrossRef] [PubMed]
  33. Tran, Y.H.; Xu, Z.; Kato, A.; Mistry, A.C.; Goya, Y.; Taira, M.; Brandt, S.J.; Hirose, S. Spliced Isoforms of LIM-Domain-Binding Protein (CLIM/NLI/Ldb) Lacking the LIM-Interaction Domain. J. Biochem. 2006, 140, 105–119. [Google Scholar] [CrossRef] [PubMed]
  34. Kadrmas, J.L.; Beckerle, M.C. The LIM Domain: From the Cytoskeleton to the Nucleus. Nat. Rev. Mol. Cell Biol. 2004, 5, 920–931. [Google Scholar] [CrossRef]
  35. Pérez-Alvarado, G.C.; Miles, C.; Michelsen, J.W.; Louis, H.A.; Winge, D.R.; Beckerle, M.C.; Summers, M.F. Structure of the Carboxy-Terminal LIM Domain from the Cysteine Rich Protein CRP. Nat. Struct. Biol. 1994, 1, 388–398. [Google Scholar] [CrossRef] [PubMed]
  36. Xu, X.; Fan, Z.; Liang, C.; Li, L.; Wang, L.; Liang, Y.; Wu, J.; Chang, S.; Yan, Z.; Lv, Z.; et al. A Signature Motif in LIM Proteins Mediates Binding to Checkpoint Proteins and Increases Tumour Radiosensitivity. Nat. Commun. 2017, 8, 14059. [Google Scholar] [CrossRef] [PubMed]
  37. Anderson, C.A.; Kovar, D.R.; Gardel, M.L.; Winkelman, J.D. LIM Domain Proteins in Cell Mechanobiology. Cytoskeleton 2021, 78, 303–311. [Google Scholar] [CrossRef]
  38. Wu, R.Y.; Gill, G.N. LIM Domain Recognition of a Tyrosine-Containing Tight Turn. J. Biol. Chem. 1994, 269, 25085–25090. [Google Scholar] [CrossRef]
  39. Dawid, I.B.; Breen, J.J.; Toyama, R. LIM Domains: Multiple Roles as Adapters and Functional Modifiers in Protein Interactions. Trends Genet. 1998, 14, 156–162. [Google Scholar] [CrossRef]
  40. Schiller, H.B.; Friedel, C.C.; Boulegue, C.; Fässler, R. Quantitative Proteomics of the Integrin Adhesome Show a Myosin II-Dependent Recruitment of LIM Domain Proteins. EMBO Rep. 2011, 12, 259–266. [Google Scholar] [CrossRef]
  41. Germain, P.; Delalande, A.; Pichon, C. Role of Muscle LIM Protein in Mechanotransduction Process. Int. J. Mol. Sci. 2022, 23, 9785. [Google Scholar] [CrossRef] [PubMed]
  42. Bouaouina, M.; Jani, K.; Long, J.Y.; Czerniecki, S.; Morse, E.M.; Ellis, S.J.; Tanentzapf, G.; Schöck, F.; Calderwood, D.A. Zasp Regulates Integrin Activation. J. Cell Sci. 2012, 125 Pt 23, 5647–5657. [Google Scholar] [CrossRef]
  43. Matthews, J.M.; Lester, K.; Joseph, S.; Curtis, D.J. LIM-Domain-Only Proteins in Cancer. Nat. Rev. Cancer 2013, 13, 111–122. [Google Scholar] [CrossRef] [PubMed]
  44. González-Morales, N.; Xiao, Y.S.; Schilling, M.A.; Marescal, O.; Liao, K.A.; Schöck, F. Myofibril Diameter Is Set by a Finely Tuned Mechanism of Protein Oligomerization in Drosophila. eLife 2019, 8, e50496. [Google Scholar] [CrossRef]
  45. She, M.; Tang, M.; Jiang, T.; Zeng, Q. The Roles of the LIM Domain Proteins in Drosophila Cardiac and Hematopoietic Morphogenesis. Front. Cardiovasc. Med. 2021, 8, 616851. [Google Scholar] [CrossRef]
  46. Jani, K.; Schöck, F. Zasp Is Required for the Assembly of Functional Integrin Adhesion Sites. J. Cell Biol. 2007, 179, 1583–1597. [Google Scholar] [CrossRef] [PubMed]
  47. Klaavuniemi, T.; Kelloniemi, A.; Ylänne, J. The ZASP-like Motif in Actinin-Associated LIM Protein Is Required for Interaction with the Alpha-Actinin Rod and for Targeting to the Muscle Z-Line. J. Biol. Chem. 2004, 279, 26402–26410. [Google Scholar] [CrossRef]
  48. Yang, N.; Higuchi, O.; Ohashi, K.; Nagata, K.; Wada, A.; Kangawa, K.; Nishida, E.; Mizuno, K. Cofilin Phosphorylation by LIM-Kinase 1 and Its Role in Rac-Mediated Actin Reorganization. Nature 1998, 393, 809–812. [Google Scholar] [CrossRef]
  49. Manetti, F. LIM Kinases Are Attractive Targets with Many Macromolecular Partners and Only a Few Small Molecule Regulators. Med. Res. Rev. 2012, 32, 968–998. [Google Scholar] [CrossRef]
  50. Yin, L.-M.; Schnoor, M.; Jun, C.-D. Structural Characteristics, Binding Partners and Related Diseases of the Calponin Homology (CH) Domain. Front. Cell Dev. Biol. 2020, 8, 342. [Google Scholar] [CrossRef]
  51. Cox, D.; Brennan, M.; Moran, N. Integrins as Therapeutic Targets: Lessons and Opportunities. Nat. Rev. Drug Discov. 2010, 9, 804–820. [Google Scholar] [CrossRef]
  52. Desgrosellier, J.S.; Cheresh, D.A. Integrins in Cancer: Biological Implications and Therapeutic Opportunities. Nat. Rev. Cancer 2010, 10, 9–22. [Google Scholar] [CrossRef]
  53. Campbell, I.D.; Humphries, M.J. Integrin Structure, Activation, and Interactions. Cold Spring Harb. Perspect. Biol. 2011, 3, a004994. [Google Scholar] [CrossRef] [PubMed]
  54. Horton, E.R.; Astudillo, P.; Humphries, M.J.; Humphries, J.D. Mechanosensitivity of Integrin Adhesion Complexes: Role of the Consensus Adhesome. Exp. Cell Res. 2016, 343, 7–13. [Google Scholar] [CrossRef] [PubMed]
  55. Ma, C.; Yao, Y.; Yue, Q.-X.; Zhou, X.-W.; Yang, P.-Y.; Wu, W.-Y.; Guan, S.-H.; Jiang, B.-H.; Yang, M.; Liu, X.; et al. Differential Proteomic Analysis of Platelets Suggested Possible Signal Cascades Network in Platelets Treated with Salvianolic Acid B. PLoS ONE 2011, 6, e14692. [Google Scholar] [CrossRef]
  56. Lv, J.; Pan, Z.; Chen, J.; Xu, R.; Wang, D.; Huang, J.; Dong, Y.; Jiang, J.; Yin, X.; Cheng, H.; et al. Phosphoproteomic Analysis Reveals Downstream PKA Effectors of AKAP Cypher/ZASP in the Pathogenesis of Dilated Cardiomyopathy. Front. Cardiovasc. Med. 2021, 8, 753072. [Google Scholar] [CrossRef] [PubMed]
  57. Cox, O.T.; O’shea, S.; Tresse, E.; Bustamante-Garrido, M.; Kiran-Deevi, R.; O’connor, R. iGF-1 Receptor and Adhesion Signaling: An Important Axis in Determining Cancer Cell Phenotype and Therapy Resistance. Front. Endocrinol. 2015, 6, 106. [Google Scholar] [CrossRef] [PubMed]
  58. Elbediwy, A.; Vanyai, H.; Diaz-De-La-Loza, M.-D.; Frith, D.; Snijders, A.P.; Thompson, B.J. Enigma Proteins Regulate YAP Mechanotransduction. J. Cell Sci. 2018, 131, jcs221788. [Google Scholar] [CrossRef]
  59. Wozniak, M.A.; Baker, B.M.; Chen, C.S.; Wilson, K.L. The Emerin-Binding Transcription Factor Lmo7 Is Regulated by Association with p130Cas at Focal Adhesions. PeerJ 2013, 1, e134. [Google Scholar] [CrossRef]
  60. Chen, M.; Sinha, M.; Luxon, B.A.; Bresnick, A.R.; O’Connor, K.L. Integrin Alpha6beta4 Controls the Expression of Genes Associated with Cell Motility, Invasion, and Metastasis, Including S100A4/Metastasin. J. Biol. Chem. 2009, 284, 1484–1494. [Google Scholar] [CrossRef]
  61. Shibue, T.; Brooks, M.W.; Weinberg, R.A. An Integrin-Linked Machinery of Cytoskeletal Regulation That Enables Experimental Tumor Initiation and Metastatic Colonization. Cancer Cell 2013, 24, 481–498. [Google Scholar] [CrossRef] [PubMed]
  62. Lui, W.-Y.; Lee, W.M.; Cheng, C.Y. Sertoli-Germ Cell Adherens Junction Dynamics in the Testis Are Regulated by RhoB GTPase via the ROCK/LIMK Signaling Pathway. Biol. Reprod. 2003, 68, 2189–2206. [Google Scholar] [CrossRef] [PubMed]
  63. Loubet, D.; Dakowski, C.; Pietri, M.; Pradines, E.; Bernard, S.; Callebert, J.; Kellermann, O.; Schneider, B.; Ardila-Osorio, H.; Mouillet-Richard, S.; et al. Neuritogenesis: The Prion Protein Controls Β1 Integrin Signaling Activity. FASEB J. 2012, 26, 678–690. [Google Scholar] [CrossRef]
  64. Peng, D.; Fu, M.; Wang, M.; Wei, Y.; Wei, X. Targeting TGF-β Signal Transduction for Fibrosis and Cancer Therapy. Mol. Cancer 2022, 21, 104. [Google Scholar] [CrossRef] [PubMed]
  65. Morin, P.; Wickman, G.; Munro, J.; Inman, G.J.; Olson, M.F. Differing Contributions of LIMK and ROCK to TGFβ-Induced Transcription, Motility and Invasion. Eur. J. Cell Biol. 2011, 90, 13–25. [Google Scholar] [CrossRef]
  66. Lee, J.; Ko, M.; Joo, C.-K. Rho Plays a Key Role in TGF-Beta1-Induced Cytoskeletal Rearrangement in Human Retinal Pigment Epithelium. J. Cell Physiol. 2008, 216, 520–526. [Google Scholar] [CrossRef]
  67. Lai, Y.-J.; Tsai, F.-C.; Chang, G.-J.; Chang, S.-H.; Huang, C.-C.; Chen, W.-J.; Yeh, Y.-H. miR-181b Targets Semaphorin 3A to Mediate TGF-β-Induced Endothelial-Mesenchymal Transition Related to Atrial Fibrillation. J. Clin. Investig. 2022, 132, e142548. [Google Scholar] [CrossRef]
