Next Article in Journal
A-Kinase Anchoring Proteins in Cardiac Myocytes and Their Roles in Regulating Calcium Cycling
Next Article in Special Issue
Engineered Synthetic STxB for Enhanced Cytosolic Delivery
Previous Article in Journal
Genetic and Chemical Controls of Sperm Fate and Spermatocyte Dedifferentiation via PUF-8 and MPK-1 in Caenorhabditis elegans
Previous Article in Special Issue
Pharmacodynamic Modeling to Evaluate the Impact of Cimetidine, an OCT2 Inhibitor, on the Anticancer Effects of Cisplatin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

RHOA Therapeutic Targeting in Hematological Cancers

by
Juliana Carvalho Santos
1,
Núria Profitós-Pelejà
1,
Salvador Sánchez-Vinces
2 and
Gaël Roué
1,*
1
Lymphoma Translational Group, Josep Carreras Leukaemia Research Institute (IJC), 08916 Badalona, Spain
2
Laboratory of Immunopharmacology and Molecular Biology, Sao Francisco University Medical School, Braganca Paulista 01246-100, São Paulo, Brazil
*
Author to whom correspondence should be addressed.
Cells 2023, 12(3), 433; https://doi.org/10.3390/cells12030433
Submission received: 4 January 2023 / Revised: 19 January 2023 / Accepted: 26 January 2023 / Published: 28 January 2023
(This article belongs to the Special Issue Innovative Drug Treatment of Cancer Cells)

Abstract

:
Primarily identified as an important regulator of cytoskeletal dynamics, the small GTPase Ras homolog gene family member A (RHOA) has been implicated in the transduction of signals regulating a broad range of cellular functions such as cell survival, migration, adhesion and proliferation. Deregulated activity of RHOA has been linked to the growth, progression and metastasis of various cancer types. Recent cancer genome-wide sequencing studies have unveiled both RHOA gain and loss-of-function mutations in primary leukemia/lymphoma, suggesting that this GTPase may exert tumor-promoting or tumor-suppressive functions depending on the cellular context. Based on these observations, RHOA signaling represents an attractive therapeutic target for the development of selective anticancer strategies. In this review, we will summarize the molecular mechanisms underlying RHOA GTPase functions in immune regulation and in the development of hematological neoplasms and will discuss the current strategies aimed at modulating RHOA functions in these diseases.

1. Introduction

RHO family proteins belonging to the Ras superfamily are highly conserved proteins found in nearly all eukaryotes. In humans, this family comprises 20 members classified in eight subgroups according to their sequence and function [1]. Among them, the Ras homolog gene family member A (RHOA) is the most studied and best characterized protein. RHOA regulates several signal transduction pathways, such as the structure of cytoskeletons, gene expression, vesicle trafficking, cell proliferation, morphogenesis, cell migration, and polarity of cells [2]. It has been shown that RHO GTPases drive cell migration through their ability to promote protrusive force associated with the extension of lamellipodium, the assembly of contractile actin and myosin filaments (stress fibers) and the cell adhesions to the extracellular matrix required to allow the body and rear of the cell to follow the movement [3].
Most Rho GTPases switch between an active GTP-bound and an inactive GDP-bound conformation. The cycling between these states is regulated by three sets of proteins, guanine nucleotide exchange factors (GEFs), GTPase activating proteins (GAPs) and guanine nucleotide dissociation inhibitors (GDIs). GEFs promote the exchange of GDP for GTP necessary for the activation of GTPase, while GAPs catalyze the GTP hydrolysis of the GTPase, thereby inactivating it. GDIs inhibit the activity of RHO proteins by holding GTPases in an inactive state in the cytosol and by preventing the action of GEFs [4]. The unbalance of this process may result in RHO-GTPase-related cell function deregulations, such as migration, polarity, proliferation and survival, which, in turn, may lead to tumorigenesis.
RHOA has long been postulated as an oncogene, mainly due to its ability to transform fibroblasts through induction of Ras [5,6]. The cancer-related role of RHOA has been mostly found in solid tumors, since increased RHOA expression was described in hepatocellular, bladder, liver, ovarian, breast, gastric, colon and lung cancer [7]. Additionally, RhoA was reported to be crucial for cell motility, suggesting an important role for this small GTPase in the invasive phenotype of the tumor cells [8,9,10]. However, human cancer genetic studies have shown that the amplification of RHOA is rare in human cancers. Instead, recent whole genome sequencing analyses have found recurrent loss-of-function RHOA hotspot mutations in gastric cancer [11], angioimmunoblastic T cell lymphoma (AITL) [12,13], adult T-cell leukemia/lymphoma (ATLL) [14], Burkitt lymphoma (BL) [15] and diffuse large B-cell lymphoma (DLBCL) [16], which suggests a rather complex picture of RHOA function depending on cancer cell type.
Another member of the family, RHOB, is generally considered to act as a tumor suppressor gene. First of all, its expression is decreased in several tumor types and its presence is inversely correlated with disease progression. Then, the enzyme has been involved in the disruption of several malignant processes, including tumor growth, cell migration and invasion [17]. Conversely, an overexpression of RHOB has been reported in T-acute lymphoblastic leukemia (T-ALL) compared to primary human T cells [18], suggesting that the functions of this GTPase might be context- and tumor-dependent.
Similar to RHOA, RHOC has also been postulated as an oncogene due to its ability to promote invasion and metastasis in several types of cancer through regulation of cell migration and proliferation [19,20]. RHOC is upregulated in T-ALL cells, in which it regulates reactive oxygen species (ROS) production and the subsequent cytoskeleton rearrangement essential for cell migration [21]. Moreover, RHOC is also implicated in the regulation of phagosome formation in macrophages through the modulation of cytoskeletal remodeling via mammalian diaphanous 1 (mDia1) [22].
Given the complex implications related to the role of individual Rho GTPases in several cancer types, this review will focus specifically on RHOA GTPase, emphasizing the recent findings on RHOA signaling in normal and cancer cell biology. We will discuss the implications of RHOA therapeutic targeting, particularly in hematological tumors.

2. Physiological Functions of RHOA in the Immune System

Several studies demonstrate that RHOA plays an important part in immune responses and is crucial to the development and activation of immune cells [23]. Figure 1 summarized the different roles of RHOA in the different immune compartments.
Hematopoietic progenitor cells (HPCs) are essential to hematopoiesis and the immune response, and their activation and differentiation are highly regulated and controlled. RHOA has an important role in both hematopoietic stem cells (HSC) and in HPCs [24]. A downregulation of RHOA signaling in vitro by an inhibitor of RHO-associated protein kinase (ROCK), Y-27632 or using small interfering RNA, highlighted some changes involved in the morphology and migration of these stem cells [25,26] and also in the engraftment and self-renewal of the HSCs [27,28]. In vivo experiments upon loss of RHOA activity demonstrated that the GTPase is essential for the multilineage hematopoiesis via regulation of actomyosin signaling, cytokinesis and programmed necrosis [29,30,31].

2.1. RHOA GTPase Signaling in Myeloid Cells

In the erythroid lineage, RHOA, other RHO GTPases and CDC42 have been implicated in progenitor commitment and proliferation. RHOA is present in the cytosol and at the membrane of erythrocytes, and its inhibition by Y-27632 results in increased erythrocyte deformability [32,33,34]. At the erythroblast stage, ROCK is found to be activated by caspase-3-mediated cleavage, implying that ROCK is a negative regulator of stress erythropoiesis [35]. Although RAC GTPases promote erythroblast enucleation [35,36], the role of RHOA in this phenomenon is still unclear and an erythroid-specific mouse model will be required to further investigate this issue.
Megakaryocytes are large, polyploid cells of the bone marrow and precursors of platelets. RHOA has been demonstrated to play a critical role in platelet development, as the cytoskeleton rearrangement is important in both the maturation and the initiation of proplatelet formation [37,38]. During cytokinesis, RHOA signaling is required for its completion as it established the actomyosin ring at the cleavage furrow, but it must be downregulated for megakaryocyte polyploidization [39].
In vitro, overexpression of active RHOA resulted in an inhibition of proplatelet formation, suggesting that RHOA is a negative regulator of this process [37,40]. An in vivo deletion of the protein, specifically in megakaryocytes and using the RHOAfl/flPF4CRE+ mice model, caused a significant macrothrombocytopenia and the appearance of larger megakaryocytes characterized by a higher mean ploidy and a defective platelet activation [41,42]. These phenomena were due to the contribution of RHOA to the activation of platelet integrins and to the structure of microtubules in spread platelets; however, the exact mechanism(s) by which RhoA contributes to platelet activation is still under debate [41].
In the myeloid lineage, polymononuclear neutrophils (PMN) are innate immune cells rapidly recruited to sites of invading pathogens and local inflammation. Directed migration towards a gradient of chemoattractant is crucial for the neutrophils. While RHOA expression is reduced in the leading edge, activated RHOA promotes myosin cytoskeleton reorganization at the rear and regulates actomyosin contraction [43,44,45,46]. Its downstream effector, ROCK, regulates cell spreading and uropod formation [47]. Francis et al. demonstrated that mice with lower RHOA activity due to the absence of the RHOA GEF, Lsc, neutrophils showed undirected migration due to a deficient pseudopod formation [48], indicating that RHOA is an essential factor for the maintenance of neutrophil polarity and efficient migration [49].
In eosinophils, the RHO–ROCK axis is necessary for migration and uropod detachment by stimulation of actomyosin filament contractions. Together with RAC and CDC42 in the leading edge, the coordination of the three GTPases are necessary for correct cell migration [50].
Conventional dendritic cells (cDCs) are a subtype of dendritic cells, which act as antigen-presenting cells (APCs). Li et al. nicely demonstrated that a loss of RhoA specifically within DCs, using the CD11cCRE+ mice model, resulted in a decrease in the number of these cells due to cytokinesis failure and increased apoptosis, identifying RhoA as an important regulator in the homeostasis and differentiation of DCs [51]. Additionally, a deficiency in the RhoA regulator Arhgef impaired the chemotaxis and migration of immature DCs in vivo [52] and affected both dendrite formation and motility [53]. Since DCs interact with T cells via the immunological synapse, Shurin G. et al. investigated the formation of these latter and found that RhoA, along with Rac1, were crucial regulators of this process and therefore regulated the DC-mediated response and antigen presentation [54].
Macrophages contribute to tissue homeostasis. Depending on microenvironmental stimuli, monocytes can undergo a polarization towards an M1 macrophage phenotype, characterized by proinflammatory and phagocytic properties, or M2 macrophages, with a more anti-inflammatory phenotype. Liu et al. demonstrated that M0 and M2 macrophages have significantly higher levels of RHOA, and that interfering with the RHOA pathway produced a different polarization and morphological adaptation of M0 and M2 macrophages [55,56]. Indeed, RhoA−/− macrophages showed a defect in their migration property, which simultaneously required the activation of RhoA in the uropod and an inhibition of the GTPase in the podosome to control and direct this migration [57].
During phagocytosis, RHOA is found to be upregulated just prior to the cell engulfment, suggesting that a RAC1/RHOA balance can modulate the critical point of phagocytosis via remodeling of the cytoskeleton [58,59].

2.2. RHOA GTPase in the Control of Lymphoid Lineage

During T cell development, a specific deletion of RhoA in the T cell lineage leads to a defective thymocyte development and a general decrease in the number of immune cells. This effect can be explained partly by the role of RHOA in the survival and the proliferation of T lymphocytes [60]. Moreover, during immune response and inflammation, T cells need to cross endothelial barriers, thereby triggering cell transmigration and polarization, two processes in which RHOA plays a crucial role [61,62]. Heasman et al. described how RHOA is activated both in the leading and rear edge of the T cell for contraction and retraction and that a loss of RHOA led to a loss of this migratory polarity [63]. As mentioned above, RHOA is important for immune synapse formation. An inactivation of RHOA in Jurkat cells inhibited the CD82-induced changes that are necessary for an activation of the T cell receptor (TCR) signaling on T lymphocytes [64].
Additionally, over the past few years, new atypical negative regulators of RHOGTPases have been identified in immune cells [65]. FAM65B, a member of the FAM65 family proteins, also known as RIPOR (RHO family interacting cell polarization regulator) proteins, has been described to bind specifically to RHOA (but not to CDC42 or RAC1) [65]. FAM65B is associated with a quiescent state of T lymphocytes, as it is expressed at high levels by naive cells, whereas activated T cells exhibit a complete loss of this factor. FAM65B was found to be responsible for RHOA activation in resting T cells and has been involved in the inhibition of T cell proliferation upon antigen recognition [66]. Interestingly, chemokine stimulation leads to FAM65B phosphorylation, thereby impeding the binding activity of this latter to RHOA. This process culminates in RHOA activation, which favors T cell motility through cytoskeletal remodeling [67].
RHOA also plays a crucial role in B cell development. In the earlier stages of the B cell differentiation process, a specific deletion of RHOA in the HSC evokes a reduction in the immature B cell and pro-/pre-B cell populations within the bone marrow. A deletion of RhoA specifically in the CD19+ cell population impaired B cell development in the spleen, as shown by a decrease in the percentage of the transitional, follicular and marginal zone B cells [68]. B-cell activating factor receptor (BAFF-R), an important regulator of cell survival, is reduced in RhoA−/− splenic cells, supporting an important role of RHOA in B cell proliferation and differentiation [68]. Finally, B cell receptor (BCR) and phosphoinositide 3-kinase (PI3K) signaling have been described to activate RHOA and, therefore, to promote cell proliferation [69], suggesting this Rho GTPase could be an important regulator of other B cell functions.

