1,276
Views
0
CrossRef citations to date
0
Altmetric
Review Article

KIF2C/MCAK a prognostic biomarker and its oncogenic potential in malignant progression, and prognosis of cancer patients: a systematic review and meta-analysis as biomarker

ORCID Icon, , , , ORCID Icon, ORCID Icon & ORCID Icon show all
Received 31 Oct 2023, Accepted 22 Jan 2024, Published online: 12 Feb 2024

Abstract

KIF2C/MCAK (KIF2C) is the most well-characterized member of the kinesin-13 family, which is critical in the regulation of microtubule (MT) dynamics during mitosis, as well as interphase. This systematic review briefly describes the important structural elements of KIF2C, its regulation by multiple molecular mechanisms, and its broad cellular functions. Furthermore, it systematically summarizes its oncogenic potential in malignant progression and performs a meta-analysis of its prognostic value in cancer patients. KIF2C was shown to be involved in multiple crucial cellular processes including cell migration and invasion, DNA repair, senescence induction and immune modulation, which are all known to be critical during the development of malignant tumors. Indeed, an increasing number of publications indicate that KIF2C is aberrantly expressed in multiple cancer entities. Consequently, we have highlighted its involvement in at least five hallmarks of cancer, namely: genome instability, resisting cell death, activating invasion and metastasis, avoiding immune destruction and cellular senescence. This was followed by a systematic search of KIF2C/MCAK’s expression in various malignant tumor entities and its correlation with clinicopathologic features. Available data were pooled into multiple weighted meta-analyses for the correlation between KIF2Chigh protein or gene expression and the overall survival in breast cancer, non-small cell lung cancer and hepatocellular carcinoma patients. Furthermore, high expression of KIF2C was correlated to disease-free survival of hepatocellular carcinoma. All meta-analyses showed poor prognosis for cancer patients with KIF2Chigh expression, associated with a decreased overall survival and reduced disease-free survival, indicating KIF2C’s oncogenic potential in malignant progression and as a prognostic marker. This work delineated the promising research perspective of KIF2C with modern in vivo and in vitro technologies to further decipher the function of KIF2C in malignant tumor development and progression. This might help to establish KIF2C as a biomarker for the diagnosis or evaluation of at least three cancer entities.

View correction statement:
Correction

1. Background

Orchestrated regulation of microtubule (MT) dynamics is critical for various cellular functions, including cell division of all eukaryotic cells [Citation1]. During cell division, the MT-based mitotic spindle is required to guide a faithful chromosome separation into two identical daughter cells [Citation2]. The dynamic MT structure consists of α- and β-tubulin polymers and its assembly/disassembly is regulated by a great number of different protein families such as motor proteins, microtubule associated proteins (MAPs) and kinases [Citation1,Citation3]. The microtubule organizing centers (MTOCs) are the main nucleation centers for MTs and are called centrosomes in animal and human cells [Citation4]. These structures typically combine two main mechanisms: the nucleation of novel MT filaments and the anchoring of minus-end MTs, which are essential for the stability of the MT cytoskeleton [Citation5]. The growth and shrinkage phases at the plus MT-ends are driven by two distinct states of MT subunits, the GTP- and GDP-bound tubulin dimers, which have different structural properties, binding kinetics as well as interaction partners [Citation6]. The model of a “GTP cap,” which prevents MT catastrophe, was established by Voter and Erickson in 1984 [Citation7] and was further corroborated by using non-hydrolyzable GTP analogs like GMPCPP [Citation8]. In line with these notions, the degradation of the “GTP cap” leads to the induction of MT catastrophe, which suggests a competition between the growing plus-end, by adding de novo GTP-tubulin subunits and the steady hydrolysis of GTP in the MT lattice [Citation8,Citation9]. This steady process of dynamic instability is a key process in spindle assembly, chromosome separation, organization of the cytoskeleton during mitosis and interphase, and the exploration of the extracellular environment (ECV) [Citation6,Citation10]. MT dynamics are vital in these cellular functions, and are orchestrated by various regulators. More than 663 MAPs are identified in eukaryotes, characterized by their ability to bind and interact with the MT lattice, tubulin heterodimers, MT plus-ends and their function to regulate MT dynamics [Citation11,Citation12], including MT stabilizing proteins such as Tau [Citation13] and cytoplasmic linker-associated protein (CLASP) [Citation14], MT modulator RBP-J interacting and tubulin-associated protein (RITA) [Citation15] and depolymerizing proteins including the highly conserved non-motile kinesin-13 family members [Citation16,Citation17].