  68. Shi, Y.; Wang, X.; Xu, Z.; He, Y.; Guo, C.; He, L.; Huan, C.; Cai, C.; Huang, J.; Zhang, J.; et al. PDLIM5 Inhibits STUB1-Mediated Degradation of SMAD3 and Promotes the Migration and Invasion of Lung Cancer Cells. J. Biol. Chem. 2020, 295, 13798–13811. [Google Scholar] [CrossRef]
  69. Cheng, H.; Chen, T.; Tor, M.; Park, D.; Zhou, Q.; Huang, J.B.; Khatib, N.; Rong, L.; Zhou, G. A High-Throughput Screening Platform Targeting PDLIM5 for Pulmonary Hypertension. J. Biomol. Screen. 2016, 21, 333–341. [Google Scholar] [CrossRef]
  70. Xie, Y.; Ostriker, A.C.; Jin, Y.; Hu, H.; Sizer, A.J.; Peng, G.; Morris, A.H.; Ryu, C.; Herzog, E.L.; Kyriakides, T.; et al. LMO7 Is a Negative Feedback Regulator of Transforming Growth Factor β Signaling and Fibrosis. Circulation 2019, 139, 679–693. [Google Scholar] [CrossRef]
  71. Nakamura, H.; Mukai, M.; Komatsu, K.; Tanaka-Okamoto, M.; Itoh, Y.; Ishizaki, H.; Tatsuta, M.; Inoue, M.; Miyoshi, J. Transforming Growth Factor-Beta1 Induces LMO7 While Enhancing the Invasiveness of Rat Ascites Hepatoma Cells. Cancer Lett. 2005, 220, 95–99. [Google Scholar] [CrossRef] [PubMed]
  72. Zhang, Q.; Wang, X.; Chen, Z.; Liu, G.; Chen, Z. Semi-Quantitative RT-PCR Analysis of LIM Mineralization Protein 1 and Its Associated Molecules in Cultured Human Dental Pulp Cells. Arch. Oral. Biol. 2007, 52, 720–726. [Google Scholar] [CrossRef]
  73. Minamide, A.; Boden, S.D.; Viggeswarapu, M.; Hair, G.A.; Oliver, C.; Titus, L. Mechanism of Bone Formation with Gene Transfer of the cDNA Encoding for the Intracellular Protein LMP-1. J. Bone Jt. Surg. Am. 2003, 85, 1030–1039. [Google Scholar] [CrossRef]
  74. Ma, J.; Guo, W.; Gao, M.; Huang, B.; Qi, Q.; Ling, Z.; Chen, Y.; Hu, H.; Zhou, H.; Yu, F.; et al. Biomimetic Matrix Fabricated by LMP-1 Gene-Transduced MC3T3-E1 Cells for Bone Regeneration. Biofabrication 2017, 9, 045010. [Google Scholar] [CrossRef]
  75. Boden, S.D. Biology of Lumbar Spine Fusion and Use of Bone Graft Substitutes: Present, Future, and next Generation. Tissue Eng. 2000, 6, 383–399. [Google Scholar] [CrossRef]
  76. Jiang, X.; Chen, Y.; Fan, X.; Zhang, H.; Kun, L. Osteogenesis and Mineralization in a Rabbit Mandibular Distraction Osteogenesis Model Is Promoted by the Human LMP-1 Gene. J. Orthop. Res. 2015, 33, 521–526. [Google Scholar] [CrossRef]
  77. Yoon, S.T.; Park, J.S.; Kim, K.S.; Li, J.; Attallah-Wasif, E.S.; Hutton, W.C.; Boden, S.D. ISSLS Prize Winner: LMP-1 Upregulates Intervertebral Disc Cell Production of Proteoglycans and BMPs in Vitro and in Vivo. Spine (Phila. Pa. 1976) 2004, 29, 2603–2611. [Google Scholar] [CrossRef]
  78. Park, J.S.; Nagata, K. [BMP and LMP-1 for intervertebral disc regeneration]. Clin. Calcium 2004, 14, 76–78. [Google Scholar] [PubMed]
  79. Mu, R.; Chen, B.; Bi, B.; Yu, H.; Liu, J.; Li, J.; He, M.; Rong, L.; Liu, B.; Liu, K.; et al. LIM Mineralization Protein-1 Enhances the Committed Differentiation of Dental Pulp Stem Cells through the ERK1/2 and P38 MAPK Pathways and BMP Signaling. Int. J. Med. Sci. 2022, 19, 1307–1319. [Google Scholar] [CrossRef]
  80. Matsuura, I.; Endo, M.; Hata, K.; Kubo, T.; Yamaguchi, A.; Saeki, N.; Yamashita, T. BMP Inhibits Neurite Growth by a Mechanism Dependent on LIM-Kinase. Biochem. Biophys. Res. Commun. 2007, 360, 868–873. [Google Scholar] [CrossRef] [PubMed]
  81. Karin, M.; Ben-Neriah, Y. Phosphorylation Meets Ubiquitination: The Control of NF- [Kappa]B Activity. Annu. Rev. Immunol. 2000, 18, 621–663. [Google Scholar] [CrossRef] [PubMed]
  82. Ono, R.; Kaisho, T.; Tanaka, T. PDLIM1 Inhibits NF-κB-Mediated Inflammatory Signaling by Sequestering the P65 Subunit of NF-κB in the Cytoplasm. Sci. Rep. 2015, 5, 18327. [Google Scholar] [CrossRef] [PubMed]
  83. Tanaka, T.; Grusby, M.J.; Kaisho, T. PDLIM2-Mediated Termination of Transcription Factor NF-kappaB Activation by Intranuclear Sequestration and Degradation of the P65 Subunit. Nat. Immunol. 2007, 8, 584–591. [Google Scholar] [CrossRef]
  84. Jodo, A.; Shibazaki, A.; Onuma, A.; Kaisho, T.; Tanaka, T. PDLIM7 Synergizes With PDLIM2 and P62/Sqstm1 to Inhibit Inflammatory Signaling by Promoting Degradation of the P65 Subunit of NF-κB. Front. Immunol. 2020, 11, 1559. [Google Scholar] [CrossRef]
  85. Guo, Q.; Xu, J.; Shi, Q.; Wu, S. PDLIM2 Protects Articular Chondrocytes from Lipopolysaccharide-Induced Apoptosis, Degeneration and Inflammatory Injury through down-Regulation of Nuclear Factor (NF)-κB Signaling. Int. Immunopharmacol. 2020, 88, 106883. [Google Scholar] [CrossRef] [PubMed]
  86. Hao, Y.-R.; Tang, F.-J.; Zhang, X.; Wang, H. Suppression of NF-κB Activation by PDLIM2 Restrains Hepatic Lipogenesis and Inflammation in High Fat Diet Induced Mice. Biochem. Biophys. Res. Commun. 2018, 503, 564–571. [Google Scholar] [CrossRef]
  87. Sun, F.; Li, L.; Yan, P.; Zhou, J.; Shapiro, S.D.; Xiao, G.; Qu, Z. Causative Role of PDLIM2 Epigenetic Repression in Lung Cancer and Therapeutic Resistance. Nat. Commun. 2019, 10, 5324. [Google Scholar] [CrossRef] [PubMed]
  88. Sun, F.; Xiao, Y.; Qu, Z. Oncovirus Kaposi Sarcoma Herpesvirus (KSHV) Represses Tumor Suppressor PDLIM2 to Persistently Activate Nuclear Factor κB (NF-κB) and STAT3 Transcription Factors for Tumorigenesis and Tumor Maintenance. J. Biol. Chem. 2015, 290, 7362–7368. [Google Scholar] [CrossRef]
  89. Fang, J.Y.; Richardson, B.C. The MAPK Signalling Pathways and Colorectal Cancer. Lancet Oncol. 2005, 6, 322–327. [Google Scholar] [CrossRef]
  90. He, H.; Li, W.; Yan, P.; Bundschuh, R.; Killian, J.A.; Labanowska, J.; Brock, P.; Shen, R.; Heerema, N.A.; de la Chapelle, A. Identification of a Recurrent LMO7-BRAF Fusion in Papillary Thyroid Carcinoma. Thyroid 2018, 28, 748–754. [Google Scholar] [CrossRef]
  91. Kang, M.; Lee, K.-H.; Lee, H.S.; Park, Y.H.; Jeong, C.W.; Ku, J.H.; Kim, H.H.; Kwak, C. PDLIM2 Suppression Efficiently Reduces Tumor Growth and Invasiveness of Human Castration-Resistant Prostate Cancer-like Cells. Prostate 2016, 76, 273–285. [Google Scholar] [CrossRef] [PubMed]
  92. Yin, H.; Zhao, J.; He, H.; Chen, Y.; Wang, Y.; Li, D.; Zhu, Q. Gga-miR-3525 Targets PDLIM3 through the MAPK Signaling Pathway to Regulate the Proliferation and Differentiation of Skeletal Muscle Satellite Cells. Int. J. Mol. Sci. 2020, 21, 5573. [Google Scholar] [CrossRef] [PubMed]
  93. He, H.; Yin, H.; Yu, X.; Zhang, Y.; Ma, M.; Li, D.; Zhu, Q. PDLIM5 Affects Chicken Skeletal Muscle Satellite Cell Proliferation and Differentiation via the P38-MAPK Pathway. Animals 2021, 11, 1016. [Google Scholar] [CrossRef] [PubMed]
  94. Yoo, J.-Y.; Jung, N.-C.; Lee, J.-H.; Choi, S.-Y.; Choi, H.-J.; Park, S.-Y.; Jang, J.-S.; Byun, S.-H.; Hwang, S.-U.; Noh, K.-E.; et al. Pdlim4 Is Essential for CCR7-JNK-Mediated Dendritic Cell Migration and F-Actin-Related Dendrite Formation. FASEB J. 2019, 33, 11035–11044. [Google Scholar] [CrossRef]
  95. Xuan, T.; Wang, D.; Lv, J.; Pan, Z.; Fang, J.; Xiang, Y.; Cheng, H.; Wang, X.; Guo, X. Downregulation of Cypher Induces Apoptosis in Cardiomyocytes via Akt/P38 MAPK Signaling Pathway. Int. J. Med. Sci. 2020, 17, 2328–2337. [Google Scholar] [CrossRef]
  96. Bongalon, S.; Dai, Y.-P.; Singer, C.A.; Yamboliev, I.A. PDGF and IL-1beta Upregulate Cofilin and LIMK2 in Canine Cultured Pulmonary Artery Smooth Muscle Cells. J. Vasc. Res. 2004, 41, 412–421. [Google Scholar] [CrossRef]
  97. Zhou, J.-K.; Fan, X.; Cheng, J.; Liu, W.; Peng, Y. PDLIM1: Structure, Function and Implication in Cancer. Cell Stress 2021, 5, 119–127. [Google Scholar] [CrossRef]
  98. Ahn, B.Y.; Saldanha-Gama, R.F.G.; Rahn, J.J.; Hao, X.; Zhang, J.; Dang, N.-H.; Alshehri, M.; Robbins, S.M.; Senger, D.L. Glioma Invasion Mediated by the P75 Neurotrophin Receptor (P75(NTR)/CD271) Requires Regulated Interaction with PDLIM1. Oncogene 2016, 35, 1411–1422. [Google Scholar] [CrossRef]
  99. Liu, Z.; Zhan, Y.; Tu, Y.; Chen, K.; Wu, C. PDZ and LIM Domain Protein 1(PDLIM1)/CLP36 Promotes Breast Cancer Cell Migration, Invasion and Metastasis through Interaction with α-Actinin. Oncogene 2015, 34, 1300–1311. [Google Scholar] [CrossRef]