3. RhoA in Hematopoietic Tissues: Oncogene or Tumor Suppressor?

RHOA was initially postulated as a pro-oncogenic signal transducer in epithelial cells [70]. Several studies strengthened this hypothesis, showing that RHOA is upregulated in solid tumors, including testicular [71], breast [72,73], colon [73], lung [73] and head and neck squamous cell carcinoma [74]. Moreover, it has been reported that an altered RHOA activity is associated with cancer progression, invasion and angiogenesis in various human cancers. Indeed, RHOA knockdown decreased tumor growth and angiogenesis in a breast cancer MDA-MB-231-xenograft mice model [73,75]. Similarly, RHOA over-activation was observed in gastric tissue and cell lines and has a proven role as a regulator of G1 to S cell cycle transition. Mechanistically, it was observed that the knockdown of mDIA1, a well-known RHOA effector, increases the intracellular levels of the cyclin-dependent kinase (CDK) inhibitors, p21Waf1/Cip1 and p27Kip1, leading to cell cycle blockade. On the other hand, the pharmacological depletion of ROCK led to the downregulation of CDK4 and CDK6 at both mRNA and protein levels [76,77]. Conversely, the functional relevance of RHOA was also shown in lung cancer, where its deletion stimulated lung adenoma formation with a more aggressive phenotype. Of special interest, RhoA and RhoC deletion significantly impaired tumor formation, suggesting that the ablation of the GTPase might lead to a pro-tumoral compensatory mechanism [78].
High-throughput sequencing technologies have identified several gain- and loss-of-function hotspot mutations within the RHOA gene, suggesting, once again, that the GTPase may have tumor-specific functions. Somatic RHOA mutations have been described in about 25% of diffuse gastric cancer patients, affecting R5, G17, Y42 and L57 residues [11,79,80]. Among them, the most frequent mutation is the oncogenic Y42C mutation, which affects the RHOA-effector interactions [81]. The enzyme is also mutated in lung and bladder cancer, with E47K being the mutational hotspot (www.cbioportal.org (accessed on 1 November 2022)). Moreover, RHOA mutation and deletions are also observed in breast cancer, however, at low frequency. It has been suggested that E40Q could be a mutational hotspot, mainly because the same alteration has been frequently described in head-and-neck squamous cell cancer [82]. Although these studies show that RHOA might be implicated in the malignant transformation of solid tumors, recent evidence has demonstrated its tumor suppressor activity in some hematological malignancies.
The development of hematological cancers is a multistep process that consists of dysregulation of cell proliferation, cell death inhibition, cell immortalization, invasion, metastasis and the accumulation of mutations in key genes that may cause primary tumor formation. In 2014, Palomero et al. [83] reported a recurrent mutation in RHOA (G17V) present in 67% of AITL and 18% of peripheral T cell lymphoma (PTCL). The glycine at the 17th position mutation disrupts RHOA interaction with GEFs due to loss of nucleotide binding, which leads to the loss of GTPase activity and inhibition of stress fiber formation (Figure 2). Subsequently, two other groups [13,84] corroborated these findings, showing that RHOA-G17V is a loss-of-function mutation able to drive AITL with high proliferative and invasive capacity. Consistently, Cortes et al. [85] have shown that RHOA-G17V mutation drives proliferation and polarization of human follicular helper T lymphocytes (Tfh), which eventually culminate in AITL development with the upregulation of ICOS (inducible co-stimulator) and activation of PI3K/mitogen-activated protein kinase (MAPK) signaling pathways.
Vav guanine nucleotide exchange factor (VAV1) is a protein stimulated by the engagement of the TCR through GEF activity and is also reported to be mutated in human AITL [86,87]. Interestingly, it has been shown that RHOA-G17V specifically binds to the DH domain of VAV1, thus modulating its function as an adaptor to regulate the TCR signaling complex in AITL [88]. Furthermore, dasatinib, a multikinase inhibitor used to treat some types of leukemias, efficiently blocks VAV1 activation and subsequent RHOA-G17V signaling [88], pointing out a new possible therapeutic approach for AITL patients.
Another potential inactivating RHOA mutation, R5Q, has also been reported in a cohort of pediatric BL patients [15]. Recent evidence based on functional assays showed that, indeed, RHOA-R5Q mutation impairs the activity of the GTPase in BL and DLBCL [16]. Additionally, based on deep sequencing of 203 ATLL samples, Nagata Y et al. have shown that ~15% of cases also harbor recurrent RHOA mutations in the GTP binding domain with different or even opposite functional consequences in terms of RHOA activity, suggesting that both loss- and gain-of-function mutations might be involved in leukemogenesis [14]. Moreover, the data published on the Cancer Genome Atlas database (TCGA; www.cancergenome.nih.gov (accessed on 1 November 2022); www.cbioportal.org (accessed on 1 November 2022)) show that the RHOA gene is rarely amplified and is often deleted in different cancer types, and that RHOA mutations are associated with a worse overall survival of non-Hodgkin lymphoma (NHL) patients. On the other hand, it has been shown that Wnt family member 16 (WNT16b) activates RHOA signaling via a non-canonical WNT pathway and that WNT5a-receptor tyrosine kinase like orphan receptor 1 (ROR1) signaling enhances the proliferation of B-cell precursor acute lymphoblastic leukemia (BCP-ALL) cells, via RHOA and signal transducer and activator of transcription 3 (STAT3) upregulation [89]. Finally, blockage of RHOA activity has been involved in selenite-induced apoptosis of leukemia cells in both in vitro and in vivo models, highlighting the role of the RHOA pathway in the survival of hematologic malignant cells [90].
Clostridium botulinum-derived toxin C3 transferase has long been used to inhibit RHOA activity by abolishing its interaction with downstream effectors. Interestingly, Cleverley S et al. [91] used transgenic mice with thymocyte-specific expression of the enzyme C3-transferase to demonstrate that the loss of RHO function in the thymus resulted in a partial blockade of pre-T cell cycle progression and, eventually, in the development of aggressive thymic lymphoblastic lymphomas. This study suggests that RHOA may act as a suppressor of lymphoid cell transformation. Importantly, mice deficient specifically in regulatory T cells (Treg)-associated RhoA displayed lymphadenopathy, high leukocyte infiltration and an increase in effector T cells and effector memory T cells. RhoA deletion in these mice resulted in systemic autoimmune responses and fatal systemic inflammatory disorders [92].
It was recently shown that miR-126-3p and miR-145-5p can repress RHOA activity in AITL through the regulation of ROCK and S1PR2, a RHOA upstream receptor. Importantly, high levels of miR-126-3p and miR-145-5p were associated with a poor clinical outcome of AITL patients [93].
A recent study showed that CD47, a cell surface transmembrane protein that inhibits phagocytosis, is associated with T-cell lymphoma tumor formation and metastasis by upregulating A-kinase anchor protein 13 (AKAP13)-mediated RHOA activation, supporting the notion that RHOA may exert an oncogenic function in this malignancy [94]. Additionally, Pan Y et al. [95] demonstrated the crucial role of RHOA activation for DLBCL amoeboid motility through activating STAT3, which is a critical hallmark of late-stage progressing disease.
Thus, the recurrent and dominant negative nature of RHOA mutations in AITL, PTCL, BL and DLBCL strongly support a tumor suppressive role for RHOA in hematological cancers. However, the current knowledge on the role of wild-type RHOA in these entities is conflicting and requires additional clarifying evidence in order to obtain a strong rationale for using RHOA-targeting therapy in these tumors.

4. Bioinformatic Prediction of RHOA Therapeutic Agents

The use of bioinformatics tools takes advantage of the abundant amount of available data obtained by different experimental techniques and the power of computational processing [96], reducing time and costs and increasing accuracy. In silico methods for therapeutic target discovery are based on genomic comparative analysis and/or network-based methods. Although the Rho family of GTPases has been linked to different cancers using bioinformatics methods [97,98], to date there are not many applications centered on the role and/or activity of RHOA in hematological cancers.
In comparative genomics [99], the DNA sequence or structure of the phenotypes of interest are used [100]. Whether it is the target gene or a gene that indirectly regulates it, these methods allow us to identify conserved sequences or specific sequences for each phenotype associated with new therapeutic targets or resistance/toxicity. In addition to a direct comparison between conditions of interest, the extensive information deposited in databases (i.e., genetic, metabolic and proteomic) can now be used to address and interpret information from new studies. For this, first, the affected pathway(s) must be defined, using databases such as KEGG [101], and the proteins/genes with differential activity between the phenotypes studied must be selected according to the objectives of the study; then, the information (i.e., sequences) of the important genes or proteins must be retrieved (i.e., using the UniProt database [102]) and some local alignment tool (i.e., BLASTp for proteins and BLASTn for nucleotides) is used to localize and quantify the differences in the compared sequences. Finally, it is necessary to evaluate and validate if the variants (i.e., SNPs) are relevant for patient stratification or targetable.
Network-based methods [103,104] provide two main advantages: integration of different types of data (i.e., clinical, omic, pharmacological) and structuring of information where elements, such as genes or proteins, are represented by nodes, and the relationship between them, qualitative or quantitative, are represented by edges. The focus of this type of analysis can be by centrality, where nodes with more edges and/or are more central in the network would be essential elements for such a network, and by dissimilarity, where elements of the network are related differently or lose relationship, indicating different states of the network for different conditions.
A large number of resources such as databases and bioinformatics tools are available to implement the previously mentioned models. UniProt [102], DEG [105] and RefSeq [106] contain sequence and annotation information for proteins, DNA or both, respectively. Platforms such as KEGG [101] and Reactome [107] contain descriptions of genes and related biomolecules represented as networks of pathways elaborated under some biological criteria (i.e., cell migration, inflammatory response). Other databases are specialized in interactions, such as STRING [108] and MINT [108] for proteins, IntAct [109] and BioGrid [110] that consider other biomolecules such as nucleotides, or miRTarBar [111] for miRNA and their targets. Other databases, such as GEO [112], DrugBank [113], GO [114] and DAVID [115] contain detailed information on molecules of interest that can help in data interpretation. Regarding data processing tools, BLAST [116] and its specialized versions offer a set of very useful analyses and results when comparing amino acid or nucleotide sequences and their functional and evolutionary relationship; MUSCLE [117] is a program that can be used to perform multiple nucleotide or protein alignments with high performance. For network-based methods, we use programs that facilitate visualization and provide structural analysis and enrichment tools, such as Cytoscape [118] and Gephi [119], or platforms that allow us to perform the same tasks with the additional feature of having direct links to databases that facilitate our exploratory analysis, such as NetworkAnalyst [120], Pathwaylinker [121], BioCyc [122], among others.
Some recent studies have applied these methods to the study of RHOA. For example, in one study, different analytical techniques were used to demonstrate variations in genomic sequence (SNV and indels) and structure (CNV), with biochemical validations and docking simulation of the loss of activity to finally integrate these data with gene expression levels and propose a characteristic pathway of AITL caused by the G17V mutation in RHOA [13]. In another study [123], consensus gene sets were defined using the gene set enrichment analysis (GSEA) approach in three collections of gene sets (KEEG, hallmarks and GO) to evidence other regulatory functions of the GTPase, such as cell growth, vesicular transport and DNA replication. Using a network approach and public data mining, Fathima and collaborators defined a list of proteins as potential anticancer drug targets, among which RHOA was identified [124]. For this aim, the study extracted a list of genes relevant in non-apoptotic cell death from literature mining and transcriptome data of some cancers deposited in NCBI’s Gene Expression (GEO, https://www.ncbi.nlm.nih.gov/geo/ (accessed on 1 July 2022)). Then, this list was used as input data for transcription factor (TF)–target and protein–protein interaction databases to then integrate the information in networks and quantify the topological relevance of each element. The most important proteins were queried in DrugBank (https://go.drugbank.com/ (accessed on 1 July 2022)) to find the list of related drugs and proteins that can be considered as new therapeutic targets. Furthermore, miRNAs that could regulate drug target genes were queried. Finally, relevant compounds were used for pathway enrichment and for structural simulation of possible drug binding sites. These types of studies not only evidenced and quantified the importance of proteins such as RHOA in cancer, but also enabled the depicting of their relationship with other biomolecules, structuring abundant information currently available and helping to plan future in vitro/in vivo studies.