2. Materials and methods

2.1. Search strategy

This in-depth systematic review about the prognostic value of KIF2C/MCAK (hereafter revered as KIF2C) follows the “Preferred Reporting Items for Systematic Reviews and Meta-Analyses” (PRISMA) guidelines [Citation18]. Databases used for the article retrieval were PubMed/NCBI, Web of Science as well as the preprint databases bioRxiv and Researchsquare. All databases were scanned with the search strategy shown in Supplemental Table 1 for Pubmed/NCBI or a strategy modified for the individual database. All literature screening was performed by both first authors and the last author of this systematic review.

2.2. Eligibility

The inclusion and exclusion criteria were defined prior to data acquisition. This systematic review is focused on studies using KIF2C/MCAK as a prognostic marker in any cancer entity, and includes publications in the English language between 1996 and 2023 (date: 05.06.2023). Abstracts from conferences and reviews were excluded, and pre-prints were intentionally included in the paper acquisition. A detailed description of the screening process can be found in Supplement Figure 1. Specific exclusion criteria concerning cancer subtypes, mutations or any other variations were not considered. The primary readout for all articles was defined as the prognostic value of KIF2C/MCAK gene or protein expression in the overall survival (OS) of cancer patients. The secondary objective was the correlation between KIF2C/MCAK and the disease-free survival (DSF) in all cancer patients, and the tertiary objective was to acquire all other clinicopathological features correlated to the expression of KIF2C/MCAK. Risk and bias calculations were not performed.

Figure 1. Illustration of KIF2C’s phosphorylation sites, domains, interaction partners and molecular functions. (A): KIF2C’s domains, functions and phosphorylation sites are depicted. The phosphorylation sites are marked as green for Aurora A, blue for Aurora B, light blue for Cdk1, red for Plk1, purple for PAK1 and black for a phosphorylation by an unknown kinase. (B): schematic illustration of KIF2C/MCAK’s molecular involvement in regulating chromosome congression/segregation, FA turnover, ciliogenesis and DSB repair. The figure was modified and updated from [Citation17]. Abbreviations: Cdk1: Cyclin-dependent kinase 1; Plk1: Polo-like kinase 1; PAK1: p21-activated kinase; EB1: end-binding protein 1; FA: focal adhesion; DSB: double-strand break; NT: NH2-terminus; CT: carboxy-terminus; NHEJ: non-homologous end joining; HR: homologous recombination.

Figure 1. Illustration of KIF2C’s phosphorylation sites, domains, interaction partners and molecular functions. (A): KIF2C’s domains, functions and phosphorylation sites are depicted. The phosphorylation sites are marked as green for Aurora A, blue for Aurora B, light blue for Cdk1, red for Plk1, purple for PAK1 and black for a phosphorylation by an unknown kinase. (B): schematic illustration of KIF2C/MCAK’s molecular involvement in regulating chromosome congression/segregation, FA turnover, ciliogenesis and DSB repair. The figure was modified and updated from [Citation17]. Abbreviations: Cdk1: Cyclin-dependent kinase 1; Plk1: Polo-like kinase 1; PAK1: p21-activated kinase; EB1: end-binding protein 1; FA: focal adhesion; DSB: double-strand break; NT: NH2-terminus; CT: carboxy-terminus; NHEJ: non-homologous end joining; HR: homologous recombination.