  100. Gupta, P.; Suman, S.; Mishra, M.; Mishra, S.; Srivastava, N.; Kumar, V.; Singh, P.K.; Shukla, Y. Autoantibodies against TYMS and PDLIM1 Proteins Detected as Circulatory Signatures in Indian Breast Cancer Patients. Proteom. Clin. Appl. 2016, 10, 564–573. [Google Scholar] [CrossRef]
  101. Huang, Z.; Zhou, J.; Wang, K.; Chen, H.; Qin, S.; Liu, J.; Luo, M.; Chen, Y.; Jiang, J.; Zhou, L.; et al. PDLIM1 Inhibits Tumor Metastasis Through Activating Hippo Signaling in Hepatocellular Carcinoma. Hepatology 2020, 71, 1643–1659. [Google Scholar] [CrossRef]
  102. Tan, Y.; Li, Y.; Zhu, H.; Wu, X.; Mei, K.; Li, P.; Yang, Q. miR-187/PDLIM1 Gets Involved in Gastric Cancer Progression and Cisplatin Sensitivity of Cisplatin by Mediating the Hippo-YAP Signaling Pathway. J. Oncol. 2022, 2022, 5456016. [Google Scholar] [CrossRef] [PubMed]
  103. Chen, H.-N.; Yuan, K.; Xie, N.; Wang, K.; Huang, Z.; Chen, Y.; Dou, Q.; Wu, M.; Nice, E.C.; Zhou, Z.-G.; et al. PDLIM1 Stabilizes the E-Cadherin/β-Catenin Complex to Prevent Epithelial-Mesenchymal Transition and Metastatic Potential of Colorectal Cancer Cells. Cancer Res. 2016, 76, 1122–1134. [Google Scholar] [CrossRef] [PubMed]
  104. Tamura, N.; Ohno, K.; Katayama, T.; Kanayama, N.; Sato, K. The PDZ-LIM Protein CLP36 Is Required for Actin Stress Fiber Formation and Focal Adhesion Assembly in BeWo Cells. Biochem. Biophys. Res. Commun. 2007, 364, 589–594. [Google Scholar] [CrossRef]
  105. Hong, S.-H. Identification of CLP36 as a Tumor Antigen That Induces an Antibody Response in Pancreatic Cancer. Cancer Res. Treat. 2005, 37, 71–77. [Google Scholar] [CrossRef]
  106. Macartney-Coxson, D.P.; A Hood, K.; Shi, H.-J.; Ward, T.; Wiles, A.; O’Connor, R.; A Hall, D.; A Lea, R.; A Royds, J.; Stubbs, R.S.; et al. Metastatic Susceptibility Locus, an 8p Hot-Spot for Tumour Progression Disrupted in Colorectal Liver Metastases: 13 Candidate Genes Examined at the DNA, mRNA and Protein Level. BMC Cancer 2008, 8, 187. [Google Scholar] [CrossRef] [PubMed]
  107. Ding, J.; Xu, Z.; Zhang, Y.; Tan, C.; Hu, W.; Wang, M.; Xu, Y.; Tang, J. Exosome-Mediated miR-222 Transferring: An Insight into NF-κB-Mediated Breast Cancer Metastasis. Exp. Cell Res. 2018, 369, 129–138. [Google Scholar] [CrossRef] [PubMed]
  108. Qu, Z.; Fu, J.; Yan, P.; Hu, J.; Cheng, S.-Y.; Xiao, G. Epigenetic Repression of PDZ-LIM Domain-Containing Protein 2: Implications for the Biology and Treatment of Breast Cancer. J. Biol. Chem. 2010, 285, 11786–11792. [Google Scholar] [CrossRef] [PubMed]
  109. Deevi, R.K.; Cox, O.T.; O’Connor, R. Essential Function for PDLIM2 in Cell Polarization in Three-Dimensional Cultures by Feedback Regulation of the Β1-Integrin-RhoA Signaling Axis. Neoplasia 2014, 16, 422–431. [Google Scholar] [CrossRef]
  110. Bowe, R.A.; Cox, O.T.; Ayllón, V.; Tresse, E.; Healy, N.C.; Edmunds, S.J.; Huigsloot, M.; O’Connor, R. PDLIM2 Regulates Transcription Factor Activity in Epithelial-to-Mesenchymal Transition via the COP9 Signalosome. Mol. Biol. Cell 2014, 25, 184–195. [Google Scholar] [CrossRef]
  111. Shi, H.; Ji, Y.; Li, W.; Zhong, Y.; Ming, Z. PDLIM2 Acts as a Cancer Suppressor Gene in Non-Small Cell Lung Cancer via the down Regulation of NF-κB Signaling. Mol. Cell. Probes 2020, 53, 101628. [Google Scholar] [CrossRef]
  112. Song, G.; Xu, J.; He, L.; Sun, X.; Xiong, R.; Luo, Y.; Hu, X.; Zhang, R.; Yue, Q.; Liu, K.; et al. Systematic Profiling Identifies PDLIM2 as a Novel Prognostic Predictor for Oesophageal Squamous Cell Carcinoma (ESCC). J. Cell. Mol. Med. 2019, 23, 5751–5761. [Google Scholar] [CrossRef]
  113. Jiang, X.; Chu, Z.; Cao, Y.; Tang, Y.; Shi, Y.; Shi, X. PDLIM2 Prevents the Malignant Phenotype of Hepatocellular Carcinoma Cells by Negatively Regulating β-Catenin. Cancer Gene Ther. 2021, 28, 1113–1124. [Google Scholar] [CrossRef] [PubMed]
  114. Wang, P.; Li, G.-Y.; Zhou, L.; Jiang, H.-L.; Yang, Y.; Wu, H.-T. Exosomes from M2 Macrophages Promoted Glycolysis in FaDu Cells by Inhibiting PDLIM2 Expression to Stabilize PFKL. Neoplasma 2022, 69, 1041–1053. [Google Scholar] [CrossRef] [PubMed]
  115. Zhao, L.; Yu, C.; Zhou, S.; Lau, W.B.; Lau, B.; Luo, Z.; Lin, Q.; Yang, H.; Xuan, Y.; Yi, T.; et al. Epigenetic Repression of PDZ-LIM Domain-Containing Protein 2 Promotes Ovarian Cancer via NOS2-Derived Nitric Oxide Signaling. Oncotarget 2016, 7, 1408–1420. [Google Scholar] [CrossRef]
  116. Oh, B.Y.; Cho, J.; Hong, H.K.; Bae, J.S.; Park, W.-Y.; Joung, J.-G.; Cho, Y.B. Exome and Transcriptome Sequencing Identifies Loss of PDLIM2 in Metastatic Colorectal Cancers. Cancer Manag. Res. 2017, 9, 581–589. [Google Scholar] [CrossRef] [PubMed]
  117. Guo, X.; Yang, Z.; Zhi, Q.; Wang, D.; Guo, L.; Li, G.; Miao, R.; Shi, Y.; Kuang, Y. Long Noncoding RNA OR3A4 Promotes Metastasis and Tumorigenicity in Gastric Cancer. Oncotarget 2016, 7, 30276–30294. [Google Scholar] [CrossRef]
  118. Shou, Y.; Robinson, D.M.; Amakye, D.D.; Rose, K.L.; Cho, Y.-J.; Ligon, K.L.; Sharp, T.; Haider, A.S.; Bandaru, R.; Ando, Y.; et al. A Five-Gene Hedgehog Signature Developed as a Patient Preselection Tool for Hedgehog Inhibitor Therapy in Medulloblastoma. Clin. Cancer Res. 2015, 21, 585–593. [Google Scholar] [CrossRef]
  119. Feng, Y.; Jiang, Y.; Wen, T.; Meng, F.; Shu, X. Identifying Potential Prognostic Markers for Muscle-Invasive Bladder Urothelial Carcinoma by Weighted Gene Co-Expression Network Analysis. Pathol. Oncol. Res. 2020, 26, 1063–1072. [Google Scholar] [CrossRef]
  120. Lu, Y.; Liu, P.; Wen, W.; Grubbs, C.J.; Townsend, R.R.; Malone, J.P.; Lubet, R.A.; You, M. Cross-Species Comparison of Orthologous Gene Expression in Human Bladder Cancer and Carcinogen-Induced Rodent Models. Am. J. Transl. Res. 2010, 3, 8–27. [Google Scholar]
  121. Stein, L.; Rothschild, J.; Luce, J.; Cowell, J.K.; Thomas, G.A.; Bogdanova, T.I.; Tronko, M.D.; Hawthorn, L.; Richter, H.; Braselmann, H.; et al. Copy Number and Gene Expression Alterations in Radiation-Induced Papillary Thyroid Carcinoma from Chernobyl Pediatric Patients. Thyroid 2010, 20, 475–487. [Google Scholar] [CrossRef] [PubMed]
  122. Jia, Y.; Shi, H.; Cao, Y.; Feng, W.; Li, M.; Li, X. PDZ and LIM Domain Protein 4 Suppresses the Growth and Invasion of Ovarian Cancer Cells via Inactivation of STAT3 Signaling. Life Sci. 2019, 233, 116715. [Google Scholar] [CrossRef] [PubMed]
  123. Vanaja, D.K.; Grossmann, M.E.; Cheville, J.C.; Gazi, M.H.; Gong, A.; Zhang, J.S.; Ajtai, K.; Burghardt, T.P.; Young, C.Y.F. Pdlim4, an Actin Binding Protein, Suppresses Prostate Cancer Cell Growth. Cancer Investig. 2009, 27, 264–272. [Google Scholar] [CrossRef] [PubMed]
  124. Vanaja, D.K.; Ballman, K.V.; Morlan, B.W.; Cheville, J.C.; Neumann, R.M.; Lieber, M.M.; Tindall, D.J.; Young, C.Y.F. PDLIM4 Repression by Hypermethylation as a Potential Biomarker for Prostate Cancer. Clin. Cancer Res. 2006, 12, 1128–1136. [Google Scholar] [CrossRef] [PubMed]
  125. Kolluru, V.; Tyagi, A.; Chandrasekaran, B.; Damodaran, C. Profiling of Differentially Expressed Genes in Cadmium-Induced Prostate Carcinogenesis. Toxicol. Appl. Pharmacol. 2019, 375, 57–63. [Google Scholar] [CrossRef] [PubMed]
  126. Vasiljević, N.; Wu, K.; Brentnall, A.R.; Kim, D.C.; Thorat, M.A.; Kudahetti, S.C.; Mao, X.; Xue, L.; Yu, Y.; Shaw, G.L.; et al. Absolute Quantitation of DNA Methylation of 28 Candidate Genes in Prostate Cancer Using Pyrosequencing. Dis. Markers 2011, 30, 151–161. [Google Scholar] [CrossRef] [PubMed]
  127. Morris, M.R.; Ricketts, C.; Gentle, D.; Abdulrahman, M.; Clarke, N.; Brown, M.; Kishida, T.; Yao, M.; Latif, F.; Maher, E.R. Identification of Candidate Tumour Suppressor Genes Frequently Methylated in Renal Cell Carcinoma. Oncogene 2010, 29, 2104–2117. [Google Scholar] [CrossRef]