5. Approaches to Modulate RHOA Signaling in Hematological Cancers

As illustrated above, the numerous pathways regulated by the RHO GTPase family define multiple aspects important for the proliferation and propagation of malignant cells. Therefore, both RHO GTPases themselves and downstream effectors are attractive targets for therapeutic drug design in cancer (Figure 2).
First, and since RHO GTPases are considered undruggable as they lack stable cavities beside the nucleotide binding pockets, several compounds have been developed, with the ability to modulate multiple stages of the GDP/GTP cycle [125,126]. As stated before, the bacterial exoenzyme C3 is a transferase extracted from Clostridium botulinum that inhibits RHOA, RHOB and RHOC by adding an ADP-ribose moiety on asparagine residues, thereby blocking the small GTPases in an inactive state and keeping them bound to GDI [127]. Since the cell penetrability of the wild-type C3 transferase was very low, and that this factor introduced covalent modifications that limited its clinical development, the cell-permeable C3 transferase Cethrin (BA-210/VX-210) was further developed. Given the importance of the Rho signaling pathway in growth cones, this compound was first shown to prevent damage after spinal cord injury [128]. In vitro studies further demonstrated that blockade of RHOA with Cethrin could enhance the adhesion of T-ALL cells in vitro, a phenomenon that required the polymerization of actin filaments, thus suggesting that activation of RHOA kinase could promote leukocyte deadhesion by inhibiting integrin- and cytoskeletal-dependent spreading of malignant leukemic cells [129].
Other compounds showed activity in preclinical models of hematological neoplasms, such as the inhibitors of RHO GTPase-GEF binding. EHop-016 is an antagonist of RAC1–VAVv2 interaction [130] that can block GTPase activity, thereby improving the antitumor effect of cisplatin in mouse models of esophageal squamous cell carcinoma [131,132]. Validating potent effects of RAC1 inhibitors in acute myeloid leukemia (AML), Hemsing and collaborators recently reported that primary AML cultures, and especially cases harboring a nucleophosmin 1 (NPM1) mutation, presented a good responsiveness to EHop-016 and to its precursor, NSC23766, as well as to three additional RAC1 inhibitors (ZINC69391, ITX3 and 1A-116), with IC50 in the low micromolar range of concentrations [133]. While EHop-016 was developed as an improved NSC23766-derived RAC1 antagonist [134], the parental compound was initially aimed at blocking a surface groove that is critical for the interaction of RAC1 and its associated GEF, Tiam1, without affecting the activity of CDC42 or RHOA [135]. In a preclinical model of chronic myeloid leukemia (CML), either sensitive or resistant to the tyrosine kinase inhibitor (TKI) imatinib, directed against the BCR-ABL fusion gene, NSC23766 displayed an antiproliferative effect in vitro and in vivo and blocked malignant cell dissemination in xenograft studies, supporting the notion that during CML pathogenesis, RHO GTPases can be activated by p210-BCR-ABL [136]. Based on the structure of NSC23766, additional blockers of RAC1 and CDC42 activation have been designed in addition to EHop-016, such as AZA1, which inhibits cellular migration of prostate cancer cells in vivo [137]. Similarly, the CDC42-targeting agent AZA197 impaired the motility of colorectal carcinoma cells [138]. In parallel, the CDC42 activity-specific small molecule inhibitors, Casin and ZCL278, specifically blocked the CDC42 interplay with its GEF, intersectin, suppressing actin-based motility and cell migration without affecting other RHO GTPase activity [139,140]. While ZCL278 was remarkably active in slowing the development of lung cancer in mice [141], the treatment of aged HSCs in a transplantation model allowed for the reestablishment of an immune system similar to that of young animals [142]. In parallel, non-competitive inhibitors, ML141 and EHT1864, targeting, respectively, CDC42 and RAC1, are able to dissociate the bound nucleotide, thus inactivating these GTPases [143]. Of note, the allosteric inhibition of CDC42 signal by the highly selective ML141 [144] revealed that the GTPase is a negative regulator of granulocyte colony-stimulating factor (G-CSF)-mediated HSC mobilization and that the inhibition of its activity by this compound can improve mobilization efficiency [145]. Like EHop-016, EHT1864 interferes with the migration of metastatic breast cancer cells [130,146]. In vitro exposure of AML cells to this agent triggered a time- and dose-dependent cell cycle arrest at the G1 phase, followed by autophagic cell death [147]. These changes were concomitant to alterations in p53 and cyclins expression levels, and to the disruption of the PI3K signaling axis. Interestingly, the combined treatment of EHT1864 and low dose daunorubicin enhanced AML cell apoptosis.
Among the best strategies to prevent the binding of RHOA with its GEFs and to suppress its GTPase activity, the leukemia-associated RHOA GEF, LARG, has been the target of a virtual screening using published protein:protein interactions. This screening identified Rhosin as a compound with the capacity to inhibit specifically the GEF-catalyzed RHOA activation, although it was not selective for RHOA-LARG, as it could also affect other RHO GEFs, such as Dbl, p115RHO GEF and PDZ-RHO GEF [148]. While Rhosin demonstrated interesting preclinical activity in chemotherapy-resistant gastric tumors, in breast cancer and in melanoma [149,150], its suitability for the treatment of blood cancers remains unknown. Y16 is another compound identified through high-throughput screening and is characterized by the capacity to inhibit RGS-containing RHO GEFs. In breast cancer models, this compound could block RHOA activity and prevent the formation of mammary spheroids [151,152].
RHO GTPase activity is tightly dependent on the anchorage of these enzymes at the plasma membrane, with this process being regulated by post-translational modifications that include isoprenylation and palmitoylation [153]. In this context, a specific and first-in-class inhibitor targeting geranylgeranyl transferase-1 (GGT-1) activity, PTX-100, is considered the only RHOA inhibitor in clinical development. Indeed, having demonstrated significant antitumor activity in mouse models of breast cancer [154], this agent was well-tolerated and achieved stable disease in a phase 1 trial in treatment-refractory advanced solid tumors [155]. Currently, PTX-100 is included in a pharmacodynamic and pharmacokinetic basket study involving cancer patients with advanced solid and hematological malignancies, including cases with relapsed/refractory PTCL or multiple myeloma (MM) (NCT03900442). The completion of this study is planned for April 2023.
In addition to this approach, RHO GTPase membrane localization can also be targeted indirectly with statins, as these molecules alter cholesterol biosynthesis by inhibiting 3-hydroxy-3-methylglutaryl-coenzyme A reductase (HMGCR), thereby reducing the intracellular levels of farnesyl pyrophosphate and geranylgeranyl pyrophosphate and retaining the RHO GTPases within the cytosol [156,157]. Statins have been used for years in the treatment of patients with cardiovascular affectations [158]. However, these agents may also target different processes involved in tumor development, such as cancer cell growth, migration, inflammation and angiogenesis, promoting, at the same time, the apoptotic death of the neoplastic cells and reducing cancer-associated mortality [159]. Therefore, this class of drugs plays a positive role in the chemoprevention of cancer and represents excellent candidates for drug repurposing in the management of patients with gynecological [160], brain [161], gastrointestinal [162] or prostate cancer [163]. In hematological neoplasms, a recent meta-analysis of 14 clinical studies did not support a potential role of statins in the prevention of cancer development, with no evidence of an association between statin use and the frequency of hematological disorders in either randomized clinical trials or observational studies [164]. Nonetheless, several preclinical studies have demonstrated the pleiotropic effects of these drugs, as they can elicit apoptosis, block angiogenesis, exert immunomodulatory effects and overcome chemoresistance in different subtypes of leukemias. Indeed, in preclinical models of CML, statins were found to exert an anti-leukemic effect and to synergize with BCR-ABL targeting agents [165,166]. In chronic lymphocytic leukemia (CLL), although some studies suggested that statins may interfere with anti-CD20 therapy due to their ability to alter the conformation of the membrane receptor [167], both lovastatin and simvastatin have been shown to activate the intrinsic apoptotic pathway [168] and to impair mitogen-induced proliferation of primary cultures [169]. However, these promising preclinical results were not validated in clinical settings. Supporting a lack of activity of statins on the natural CLL clinical course, neither time to first treatment of newly diagnosed CLL patients nor the efficacy of anti-CD20 therapy, overall survival or treatment-free survival were dependent on whether or not patients were receiving a statin at diagnosis [170,171]. Among the B-cell subtypes of NHL, and in particular, in DLBCL, which accounts for the most prevalent B-NHL entity, the added values of statins are controversial. Indeed, while patients treated with high-dose statins associated with a standard immunochemotherapeutic (R-CHOP) regimen experienced a significantly higher complete response (CR) rate when compared to patients receiving R-CHOP monotherapy, with an improved 6-years progression free survival (PFS) [172], its seems that this assumption may be ethnicity-dependent. Indeed, in a recent meta-analysis collecting the data from almost 10.000 DLBCL patients, a subgroup analysis failed to show that statin use was associated with overall survival (OS) of patients from Asia, whereas statins were found to be associated with a significantly improved OS of patients from Western countries [173]. Very recently, it has been described that simvastatin exerts single agent activity in in vitro and in vivo models of mantle cell lymphoma (MCL), a rare but very aggressive subtype of B-NHL. In this study, simvastatin exhibited a pleiotropic effect against several MCL hallmarks, including deregulated cell cycle progression, apoptosis, migration and invasion, warranting further clinical evaluation of this agent in clinical settings [174]. Finally, AML is probably the entity that may benefit the most from statin-based therapies. In preclinical models of the disease, statins have been shown to exert single agent activity and to improve the activity of conventional cytotoxic therapy through different mechanisms that involved the modulation of isoprenylation, cholesterol metabolism, leukemic cell differentiation, apoptotic signaling, regulation of protein kinase C (PKC) signaling, intracellular accumulation of intracellular Fms-related receptor tyrosine kinase 3-internal tandem duplication (FLT3-ITD) and consequent blockade of MAPK and AKT signaling [175]. In a phase 1 study assessing the combination of pravastatin with idarubicin-high dose AraC (Ida-HIDAC) including a significant percentage of AML patients with unfavorable cytogenetic/molecular features and/or in salvage therapy, the maximum tolerated dose (MTD) was not reached and the response rate was 30%, while approximately 25% of patients were able to proceed to allogeneic HSC transplant [176]. These encouraging results led to a phase 2 tevaluation of this combination, which demonstrated an impressive rate of CR and CR with incomplete count recovery (CRi) of 75% in this subgroup of patients [177].
Considering the possibility to target RHO GTPase downstream signaling, and in particular the cancer cell metastatic properties, ROCK has emerged as the most studied downstream effector. The antitumor activity of the ROCK inhibitors Y27632 and fasudil have been well characterized and studied preclinically in different types of solid tumors [178,179]. Fasudil exhibits antagonistic activity against serine–threonine kinases, including ROCK [180], to which it binds through the ATP-binding pocket. Fasudil is, so far, the only approved RHO GTPase inhibitor, although in a cancer-unrelated indication (i.e., for the prevention of vasospasm after subarachnoid hemorrhage in Japan). This clinically safe drug demonstrated the capacity to partially control white cell and monocyte counts and to prolong the survival of some animals in a mouse model of Cbl/Cbl-b double knockout (DKO)-associated myeloproliferative disorder [181]. Y-27632 activity on the regulation of the cytoskeleton was first demonstrated in inflammatory diseases [182] and then exploited later on to mitigate WNT5A-dependent cell migration in preclinical models of classical Hodgkin’s disease (cHL) [183]. More selective inhibitors of ROCK include OXA-06 and PT262, RKI-1447 CCT129253 and AT13148, but their evaluation as anti-metastatic agents has been restricted so far to non-hematological diseases [184,185,186,187]. Of note, due to controversial reports on the possible pro-tumoral activity of some of these agents, the sole phase 1 dose-escalation trial carried out in the last years (NCT01585701) is evaluating AT13148 for the treatment of cancer patients with advanced disease. At the maximum dose of 240mg, the most common drug-related toxicities in this study so far have been nausea, anorexia, headache and hypotension. At the present time, data collection has been completed and patient outcome is under evaluation.
Another downstream effector that has been extensively studied to counteract RHO GTPase signaling is p21 (RAC1) activated kinase 1 (PAK1) [188]. IPA-3 is a non-competitive antagonist designed to target PAKs independently of ATP by adding a sulfhydryl moiety to the N-terminal region, thereby promoting the inhibitory conformation [189]. Using a primary culture of progenitor cells and normal hematopoietic counterparts isolated from the bone marrow of newly diagnosed CML patients in chronic phase and CML-blast phase cell lines, IPA-3-mediated targeting of PAK1 and/or PAK2 enhanced the antiproliferative and pro-apoptotic effect of imatinib, without affecting normal counterparts [190]. Due to the unfavorable pharmacological properties of IPA-3, several allosteric inhibitors of PAKs have been designed [191]. Among these molecules, the ATP-competitive PF-3758309 showed promising results in different in vitro and in vivo cancer models, especially in murine and human models of AML and myeloproliferative neoplasms [192]. In this study, the authors demonstrated that the focal adhesion kinase (FAK)/RAC1-TIAM1/PAK1 axis plays a crucial role in the transformation induced by FLT3-ITD and the oncogenic form of KIT (KITD816V), and that targeting PAK1 in oncogene-bearing cells in vitro or in vivo inhibits the presence of active STAT5 in the nuclear compartment. This phenomenon delayed the onset of leukemia by repressing the expression of STAT5-responsive genes. PF-3758309 entered into clinical evaluation in 2009, but the study was closed prematurely due to pharmacokinetic issues [193]. A second molecule, called FRAX597, identified by high-throughput screening and showing potent antitumor activity in preclinical models of Ras-dependent cancers [194], also exhibited potent leukemia inhibitory effects in in vitro and in vivo models of AML, affecting both the leukemic blasts and the more immature, leukemic stem cell (LSC)-enriched populations. The effect of the drug was mediated through the induction of AML cell differentiation and apoptosis, and the modulation of the MYC transcriptional network [195].

6. Conclusions

The recent deciphering of activating or inactivating mutations affecting the RHOA GTPase and/or its complexes networks in hematological cancers strongly support the notion that, similarly to solid tumors, RHOA can exert pro- or anti-tumorigenic functions depending on the cell context and tumor type. In this sense, in determined B-NHL subtypes rather than in chronic or acute leukemias, the enzyme may prevent the development and dissemination of the malignant cells. In order to unravel the identity of key therapeutic targets involved in aberrant RHOA signaling, it will be necessary to achieve a better understanding of the regulating proteins and of the mechanisms underlying the selective recruitment of RHOA downstream effectors in individual disease subtypes. Since a few ATP competitive and non-competitive inhibitors have been developed and entered into the clinic following standard drug development procedures, the implementation of new bioinformatics tools such as networks-based methods and multiomic analyses may be crucial for the design of improved drug candidates that could reasonably be tested in determined hematological diseases.