2.3. Data extraction

The obtained data were extracted into two independent tables. includes the correlation between KIF2C/MCAK expression and the clinicopathological features of eleven cancer entities. summarizes molecular mechanisms involved in KIF2C oncogenic potential, separated into tumor entities. This systematic review is focused on studies using KIF2C/MCAK as a prognostic marker in any cancer entity. Third the data concerning the OS/DSF and the expression of KIF2C/MCAK was extracted into a meta-analysis visualized in by forest plots portraying the hazard ratio of both clinicopathological features. To prevent a selection bias, all studies were included regardless of their statistical significance. Weighted scores (ws), based on the patient cohort of the studies, were generated and used to calculate the mean OS/DFS value for each meta-analysis.

Table 1. Correlation between KIF2C/MCAK expression and the clinicopathologic features of various cancer entities.

Table 2. Molecular mechanisms involved in KIF2Cs oncogenic potential in multiple cancer entities/cell lines.

3. Kinesin-13 family

Kinesins are ATP-driven motor proteins with the ability to move across the MT lattice, mostly toward MT plus-ends, gaining the required energy by hydrolysis of ATP subunits [Citation19]. They fulfill multiple functions such as the intracellular transport of macromolecules, vesicles or cell organelles, regulating MT dynamics, MT morphogenesis, chromosome segregation, spindle formation and elongation [Citation20]. The entire kinesin superfamily is highly diverse and has been subdivided into 14 distinct families based on phylogeny, structure and sequence homology [Citation21]. While many kinesins are MT polymerizing or stabilizing proteins such as kinesin-5 [Citation22], this superfamily also contains proteins with the ability to induce MT depolymerization, the so-called kinesin-13 family, also known as middle motor domain type (m-type) [Citation19]. Members of the kinesin-13 family depolymerize MTs by using energy from ATP hydrolysis and remove α- and β-tubulin subunits from the MT polymer [Citation23]. Four different members of the kinesin-13 family, namely KIF2A, KIF2B, KIF2C (also known as mitotic centromere-associated kinesin) and KIF24 have been described in the human genome [Citation24]. They have functions in spindle assembly, chromosome segregation, MT-kinetochore attachment, migration and cilia formation [Citation25]. The potent depolymerization activity of the kinesin-13 family depends on the effective direct targeting of MT-ends or by one-dimensional diffusion on the MT lattice [Citation17]. This specific targeting of the MT-end is supported by the positively charged and highly conserved neck domain of the kinesin-13 members [Citation26]. At the MT-ends, these proteins induce MT catastrophe through their enzymatic ability to cause bending of a single tubulin protofilament, leading to the release of α- and β-tubulin dimers [Citation25].

The best described member of the kinesin-13 family is KIF2C [Citation27]. It plays an essential role in the modulation of MT dynamics. KIF2C is found frequently at centromeres, kinetochore MTs and astral MTs [Citation28–30]. The protein has various functions during mitosis, in spindle assembly, chromosome congression, kinetochore-MT attachment and chromosome segregation [Citation31–34]. In addition to its roles during mitosis, KIF2C is important in cytoskeleton remodeling, especially in the regulation of MT dynamics during migration, invasion and focal adhesion (FA) turnover [Citation35–38]. Recently, KIF2C was described to have a negative role in ciliogenesis [Citation39] as well as in DNA double strand break dynamics and repair [Citation40]. KIF2A, is predominantly located at centrosomes where it affects the spindles dynamics and is therefore essential for both bipolar spindle assembly and chromosome movement [Citation24,Citation41,Citation42]. Additionally, KIF2A contributes to pole-directed chromosome movement in anaphase by disassembling MTs at their minus-ends at the spindle poles in association with a poleward MT flux [Citation43].

Interestingly, KIF2A also plays a role in meiotic cells. It has been shown that KIF2A localizes to the meiotic spindle, specifically to the spindle poles and inner centromeres in metaphase, and was translocated to the midbody in telophase. Depletion of KIF2A led to severely defective spindles [Citation44]. Besides its function during cell division, KIF2A is essentially involved in primary cilia disassembly [Citation45] and in cortical neuron migration and differentiation [Citation46]. KIF2B was reported to promote faithful chromosome segregation [Citation47,Citation48]. It is expressed at low levels in almost all human tissue types except testis. [Citation49,Citation50]. KIF24, the fourth member of the family, shares significant homologies with other members of the kinesin-13 subfamily and is involved in MT depolymerization activity in cilium formation [Citation51], clustering of supernumerary centrosomes [Citation52], disassembly and maturation of centrioles [Citation53,Citation54]. Altogether, these data emphasize the crucial roles of the kinesin-13 family in regulating MT dynamics through all phases of the cell cycle. Interestingly, most of these functions are known to be highly deregulated during cancer development and progression.