  128. Patai, V.; Barták, B.K.; Péterfia, B.; Micsik, T.; Horváth, R.; Sumánszki, C.; Péter, Z.; Valcz, G.; Kalmár, A.; Tóth, K.; et al. Comprehensive DNA Methylation and Mutation Analyses Reveal a Methylation Signature in Colorectal Sessile Serrated Adenomas. Pathol. Oncol. Res. 2017, 23, 589–594. [Google Scholar] [CrossRef]
  129. Feng, W.; Orlandi, R.; Zhao, N.; Carcangiu, M.L.; Tagliabue, E.; Xu, J.; Bast, R.C.; Yu, Y. Tumor Suppressor Genes Are Frequently Methylated in Lymph Node Metastases of Breast Cancers. BMC Cancer 2010, 10, 378. [Google Scholar] [CrossRef]
  130. Xu, J.; Shetty, P.B.; Feng, W.; Chenault, C.; Bast, R.C.; Issa, J.-P.J.; Hilsenbeck, S.G.; Yu, Y. Methylation of HIN-1, RASSF1A, RIL and CDH13 in Breast Cancer Is Associated with Clinical Characteristics, but Only RASSF1A Methylation Is Associated with Outcome. BMC Cancer 2012, 12, 243. [Google Scholar] [CrossRef]
  131. Kravchenko, D.S.; Ivanova, A.E.; Podshivalova, E.S.; Chumakov, S.P. PDLIM4/RIL-Mediated Regulation of Src and Malignant Properties of Breast Cancer Cells. Oncotarget 2020, 11, 22–30. [Google Scholar] [CrossRef]
  132. Liu, X.; Chen, L.; Huang, H.; Lv, J.-M.; Chen, M.; Qu, F.-J.; Pan, X.-W.; Li, L.; Yin, L.; Cui, X.-G.; et al. High Expression of PDLIM5 Facilitates Cell Tumorigenesis and Migration by Maintaining AMPK Activation in Prostate Cancer. Oncotarget 2017, 8, 98117–98134. [Google Scholar] [CrossRef]
  133. Shui, I.M.; Lindström, S.; Kibel, A.S.; Berndt, S.I.; Campa, D.; Gerke, T.; Penney, K.L.; Albanes, D.; Berg, C.; Bueno-De-Mesquita, H.B.; et al. Prostate Cancer (PCa) Risk Variants and Risk of Fatal PCa in the National Cancer Institute Breast and Prostate Cancer Cohort Consortium. Eur. Urol. 2014, 65, 1069–1075. [Google Scholar] [CrossRef] [PubMed]
  134. Yan, Y.; Tsukamoto, O.; Nakano, A.; Kato, H.; Kioka, H.; Ito, N.; Higo, S.; Yamazaki, S.; Shintani, Y.; Matsuoka, K.; et al. Augmented AMPK Activity Inhibits Cell Migration by Phosphorylating the Novel Substrate Pdlim5. Nat. Commun. 2015, 6, 6137. [Google Scholar] [CrossRef] [PubMed]
  135. Wei, X.; Zhang, Y.; Yu, S.; Li, S.; Jiang, W.; Zhu, Y.; Xu, Y.; Yang, C.; Tian, G.; Mi, J.; et al. PDLIM5 Identified by Label-Free Quantitative Proteomics as a Potential Novel Biomarker of Papillary Thyroid Carcinoma. Biochem. Biophys. Res. Commun. 2018, 499, 338–344. [Google Scholar] [CrossRef] [PubMed]
  136. Zhang, Y.; Lv, W.; Li, H.; Dong, T.; Wu, H.; Su, C.; Shu, H.; Nie, F. Exploring the Relationship between Autophagy and Gefitinib Resistance in NSCLC by Silencing PDLIM5 Using Ultrasound-Targeted Microbubble Destruction Technology. Cancer Cell Int. 2022, 22, 293. [Google Scholar] [CrossRef]
  137. Edlund, K.; Lindskog, C.; Saito, A.; Berglund, A.; Pontén, F.; Göransson-Kultima, H.; Isaksson, A.; Jirström, K.; Planck, M.; Johansson, L.; et al. CD99 Is a Novel Prognostic Stromal Marker in Non-Small Cell Lung Cancer. Int. J. Cancer 2012, 131, 2264–2273. [Google Scholar] [CrossRef] [PubMed]
  138. Wu, H.; Lv, W.-H.; Zhu, Y.-Y.; Jia, Y.-Y.; Nie, F. Ultrasound-Mediated Mesoporous Silica Nanoparticles Loaded with PDLIM5 siRNA Inhibit Gefitinib Resistance in NSCLC Cells by Attenuating EMT. Eur. J. Pharm. Sci. 2023, 182, 106372. [Google Scholar] [CrossRef] [PubMed]
  139. Yamashita, Y.; Matsuura, T.; Kurosaki, T.; Amakusa, Y.; Kinoshita, M.; Ibi, T.; Sahashi, K.; Ohno, K. LDB3 Splicing Abnormalities Are Specific to Skeletal Muscles of Patients with Myotonic Dystrophy Type 1 and Alter Its PKC Binding Affinity. Neurobiol. Dis. 2014, 69, 200–205. [Google Scholar] [CrossRef]
  140. Yu, H.; Yuan, C.; Westenbroek, R.E.; Catterall, W.A. The AKAP Cypher/Zasp Contributes to β-Adrenergic/PKA Stimulation of Cardiac CaV1.2 Calcium Channels. J. Gen. Physiol. 2018, 150, 883–889. [Google Scholar] [CrossRef]
  141. Leung, M.-C.; Hitchen, P.G.; Ward, D.G.; Messer, A.E.; Marston, S.B. Z-Band Alternatively Spliced PDZ Motif Protein (ZASP) Is the Major O-Linked β-N-Acetylglucosamine-Substituted Protein in Human Heart Myofibrils. J. Biol. Chem. 2013, 288, 4891–4898. [Google Scholar] [CrossRef]
  142. Cui, L.; Cheng, Z.; Hu, K.; Pang, Y.; Liu, Y.; Qian, T.; Quan, L.; Dai, Y.; Pang, Y.; Ye, X.; et al. Prognostic Value of the PDLIM Family in Acute Myeloid Leukemia. Am. J. Transl. Res. 2019, 11, 6124–6131. [Google Scholar] [PubMed]
  143. Kales, S.C.; Nau, M.M.; Merchant, A.S.; Lipkowitz, S. Enigma Prevents Cbl-c-Mediated Ubiquitination and Degradation of RETMEN2A. PLoS ONE 2014, 9, e87116. [Google Scholar] [CrossRef]
  144. Tabariès, S.; McNulty, A.; Ouellet, V.; Annis, M.G.; Dessureault, M.; Vinette, M.; Hachem, Y.; Lavoie, B.; Omeroglu, A.; Simon, H.-G.; et al. Afadin Cooperates with Claudin-2 to Promote Breast Cancer Metastasis. Genes Dev. 2019, 33, 180–193. [Google Scholar] [CrossRef] [PubMed]
  145. Jung, C.-R.; Lim, J.H.; Choi, Y.; Kim, D.-G.; Kang, K.J.; Noh, S.-M.; Im, D.-S. Enigma Negatively Regulates P53 through MDM2 and Promotes Tumor Cell Survival in Mice. J. Clin. Investig. 2010, 120, 4493–4506. [Google Scholar] [CrossRef]
  146. Firek, A.A.; Perez, M.C.; Gonda, A.; Lei, L.; Munir, I.; Simental, A.A.; Carr, F.E.; Becerra, B.J.; De Leon, M.; Khan, S. Pathologic Significance of a Novel Oncoprotein in Thyroid Cancer Progression. Head Neck 2017, 39, 2459–2469. [Google Scholar] [CrossRef] [PubMed]
  147. Borrello, M.G.; Mercalli, E.; Perego, C.; Degl’Innocenti, D.; Ghizzoni, S.; Arighi, E.; Eroini, B.; Rizzetti, M.G.; A Pierotti, M. Differential Interaction of Enigma Protein with the Two RET Isoforms. Biochem. Biophys. Res. Commun. 2002, 296, 515–522. [Google Scholar] [CrossRef]
  148. Kim, Y.J.; Hwang, H.-J.; Kang, J.G.; Kim, C.S.; Ihm, S.-H.; Choi, M.G.; Lee, S.J. Enigma Plays Roles in Survival of Thyroid Carcinoma Cells through PI3K/AKT Signaling and Survivin. Anticancer Res. 2018, 38, 3515–3525. [Google Scholar] [CrossRef]
  149. Lu, Y.; Jin, Z.; Hou, J.; Wu, X.; Yu, Z.; Yao, L.; Pan, T.; Chang, X.; Yu, B.; Li, J.; et al. Calponin 1 Increases Cancer-Associated Fibroblasts-Mediated Matrix Stiffness to Promote Chemoresistance in Gastric Cancer. Matrix Biol. 2023, 115, 1–15. [Google Scholar] [CrossRef]
  150. Liu, H.; Huang, L.; Zhang, Z.; Zhang, Z.; Yu, Z.; Chen, X.; Chen, Z.; Zen, Y.; Yang, D.; Han, Z.; et al. LIM Mineralization Protein-1 Inhibits the Malignant Phenotypes of Human Osteosarcoma Cells. Int. J. Mol. Sci. 2014, 15, 7037–7048. [Google Scholar] [CrossRef]
  151. Hu, Q.; Guo, C.; Li, Y.; Aronow, B.J.; Zhang, J. LMO7 Mediates Cell-Specific Activation of the Rho-Myocardin-Related Transcription Factor-Serum Response Factor Pathway and Plays an Important Role in Breast Cancer Cell Migration. Mol. Cell. Biol. 2011, 31, 3223–3240. [Google Scholar] [CrossRef]
  152. Nakamura, H.; Hori, K.; Tanaka-Okamoto, M.; Higashiyama, M.; Itoh, Y.; Inoue, M.; Morinaka, S.; Miyoshi, J. Decreased Expression of LMO7 and Its Clinicopathological Significance in Human Lung Adenocarcinoma. Exp. Ther. Med. 2011, 2, 1053–1057. [Google Scholar] [CrossRef]
  153. Tanaka-Okamoto, M.; Hori, K.; Ishizaki, H.; Hosoi, A.; Itoh, Y.; Wei, M.; Wanibuchi, H.; Mizoguchi, A.; Nakamura, H.; Miyoshi, J. Increased Susceptibility to Spontaneous Lung Cancer in Mice Lacking LIM-Domain Only 7. Cancer Sci. 2009, 100, 608–616. [Google Scholar] [CrossRef] [PubMed]
  154. Wu, H.; Zhou, J.; Mei, S.; Wu, D.; Mu, Z.; Chen, B.; Xie, Y.; Ye, Y.; Liu, J. Circulating Exosomal microRNA-96 Promotes Cell Proliferation, Migration and Drug Resistance by Targeting LMO7. J. Cell. Mol. Med. 2017, 21, 1228–1236. [Google Scholar] [CrossRef] [PubMed]
  155. Yang, Y.; Zheng, H.; Li, Z.; Shi, S.; Zhong, L.; Gong, L.; Lan, B. LMO7-ALK Fusion in a Lung Adenocarcinoma Patient With Crizotinib: A Case Report. Front. Oncol. 2022, 12, 841493. [Google Scholar] [CrossRef]
  156. Mixed Responses to First-Line Alectinib in Non-Small Cell Lung Cancer Patients with Rare ALK Gene Fusions: A Case Series and Literature Review—PubMed. Available online: https://pubmed.ncbi.nlm.nih.gov/34541785/ (accessed on 4 May 2023).