Author Contributions

Conceptualization, J.C.S. and G.R.; methodology, J.C.S., N.P.-P. and S.S.-V.; software, S.S.-V.; validation, J.C.S. and G.R.; investigation, J.C.S.; resources, J.C.S., N.P.-P. and S.S.-V.; writing—original draft preparation, J.C.S., N.P.-P., S.S.-V. and G.R.; writing—review and editing, J.C.S., N.P.-P., S.S.-V. and G.R.; supervision, G.R.; project administration, G.R.; funding acquisition, G.R. All authors have read and agreed to the published version of the manuscript.

Funding

Research at Roué lab was co-financed by the European Regional Development Fund (ERDF) through the Interreg V-A Spain-France-Andorra (POCTEFA) program (EFA360/19) and the Spanish Ministry of Science and Innovation (grant PID2021-123039OB-C21). J.C.S. holds a Sara Borrell research contract from Instituto de Salud Carlos III (CD19/00228).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank Josep Carreras Leukaemia Research Institute and CERCA Programme/Generalitat de Catalunya for institutional support.

Conflicts of Interest

G.R. received research funding from Onconova Therapeutics. Remaining authors declare no conflict of interest.

References

  1. Vega, F.M.; Ridley, A.J. Rho GTPases in cancer cell biology. FEBS Lett. 2008, 582, 2093–2101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Etienne-Manneville, S.; Hall, A. Rho GTPases in cell biology. Nature 2002, 420, 629–635. [Google Scholar] [CrossRef] [PubMed]
  3. Ridley, A.J. Rho GTPases and cell migration. J. Cell Sci. 2001, 114, 2713–2722. [Google Scholar] [CrossRef] [PubMed]
  4. Hodge, R.G.; Ridley, A.J. Regulating Rho GTPases and their regulators. Nat. Rev. Mol. Cell Biol. 2016, 17, 496–510. [Google Scholar] [CrossRef] [PubMed]
  5. Qiu, R.G.; Chen, J.; Mccormick, F.; Symons, M. A role for Rho in Ras transformation. Proc. Natl. Acad. Sci. USA 1995, 92, 11781–11785. [Google Scholar] [CrossRef] [Green Version]
  6. Khosravi-Far, R.; Solski, P.A.; Clark, G.J.; Kinch, M.S.; Der, C.J. Activation of Rac1, RhoA, and mitogen-activated protein kinases is required for Ras transformation. Mol. Cell. Biol. 1995, 15, 6443–6453. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Karlsson, R.; Pedersen, E.D.; Wang, Z.; Brakebusch, C. Rho GTPase function in tumorigenesis. Biochim. Biophys. Acta-Rev. Cancer 2009, 1796, 91–98. [Google Scholar] [CrossRef] [PubMed]
  8. Murray, D.; Horgan, G.; MacMathuna, P.; Doran, P. NET1-mediated RhoA activation facilitates lysophosphatidic acid-induced cell migration and invasion in gastric cancer. Br. J. Cancer 2008, 99, 1322–1329. [Google Scholar] [CrossRef] [Green Version]
  9. Joshi, R.; Qin, L.; Cao, X.; Zhong, S.; Voss, C.; Min, W.; Li, S.S.C. DLC1 SAM domain-binding peptides inhibit cancer cell growth and migration by inactivating RhoA. J. Biol. Chem. 2020, 295, 645–656. [Google Scholar] [CrossRef]
  10. Kaida, T.; Nitta, H.; Kitano, Y.; Yamamura, K.; Arima, K.; Izumi, D.; Higashi, T.; Kurashige, J.; Imai, K.; Hayashi, H.; et al. C5α receptor (CD88) promotes motility and invasiveness of gastric cancer by activating RhoA. Oncotarget 2016, 7, 84798–84809. [Google Scholar] [CrossRef]
  11. Wang, K.; Yuen, S.T.; Xu, J.; Lee, S.P.; Yan, H.H.N.; Shi, S.T.; Siu, H.C.; Deng, S.; Chu, K.M.; Law, S.; et al. Whole-genome sequencing and comprehensive molecular profiling identify new driver mutations in gastric cancer. Nat. Genet. 2014, 46, 573–582. [Google Scholar] [CrossRef] [PubMed]
  12. Sakata-Yanagimoto, M.; Enami, T.; Yoshida, K.; Shiraishi, Y.; Ishii, R.; Miyake, Y.; Muto, H.; Tsuyama, N.; Sato-Otsubo, A.; Okuno, Y.; et al. Somatic RHOA mutation in angioimmunoblastic T cell lymphoma. Nat. Genet. 2014, 46, 171–175. [Google Scholar] [CrossRef] [PubMed]
  13. Yoo, H.Y.; Sung, M.K.; Lee, S.H.; Kim, S.; Lee, H.; Park, S.; Kim, S.C.; Lee, B.; Rho, K.; Lee, J.E.; et al. A recurrent inactivating mutation in RHOA GTPase in angioimmunoblastic T cell lymphoma. Nat. Genet. 2014, 46, 371–375. [Google Scholar] [CrossRef] [PubMed]
  14. Nagata, Y.; Kontani, K.; Enami, T.; Kataoka, K.; Ishii, R.; Totoki, Y.; Kataoka, T.R.; Hirata, M.; Aoki, K.; Nakano, K.; et al. Variegated RHOA mutations in adult T-cell leukemia/lymphoma. Blood 2016, 127, 596–604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Rohde, M.; Richter, J.; Schlesner, M.; Betts, M.J.; Claviez, A.; Bonn, B.R.; Zimmermann, M.; Damm-Welk, C.; Russell, R.B.; Borkhardt, A.; et al. Recurrent RHOA mutations in pediatric Burkitt lymphoma treated according to the NHL-BFM protocols. Genes Chromosom. Cancer 2014, 53, 911–916. [Google Scholar] [CrossRef]
  16. O’Hayre, M.; Inoue, A.; Kufareva, I.; Wang, Z.; Mikelis, C.M.; Drummond, R.A.; Avino, S.; Finkel, K.; Kalim, K.W.; Dipasquale, G.; et al. Inactivating mutations in GNA13 and RHOA in Burkitt’s lymphoma and diffuse large B-cell lymphoma: A tumor suppressor function for the Gα13/RhoA axis in B cells. Oncogene 2016, 35, 3771–3780. [Google Scholar] [CrossRef]
  17. Huang, M.; Prendergast, G.C. RhoB in cancer suppression. Histol. Histopathol. 2006, 21, 213–218. [Google Scholar] [CrossRef]
  18. Bhavsar, P.J.; Infante, E.; Khwaja, A.; Ridley, A.J. Analysis of Rho GTPase expression in T-ALL identifies RhoU as a target for Notch involved in T-ALL cell migration. Oncogene 2013, 32, 198–208. [Google Scholar] [CrossRef] [Green Version]
  19. Lang, S.; Busch, H.; Boerries, M.; Brummer, T.; Timme, S.; Lassmann, S.; Aktories, K.; Schmidt, G. Specific role of RhoC in tumor invasion and metastasis. Oncotarget 2017, 8, 87364–87378. [Google Scholar] [CrossRef] [Green Version]
  20. Ridley, A.J. RhoA, RhoB and RhoC have different roles in cancer cell migration. J. Microsc. 2013, 251, 242–249. [Google Scholar] [CrossRef]
  21. Luo, J.; Li, D.; Wei, D.; Wang, X.; Wang, L.; Zeng, X. RhoA and RhoC are involved in stromal cell-derived factor-1-induced cell migration by regulating F-actin redistribution and assembly. Mol. Cell. Biochem. 2017, 436, 13–21. [Google Scholar] [CrossRef] [PubMed]
  22. Egami, Y.; Kawai, K.; Araki, N. RhoC regulates the actin remodeling required for phagosome formation during FcγR-mediated phagocytosis. J. Cell Sci. 2017, 130, 4168–4179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Bros, M.; Haas, K.; Moll, L.; Grabbe, S. RhoA as a Key Regulator of Innate and Adaptive Immunity. Cells 2019, 8, 733. [Google Scholar] [CrossRef] [Green Version]
  24. Williams, D.A.; Zheng, Y.; Cancelas, J.A. Rho GTPases and Regulation of Hematopoietic Stem Cell Localization. Methods Enzymol. 2008, 439, 365–393. [Google Scholar] [CrossRef] [PubMed]
  25. Fonseca, A.V.; Freund, D.; Bornhäuser, M.; Corbeil, D. Polarization and migration of hematopoietic stem and progenitor cells rely on the RhoA/ROCK I pathway and an active reorganization of the microtubule network. J. Biol. Chem. 2010, 285, 31661–31671. [Google Scholar] [CrossRef] [Green Version]
  26. Fonseca, A.V.; Corbeil, D. The hematopoietic stem cell polarization and migration: A dynamic link between rhoa signaling pathway, microtubule network and ganglioside-based membrane microdomains. Commun. Integr. Biol. 2011, 4, 201–204. [Google Scholar] [CrossRef]
  27. Ghiaur, G.; Lee, A.; Bailey, J.; Cancelas, J.A.; Zheng, Y.; Williams, D.A. Inhibition of RhoA GTPase activity enhances hematopoietic stem and progenitor cell proliferation and engraftment. Blood 2006, 108, 2087–2094. [Google Scholar] [CrossRef] [Green Version]
  28. Jaganathan, B.G.; Anjos-Afonso, F.; Kumar, A.; Bonnet, D. Active RHOA favors retention of human hematopoietic stem/progenitor cells in their niche. J. Biomed. Sci. 2013, 20, 1–8. [Google Scholar] [CrossRef] [Green Version]
  29. Zhou, X.; Florian, M.C.; Arumugam, P.; Chen, X.; Cancelas, J.A.; Lang, R.; Malik, P.; Geiger, H.; Zheng, Y. RhoA GTPase controls cytokinesis and programmed necrosis of hematopoietic progenitors. J. Exp. Med. 2013, 210, 2371–2385. [Google Scholar] [CrossRef] [Green Version]
  30. Nayak, R.C.; Chang, K.H.; Vaitinadin, N.S.; Cancelas, J.A. Rho GTPases control specific cytoskeleton-dependent functions of hematopoietic stem cells. Immunol. Rev. 2013, 256, 255–268. [Google Scholar] [CrossRef] [PubMed]
  31. Cai, X.; Zhou, X.; Zheng, Y. In vivo rescue assay of rhoa-deficient hematopoietic stem and progenitor cells. Methods Mol. Biol. 2018, 1821, 247–256. [Google Scholar] [CrossRef] [PubMed]
  32. Boukharov, A.A.; Cohen, C.M. Guanine nucleotide-dependent translocation of RhoA from cytosol to high affinity membrane binding sites in human erythrocytes. Biochem. J. 1998, 330, 1391–1398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Thuet, K.M.; Bowles, E.A.; Ellsworth, M.L.; Sprague, R.S.; Stephenson, A.H. The Rho kinase inhibitor Y-27632 increases erythrocyte deformability and low oxygen tension-induced ATP release. Am. J. Physiol.-Hear. Circ. Physiol. 2011, 301, 1891–1896. [Google Scholar] [CrossRef] [Green Version]
  34. Kalfa, T.A.; Pushkaran, S.; Mohandas, N.; Hartwig, J.H.; Fowler, V.M.; Johnson, J.F.; Joiner, C.H.; Williams, D.A.; Zheng, Y. Rac GTPases regulate the morphology and deformability of the erythrocyte cytoskeleton. Blood 2006, 108, 3637–3645. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Gabet, A.S.; Coulon, S.; Fricot, A.; Vandekerckhove, J.; Chang, Y.; Ribeil, J.A.; Lordier, L.; Zermati, Y.; Asnafi, V.; Belaid, Z.; et al. Caspase-activated ROCK-1 allows erythroblast terminal maturation independently of cytokine-induced Rho signaling. Cell Death Differ. 2011, 18, 678–689. [Google Scholar] [CrossRef]
  36. Kalfa, T.A.; Zheng, Y. Rho GTPases in Erythroid Maturation. Curr. Opin. Hematol. 2014, 21, 165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Pleines, I.; Cherpokova, D.; Bender, M. Rho GTPases and their downstream effectors in megakaryocyte biology. Platelets 2019, 30, 9–16. [Google Scholar] [CrossRef] [PubMed]
  38. Smith, E.C.; Teixeira, A.M.; Chen, R.C.; Wang, L.; Gao, Y.; Hahn, K.L.; Krause, D.S. Induction of megakaryocyte differentiation drives nuclear accumulation and transcriptional function of MKL1 via actin polymerization and RhoA activation. Blood 2013, 121, 1094–1101. [Google Scholar] [CrossRef] [Green Version]
  39. Gao, Y.; Smith, E.; Ker, E.; Campbell, P.; Cheng, E.C.; Zou, S.; Lin, S.; Wang, L.; Halene, S.; Krause, D.S. Role of RhoA-Specific Guanine Exchange Factors in Regulation of Endomitosis in Megakaryocytes. Dev. Cell 2012, 22, 573–584. [Google Scholar] [CrossRef] [Green Version]