3.1. KIF2C domains, structure, localization and mitotic regulation

The protein structure of KIF2C can be divided into four functional units (). The N-terminal domain contains regions important for dimerization, centromere binding, generation of anchoring forces and the binding to EB1/EB3 with its SxIP motif, while the C-terminal tail regulates KIF2C’s tubulin interaction, stability, conformation and enzymatic activity [Citation17,Citation55]. The stability of KIF2C is mainly regulated by its phosphorylation at S621 on the C-terminus, which facilitates its recognition by the ubiquitin/proteasome dependent APC (anaphase-promoting complex)/C (cyclosome) pathway leading to its D-box dependent degradation in mitosis [Citation29]. Moreover, a study using deuterium exchange mass spectrometry showed that the closed N-terminal form of KIF2C imparts the outward curvature and lateral separation of α/β-MT dimers [Citation56]. KIF2C’s neck region is responsible for its enzymatic activity and conformation changes. Enzymatic regulation is highly dependent on the positively charged neck structure, which can be modulated by several phosphorylation sites including the highly conserved phosphorylation site S192 for multiple kinases, including Aurora A and B [Citation37,Citation57]. The X-ray crystallographic structure of KIF2C with additionally docked tubulin dimers has revealed that the neck and KVD motif of KIF2C are able to directly interact with the α-tubulin distal end [Citation16,Citation58]. The catalytic core domain generates the energy required for MT depolymerization by ATP hydrolysis and is responsible for a robust MT binding [Citation59]. At the C-terminus is a domain that, like the N-terminus, is responsible for the conformational change and stable tubulin interaction [Citation17]. KIF2C functions in a dimeric state, which increases its robust MT depolymerization activity, and the dimeric state also facilitates attachment to MT-ends [Citation26,Citation59–62]. KIF2C interacts directly with MTs on multiple sites, including its helix a4, the loop 2 structure with its KVD finger motif, the loop 8 in the motor domain and the positively charged neck region [Citation58,Citation63,Citation64].