  157. Tzeng, Y.-W.; Li, D.-Y.; Chen, Y.; Yang, C.-H.; Chang, C.-Y.; Juang, Y.-L. LMO7 Exerts an Effect on Mitosis Progression and the Spindle Assembly Checkpoint. Int. J. Biochem. Cell Biol. 2018, 94, 22–30. [Google Scholar] [CrossRef]
  158. Liu, X.; Yuan, H.; Zhou, J.; Wang, Q.; Qi, X.; Bernal, C.; Avella, D.; Kaifi, J.T.; Kimchi, E.T.; Timothy, P.; et al. LMO7 as an Unrecognized Factor Promoting Pancreatic Cancer Progression and Metastasis. Front. Cell Dev. Biol. 2021, 9, 647387. [Google Scholar] [CrossRef]
  159. Takahashi, M.; Lio, C.-W.J.; Campeau, A.; Steger, M.; Ay, F.; Mann, M.; Gonzalez, D.J.; Jain, M.; Sharma, S. The Tumor Suppressor Kinase DAPK3 Drives Tumor-Intrinsic Immunity through the STING-IFN-β Pathway. Nat. Immunol. 2021, 22, 485–496. [Google Scholar] [CrossRef] [PubMed]
  160. 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. [Google Scholar] [CrossRef]
  161. Maekawa, M.; Ishizaki, T.; Boku, S.; Watanabe, N.; Fujita, A.; Iwamatsu, A.; Obinata, T.; Ohashi, K.; Mizuno, K.; Narumiya, S. Signaling from Rho to the Actin Cytoskeleton through Protein Kinases ROCK and LIM-Kinase. Science 1999, 285, 895–898. [Google Scholar] [CrossRef]
  162. Ohashi, K.; Nagata, K.; Maekawa, M.; Ishizaki, T.; Narumiya, S.; Mizuno, K. Rho-Associated Kinase ROCK Activates LIM-Kinase 1 by Phosphorylation at Threonine 508 within the Activation Loop. J. Biol. Chem. 2000, 275, 3577–3582. [Google Scholar] [CrossRef]
  163. Suyama, E.; Wadhwa, R.; Kawasaki, H.; Yaguchi, T.; Kaul, S.C.; Nakajima, M.; Taira, K. LIM Kinase-2 Targeting as a Possible Anti-Metastasis Therapy. J. Gene Med. 2004, 6, 357–363. [Google Scholar] [CrossRef]
  164. Yoshioka, K.; Foletta, V.; Bernard, O.; Itoh, K. A Role for LIM Kinase in Cancer Invasion. Proc. Natl. Acad. Sci. USA 2003, 100, 7247–7252. [Google Scholar] [CrossRef]
  165. You, T.; Gao, W.; Wei, J.; Jin, X.; Zhao, Z.; Wang, C.; Li, Y. Overexpression of LIMK1 Promotes Tumor Growth and Metastasis in Gastric Cancer. Biomed. Pharmacother. 2015, 69, 96–101. [Google Scholar] [CrossRef]
  166. Li, X.; Ke, Q.; Li, Y.; Liu, F.; Zhu, G.; Li, F. DGCR6L, a Novel PAK4 Interaction Protein, Regulates PAK4-Mediated Migration of Human Gastric Cancer Cell via LIMK1. Int. J. Biochem. Cell Biol. 2010, 42, 70–79. [Google Scholar] [CrossRef]
  167. Zhang, J.; Zhang, H.-Y.; Wang, J.; You, L.-H.; Zhou, R.-Z.; Zhao, D.-M.; Cheng, M.-S.; Li, F. GL-1196 Suppresses the Proliferation and Invasion of Gastric Cancer Cells via Targeting PAK4 and Inhibiting PAK4-Mediated Signaling Pathways. Int. J. Mol. Sci. 2016, 17, 470. [Google Scholar] [CrossRef] [PubMed]
  168. Zhang, H.Y.; Zhang, J.; Hao, C.Z.; Zhou, Y.; Wang, J.; Cheng, M.S.; Zhao, D.M.; Li, F. LC-0882 Targets PAK4 and Inhibits PAK4-Related Signaling Pathways to Suppress the Proliferation and Invasion of Gastric Cancer Cells. Am. J. Transl. Res. 2017, 9, 2736–2747. [Google Scholar] [PubMed]
  169. Zhang, J.; Wang, J.; Guo, Q.; Wang, Y.; Zhou, Y.; Peng, H.; Cheng, M.; Zhao, D.; Li, F. LCH-7749944, a Novel and Potent P21-Activated Kinase 4 Inhibitor, Suppresses Proliferation and Invasion in Human Gastric Cancer Cells. Cancer Lett. 2012, 317, 24–32. [Google Scholar] [CrossRef] [PubMed]
  170. Su, B.; Su, J.; Zeng, Y.; Liu, F.; Xia, H.; Ma, Y.-H.; Zhou, Z.-G.; Zhang, S.; Yang, B.-M.; Wu, Y.-H.; et al. Diallyl Disulfide Suppresses Epithelial-Mesenchymal Transition, Invasion and Proliferation by Downregulation of LIMK1 in Gastric Cancer. Oncotarget 2016, 7, 10498–10512. [Google Scholar] [CrossRef]
  171. Guo, B.; Zhang, T.; Su, J.; Wang, K.; Li, X. Oxymatrine Targets EGFR(p-Tyr845) and Inhibits EGFR-Related Signaling Pathways to Suppress the Proliferation and Invasion of Gastric Cancer Cells. Cancer Chemother. Pharmacol. 2015, 75, 353–363. [Google Scholar] [CrossRef]
  172. Li, H.; Chen, C. Quercetin Has Antimetastatic Effects on Gastric Cancer Cells via the Interruption of uPA/uPAR Function by Modulating NF-Κb, PKC-δ, ERK1/2, and AMPKα. Integr. Cancer Ther. 2018, 17, 511–523. [Google Scholar] [CrossRef]
  173. Zeng, Y.; Ren, M.; Li, Y.; Liu, Y.; Chen, C.; Su, J.; Su, B.; Xia, H.; Liu, F.; Jiang, H.; et al. Knockdown of RhoGDI2 Represses Human Gastric Cancer Cell Proliferation, Invasion and Drug Resistance via the Rac1/Pak1/LIMK1 Pathway. Cancer Lett. 2020, 492, 136–146. [Google Scholar] [CrossRef]
  174. Kang, X.; Li, W.; Liu, W.; Liang, H.; Deng, J.; Wong, C.C.; Zhao, S.; Kang, W.; To, K.F.; Chiu, P.W.Y.; et al. LIMK1 Promotes Peritoneal Metastasis of Gastric Cancer and Is a Therapeutic Target. Oncogene 2021, 40, 3422–3433. [Google Scholar] [CrossRef]
  175. Liu, X.; Song, Q.; Wang, D.; Liu, Y.; Zhang, Z.; Fu, W. LIMK1: A Promising Prognostic and Immune Infiltration Indicator in Colorectal Cancer. Oncol. Lett. 2022, 24, 234. [Google Scholar] [CrossRef]
  176. Su, J.; Zhou, Y.; Pan, Z.; Shi, L.; Yang, J.; Liao, A.; Liao, Q.; Su, Q. Downregulation of LIMK1-ADF/Cofilin by DADS Inhibits the Migration and Invasion of Colon Cancer. Sci. Rep. 2017, 7, 45624. [Google Scholar] [CrossRef]
  177. Liao, Q.; Li, R.; Zhou, R.; Pan, Z.; Xu, L.; Ding, Y.; Zhao, L. LIM Kinase 1 Interacts with Myosin-9 and Alpha-Actinin-4 and Promotes Colorectal Cancer Progression. Br. J. Cancer 2017, 117, 563–571. [Google Scholar] [CrossRef]
  178. Sun, X.; Li, S.; Lin, H. LIMK1 Interacts with STK25 to Regulate EMT and Promote the Proliferation and Metastasis of Colorectal Cancer. J. Oncol. 2022, 2022, 3963883. [Google Scholar] [CrossRef] [PubMed]
  179. Zhou, Y.; Su, J.; Shi, L.; Liao, Q.; Su, Q. DADS Downregulates the Rac1-ROCK1/PAK1-LIMK1-ADF/Cofilin Signaling Pathway, Inhibiting Cell Migration and Invasion. Oncol. Rep. 2013, 29, 605–612. [Google Scholar] [CrossRef] [PubMed]
  180. Sheng, N.; Tan, G.; You, W.; Chen, H.; Gong, J.; Chen, D.; Zhang, H.; Wang, Z. MiR-145 Inhibits Human Colorectal Cancer Cell Migration and Invasion via PAK4-Dependent Pathway. Cancer Med. 2017, 6, 1331–1340. [Google Scholar] [CrossRef]
  181. Chen, Y.; Chen, G.; Zhang, B.; Liu, C.; Yu, Y.; Jin, Y. miR-27b-3p Suppresses Cell Proliferation, Migration and Invasion by Targeting LIMK1 in Colorectal Cancer. Int. J. Clin. Exp. Pathol. 2017, 10, 9251–9261. [Google Scholar] [PubMed]
  182. Zhu, Q.; Wu, Y.; Yang, M.; Wang, Z.; Zhang, H.; Jiang, X.; Chen, M.; Jin, T.; Wang, T. IRX5 Promotes Colorectal Cancer Metastasis by Negatively Regulating the Core Components of the RHOA Pathway. Mol. Carcinog. 2019, 58, 2065–2076. [Google Scholar] [CrossRef]
  183. Hu, Y.H.; Lu, Y.X.; Zhang, Z.Y.; Zhang, J.M.; Zhang, W.J.; Zheng, L.; Lin, W.H.; Zhang, W.; Li, X.N. SSH3 Facilitates Colorectal Cancer Cell Invasion and Metastasis by Affecting Signaling Cascades Involving LIMK1/Rac1. Am. J. Cancer Res. 2019, 9, 1061–1073. [Google Scholar]
  184. Croft, D.R.; Coleman, M.L.; Li, S.; Robertson, D.; Sullivan, T.; Stewart, C.L.; Olson, M.F. Actin-Myosin–Based Contraction Is Responsible for Apoptotic Nuclear Disintegration. J. Cell Biol. 2005, 168, 245–255. [Google Scholar] [CrossRef] [PubMed]
  185. McConnell, B.V.; Koto, K.; Gutierrez-Hartmann, A. Nuclear and Cytoplasmic LIMK1 Enhances Human Breast Cancer Progression. Mol. Cancer 2011, 10, 75. [Google Scholar] [CrossRef] [PubMed]
  186. Bagheri-Yarmand, R.; Mazumdar, A.; Sahin, A.A.; Kumar, R. LIM Kinase 1 Increases Tumor Metastasis of Human Breast Cancer Cells via Regulation of the Urokinase-Type Plasminogen Activator System. Int. J. Cancer 2006, 118, 2703–2710. [Google Scholar] [CrossRef] [PubMed]
  187. Lagoutte, E.; Villeneuve, C.; Lafanechère, L.; Wells, C.M.; Jones, G.E.; Chavrier, P.; Rossé, C. LIMK Regulates Tumor-Cell Invasion and Matrix Degradation through Tyrosine Phosphorylation of MT1-MMP. Sci. Rep. 2016, 6, 24925. [Google Scholar] [CrossRef] [PubMed]
  188. Yan, L.; Li, H.; An, W.; Wei, W.; Zhang, X.; Wang, L. Mex-3 RNA Binding MEX3A Promotes the Proliferation and Migration of Breast Cancer Cells via Regulating RhoA/ROCK1/LIMK1 Signaling Pathway. Bioengineered 2021, 12, 5850–5858. [Google Scholar] [CrossRef] [PubMed]
  189. Shi, W.; Ma, D.; Cao, Y.; Hu, L.; Liu, S.; Yan, D.; Zhang, S.; Zhang, G.; Wang, Z.; Wu, J.; et al. SphK2/S1P Promotes Metastasis of Triple-Negative Breast Cancer through the PAK1/LIMK1/Cofilin1 Signaling Pathway. Front. Mol. Biosci. 2021, 8, 598218. [Google Scholar] [CrossRef]
  190. Li, H.; Zhang, B.; Liu, Y.; Yin, C. EBP50 Inhibits the Migration and Invasion of Human Breast Cancer Cells via LIMK/Cofilin and the PI3K/Akt/mTOR/MMP Signaling Pathway. Med. Oncol. 2014, 31, 162. [Google Scholar] [CrossRef]