  40. Chang, Y.; Auradé, F.; Larbret, F.; Zhang, Y.; Le Couedic, J.P.; Momeux, L.; Larghero, J.; Bertoglio, J.; Louache, F.; Cramer, E.; et al. Proplatelet formation is regulated by the Rho/ROCK pathway. Blood 2007, 109, 4229–4236. [Google Scholar] [CrossRef]
  41. Pleines, I.; Hagedorn, I.; Gupta, S.; May, F.; Chakarova, L.; Van Hengel, J.; Offermanns, S.; Krohne, G.; Kleinschnitz, C.; Brakebusch, C.; et al. Megakaryocyte-specific RhoA deficiency causes macrothrombocytopenia and defective platelet activation in hemostasis and thrombosis. Blood 2012, 119, 1054–1063. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Suzuki, A.; Shin, J.W.; Wang, Y.; Min, S.H.; Poncz, M.; Choi, J.K.; Discher, D.E.; Carpenter, C.L.; Lian, L.; Zhao, L.; et al. RhoA Is Essential for Maintaining Normal Megakaryocyte Ploidy and Platelet Generation. PLoS ONE 2013, 8, e69315. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Mulloy, J.C.; Cancelas, J.A.; Filippi, M.D.; Kalfa, T.A.; Guo, F.; Zheng, Y. Rho GTPases in hematopoiesis and hemopathies. Blood 2010, 115, 936–947. [Google Scholar] [CrossRef]
  44. Xu, J.; Wang, F.; Van Keymeulen, A.; Herzmark, P.; Straight, A.; Kelly, K.; Takuwa, Y.; Sugimoto, N.; Mitchison, T.; Bourne, H.R. Divergent Signals and Cytoskeletal Assemblies Regulate Self-Organizing Polarity in Neutrophils. Cell 2003, 114, 201–214. [Google Scholar] [CrossRef]
  45. Van Keymeulen, A.; Wong, K.; Knight, Z.A.; Govaerts, C.; Hahn, K.M.; Shokat, K.M.; Bourne, H.R. To stabilize neutrophil polarity, PIP3 and Cdc42 augment RhoA activity at the back as well as signals at the front. J. Cell Biol. 2006, 174, 437–445. [Google Scholar] [CrossRef]
  46. Kilian, L.S.; Frank, D.; Rangrez, A.Y. RhoA Signaling in Immune Cell Response and Cardiac Disease. Cells 2021, 10, 1681. [Google Scholar] [CrossRef]
  47. Filippi, M.D.; Szczur, K.; Harris, C.E.; Berclaz, P.Y. Rho GTPase Rac1 is critical for neutrophil migration into the lung. Blood 2007, 109, 1257–1264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Francis, S.A.; Shen, X.; Young, J.B.; Kaul, P.; Lerner, D.J. Rho GEF Lsc is required for normal polarization, migration, and adhesion of formyl-peptide–stimulated neutrophils. Blood 2006, 107, 1627–1635. [Google Scholar] [CrossRef] [Green Version]
  49. Hind, L.E.; Vincent, W.J.B.; Huttenlocher, A. Leading from the Back: The Role of the Uropod in Neutrophil Polarization and Migration. Dev. Cell 2016, 38, 161–169. [Google Scholar] [CrossRef] [Green Version]
  50. Alblas, J.; Ulfman, L.; Hordijk, P.; Koenderman, L. Activation of RhoA and ROCK Are Essential for Detachment of Migrating Leukocytes. Mol. Biol. Cell 2001, 12, 2137. [Google Scholar] [CrossRef]
  51. Li, S.; Dislich, B.; Brakebusch, C.H.; Lichtenthaler, S.F.; Brocker, T. Control of Homeostasis and Dendritic Cell Survival by the GTPase RhoA. J. Immunol. 2015, 195, 4244–4256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Wang, Z.; Kumamoto, Y.; Wang, P.; Gan, X.; Lehmann, D.; Smrcka, A.V.; Cohn, L.; Iwasaki, A.; Li, L.; Wu, D. Regulation of immature dendritic cell migration by RHoA guanine nucleotide exchange factor Arhgef5. J. Biol. Chem. 2009, 284, 28599–28606. [Google Scholar] [CrossRef] [Green Version]
  53. Swetman, C.A.; Leverrier, Y.; Garg, R.; Gan, C.H.V.; Ridley, A.J.; Katz, D.R.; Chain, B.M. Extension, retraction and contraction in the formation of a dendritic cell dendrite: Distinct roles for Rho GTPases. Eur. J. Immunol. 2002, 32, 2074–2083. [Google Scholar] [CrossRef] [PubMed]
  54. Shurin, G.V.; Tourkova, I.L.; Chatta, G.S.; Schmidt, G.; Wei, S.; Djeu, J.Y.; Shurin, M.R. Small Rho GTPases Regulate Antigen Presentation in Dendritic Cells. J. Immunol. 2005, 174, 3394–3400. [Google Scholar] [CrossRef] [Green Version]
  55. Liu, Y.; Chen, W.; Minze, L.J.; Kubiak, J.Z.; Li, X.C.; Ghobrial, R.M.; Kloc, M. Dissonant response of M0/M2 and M1 bone-marrow-derived macrophages to RhoA pathway interference. Cell Tissue Res. 2016, 366, 707–720. [Google Scholar] [CrossRef]
  56. Chen, W.; Zhao, Y.; Li, X.C.; Kubiak, J.Z.; Ghobrial, R.M.; Kloc, M. Rho-specific Guanine nucleotide exchange factors (Rho-GEFs) inhibition affects macrophage phenotype and disrupts Golgi complex. Int. J. Biochem. Cell Biol. 2017, 93, 12–24. [Google Scholar] [CrossRef] [PubMed]
  57. Königs, V.; Jennings, R.; Vogl, T.; Horsthemke, M.; Bachg, A.C.; Xu, Y.; Grobe, K.; Brakebusch, C.; Schwab, A.; Bähler, M.; et al. Mouse Macrophages completely lacking Rho subfamily GTPases (RhoA, RhoB, and RhoC) have severe lamellipodial retraction defects, but robust chemotactic navigation and altered motility. J. Biol. Chem. 2014, 289, 30772–30784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Kim, S.Y.; Kim, S.; Bae, D.J.; Park, S.Y.; Lee, G.Y.; Park, G.M.; Kim, I.S. Coordinated balance of Rac1 and RhoA plays key roles in determining phagocytic appetite. PLoS ONE 2017, 12, e0174603. [Google Scholar] [CrossRef] [Green Version]
  59. Kim, J.G.; Moon, M.Y.; Kim, H.J.; Li, Y.; Song, D.K.; Kim, J.S.; Lee, J.Y.; Kim, J.; Kim, S.C.; Park, J.B. Ras-related GTPases Rap1 and RhoA collectively induce the phagocytosis of serum-opsonized zymosan particles in macrophages. J. Biol. Chem. 2012, 287, 5145–5155. [Google Scholar] [CrossRef] [Green Version]
  60. Zhang, S.; Konstantinidis, D.G.; Yang, J.-Q.; Mizukawa, B.; Kalim, K.; Lang, R.A.; Kalfa, T.A.; Zheng, Y.; Guo, F. Gene Targeting RhoA Reveals Its Essential Role in Coordinating Mitochondrial Function and Thymocyte Development. J. Immunol. 2014, 193, 5973–5982. [Google Scholar] [CrossRef]
  61. Heasman, S.J.; Ridley, A.J. Multiple roles for RhoA during T cell transendothelial migration. Small GTPases 2010, 1, 174–179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Vicente-Manzanares, M.; Rey, M.; Pérez-Martínez, M.; Yáñez-Mó, M.; Sancho, D.; Cabrero, J.R.; Barreiro, O.; de la Fuente, H.; Itoh, K.; Sánchez-Madrid, F. The RhoA Effector mDia Is Induced During T Cell Activation and Regulates Actin Polymerization and Cell Migration in T Lymphocytes. J. Immunol. 2003, 171, 1023–1034. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Heasman, S.J.; Carlin, L.M.; Cox, S.; Ng, T.; Ridley, A.J. Coordinated RhoA signaling at the leading edge and uropod is required for T cell transendothelial migration. J. Cell Biol. 2010, 190, 553–563. [Google Scholar] [CrossRef] [Green Version]
  64. Delaguillaumie, A.; Lagaudrière-Gesbert, C.; Popoff, M.R.; Conjeaud, H. Rho GTPase link cytoskeletal rearrangements and activation processes induced via the tetraspanin CD82 in T lymphocytes. J. Cell Sci. 2002, 115, 433–443. [Google Scholar] [CrossRef]
  65. El Masri, R.; Delon, J. RHO GTPases: From new partners to complex immune syndromes. Nat. Rev. Immunol. 2021, 21, 499–513. [Google Scholar] [CrossRef] [PubMed]
  66. Froehlich, J.; Versapuech, M.; Megrelis, L.; Largeteau, Q.; Meunier, S.; Tanchot, C.; Bismuth, G.; Delon, J.; Mangeney, M. FAM65B controls the proliferation of transformed and primary T cells. Oncotarget 2016, 7, 63215–63225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Megrelis, L.; El Ghoul, E.; Moalli, F.; Versapuech, M.; Cassim, S.; Ruef, N.; Stein, J.V.; Mangeney, M.; Delon, J. Fam65b phosphorylation relieves tonic RhoA inhibition during T cell migration. Front. Immunol. 2018, 9, 2001. [Google Scholar] [CrossRef]
  68. Zhang, S.; Zhou, X.; Lang, R.A.; Guo, F. RhoA of the Rho Family Small GTPases Is Essential for B Lymphocyte Development. PLoS ONE 2012, 7, e33773. [Google Scholar] [CrossRef] [Green Version]
  69. Saci, A.; Carpenter, C.L. RhoA GTPase regulates B cell receptor signaling. Mol. Cell 2005, 17, 205–214. [Google Scholar] [CrossRef]
  70. Madaule, P.; Axel, R. A novel ras-related gene family. Cell 1985, 41, 31–40. [Google Scholar] [CrossRef]
  71. Kamai, T.; Yamanishi, T.; Shirataki, H.; Takagi, K.; Asami, H.; Ito, Y.; Yoshida, K.I. Overexpression of RhoA, Rac1, and Cdc42 GTPases is associated with progression in testicular cancer. Clin. Cancer Res. 2004, 10, 4799–4805. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Fritz, G.; Brachetti, C.; Bahlmann, F.; Schmidt, M.; Kaina, B. Rho GTPases in human breast tumours: Expression and mutation analyses and correlation with clinical parameters. Br. J. Cancer 2002, 87, 635–644. [Google Scholar] [CrossRef] [PubMed]
  73. Fritz, G.; Just, I.; Kaina, B. Rho GTPases are over-expressed in human tumors. Int. J. Cancer 1999, 81, 682–687. [Google Scholar] [CrossRef]
  74. Abraham, M.T.; Kuriakose, M.A.; Sacks, P.G.; Yee, H.; Chiriboga, L.; Bearer, E.L.; Delacure, M.D. Motility-related proteins as markers for head and neck squamous cell cancer. Laryngoscope 2001, 111, 1285–1289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Pillé, J.Y.; Denoyelle, C.; Varet, J.; Bertrand, J.R.; Soria, J.; Opolon, P.; Lu, H.; Pritchard, L.L.; Vannier, J.P.; Malvy, C.; et al. Anti-RhoA and Anti-RhoC siRNAs inhibit the proliferation and invasiveness of MDA-MB-231 breast cancer cells in vitro and in vivo. Mol. Ther. 2005, 11, 267–274. [Google Scholar] [CrossRef]
  76. Liu, N.; Bi, F.; Pan, Y.; Sun, L.; Xue, V.; Shi, Y.; Yao, X.; Zheng, Y.; Fan, D. Reversal of the malignant phenotype of gastric cancer cells by inhibition of RhoA expression and activity. Clin. Cancer Res. 2004, 10, 6239–6247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Zhang, S.; Tang, Q.; Xu, F.; Xue, Y.; Zhen, Z.; Deng, Y.; Liu, M.; Chen, J.; Liu, S.; Qiu, M.; et al. RhoA regulates G1-S progression of gastric cancer cells by modulation of multiple INK4 family tumor suppressors. Mol. Cancer Res. 2009, 7, 570–580. [Google Scholar] [CrossRef] [Green Version]
  78. Zandvakili, I.; Davis, A.K.; Hu, G.; Zheng, Y. Loss of RhoA exacerbates, rather than dampens, oncogenic K-Ras induced lung adenoma formation in mice. PLoS ONE 2015, 10, e0127923. [Google Scholar] [CrossRef] [Green Version]
  79. Bass, A.J.; Thorsson, V.; Shmulevich, I.; Reynolds, S.M.; Miller, M.; Bernard, B.; Hinoue, T.; Laird, P.W.; Curtis, C.; Shen, H.; et al. Comprehensive molecular characterization of gastric adenocarcinoma. Nature 2014, 513, 202–209. [Google Scholar] [CrossRef] [Green Version]
  80. Kakiuchi, M.; Nishizawa, T.; Ueda, H.; Gotoh, K.; Tanaka, A.; Hayashi, A.; Yamamoto, S.; Tatsuno, K.; Katoh, H.; Watanabe, Y.; et al. Recurrent gain-of-function mutations of RHOA in diffuse-type gastric carcinoma. Nat. Genet. 2014, 46, 583–587. [Google Scholar] [CrossRef]
  81. Zhang, H.; Schaefer, A.; Wang, Y.; Hodge, R.G.; Blake, D.R.; Diehl, J.N.; Papageorge, A.G.; Stachler, M.D.; Liao, J.; Zhou, J.; et al. Gain-of-function RHOA mutations promote focal adhesion kinase activation and dependency in diffuse gastric cancer. Cancer Discov. 2020, 10, 288–305. [Google Scholar] [CrossRef] [PubMed]