All of these functional domains are regulated by multiple phosphorylation sites, which modulate both the activity and localization of KIF2C, as shown by multiple investigations using phospho-antibodies and site-directed mutagenesis [Citation26,Citation64]. Key players in this phosphorylation events are the mitotic kinases Aurora A/B, Plk1 (Polo-like kinase 1), Cdk1 (Cyclin-dependent kinase 1), and PAK1 (p21-activated kinase 1) [Citation17]. One of the major phosphorylation sites regulating the depolymerization activity of KIF2C is S192, which is phosphorylated by Aurora A/B and PAK1 [Citation37,Citation65,Citation66]. Both phosphorylation and dephosphorylation of this site affect the catalytic activity of human KIF2C in cancer and normal cells [Citation37,Citation57]. Furthermore, interference with this phosphorylation site partly arrests cancer cells in metaphase, and induces congression and segregation defects [Citation37]. In support, Aurora B also phosphorylates the conserved serine S196 in Xenopus leavis. This phosphorylation leads to increased localization of KIF2C near the centromere, which downregulates its depolymerization activity, especially in the crucial kinetochore-centromere region [Citation67,Citation68]. Functionally, the Aurora B phosphorylation at S196 in the neck region triggers a change in the conformation of KIF2C and reduces the interaction between the C-terminus and the neck region, which significantly impairs its MT affinity [Citation69]. Additionally, Aurora B regulates the localization of KIF2C’s subcellular localization by phosphorylation of five different sites (S92, S106, S108, S112 and S186) in hamster KIF2C [Citation30]. While the localization of the phospho-mimetic KIF2C 5xE is increased in the centromere region and causes aberrant mitotic spindles, the non-phosphorylatable KIF2C 5xA is localized predominantly in the kinetochore region with reduced binding to TIP150 (tracking protein of 150 KDa) [Citation30]. Like Aurora B, Aurora A targets the same phosphorylation sites and phosphorylates primarily 5 sites (S95, S109, S111, S115, S192) in the N-terminus and neck region of KIF2C [Citation70]. Further, KIF2C is co-localized with nuclear mitotic apparatus (NuMA) protein and Xenopus microtubule-associated protein (XMAP215) at the center of Ran-asters, where its activity is regulated by Aurora A-dependent phosphorylation of S196, resulting in proper pole focusing [Citation70]. In addition, the Aurora A phosphorylation of the site S719 increases KIF2C spindle pole focusing ability [Citation70,Citation71]. Another key kinase in the regulation of KIF2C is Plk1, which phosphorylates the site S632/S633 and increases the activity of KIF2C during the early phases of mitosis [Citation72]. Plk1 also phosphorylates S621 of KIF2C without a required priming phosphorylation by another kinase. This phosphorylation facilitates the recognition by the Cdc20 subunit of the APC/C complex, leading to its D-box dependent degradation by the proteasome at the transition from metaphase to anaphase [Citation29]. The fourth important mitotic kinase in regulating KIF2C is Cdk1, which phosphorylates T537 in the core domain of KIF2C and thereby enhances the catalytic activity of KIF2C [Citation73]. Interfering with this phosphorylation generates substantial mitotic defects during chromosome congression and segregation [Citation73]. A T537E mutant disrupts the recognition of the microtubule end, decreasing its capacity to stimulate ADP dissociation at the MT minus-end [Citation74]. Another regulatory mechanism of KIF2C is the binding to MTs, as well as tip tracking proteins such as EB1 and TIP150. Phosphorylation of the neck domain abolishes these interactions, highly reducing KIF2C’s depolymerization activity, especially at the MT plus-ends [Citation34,Citation75]. Additionally, binding to the MAP nucleolar spindle-associated protein (NuSAP) significantly enhances the activity of KIF2C on kinetochore MTs [Citation76]. In summary, these observations highlight the crucial importance of various mitotic proteins in regulating the fine-tuned localization and activity of KIF2C via diverse molecular mechanisms during mitosis and beyond.

3.2. KIF2C regulation of cell polarization, migration and invasion

The process of cell polarization, migration and invasion is highly dynamic and involves all three filament types of the cytoskeleton: actin microfilaments, intermediate filaments and MTs [Citation77]. Key functions of MTs in these processes are highly complex and include three biological mechanisms. The long-distance intracellular transport of proteins and vesicles, the delivery of new membrane components and signaling molecules essential for the assembly of structures transmitting traction forces and the recycling of adhesion components as well as receptors [Citation77,Citation78]. MTs are physically connected to FA sites, which are specialized structures within the cell, where various integrin receptors interact with the ECM (extracellular matrix) components and the actin cytoskeleton inside the cells [Citation79]. MTs facilitate not only FA assembly by transporting vesicles and structure proteins to the forming FAs, but are also essential for the assembly of FAs by transporting autophagosomes and other components to the retrenching FAs [Citation77,Citation80].

It is therefore hardly surprising that KIF2C has been linked to MT dynamics during interphase, polarization of ECs, assembly and disassembly of FAs and migration/invasion of cancer as well as benign cells [Citation35–37,Citation57]. In detail, KIF2C has been shown to modulate the MT growth speed and branching in compliant 2D and 3D ECM in ECs [Citation81]. Mechanistically, the KIF2C-RAC1 (Rac Family Small GTPase 1) pathway is regulated by Aurora A. The kinase ensures and regulates EC polarization and directional migration by regulating MT dynamics through KIF2C on the leading edge of cells () [Citation35]. In detail, the activity of RAC1 on the trailing edge is able to enhance the kinase activity of Aurora A, which in turn is able to modulate KIF2C’s MT depolymerization activity as shown in mitosis by phosphorylation of S196 [Citation35]. In line with this, the immortalization of human bronchial epithelial cells by expressing K-Ras (Kirsten rat sarcoma virus) and knockdown of p53, resulted in a simultaneous expression of KIF2C and increased the dynamic instability of MTs and enhanced cell migration [Citation82]. Moreover, KIF2C was shown to regulate cell polarity, protrusion formation, centrosome reorientation during migration and the FA turnover by regulating the plus-tip dynamics in human cervical cancer cells and retinal pigment epithelial cells [Citation36,Citation57]. Overall, these reports indicate KIF2C’s involvement in cell migration and invasion of cancer cells [Citation36,Citation37,Citation57,Citation83].