  191. Shahi, P.; Wang, C.-Y.; Chou, J.; Hagerling, C.; Velozo, H.G.; Ruderisch, A.; Yu, Y.; Lai, M.-D.; Werb, Z. GATA3 Targets Semaphorin 3B in Mammary Epithelial Cells to Suppress Breast Cancer Progression and Metastasis. Oncogene 2017, 36, 5567–5575. [Google Scholar] [CrossRef]
  192. Fu, J.; Yu, J.; Chen, J.; Xu, H.; Luo, Y.; Lu, H. In Vitro Inhibitory Properties of Sesquiterpenes from Chloranthus Serratus on Cell Motility via Down-Regulation of LIMK1 Activation in Human Breast Cancer. Phytomedicine 2018, 49, 23–31. [Google Scholar] [CrossRef]
  193. Zhao, J.; Li, D.; Fang, L. MiR-128-3p Suppresses Breast Cancer Cellular Progression via Targeting LIMK1. Biomed. Pharmacother. 2019, 115, 108947. [Google Scholar] [CrossRef] [PubMed]
  194. Li, D.; Wang, H.; Song, H.; Xu, H.; Zhao, B.; Wu, C.; Hu, J.; Wu, T.; Xie, D.; Zhao, J.; et al. The microRNAs miR-200b-3p and miR-429-5p Target the LIMK1/CFL1 Pathway to Inhibit Growth and Motility of Breast Cancer Cells. Oncotarget 2017, 8, 85276–85289. [Google Scholar] [CrossRef] [PubMed]
  195. Li, D.; Song, H.; Wu, T.; Xie, D.; Hu, J.; Zhao, J.; Shen, Q.; Fang, L. MiR-519d-3p Suppresses Breast Cancer Cell Growth and Motility via Targeting LIM Domain Kinase 1. Mol. Cell. Biochem. 2018, 444, 169–178. [Google Scholar] [CrossRef]
  196. Li, D.; Hu, J.; Song, H.; Xu, H.; Wu, C.; Zhao, B.; Xie, D.; Wu, T.; Zhao, J.; Fang, L. miR-143-3p Targeting LIM Domain Kinase 1 Suppresses the Progression of Triple-Negative Breast Cancer Cells. Am. J. Transl. Res. 2017, 9, 2276–2285. [Google Scholar] [PubMed]
  197. Davila, M.; Jhala, D.; Ghosh, D.; E Grizzle, W.; Chakrabarti, R. Expression of LIM Kinase 1 Is Associated with Reversible G1/S Phase Arrest, Chromosomal Instability and Prostate Cancer. Mol. Cancer 2007, 6, 40. [Google Scholar] [CrossRef]
  198. Limonta, P.; Moretti, R.M.; Mai, S.; Marelli, M.M.; Rizzi, F.; Bettuzzi, S. Molecular Mechanisms of the Antimetastatic Activity of Nuclear Clusterin in Prostate Cancer Cells. Int. J. Oncol. 2011, 39, 225–234. [Google Scholar] [CrossRef]
  199. Cai, S.; Chen, R.; Li, X.; Cai, Y.; Ye, Z.; Li, S.; Li, J.; Huang, H.; Peng, S.; Wang, J.; et al. Downregulation of microRNA-23a Suppresses Prostate Cancer Metastasis by Targeting the PAK6-LIMK1 Signaling Pathway. Oncotarget 2015, 6, 3904–3917. [Google Scholar] [CrossRef]
  200. Ngalame, N.N.; Makia, N.L.; Waalkes, M.P.; Tokar, E.J. Mitigation of Arsenic-Induced Acquired Cancer Phenotype in Prostate Cancer Stem Cells by miR-143 Restoration. Toxicol. Appl. Pharmacol. 2016, 312, 11–18. [Google Scholar] [CrossRef]
  201. Bhardwaj, A.; Srivastava, S.K.; Singh, S.; Arora, S.; Tyagi, N.; Andrews, J.; McClellan, S.; Carter, J.E.; Singh, A.P. CXCL12/CXCR4 Signaling Counteracts Docetaxel-Induced Microtubule Stabilization via P21-Activated Kinase 4-Dependent Activation of LIM Domain Kinase 1. Oncotarget 2014, 5, 11490–11500. [Google Scholar] [CrossRef]
  202. Mardilovich, K.; Gabrielsen, M.; McGarry, L.; Orange, C.; Patel, R.; Shanks, E.; Edwards, J.; Olson, M.F. Elevated LIM Kinase 1 in Nonmetastatic Prostate Cancer Reflects Its Role in Facilitating Androgen Receptor Nuclear Translocation. Mol. Cancer Ther. 2015, 14, 246–258. [Google Scholar] [CrossRef]
  203. Lu, G.; Zhou, Y.; Zhang, C.; Zhang, Y. Upregulation of LIMK1 Is Correlated with Poor Prognosis and Immune Infiltrates in Lung Adenocarcinoma. Front. Genet. 2021, 12, 671585. [Google Scholar] [CrossRef] [PubMed]
  204. Cai, S.; Ye, Z.; Wang, X.; Pan, Y.; Weng, Y.; Lao, S.; Wei, H.; Li, L. Overexpression of P21-Activated Kinase 4 Is Associated with Poor Prognosis in Non-Small Cell Lung Cancer and Promotes Migration and Invasion. J. Exp. Clin. Cancer Res. 2015, 34, 48. [Google Scholar] [CrossRef] [PubMed]
  205. Guo, B.; Li, X.; Song, S.; Chen, M.; Cheng, M.; Zhao, D.; Li, F. (-)-β-Hydrastine Suppresses the Proliferation and Invasion of Human Lung Adenocarcinoma Cells by Inhibiting PAK4 Kinase Activity. Oncol. Rep. 2016, 35, 2246–2256. [Google Scholar] [CrossRef] [PubMed]
  206. Zhang, M.; Wang, R.; Tian, J.; Song, M.; Zhao, R.; Liu, K.; Zhu, F.; Shim, J.; Dong, Z.; Lee, M. Targeting LIMK1 with Luteolin Inhibits the Growth of Lung Cancer In Vitro and In Vivo. J. Cell. Mol. Med. 2021, 25, 5560–5571. [Google Scholar] [CrossRef] [PubMed]
  207. Kang, C.G.; Im, E.; Lee, H.-J.; Lee, E.-O. Plumbagin Reduces Osteopontin-Induced Invasion through Inhibiting the Rho-Associated Kinase Signaling Pathway in A549 Cells and Suppresses Osteopontin-Induced Lung Metastasis in BalB/c Mice. Bioorg Med. Chem. Lett. 2017, 27, 1914–1918. [Google Scholar] [CrossRef] [PubMed]
  208. Wan, L.; Zhang, L.; Fan, K.; Wang, J. MiR-27b Targets LIMK1 to Inhibit Growth and Invasion of NSCLC Cells. Mol. Cell. Biochem. 2014, 390, 85–91. [Google Scholar] [CrossRef]
  209. Chen, Q.; Jiao, D.; Hu, H.; Song, J.; Yan, J.; Wu, L.; Xu, L.-Q. Downregulation of LIMK1 Level Inhibits Migration of Lung Cancer Cells and Enhances Sensitivity to Chemotherapy Drugs. Oncol. Res. 2013, 20, 491–498. [Google Scholar] [CrossRef]
  210. Yang, J.Z.; Huang, L.H.; Chen, R.; Meng, L.J.; Gao, Y.Y.; Ji, Q.Y.; Wang, Y. LIM Kinase 1 Serves an Important Role in the Multidrug Resistance of Osteosarcoma Cells. Oncol. Lett. 2018, 15, 250–256. [Google Scholar] [CrossRef]
  211. Zhang, H.-S.; Zhao, J.-W.; Wang, H.; Zhang, H.-Y.; Ji, Q.-Y.; Meng, L.-J.; Xing, F.-J.; Yang, S.-T.; Wang, Y. LIM Kinase 1 Is Required for Insulin-dependent Cell Growth of Osteosarcoma Cell Lines. Mol. Med. Rep. 2014, 9, 103–108. [Google Scholar] [CrossRef]
  212. Yoshizawa, M.; Nakamura, S.; Sugiyama, Y.; Tamai, S.; Ishida, Y.; Sueyoshi, M.; Toda, Y.; Hosogi, S.; Yano, Y.; Ashihara, E. 6-Hydroxythiobinupharidine Inhibits Migration of LM8 Osteosarcoma Cells by Decreasing Expression of LIM Domain Kinase 1. Anticancer Res. 2019, 39, 6507–6513. [Google Scholar] [CrossRef] [PubMed]
  213. Zhang, H.; Wang, Y.; Xing, F.; Wang, J.; Wang, H.; Yang, Y.; Gao, Z.; Wang, Y. Overexpression of LIMK1 Promotes Migration Ability of Multidrug-Resistant Osteosarcoma Cells. Oncol. Res. 2011, 19, 501–509. [Google Scholar] [CrossRef]
  214. Chhavi; Saxena, M.; Singh, S.; Negi, M.; Srivastava, A.; Trivedi, R.; Singh, U.; Pant, M.; Bhatt, M. Expression Profiling of G2/M Phase Regulatory Proteins in Normal, Premalignant and Malignant Uterine Cervix and Their Correlation with Survival of Patients. J. Cancer Res. Ther. 2010, 6, 167–171. [Google Scholar] [CrossRef]
  215. Yang, X.; Du, H.; Bian, W.; Li, Q.; Sun, H. FOXD3-AS1/miR-128-3p/LIMK1 Axis Regulates Cervical Cancer Progression. Oncol. Rep. 2021, 45, 62. [Google Scholar] [CrossRef] [PubMed]
  216. Xu, Y.; Zheng, Y.; Duan, Y.; Ma, L.; Nan, P. MicroRNA-125a-5p Targets LIM Kinase 1 to Inhibit Cisplatin Resistance of Cervical Cancer Cells. Oncol. Lett. 2021, 21, 392. [Google Scholar] [CrossRef] [PubMed]
  217. Ding, B.; Lou, W.; Fan, W.; Pan, J. Exosomal miR-374c-5p derived from mesenchymal stem cells suppresses epithelial-mesenchymal transition of hepatocellular carcinoma via the LIMK1-Wnt/β-catenin axis. Environ. Toxicol. 2023, 38, 1038–1052. [Google Scholar] [CrossRef]
  218. Wang, D.; Xing, N.; Yang, T.; Liu, J.; Zhao, H.; He, J.; Ai, Y.; Yang, J. Exosomal lncRNA H19 Promotes the Progression of Hepatocellular Carcinoma Treated with Propofol via miR-520a-3p/LIMK1 Axis. Cancer Med. 2020, 9, 7218–7230. [Google Scholar] [CrossRef]
  219. Pan, Z.; Liu, C.; Zhi, Y.; Xie, Z.; Wu, L.; Jiang, M.; Zhang, Y.; Zhou, R.; Zhao, L. LIMK1 Nuclear Translocation Promotes Hepatocellular Carcinoma Progression by Increasing P-ERK Nuclear Shuttling and by Activating c-Myc Signalling upon EGF Stimulation. Oncogene 2021, 40, 2581–2595. [Google Scholar] [CrossRef]
  220. Lee, C.; Siu, A.; Chang, J.; Dang, D.; Ramos, D.M. Alphavbeta3 Suppresses the RhoA-LIMK1 Pathway in K1735 Melanoma. J. Calif. Dent. Assoc. 2012, 40, 921–927. [Google Scholar] [CrossRef]
  221. Shi, B.; Ma, C.; Liu, G.; Guo, Y. MiR-106a Directly Targets LIMK1 to Inhibit Proliferation and EMT of Oral Carcinoma Cells. Cell Mol. Biol. Lett. 2019, 24, 1. [Google Scholar] [CrossRef]
  222. Peng, T.; Wang, T.; Liu, G.; Zhou, L.; Teng, Z. Effects of miR-373 Inhibition on Glioblastoma Growth by Reducing Limk1 In Vitro. J. Immunol. Res. 2020, 2020, 7671502. [Google Scholar] [CrossRef]
  223. Xiong, Y.; Zhang, L.; Kebebew, E. MiR-20a Is Upregulated in Anaplastic Thyroid Cancer and Targets LIMK1. PLoS ONE 2014, 9, e96103. [Google Scholar] [CrossRef]