  82. Lawrence, M.S.; Stojanov, P.; Mermel, C.H.; Robinson, J.T.; Garraway, L.A.; Golub, T.R.; Meyerson, M.; Gabriel, S.B.; Lander, E.S.; Getz, G. Discovery and saturation analysis of cancer genes across 21 tumour types. Nature 2014, 505, 495–501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Palomero, T.; Couronné, L.; Khiabanian, H.; Kim, M.Y.; Ambesi-Impiombato, A.; Perez-Garcia, A.; Carpenter, Z.; Abate, F.; Allegretta, M.; Haydu, J.E.; et al. Recurrent mutations in epigenetic regulators, RHOA and FYN kinase in peripheral T cell lymphomas. Nat. Genet. 2014, 46, 166–170. [Google Scholar] [CrossRef] [PubMed]
  84. Manso, R.; Rojo, F.; Rodríguez-Pinilla, S.M.; Sánchez-Beato, M.; Gómez, S.; Monsalvo, S.; Llamas, P.; Cereceda, L.; Piris, M.A.; Mollejo, M.; et al. The RHOA G17V gene mutation occurs frequently in peripheral T-cell lymphoma and is associated with a characteristic molecular signature. Blood 2014, 123, 2893–2894. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Cortes, J.R.; Ambesi-Impiombato, A.; Couronné, L.; Quinn, S.A.; Kim, C.S.; da Silva Almeida, A.C.; West, Z.; Belver, L.; Martin, M.S.; Scourzic, L.; et al. RHOA G17V Induces T Follicular Helper Cell Specification and Promotes Lymphomagenesis. Cancer Cell 2018, 33, 259–273.e7. [Google Scholar] [CrossRef] [Green Version]
  86. Abate, F.; Da Silva-Almeida, A.C.; Zairis, S.; Robles-Valero, J.; Couronne, L.; Khiabanian, H.; Quinn, S.A.; Kim, M.Y.; Laginestra, M.A.; Kim, C.; et al. Activating mutations and translocations in the guanine exchange factor VAV1 in peripheral T-cell lymphomas. Proc. Natl. Acad. Sci. USA 2017, 114, 764–769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Vallois, D.; Dobay, M.P.D.; Morin, R.D.; Lemonnier, F.; Missiaglia, E.; Juilland, M.; Iwaszkiewicz, J.; Fataccioli, V.; Bisig, B.; Roberti, A.; et al. Activating mutations in genes related to TCR signaling in angioimmunoblastic and other follicular helper T-cell-derived lymphomas. Blood 2016, 128, 1490–1502. [Google Scholar] [CrossRef] [Green Version]
  88. Fujisawa, M.; Sakata-Yanagimoto, M.; Nishizawa, S.; Komori, D.; Gershon, P.; Kiryu, M.; Tanzima, S.; Fukumoto, K.; Enami, T.; Muratani, M.; et al. Activation of RHOA-VAV1 signaling in angioimmunoblastic T-cell lymphoma. Leukemia 2018, 32, 694–702. [Google Scholar] [CrossRef] [Green Version]
  89. Karvonen, H.; Perttilä, R.; Niininen, W.; Hautanen, V.; Barker, H.; Murumägi, A.; Heckman, C.A.; Ungureanu, D. Wnt5a and ROR1 activate non-canonical Wnt signaling via RhoA in TCF3-PBX1 acute lymphoblastic leukemia and highlight new treatment strategies via Bcl-2 co-targeting. Oncogene 2019, 38, 3288–3300. [Google Scholar] [CrossRef]
  90. Li, F.; Jiang, Q.; Shi, K.J.; Luo, H.; Yang, Y.; Xu, C.M. RhoA modulates functional and physical interaction between ROCK1 and Erk1/2 in selenite-induced apoptosis of leukaemia cells. Cell Death Dis. 2013, 4, e708. [Google Scholar] [CrossRef]
  91. Cleverley, S.C.; Costello, P.S.; Henning, S.W.; Cantrell, D.A. Loss of Rho function in the thymus is accompanied by the development of thymic lymphoma. Oncogene 2000, 19, 13–20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Kalim, K.W.; Yang, J.Q.; Modur, V.; Nguyen, P.; Li, Y.; Zheng, Y.; Guo, F. Graded RhoA GTPase Expression in Treg Cells Distinguishes Tumor Immunity From Autoimmunity. Front. Immunol. 2021, 12, 726393. [Google Scholar] [CrossRef] [PubMed]
  93. Lone, W.; Bouska, A.; Sharma, S.; Amador, C.; Saumyaranjan, M.; Herek, T.A.; Heavican, T.B.; Yu, J.; Lim, S.T.; Ong, C.K.; et al. Genome-wide miRNA expression profiling of molecular subgroups of peripheral T-cell lymphoma. Clin. Cancer Res. 2021, 27, 6039–6053. [Google Scholar] [CrossRef]
  94. Kitai, Y.; Ishiura, M.; Saitoh, K.; Matsumoto, N.; Owashi, K.; Yamada, S.; Muromoto, R.; Kashiwakura, J.I.; Oritani, K.; Matsuda, T. CD47 promotes T-cell lymphoma metastasis by up-regulating AKAP13-mediated RhoA activation. Int. Immunol. 2021, 33, 273–280. [Google Scholar] [CrossRef]
  95. Pan, Y.R.; Chen, C.C.; Chan, Y.T.; Wang, H.J.; Chien, F.T.; Chen, Y.L.; Liu, J.L.; Yang, M.H. STAT3-coordinated migration facilitates the dissemination of diffuse large B-cell lymphomas. Nat. Commun. 2018, 9, 1–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Zheng, H.; Zhang, G.; Zhang, L.; Wang, Q.; Li, H.; Han, Y.; Xie, L.; Yan, Z.; Li, Y.; An, Y.; et al. Comprehensive Review of Web Servers and Bioinformatics Tools for Cancer Prognosis Analysis. Front. Oncol. 2020, 10, 68. [Google Scholar] [CrossRef] [PubMed]
  97. Bayo, J.; Fiore, E.J.; Dominguez, L.M.; Cantero, M.J.; Ciarlantini, M.S.; Malvicini, M.; Atorrasagasti, C.; Garcia, M.G.; Rossi, M.; Cavasotto, C.; et al. Bioinformatic analysis of RHO family of GTPases identifies RAC1 pharmacological inhibition as a new therapeutic strategy for hepatocellular carcinoma. Gut 2021, 70, 1362–1374. [Google Scholar] [CrossRef] [PubMed]
  98. Aspenström, P. Activated Rho GTPases in Cancer—The Beginning of a New Paradigm. Int. J. Mol. Sci. 2018, 19, 3949. [Google Scholar] [CrossRef] [Green Version]
  99. Jakobsen, N.A.; Vyas, P. From genomics to targeted treatment in haematological malignancies: A focus on acute myeloid leukaemia. Clin. Med. 2018, 18, s47–s53. [Google Scholar] [CrossRef]
  100. Hardison, R.C. Comparative Genomics. PLoS Biol. 2003, 1, e58. [Google Scholar] [CrossRef]
  101. Kanehisa, M.; Furumichi, M.; Sato, Y.; Ishiguro-Watanabe, M.; Tanabe, M. KEGG: Integrating viruses and cellular organisms. Nucleic Acids Res. 2021, 49, D545–D551. [Google Scholar] [CrossRef] [PubMed]
  102. Bateman, A. UniProt: A worldwide hub of protein knowledge. Nucleic Acids Res. 2019, 47, D506. [Google Scholar] [CrossRef] [Green Version]
  103. Madeddu, L.; Stilo, G. Deep learning methods for network biology. In Deep Learning in Biology and Medicine; Bacciu, D., Lisboa, P.J.G., Vellido, A., Eds.; World Scientific Publishing Group: London, UK, 2022. [Google Scholar] [CrossRef]
  104. Rocca, A.; Kholodenko, B.N. Can Systems Biology Advance Clinical Precision Oncology? Cancers 2021, 13, 6312. [Google Scholar] [CrossRef] [PubMed]
  105. Luo, H.; Lin, Y.; Gao, F.; Zhang, C.T.; Zhang, R. DEG 10, an update of the database of essential genes that includes both protein-coding genes and noncoding genomic elements. Nucleic Acids Res. 2014, 42, D574. [Google Scholar] [CrossRef] [Green Version]
  106. Pruitt, K.D.; Tatusova, T.; Maglott, D.R. NCBI reference sequences (RefSeq): A curated non-redundant sequence database of genomes, transcripts and proteins. Nucleic Acids Res. 2007, 35, D61. [Google Scholar] [CrossRef] [Green Version]
  107. Gillespie, M.; Jassal, B.; Stephan, R.; Milacic, M.; Rothfels, K.; Senff-Ribeiro, A.; Griss, J.; Sevilla, C.; Matthews, L.; Gong, C.; et al. The reactome pathway knowledgebase 2022. Nucleic Acids Res. 2022, 50, D687–D692. [Google Scholar] [CrossRef]
  108. Szklarczyk, D.; Gable, A.L.; Lyon, D.; Junge, A.; Wyder, S.; Huerta-Cepas, J.; Simonovic, M.; Doncheva, N.T.; Morris, J.H.; Bork, P.; et al. STRING v11: Protein–protein association networks with increased coverage, supporting functional discovery in genome-wide experimental datasets. Nucleic Acids Res. 2019, 47, D607. [Google Scholar] [CrossRef] [Green Version]
  109. Orchard, S.; Ammari, M.; Aranda, B.; Breuza, L.; Briganti, L.; Broackes-Carter, F.; Campbell, N.H.; Chavali, G.; Chen, C.; Del-Toro, N.; et al. The MIntAct project—IntAct as a common curation platform for 11 molecular interaction databases. Nucleic Acids Res. 2014, 42, D358. [Google Scholar] [CrossRef] [Green Version]
  110. Oughtred, R.; Rust, J.; Chang, C.; Breitkreutz, B.J.; Stark, C.; Willems, A.; Boucher, L.; Leung, G.; Kolas, N.; Zhang, F.; et al. The BioGRID database: A comprehensive biomedical resource of curated protein, genetic, and chemical interactions. Protein Sci. 2021, 30, 187. [Google Scholar] [CrossRef]
  111. Huang, H.Y.; Lin, Y.C.D.; Li, J.; Huang, K.Y.; Shrestha, S.; Hong, H.C.; Tang, Y.; Chen, Y.G.; Jin, C.N.; Yu, Y.; et al. miRTarBase 2020: Updates to the experimentally validated microRNA–target interaction database. Nucleic Acids Res. 2020, 48, D148. [Google Scholar] [CrossRef] [Green Version]
  112. Barrett, T.; Wilhite, S.E.; Ledoux, P.; Evangelista, C.; Kim, I.F.; Tomashevsky, M.; Marshall, K.A.; Phillippy, K.H.; Sherman, P.M.; Holko, M.; et al. NCBI GEO: Archive for functional genomics data sets—Update. Nucleic Acids Res. 2013, 41, D991. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Wishart, D.S.; Feunang, Y.D.; Guo, A.C.; Lo, E.J.; Marcu, A.; Grant, J.R.; Sajed, T.; Johnson, D.; Li, C.; Sayeeda, Z.; et al. DrugBank 5.0: A major update to the DrugBank database for 2018. Nucleic Acids Res. 2018, 46, D1074–D1082. [Google Scholar] [CrossRef] [PubMed]
  114. Carbon, S.; Douglass, E.; Good, B.M.; Unni, D.R.; Harris, N.L.; Mungall, C.J.; Basu, S.; Chisholm, R.L.; Dodson, R.J.; Hartline, E.; et al. The Gene Ontology resource: Enriching a GOld mine. Nucleic Acids Res. 2021, 49, D325. [Google Scholar] [CrossRef]
  115. Huang, D.W.; Sherman, B.T.; Lempicki, R.A. Systematic and integrative analysis of large gene lists using DAVID bioinformatics resources. Nat. Protoc. 2009, 4, 44–57. [Google Scholar] [CrossRef]
  116. Boratyn, G.M.; Camacho, C.; Cooper, P.S.; Coulouris, G.; Fong, A.; Ma, N.; Madden, T.L.; Matten, W.T.; McGinnis, S.D.; Merezhuk, Y.; et al. BLAST: A more efficient report with usability improvements. Nucleic Acids Res. 2013, 41, W29–W33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Edgar, R.C. MUSCLE: Multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 2004, 32, 1792. [Google Scholar] [CrossRef] [Green Version]
  118. Shannon, P.; Markiel, A.; Ozier, O.; Baliga, N.S.; Wang, J.T.; Ramage, D.; Amin, N.; Schwikowski, B.; Ideker, T. Cytoscape: A software Environment for integrated models of biomolecular interaction networks. Genome Res. 2003, 13, 2498–2504. [Google Scholar] [CrossRef]
  119. Bastian, M.; Heymann, S.; Jacomy, M. Gephi: An Open Source Software for Exploring and Manipulating Networks. Proc. Int. AAAI Conf. Web Soc. Media 2009, 3, 361–362. [Google Scholar] [CrossRef]
  120. Zhou, G.; Soufan, O.; Ewald, J.; Hancock, R.E.W.; Basu, N.; Xia, J. NetworkAnalyst 3.0: A visual analytics platform for comprehensive gene expression profiling and meta-analysis. Nucleic Acids Res. 2019, 47, W234. [Google Scholar] [CrossRef]