  224. Zhou, J.; Liu, R.; Luo, C.; Zhou, X.; Xia, K.; Chen, X.; Zhou, M.; Zou, Q.; Cao, P.; Cao, K. MiR-20a Inhibits Cutaneous Squamous Cell Carcinoma Metastasis and Proliferation by Directly Targeting LIMK1. Cancer Biol. Ther. 2014, 15, 1340–1349. [Google Scholar] [CrossRef] [PubMed]
  225. Wang, J.; He, Z.; Xu, J.; Chen, P.; Jiang, J. Long Noncoding RNA LINC00941 Promotes Pancreatic Cancer Progression by Competitively Binding miR-335-5p to Regulate ROCK1-Mediated LIMK1/Cofilin-1 Signaling. Cell Death Dis. 2021, 12, 36. [Google Scholar] [CrossRef] [PubMed]
  226. 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, 1067–1073. [Google Scholar] [CrossRef] [PubMed]
  227. Chen, P.; Zeng, M.; Zhao, Y.; Fang, X. Upregulation of Limk1 Caused by microRNA-138 Loss Aggravates the Metastasis of Ovarian Cancer by Activation of Limk1/Cofilin Signaling. Oncol. Rep. 2014, 32, 2070–2076. [Google Scholar] [CrossRef]
  228. Zhang, Y.; Li, A.; Shi, J.; Fang, Y.; Gu, C.; Cai, J.; Lin, C.; Zhao, L.; Liu, S. Imbalanced LIMK1 and LIMK2 Expression Leads to Human Colorectal Cancer Progression and Metastasis via Promoting β-Catenin Nuclear Translocation. Cell Death Dis. 2018, 9, 749. [Google Scholar] [CrossRef]
  229. Lourenço, F.C.; Munro, J.; Brown, J.; Cordero, J.; Stefanatos, R.; Strathdee, K.; Orange, C.; Feller, S.M.; Sansom, O.J.; Vidal, M.; et al. Reduced LIMK2 Expression in Colorectal Cancer Reflects Its Role in Limiting Stem Cell Proliferation. Gut 2014, 63, 480–493. [Google Scholar] [CrossRef]
  230. Aggelou, H.; Chadla, P.; Nikou, S.; Karteri, S.; Maroulis, I.; Kalofonos, H.P.; Papadaki, H.; Bravou, V. LIMK/Cofilin Pathway and Slingshot Are Implicated in Human Colorectal Cancer Progression and Chemoresistance. Virchows Arch. 2018, 472, 727–737. [Google Scholar] [CrossRef]
  231. Zhang, Y.; Liu, X.; Li, Q.; Zhang, Y. lncRNA LINC00460 Promoted Colorectal Cancer Cells Metastasis via miR-939-5p Sponging. Cancer Manag. Res. 2019, 11, 1779–1789. [Google Scholar] [CrossRef]
  232. Nikhil, K.; Chang, L.; Viccaro, K.; Jacobsen, M.; McGuire, C.; Satapathy, S.R.; Tandiary, M.; Broman, M.M.; Cresswell, G.; He, Y.J.; et al. Identification of LIMK2 as a Therapeutic Target in Castration Resistant Prostate Cancer. Cancer Lett. 2019, 448, 182–196. [Google Scholar] [CrossRef] [PubMed]
  233. Niu, Y.; Ma, F.; Huang, W.; Fang, S.; Li, M.; Wei, T.; Guo, L. Long Non-Coding RNA TUG1 Is Involved in Cell Growth and Chemoresistance of Small Cell Lung Cancer by Regulating LIMK2b via EZH2. Mol. Cancer 2017, 16, 5. [Google Scholar] [CrossRef] [PubMed]
  234. Su, Y.; Xu, B.; Shen, Q.; Lei, Z.; Zhang, W.; Hu, T. LIMK2 Is a Novel Prognostic Biomarker and Correlates With Tumor Immune Cell Infiltration in Lung Squamous Cell Carcinoma. Front. Immunol. 2022, 13, 788375. [Google Scholar] [CrossRef]
  235. Ren, T.; Zheng, B.; Huang, Y.; Wang, S.; Bao, X.; Liu, K.; Guo, W. Osteosarcoma Cell Intrinsic PD-L2 Signals Promote Invasion and Metastasis via the RhoA-ROCK-LIMK2 and Autophagy Pathways. Cell Death Dis. 2019, 10, 261. [Google Scholar] [CrossRef]
  236. Sooreshjani, M.A.; Nikhil, K.; Kamra, M.; Nguyen, D.N.; Kumar, D.; Shah, K. LIMK2-NKX3.1 Engagement Promotes Castration-Resistant Prostate Cancer. Cancers 2021, 13, 2324. [Google Scholar] [CrossRef] [PubMed]
  237. Malvi, P.; Janostiak, R.; Chava, S.; Manrai, P.; Yoon, E.; Singh, K.; Harigopal, M.; Gupta, R.; Wajapeyee, N. LIMK2 Promotes the Metastatic Progression of Triple-Negative Breast Cancer by Activating SRPK1. Oncogenesis 2020, 9, 77. [Google Scholar] [CrossRef]
  238. Xu, M.; Wang, F.; Li, G.; Wang, X.; Fang, X.; Jin, H.; Chen, Z.; Zhang, J.; Fu, L. MED12 Exerts an Emerging Role in Actinmediated Cytokinesis via LIMK2/Cofilin Pathway in NSCLC. Mol. Cancer 2019, 18, 93. [Google Scholar] [CrossRef]
  239. Wang, S.; Ren, T.; Jiao, G.; Huang, Y.; Bao, X.; Zhang, F.; Liu, K.; Zheng, B.; Sun, K.; Guo, W. BMPR2 Promotes Invasion and Metastasis via the RhoA-ROCK-LIMK2 Pathway in Human Osteosarcoma Cells. Oncotarget 2017, 8, 58625–58641. [Google Scholar] [CrossRef]
  240. Po’Uha, S.T.; Shum, M.S.Y.; 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, 597–607. [Google Scholar] [CrossRef]
  241. Gamell, C.; Schofield, A.V.; Suryadinata, R.; Sarcevic, B.; Bernard, O. LIMK2 Mediates Resistance to Chemotherapeutic Drugs in Neuroblastoma Cells through Regulation of Drug-Induced Cell Cycle Arrest. PLoS ONE 2013, 8, e72850. [Google Scholar] [CrossRef]
  242. Vlecken, D.H.; Bagowski, C.P. LIMK1 and LIMK2 Are Important for Metastatic Behavior and Tumor Cell-Induced Angiogenesis of Pancreatic Cancer Cells. Zebrafish 2009, 6, 433–439. [Google Scholar] [CrossRef] [PubMed]
  243. Rak, R.; Haklai, R.; Elad-Tzfadia, G.; Wolfson, H.J.; Carmeli, S.; Kloog, Y. Novel LIMK2 Inhibitor Blocks Panc-1 Tumor Growth in a Mouse Xenograft Model. Oncoscience 2014, 1, 39–48. [Google Scholar] [CrossRef] [PubMed]
  244. Wang, W.; Yang, C.; Nie, H.; Qiu, X.; Zhang, L.; Xiao, Y.; Zhou, W.; Zeng, Q.; Zhang, X.; Wu, Y.; et al. LIMK2 Acts as an Oncogene in Bladder Cancer and Its Functional SNP in the microRNA-135a Binding Site Affects Bladder Cancer Risk. Int. J. Cancer 2019, 144, 1345–1355. [Google Scholar] [CrossRef]
  245. Hsu, F.-F.; Lin, T.-Y.; Chen, J.-Y.; Shieh, S.-Y. P53-Mediated Transactivation of LIMK2b Links Actin Dynamics to Cell Cycle Checkpoint Control. Oncogene 2010, 29, 2864–2876. [Google Scholar] [CrossRef]
  246. Nikhil, K.; Haymour, H.S.; Kamra, M.; Shah, K. Phosphorylation-Dependent Regulation of SPOP by LIMK2 Promotes Castration-Resistant Prostate Cancer. Br. J. Cancer 2021, 124, 995–1008. [Google Scholar] [CrossRef] [PubMed]
  247. Johnson, E.O.; Chang, K.-H.; Ghosh, S.; Venkatesh, C.; Giger, K.; Low, P.S.; Shah, K. LIMK2 Is a Crucial Regulator and Effector of Aurora-A-Kinase-Mediated Malignancy. J. Cell Sci. 2012, 125 Pt 5, 1204–1216. [Google Scholar] [CrossRef] [PubMed]
  248. Nikhil, K.; Kamra, M.; Raza, A.; Shah, K. Negative Cross Talk between LIMK2 and PTEN Promotes Castration Resistant Prostate Cancer Pathogenesis in Cells and in Vivo. Cancer Lett. 2021, 498, 1–18. [Google Scholar] [CrossRef]
  249. The Cause of Cancer. JAMA 2021, 325, 311. [CrossRef]
  250. Lasorsa, F.; Rutigliano, M.; Milella, M.; Ferro, M.; Pandolfo, S.D.; Crocetto, F.; Autorino, R.; Battaglia, M.; Ditonno, P.; Lucarelli, G. Cancer Stem Cells in Renal Cell Carcinoma: Origins and Biomarkers. Int. J. Mol. Sci. 2023, 24, 13179. [Google Scholar] [CrossRef]
  251. Aveta, A.; Cilio, S.; Contieri, R.; Spena, G.; Napolitano, L.; Manfredi, C.; Franco, A.; Crocerossa, F.; Cerrato, C.; Ferro, M.; et al. Urinary MicroRNAs as Biomarkers of Urological Cancers: A Systematic Review. Int. J. Mol. Sci. 2023, 24, 10846. [Google Scholar] [CrossRef]
  252. Gao, J.; Hou, B.; Zhu, Q.; Yang, L.; Jiang, X.; Zou, Z.; Li, X.; Xu, T.; Zheng, M.; Chen, Y.-H.; et al. Engineered Bioorthogonal POLY-PROTAC Nanoparticles for Tumour-Specific Protein Degradation and Precise Cancer Therapy. Nat. Commun. 2022, 13, 4318. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The 3D structure of ALP subfamily. (A) The 3D structure of PDLIM1. (B) The 3D structure of PDLIM2. (C) The 3D structure of PDLIM3. (D) The 3D structure of PDLIM4.
Figure 1. The 3D structure of ALP subfamily. (A) The 3D structure of PDLIM1. (B) The 3D structure of PDLIM2. (C) The 3D structure of PDLIM3. (D) The 3D structure of PDLIM4.
Cancers 15 05042 g001
Figure 2. The 3D structure of Enigma subfamily. (A) The 3D structure of PDLIM5. (B) The 3D structure of PDLIM6. (C) The 3D structure of PDLIM7.
Figure 2. The 3D structure of Enigma subfamily. (A) The 3D structure of PDLIM5. (B) The 3D structure of PDLIM6. (C) The 3D structure of PDLIM7.
Cancers 15 05042 g002
Figure 3. The 3D structure of LIM Kinase and LMO7. (A) The 3D structure of LIMK1. (B) The 3D structure of LIMK2. (C) The 3D structure of LMO7.
Figure 3. The 3D structure of LIM Kinase and LMO7. (A) The 3D structure of LIMK1. (B) The 3D structure of LIMK2. (C) The 3D structure of LMO7.
Cancers 15 05042 g003
Figure 4. PDLIMs and integrin signaling pathway.
Figure 4. PDLIMs and integrin signaling pathway.
Cancers 15 05042 g004
Figure 5. PDLIMs and TGF-β signaling pathway.
Figure 5. PDLIMs and TGF-β signaling pathway.
Cancers 15 05042 g005
Figure 6. PDLIMs and MAPK and NF-κB signaling pathway.
Figure 6. PDLIMs and MAPK and NF-κB signaling pathway.
Cancers 15 05042 g006
Table 1. ALP subfamily in tumor development and progression.
Table 1. ALP subfamily in tumor development and progression.
PDLIMsTumor TypeExpressionEffectReference
PDLIM1GliomaExogenous suppressionInhibits tumor invasion Ahn et al., 2016 [98]
Breast cancerIncreasedPromotes tumor migration and invasionLiu et al., 2015 [99]; Gupta et al., 2016 [100]
Pancreatic cancerIncreasedAs a tumor antigen that induces antibody responseHong 2005 [105]
Hepatocellular carcinomaReducedPromotes tumor migrationHuang et al., 2020 [101]
Gastric cancerReducedPromotes tumor progression and cisplatin sensitivityTan et al., 2022 [102]
Colorectal CancerReducedInhibits tumor metastasis formation and EMTChen et al., 2016 [103]
ChoriocarcinomaExogenous suppressionInhibits tumor cell actin stress fiber formation and focal adhesion assemblyTamura et al., 2007 [104]
PDLIM2Breast cancerExogenous suppressionPromotes tumor progression and lymphatic metastasisDing 2018 [107]; Qu et al., 2010 [108]
Breast cancerIncreasedPromotes cell migration, cytoskeletal polarization, and EMTDeevi, Cox, and O’Connor 2014 [109]; Bowe et al., 2014 [110]
Lung cancerReducedPromotes tumor progression and therapeutic resistanceSun et al., 2019 [87]; Shi et al., 2020 [111]
Esophageal squamous cell carcinomaHigh PDLIM2 expression group has longer overall survivalSong et al., 2019 [112]
Hepatocellular carcinomaExogenous overexpressionInhibits tumor malignant phenotypeJiang et al., 2021 [113]
Laryngeal squamous cell carcinomaReducedPromotes tumor cell proliferationWang et al., 2022 [114]
Ovarian cancerReducedPromotes tumor pathogenesisZhao et al., 2016 [115]
Metastatic colorectal cancerReducedPromotes tumor metastasisOh et al., 2017 [116]
Castration-Resistant Prostate CancerIncreasedPromotes tumor growth and invasionKang et al., 2016 [91]
Gastric cancerExogenous activationPromotes tumorigenicity and metastasisGuo et al., 2016 [117]
PDLIM3MedulloblastomaIncreasedUnknownShou et al., 2015 [118]
Invasive bladder urothelium carcinomaIncreasedAssociated with the unfavorable survivalFeng et al., 2020 [119]
Bladder cancerReducedUnknownLu et al., 2010 [120]
Thyroid Papillary carcinomaReducedUnknownStein et al., 2010 [121]
PDLIM4Ovarian cancerReducedAssociated with aggressive tumor features and poor prognosis Jia et al., 2019 [122]
Prostatic carcinomaReducedPromotes carcinogenesisVanaja et al., 2009 [123]; Vanaja et al., 2006 [124]; Kolluru et al., 2019 [125]; Vasiljević et al., 2011 [126]
Thyroid carcinomaReducedUnknown Patai et al., 2017 [128]
Kidney cancerReducedUnknown Morris et al., 2010 [127]
Breast cancerReducedPromotes tumor progressionFeng et al., 2010 [129]
Xu et al., 2012 [130]
Breast cancerExogenous overexpressionPromotes tumor metastasisKravchenko et al., 2020 [131]
Table 2. Enigma subfamily in tumor development and progression.
Table 2. Enigma subfamily in tumor development and progression.
PDLIMsTumor TypeExpressionEffectReference
PDLIM5Prostatic carcinomaIncreasedPromotes tumorigenesis and migrationLiu et al., 2017 [132]; Shui et al., 2014 [133]
Thyroid Papillary carcinomaIncreasedPromotes tumor migration, invasion and proliferation Wei et al., 2018 [135]
Lung cancerIncreasedPromotes tumor migration and invasionShi et al., 2020 [68]
Edlund et al., 2012 [137]
Zhang et al., 2022 [136]
Wu et al., 2023 [138]
PDLIM6Unknown
PDLIM7Acute myeloid leukemiaIncreasedAn independent risk factor for EFS and OSCui et al., 2019 [142]
Breast cancerIncreasedCorrelates with a poor outcomeKales et al., 2014 [143]
Tabariès et al., 2019 [144]
Colorectal cancerExogenous overexpressionPromotes tumor cell survivalJung et al., 2010 [145]
Hepatocellular carcinomaExogenous overexpressionPromotes tumor cell survivalJung et al., 2010 [145]
Thyroid carcinomaIncreasedPromotes carcinogenesisFirek et al., 2017 [146]
Borrello et al., 2002 [147]
Kim et al., 2018 [148]
Gastric cancerDegradation InhibitedContributes to 5-Fu resistance in GC cellsLu et al., 2023 [149]
OsteosarcomaReducedPromotes tumor malignant phenotypesLiu et al., 2014 [150]
Table 3. LIM Kinase and LMO7 in tumor development and progression.
Table 3. LIM Kinase and LMO7 in tumor development and progression.
PDLIMsTumor TypeExpressionEffectReference
LIMK1Gastric cancerIncreasedPromotes tumor growth and metastasisYou et al., 2015 [165]; Li et al., 2010 [166]; Zhang et al., 2016 [167]; H.-Y. Zhang et al., 2017 [168]; J. Zhang et al., 2012 [169]; Su et al., 2016 [170]; Guo et al., 2015 [171]; Li and Chen 2018 [172]; Zeng et al., 2020 [173]; Kang et al., 2021 [174]
Colorectal cancerIncreasedPromoted tumor developmentLiu et al., 2022 [175]; Su et al., 2017 [176]; Liao et al., 2017 [177]; Sun, Li, and Lin 2022 [178]; Zhou et al., 2013 [179]; Sheng et al., 2017 [180]; Chen et al., 2017 [181]
Colorectal cancerExogenous suppressionPromoted tumor developmentZhu et al., 2019 [182]; Hu et al., 2019 [183]
Breast cancerIncreasedPromoted tumor developmentCroft et al., 2005 [184]; McConnell, Koto, and Gutierrez-Hartmann 2011 [185]; Bagheri-Yarmand et al., 2006 [186]; Lagoutte et al., 2016 [187]; Yan et al., 2021 [188]; Shi et al., 2021 [189]; Li et al., 2014 [190]; Shahi et al., [191]; Fu et al., 2018 [192]; Zhao, Li, and Fang 2019 [193]; Li, Hu, et al., 2017 [194]; Li, Wang, et al., 2017 [195]; Li et al., 2018, 1 [196] J.
Prostatic carcinomaExogenous overexpressionPromoted tumor developmentDavila et al., 2007 [197]; Moretti et al., 2011 [198]; Cai et al., 2015 [199]; Ngalame et al., 2016 [200]; Bhardwaj et al., 2014, 1 [201]; Mardilovich et al., 2015 [202]
Lung cancerIncreasedPromoted tumor developmentLu et al., 2021 [203]; Cai et al., 2015 [204]; Guo et al., 2016 [205]; Zhang et al., 2021 [206]; Kang et al., 2017 [207]; Wan et al., 2014 [208]; Chen et al., 2013 [209]
OsteosarcomaIncreasedPromoted tumor developmentYang et al., 2018 [210]; H.-S. Zhang et al., 2014 [211]; Yoshizawa et al., 2019, 1 [212]; H. Zhang et al., 2011 [213]
Uterine cervix carcinomaIncreasedPromoted tumor developmentChhavi et al., 2010 [214]; Yang et al., 2021 [215]; Xu et al., 2021 [216]
Hepatocellular carcinomaExogenous overexpressionPromoted tumor developmentDing et al., 2023 [217]; Wang et al., 2020 [218]; Pan et al., 2021 [219]
MelanomaExogenous suppressionPromotes tumor invasion Lee et al., 2012 [220]
Oral cancerExogenous suppressionInhibits tumor proliferation, migration and EMTShi et al., 2019 [221]
Thyroid Papillary carcinomaExogenous suppressionInhibits tumor progression Xiong, Zhang, and Kebebew 2014 [223]
Cutaneous squamous cell carcinomaExogenous suppressionInhibits tumor progression Zhou et al., 2014 [224]
GlioblastomaExogenous suppressionInhibits tumor progression Peng et al., 2020 [222]
Pancreatic adenocarcinomaExogenous activationPromotes tumor growth and metastasisWang et al., 2021 [225]
Ovarian cancerIncreasedPromotes tumor migrationZhang, Gan, and Zhou 2012 [226]; Chen et al., 2014 [227]
LIMK2Colorectal cancerExogenous overexpressionInhibits tumor cell proliferationZhang et al., 2019 [231]
Colorectal cancerIncreasedPromotes tumor developmentAggelou et al., 2018 [230]
Colorectal cancerReducedPromotes tumor developmentYue Zhang et al., 2018 [228]; Lourenço et al., 2014 [229]
Castration resistant prostate cancerIncreasedPromotes tumor initiation, progression, and poor prognosisNikhil et al., 2019 [232]
Prostate cancerLIMK regulates other moleculesPromotes tumor progression and drug resistanceNikhil, Kamra, et al., 2021 [246]; Nikhil et al., 2021 [247]; Sooreshjani et al., 2021 [248]
Triple negative breast cancerLIMK regulates other moleculesPromotes tumor migrationMalvi et al., 2020 [237]
Breast cancerExogenous suppressionInhibits tumor migrationShahi et al., 2017 [191]; Malvi et al., 2020 [237]
Non-small-cell lung cancerExogenous activationInhibits tumor proliferationXu et al., 2019 [238]
Non-small-cell lung cancerExogenous overexpressionPromotes tumor proliferation and drug resistanceNiu et al., 2017 [233]; Su et al., 2022 [234]
OsteosarcomaExogenous activationPromotes tumor migrationRen et al., 2019 [235]; Wang et al., 2017 [239]
NeuroblastomaIncreasedPromotes tumor drug resistancePo’uha et al., 2010 [240]; Gamell et al., 2013 [241]
Pancreatic adenocarcinomaExogenous suppressionInhibits tumor developmentRak et al., 2014 [243]
Bladder cancerIncreasedPromotes tumor proliferation, migration, and invasionWang et al., 2019 [244]
Thyroid carcinomaReducedPromotes tumor migrationHsu et al., 2010 [245]
Esophageal cancerReducedPromotes tumor migrationHsu et al., 2010 [245]
LMO7Breast cancerIncreasedPromotes tumor migrationHu et al., 2011 [151]; Tanaka-Okamoto et al., 2009, 7 [152]; Nakamura et al., 2011 [153]
Lung cancerExogenous suppressionPromoted tumor developmentWu et al., 2017 [154]; Yang et al., 2022 [155]; Li et al., 2021 [156]
Cervical carcinomaExogenous overexpressionInhibits tumor cell proliferationTzeng et al., 2018 [157]
Colorectal cancerExogenous overexpressionInhibits tumor cell proliferationTzeng et al., 2018 [157]
Pancreatic cancerIncreasedPromotes tumor progression and metastasisLin et al., 2021 [158]
Hepatocellular carcinomaExogenous overexpressionPromotes tumor invasionNakamura et al., 2005 [72]
Table 4. Abbreviations in this article.
Table 4. Abbreviations in this article.
ATCUndifferentiated thyroid cancer
BCBreast cancer
BCABladder carcinoma
BRCABreast invasive carcinoma
CCChoriocarcinoma
CCACervical carcinoma
CRCColorectal cancer
EMTEpithelial–mesenchymal transition
ECEsophageal cancer
FVPTCFollicular papillary thyroid carcinoma
GCGastric cancer
GCCHepatocarcinoma
LSCCLaryngeal squamous cell carcinoma
LUADLung adenocarcinoma
LCLung cancer
NSCLCNon-small-cell lung cancer
OSOsteosarcoma
OVOvarian Cancer
PAADPancreatic adenocarcinoma
PTCPapillary thyroid carcinoma
PDTCPoorly differentiated thyroid carcinoma
PCProstatic carcinoma
THCAThyroid carcinoma
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jiang, X.; Xu, Z.; Jiang, S.; Wang, H.; Xiao, M.; Shi, Y.; Wang, K. PDZ and LIM Domain-Encoding Genes: Their Role in Cancer Development. Cancers 2023, 15, 5042. https://doi.org/10.3390/cancers15205042

AMA Style

Jiang X, Xu Z, Jiang S, Wang H, Xiao M, Shi Y, Wang K. PDZ and LIM Domain-Encoding Genes: Their Role in Cancer Development. Cancers. 2023; 15(20):5042. https://doi.org/10.3390/cancers15205042

Chicago/Turabian Style

Jiang, Xinyuan, Zhiyong Xu, Sujing Jiang, Huan Wang, Mingshu Xiao, Yueli Shi, and Kai Wang. 2023. "PDZ and LIM Domain-Encoding Genes: Their Role in Cancer Development" Cancers 15, no. 20: 5042. https://doi.org/10.3390/cancers15205042

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

Article Metrics

Back to TopTop