  121. Farkas, I.J.; Szántó-Várnagy, Á.; Korcsmáros, T. Linking proteins to signaling pathways for experiment design and evaluation. PLoS ONE 2012, 7, e36202. [Google Scholar] [CrossRef] [Green Version]
  122. Karp, P.D.; Billington, R.; Caspi, R.; Fulcher, C.A.; Latendresse, M.; Kothari, A.; Keseler, I.M.; Krummenacker, M.; Midford, P.E.; Ong, Q.; et al. The BioCyc collection of microbial genomes and metabolic pathways. Brief. Bioinform. 2019, 20, 1085. [Google Scholar] [CrossRef]
  123. Chun, K.H. Discovery of Cellular RhoA Functions by the Integrated Application of Gene Set Enrichment Analysis. Biomol. Ther. 2022, 30, 98–116. [Google Scholar] [CrossRef] [PubMed]
  124. Fathima, S.; Sinha, S.; Donakonda, S. Network analysis identifies drug targets and small molecules to modulate apoptosis resistant cancers. Cancers 2021, 13, 851. [Google Scholar] [CrossRef]
  125. Maldonado, M.d.M.; Dharmawardhane, S. Targeting Rac and Cdc42 GTPases in Cancer. Cancer Res. 2018, 78, 3101–3111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Jansen, S.; Gosens, R.; Wieland, T.; Schmidt, M. Paving the Rho in cancer metastasis: Rho GTPases and beyond. Pharmacol. Ther. 2018, 183, 1–21. [Google Scholar] [CrossRef]
  127. Vogelsgesang, M.; Pautsch, A.; Aktories, K. C3 exoenzymes, novel insights into structure and action of Rho-ADP-ribosylating toxins. Naunyn. Schmiedebergs. Arch. Pharmacol. 2007, 374, 347–360. [Google Scholar] [CrossRef]
  128. Lord-Fontaine, S.; Yang, F.; Diep, Q.; Dergham, P.; Munzer, S.; Tremblay, P.; McKerracher, L. Local Inhibition of Rho Signaling by Cell-Permeable Recombinant Protein BA-210 Prevents Secondary Damage and Promotes Functional Recovery following Acute Spinal Cord Injury. J. Neurotrauma 2008, 25, 1309–1322. [Google Scholar] [CrossRef] [PubMed]
  129. Liu, L.; Schwartz, B.R.; Lin, N.; Winn, R.K.; Harlan, J.M. Requirement for RhoA kinase activation in leukocyte de-adhesion. J. Immunol. 2002, 169, 2330–2336. [Google Scholar] [CrossRef] [Green Version]
  130. Montalvo-Ortiz, B.L.; Castillo-Pichardo, L.; Hernández, E.; Humphries-Bickley, T.; De La Mota-Peynado, A.; Cubano, L.A.; Vlaar, C.P.; Dharmawardhane, S. Characterization of EHop-016, Novel Small Molecule Inhibitor of Rac GTPase. J. Biol. Chem. 2012, 287, 13228–13238. [Google Scholar] [CrossRef]
  131. Zeng, R.-J.; Zheng, C.-W.; Gu, J.-E.; Zhang, H.-X.; Xie, L.; Xu, L.-Y.; Li, E.-M. RAC1 inhibition reverses cisplatin resistance in esophageal squamous cell carcinoma and induces downregulation of glycolytic enzymes. Mol. Oncol. 2019, 13, 2010–2030. [Google Scholar] [CrossRef] [Green Version]
  132. Ito, M.A.; Kojima, E.; Yanagihara, Y.; Yoshida, K.; Matsuoka, I. Differential Effects of Gq Protein-Coupled Uridine Receptor Stimulation on IL-8 Production in 1321N1 Human Astrocytoma Cells. Biol. Pharm. Bull. 2022, 45, 691–697. [Google Scholar] [CrossRef]
  133. Hemsing, A.L.; Rye, K.P.; Hatfield, K.J.; Reikvam, H. NPM1-Mutated Patient-Derived AML Cells Are More Vulnerable to Rac1 Inhibition. Biomedicines 2022, 10, 1881. [Google Scholar] [CrossRef] [PubMed]
  134. Akbar, H.; Cancelas, J.; Williams, D.A.; Zheng, J.; Zheng, Y. Rational design and applications of a Rac GTPase-specific small molecule inhibitor. Methods Enzymol. 2006, 406, 554–565. [Google Scholar] [CrossRef] [PubMed]
  135. Gao, Y.; Dickerson, J.B.; Guo, F.; Zheng, J.; Zheng, Y. Rational design and characterization of a Rac GTPase-specific small molecule inhibitor. Proc. Natl. Acad. Sci. USA 2004, 101, 7618–7623. [Google Scholar] [CrossRef] [Green Version]
  136. Thomas, E.K.; Cancelas, J.A.; Chae, H.-D.; Cox, A.D.; Keller, P.J.; Perrotti, D.; Neviani, P.; Druker, B.J.; Setchell, K.D.R.; Zheng, Y.; et al. Rac Guanosine Triphosphatases Represent Integrating Molecular Therapeutic Targets for BCR-ABL-Induced Myeloproliferative Disease. Cancer Cell 2007, 12, 467–478. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Zins, K.; Lucas, T.; Reichl, P.; Abraham, D.; Aharinejad, S. A Rac1/Cdc42 GTPase-Specific Small Molecule Inhibitor Suppresses Growth of Primary Human Prostate Cancer Xenografts and Prolongs Survival in Mice. PLoS ONE 2013, 8, e74924. [Google Scholar] [CrossRef] [Green Version]
  138. Zins, K.; Gunawardhana, S.; Lucas, T.; Abraham, D.; Aharinejad, S. Targeting Cdc42 with the small molecule drug AZA197 suppresses primary colon cancer growth and prolongs survival in a preclinical mouse xenograft model by downregulation of PAK1 activity. J. Transl. Med. 2013, 11, 295. [Google Scholar] [CrossRef] [Green Version]
  139. Florian, M.C.; Dörr, K.; Niebel, A.; Daria, D.; Schrezenmeier, H.; Rojewski, M.; Filippi, M.-D.; Hasenberg, A.; Gunzer, M.; Scharffetter-Kochanek, K.; et al. Cdc42 Activity Regulates Hematopoietic Stem Cell Aging and Rejuvenation. Cell Stem Cell 2012, 10, 520–530. [Google Scholar] [CrossRef] [Green Version]
  140. Friesland, A.; Zhao, Y.; Chen, Y.-H.; Wang, L.; Zhou, H.; Lu, Q. Small molecule targeting Cdc42–intersectin interaction disrupts Golgi organization and suppresses cell motility. Proc. Natl. Acad. Sci. USA 2013, 110, 1261–1266. [Google Scholar] [CrossRef]
  141. Aguilar, B.J.; Zhao, Y.; Zhou, H.; Huo, S.; Chen, Y.-H.; Lu, Q. Inhibition of Cdc42–intersectin interaction by small molecule ZCL367 impedes cancer cell cycle progression, proliferation, migration, and tumor growth. Cancer Biol. Ther. 2019, 20, 740–749. [Google Scholar] [CrossRef]
  142. Leins, H.; Mulaw, M.; Eiwen, K.; Sakk, V.; Liang, Y.; Denkinger, M.; Geiger, H.; Schirmbeck, R. Aged murine hematopoietic stem cells drive aging-associated immune remodeling. Blood 2018, 132, 565–576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Shutes, A.; Onesto, C.; Picard, V.; Leblond, B.; Schweighoffer, F.; Der, C.J. Specificity and Mechanism of Action of EHT 1864, a Novel Small Molecule Inhibitor of Rac Family Small GTPases. J. Biol. Chem. 2007, 282, 35666–35678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Hong, L.; Kenney, S.R.; Phillips, G.K.; Simpson, D.; Schroeder, C.E.; Nöth, J.; Romero, E.; Swanson, S.; Waller, A.; Strouse, J.J.; et al. Characterization of a Cdc42 Protein Inhibitor and Its Use as a Molecular Probe. J. Biol. Chem. 2013, 288, 8531–8543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Chen, C.; Song, X.; Ma, S.; Wang, X.; Xu, J.; Zhang, H.; Wu, Q.; Zhao, K.; Cao, J.; Qiao, J.; et al. Cdc42 inhibitor ML141 enhances G-CSF-induced hematopoietic stem and progenitor cell mobilization. Int. J. Hematol. 2015, 101, 5–12. [Google Scholar] [CrossRef] [PubMed]
  146. Katz, E.; Sims, A.H.; Sproul, D.; Caldwell, H.; Michael Dixon, J.; Meehan, R.R.; Harrison, D.J. Targeting of Rac GTPases blocks the spread of intact human breast cancer. Oncotarget 2012, 3, 608–619. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Ramos, D.F.V.; Mancuso, R.I.; Contieri, B.; Duarte, A.; Paiva, L.; de Melo Carrilho, J.; Saad, S.T.O.; Lazarini, M. Rac GTPases in acute myeloid leukemia cells: Expression profile and biological effects of pharmacological inhibition. Toxicol. Appl. Pharmacol. 2022, 442, 115990. [Google Scholar] [CrossRef]
  148. Shang, X.; Marchioni, F.; Sipes, N.; Evelyn, C.R.; Jerabek-Willemsen, M.; Duhr, S.; Seibel, W.; Wortman, M.; Zheng, Y. Rational Design of Small Molecule Inhibitors Targeting RhoA Subfamily Rho GTPases. Chem. Biol. 2012, 19, 699–710. [Google Scholar] [CrossRef] [Green Version]
  149. Yoon, C.; Cho, S.-J.; Aksoy, B.A.; Park, D.J.; Schultz, N.; Ryeom, S.W.; Yoon, S.S. Chemotherapy Resistance in Diffuse-Type Gastric Adenocarcinoma Is Mediated by RhoA Activation in Cancer Stem-Like Cells. Clin. Cancer Res. 2016, 22, 971–983. [Google Scholar] [CrossRef] [Green Version]
  150. Tsubaki, M.; Genno, S.; Takeda, T.; Matsuda, T.; Kimura, N.; Yamashita, Y.; Morii, Y.; Shimomura, K.; Nishida, S. Rhosin Suppressed Tumor Cell Metastasis through Inhibition of Rho/YAP Pathway and Expression of RHAMM and CXCR4 in Melanoma and Breast Cancer Cells. Biomedicines 2021, 9, 35. [Google Scholar] [CrossRef]
  151. Evelyn, C.R.; Ferng, T.; Rojas, R.J.; Larsen, M.J.; Sondek, J.; Neubig, R.R. High-Throughput Screening for Small-Molecule Inhibitors of LARG-Stimulated RhoA Nucleotide Binding via a Novel Fluorescence Polarization Assay. SLAS Discov. 2000, 14, 161–172. [Google Scholar] [CrossRef] [Green Version]
  152. Shang, X.; Marchioni, F.; Evelyn, C.R.; Sipes, N.; Zhou, X.; Seibel, W.; Wortman, M.; Zheng, Y. Small-molecule inhibitors targeting G-protein–coupled Rho guanine nucleotide exchange factors. Proc. Natl. Acad. Sci. USA 2013, 110, 3155–3160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Mitin, N.; Roberts, P.J.; Chenette, E.J.; Der, C.J. Posttranslational lipid modification of rho family small GTPases. Methods Mol. Biol. 2012, 827, 87–95. [Google Scholar] [CrossRef] [PubMed]
  154. Aspenström, P.; Ruusala, A.; Pacholsky, D. Taking Rho GTPases to the next level: The cellular functions of atypical Rho GTPases. Exp. Cell Res. 2007, 313, 3673–3679. [Google Scholar] [CrossRef]
  155. Karasic, T.B.; Chiorean, E.G.; Sebti, S.M.; O’Dwyer, P.J. A Phase I Study of GGTI-2418 (Geranylgeranyl Transferase I Inhibitor) in Patients with Advanced Solid Tumors. Target. Oncol. 2019, 14, 613–618. [Google Scholar] [CrossRef] [PubMed]
  156. Denoyelle, C.; Vasse, M.; Körner, M.; Mishal, Z.; Ganné, F.; Vannier, J.-P.; Soria, J.; Soria, C. Cerivastatin, an inhibitor of HMG-CoA reductase, inhibits the signaling pathways involved in the invasiveness and metastatic properties of highly invasive breast cancer cell lines: An in vitro study. Carcinogenesis 2001, 22, 1139–1148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Gazzerro, P.; Proto, M.C.; Gangemi, G.; Malfitano, A.M.; Ciaglia, E.; Pisanti, S.; Santoro, A.; Laezza, C.; Bifulco, M. Pharmacological Actions of Statins: A Critical Appraisal in the Management of Cancer. Pharmacol. Rev. 2012, 64, 102–146. [Google Scholar] [CrossRef]
  158. Tobert, J.A. Lovastatin and beyond: The history of the HMG-CoA reductase inhibitors. Nat. Rev. Drug Discov. 2003, 2, 517–526. [Google Scholar] [CrossRef]
  159. Beckwitt, C.H.; Brufsky, A.; Oltvai, Z.N.; Wells, A. Statin drugs to reduce breast cancer recurrence and mortality. Breast Cancer Res. 2018, 20, 144. [Google Scholar] [CrossRef]
  160. Markowska, A.; Antoszczak, M.; Markowska, J.; Huczyński, A. Statins: HMG-CoA Reductase Inhibitors as Potential Anticancer Agents against Malignant Neoplasms in Women. Pharmaceuticals 2020, 13, 422. [Google Scholar] [CrossRef]
  161. Afshari, A.R.; Mollazadeh, H.; Henney, N.C.; Jamialahmad, T.; Sahebkar, A. Effects of statins on brain tumors: A review. Semin. Cancer Biol. 2021, 73, 116–133. [Google Scholar] [CrossRef]
  162. Gachpazan, M.; Kashani, H.; Khazaei, M.; Hassanian, S.M.; Rezayi, M.; Asgharzadeh, F.; Ghayour-Mobarhan, M.; Ferns, G.A.; Avan, A. The Impact of Statin Therapy on the Survival of Patients with Gastrointestinal Cancer. Curr. Drug Targets 2018, 20, 738–747. [Google Scholar] [CrossRef] [PubMed]
  163. Papadopoulos, G.; Delakas, D.; Nakopoulou, L.; Kassimatis, T. Statins and prostate cancer: Molecular and clinical aspects. Eur. J. Cancer 2011, 47, 819–830. [Google Scholar] [CrossRef] [PubMed]
  164. Bonovas, S.; Filioussi, K.; Tsantes, A.; Sitaras, N.M. Use of statins and risk of haematological malignancies: A meta-analysis of six randomized clinical trials and eight observational studies. Br. J. Clin. Pharmacol. 2007, 64, 255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Chen, R.; Xiao, W.; Li, D.; Mu, S. Combination of simvastatin and imatinib sensitizes the CD34+ cells in K562 to cell death. Med. Oncol. 2011, 28, 528–531. [Google Scholar] [CrossRef]
  166. Yang, Y.C.; Huang, W.F.; Chuan, L.M.; Xiao, D.W.; Zeng, Y.L.; Zhou, D.A.; Xu, G.Q.; Liu, W.; Huang, B.; Hu, Q. In vitro and in vivo study of cell growth inhibition of simvastatin on chronic myelogenous leukemia cells. Chemotherapy 2008, 54, 438–446. [Google Scholar] [CrossRef]
  167. Winiarska, M.; Bil, J.; Wilczek, E.; Wilczynski, G.M.; Lekka, M.; Engelberts, P.J.; Mackus, W.J.M.; Gorska, E.; Bojarski, L.; Stoklosa, T.; et al. Statins Impair Antitumor Effects of Rituximab by Inducing Conformational Changes of CD20. PLoS Med. 2008, 5, e64. [Google Scholar] [CrossRef] [Green Version]
  168. Chapman-Shimshoni, D.; Yuklea, M.; Radnay, J.; Shapiro, H.; Lishner, M. Simvastatin induces apoptosis of B-CLL cells by activation of mitochondrial caspase 9. Exp. Hematol. 2003, 31, 779–783. [Google Scholar] [CrossRef]
  169. Vitols, S.; Angelin, B.; Juliusson, G. Simvastatin impairs mitogen-induced proliferation of malignant B-lymphocytes from humans—In vitro and in vivo studies. Lipids 1997, 32, 255–262. [Google Scholar] [CrossRef]
  170. Shanafelt, T.D.; Rabe, K.G.; Kay, N.E.; Zent, C.S.; Call, T.G.; Slager, S.L.; Bowen, D.A.; Schwager, S.M.; Nowakowski, G.S. Statin and non-steroidal anti-inflammatory drug use in relation to clinical outcome among patients with Rai stage 0 chronic lymphocytic leukemia. Leuk. Lymphoma 2010, 51, 1233–1240. [Google Scholar] [CrossRef]
  171. Friedman, D.R.; Magura, L.A.; Warren, H.A.C.; Harrison, J.D.; Diehl, L.F.; Weinberg, J.B. Statin use and need for therapy in chronic lymphocytic leukemia. Leuk. Lymphoma 2010, 51, 2295–2298. [Google Scholar] [CrossRef]
  172. Gouni, S.; Strati, P.; Toruner, G.; Aradhya, A.; Landgraf, R.; Bilbao, D.; Vega, F.; Agarwal, N.K. Statins enhance the chemosensitivity of R-CHOP in diffuse large B-cell lymphoma. Leuk. Lymphoma 2022, 63, 1302–1313. [Google Scholar] [CrossRef] [PubMed]
  173. Li, Y.; Zhou, H.; Zou, L. Influence of Statins on the Survival Outcomes of Patients with Diffuse Large B Cell Lymphoma: A Systematic Review and Meta-Analysis. Int. J. Clin. Pract. 2022, 2022, 5618290. [Google Scholar] [CrossRef] [PubMed]
  174. Santos, J.C.; Profitós-Pelejà, N.; Ribeiro, M.L.; Roué, G. Antitumor Activity of Simvastatin in Preclinical Models of Mantle Cell Lymphoma. Cancers 2022, 14, 5601. [Google Scholar] [CrossRef] [PubMed]
  175. Van Besien, H.; Sassano, A.; Altman, J.K.; Platanias, L.C. Antileukemic properties of 3-hydroxy-3-methylglutaryl-coenzyme A reductase inhibitors. Leuk. Lymphoma 2013, 54, 2601–2605. [Google Scholar] [CrossRef]
  176. Kornblau, S.M.; Banker, D.E.; Stirewalt, D.; Shen, D.; Lemker, E.; Verstovsek, S.; Estrov, Z.; Faderl, S.; Cortes, J.; Beran, M.; et al. Blockade of adaptive defensive changes in cholesterol uptake and synthesis in AML by the addition of pravastatin to idarubicin + high-dose Ara-C: A phase 1 study. Blood 2007, 109, 2999–3006. [Google Scholar] [CrossRef] [PubMed]
  177. Advani, A.S.; Mcdonough, S.; Copelan, E.; Willman, C.; Mulford, D.A.; List, A.F.; Sekeres, M.A.; Othus, M.; Appelbaum, F.R. SWOG0919: A Phase 2 study of idarubicin and cytarabine in combination with pravastatin for relapsed acute myeloid leukaemia. Br. J. Haematol. 2014, 167, 233–237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Kale, V.P.; Hengst, J.A.; Desai, D.H.; Amin, S.G.; Yun, J.K. The regulatory roles of ROCK and MRCK kinases in the plasticity of cancer cell migration. Cancer Lett. 2015, 361, 185–196. [Google Scholar] [CrossRef] [PubMed]
  179. Rath, N.; Olson, M.F. Rho-associated kinases in tumorigenesis: Re-considering ROCK inhibition for cancer therapy. EMBO Rep. 2012, 13, 900–908. [Google Scholar] [CrossRef] [Green Version]
  180. Nagumo, H.; Sasaki, Y.; Ono, Y.; Okamoto, H.; Seto, M.; Takuwa, Y. Rho kinase inhibitor HA-1077 prevents Rho-mediated myosin phosphatase inhibition in smooth muscle cells. Am. J. Physiol. Physiol. 2000, 278, C57–C65. [Google Scholar] [CrossRef]
  181. William, B.M.; An, W.; Feng, D.; Nadeau, S.; Mohapatra, B.C.; Storck, M.A.; Band, V.; Band, H. Fasudil, a clinically safe ROCK inhibitor, decreases disease burden in a Cbl/Cbl-b deficiency-driven murine model of myeloproliferative disorders. Hematology 2016, 21, 218–224. [Google Scholar] [CrossRef] [Green Version]
  182. Narumiya, S.; Ishizaki, T.; Uehata, M. Use and properties of ROCK-specific inhibitor Y-27632. Methods Enzymol. 2000, 325, 273–284. [Google Scholar] [CrossRef]
  183. Linke, F.; Zaunig, S.; Nietert, M.M.; VonBonin, F.; Lutz, S.; Dullin, C.; Janovská, P.; Beissbarth, T.; Alves, F.; Klapper, W.; et al. WNT5A: A motility-promoting factor in Hodgkin lymphoma. Oncogene 2017, 36, 13–23. [Google Scholar] [CrossRef] [PubMed]
  184. Tsai, C.C.; Liu, H.F.; Hsu, K.C.; Yang, J.M.; Chen, C.; Liu, K.K.; Hsu, T.S.; Chao, J.I. 7-Chloro-6-piperidin-1-yl-quinoline-5,8-dione (PT-262), a novel ROCK inhibitor blocks cytoskeleton function and cell migration. Biochem. Pharmacol. 2011, 81, 856–865. [Google Scholar] [CrossRef] [PubMed]
  185. Vigil, D.; Kim, T.Y.; Plachco, A.; Garton, A.J.; Castaldo, L.; Pachter, J.A.; Dong, H.; Chen, X.; Tokar, B.; Campbell, S.L.; et al. ROCK1 and ROCK2 Are Required for Non-Small Cell Lung Cancer Anchorage-Independent Growth and Invasion. Cancer Res. 2012, 72, 5338–5347. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Patel, R.A.; Forinash, K.D.; Pireddu, R.; Sun, Y.; Sun, N.; Martin, M.P.; Schönbrunn, E.; Lawrence, N.J.; Sebti, S.M. RKI-1447 Is a Potent Inhibitor of the Rho-Associated ROCK Kinases with Anti-Invasive and Antitumor Activities in Breast Cancer. Cancer Res. 2012, 72, 5025–5034. [Google Scholar] [CrossRef] [Green Version]
  187. Sadok, A.; McCarthy, A.; Caldwell, J.; Collins, I.; Garrett, M.D.; Yeo, M.; Hooper, S.; Sahai, E.; Kuemper, S.; Mardakheh, F.K.; et al. Rho Kinase Inhibitors Block Melanoma Cell Migration and Inhibit Metastasis. Cancer Res. 2015, 75, 2272–2284. [Google Scholar] [CrossRef] [Green Version]
  188. Semenova, G.; Chernoff, J. Targeting PAK1. Biochem. Soc. Trans. 2017, 45, 79–88. [Google Scholar] [CrossRef]
  189. Deacon, S.W.; Beeser, A.; Fukui, J.A.; Rennefahrt, U.E.E.; Myers, C.; Chernoff, J.; Peterson, J.R. An Isoform-Selective, Small-Molecule Inhibitor Targets the Autoregulatory Mechanism of p21-Activated Kinase. Chem. Biol. 2008, 15, 322–331. [Google Scholar] [CrossRef] [Green Version]
  190. Flis, S.; Bratek, E.; Chojnacki, T.; Piskorek, M.; Skorski, T. Simultaneous inhibition of BCR-ABL1 tyrosine kinase and PAK1/2 serine/threonine kinase exerts synergistic effect against chronic myeloid leukemia cells. Cancers 2019, 11, 1544. [Google Scholar] [CrossRef] [PubMed]
  191. Kim, D.J.; Choi, C.K.; Lee, C.S.; Park, M.H.; Tian, X.; Kim, N.D.; Lee, K.I.; Choi, J.K.; Ahn, J.H.; Shin, E.Y.; et al. Small molecules that allosterically inhibit p21-activated kinase activity by binding to the regulatory p21-binding domain. Exp. Mol. Med. 2016, 48, e229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Chatterjee, A.; Ghosh, J.; Ramdas, B.; Mali, R.S.; Martin, H.; Kobayashi, M.; Vemula, S.; Canela, V.H.; Waskow, E.R.; Visconte, V.; et al. Regulation of Stat5 by FAK and PAK1 in Oncogenic FLT3- and KIT-Driven Leukemogenesis. Cell Rep. 2014, 9, 1333–1348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Murray, B.W.; Guo, C.; Piraino, J.; Westwick, J.K.; Zhang, C.; Lamerdin, J.; Dagostino, E.; Knighton, D.; Loi, C.M.; Zager, M.; et al. Small-molecule p21-activated kinase inhibitor PF-3758309 is a potent inhibitor of oncogenic signaling and tumor growth. Proc. Natl. Acad. Sci. USA 2010, 107, 9446–9451. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Chow, H.Y.; Jubb, A.M.; Koch, J.N.; Jaffer, Z.M.; Stepanova, D.; Campbell, D.A.; Duron, S.G.; O’Farrell, M.; Cai, K.Q.; Klein-Szanto, A.J.P.; et al. p21-Activated kinase 1 is required for efficient tumor formation and progression in a Ras-mediated skin cancer model. Cancer Res. 2012, 72, 5966–5975. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Pandolfi, A.; Stanley, R.F.; Yu, Y.; Bartholdy, B.; Pendurti, G.; Gritsman, K.; Boultwood, J.; Chernoff, J.; Verma, A.; Steidl, U. PAK1 is a therapeutic target in acute myeloid leukemia and myelodysplastic syndrome. Blood 2015, 126, 1118–1127. [Google Scholar] [CrossRef]
Figure 1. Functions of RhoA signaling in the immune system.
Figure 1. Functions of RhoA signaling in the immune system.
Cells 12 00433 g001
Figure 2. RhoA alterations and therapeutic targeting RhoA-associated signaling in hematological cancers.
Figure 2. RhoA alterations and therapeutic targeting RhoA-associated signaling in hematological cancers.
Cells 12 00433 g002
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

Santos, J.C.; Profitós-Pelejà, N.; Sánchez-Vinces, S.; Roué, G. RHOA Therapeutic Targeting in Hematological Cancers. Cells 2023, 12, 433. https://doi.org/10.3390/cells12030433

AMA Style

Santos JC, Profitós-Pelejà N, Sánchez-Vinces S, Roué G. RHOA Therapeutic Targeting in Hematological Cancers. Cells. 2023; 12(3):433. https://doi.org/10.3390/cells12030433

Chicago/Turabian Style

Santos, Juliana Carvalho, Núria Profitós-Pelejà, Salvador Sánchez-Vinces, and Gaël Roué. 2023. "RHOA Therapeutic Targeting in Hematological Cancers" Cells 12, no. 3: 433. https://doi.org/10.3390/cells12030433

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