4. Altered KIF2C is involved in several hallmarks of cancer

The hallmarks of cancer propose a panel of functional alterations that a cell has to acquire to change its phenotype from a benign cell to a neoplastic growth state and its ability to form malignant tumors [Citation84,Citation85]. These hallmarks include eight characteristics and six capabilities, including (a) genome instability, (b) resistance to cell death, (c) activating invasion/metastasis and (d) avoiding immune destruction, (e) cellular senescence [Citation84,Citation86], all of which can be associated with the deficient function of KIF2C induced by its deregulation () and will be summarized below. This altered expression or activity is either induced directly by an altered gene expression, changed DNA methylation, abnormal mRNA regulation such as long non-coding RNAs (lncRNAs) or micro RNAs (miRNAs), or indirectly by post-translational modifications via overexpressed kinases such as Aurora A, Aurora B, Plk1 or Cdk1.

Figure 2. The model depicts how deregulated KIF2C/MCAK is associated with some of the fourteen known and discussed hallmarks of cancer including acquired capabilities and enabling characteristics [Citation84]. These hallmarks encompass the following terms: deregulated cellular metabolism, sustaining proliferative signaling, resisting cell death, evading growth suppressors, polymorphic microbiomes, enabling replicative immortality, avoiding immune destruction, nonmutational epigenetic reprogramming, unlocking phenotypic plasticity, activating invasion and metastasis, tumor-promoting inflammation, cellular senescence, inducing or accessing vasculature and genome instability and mutation. Deregulated KIF2C may be associated with at least five of these hallmarks by promoting migration and invasion, disrupting MT dynamics, impairing the immune response, inducing senescence and increasing therapy resistance to MT interfering agents.

Figure 2. The model depicts how deregulated KIF2C/MCAK is associated with some of the fourteen known and discussed hallmarks of cancer including acquired capabilities and enabling characteristics [Citation84]. These hallmarks encompass the following terms: deregulated cellular metabolism, sustaining proliferative signaling, resisting cell death, evading growth suppressors, polymorphic microbiomes, enabling replicative immortality, avoiding immune destruction, nonmutational epigenetic reprogramming, unlocking phenotypic plasticity, activating invasion and metastasis, tumor-promoting inflammation, cellular senescence, inducing or accessing vasculature and genome instability and mutation. Deregulated KIF2C may be associated with at least five of these hallmarks by promoting migration and invasion, disrupting MT dynamics, impairing the immune response, inducing senescence and increasing therapy resistance to MT interfering agents.

4.1. KIF2C, genome instability and mutation

MT dynamics during metaphase and anaphase is the driving force to ensure a flawless chromosome separation, which safeguards genome stability in cells. The kinetochore–MT interface is crucial to prevent congression defects as well as segregation errors [Citation87]. KIF2Cs activity and functions are precisely regulated by various kinases and its malfunction results in aberrant spindles, defective chromosome congression and segregation, leading to an error-prone mitosis, which is associated with chromosomal instability (CIN), a known hallmark of oncogenesis [Citation83]. Indeed, 3T3 fibroblasts expressing a nonfunctional KIF2C demonstrated a high degree of CIN [Citation88]. Interestingly, p53 knockout in these cells reduced proliferation, and mice with KIF2Cdeficient/p53KO xenograft tumors had a highly increased progression-free survival (PFS) compared to KIF2CWT/p53KO [Citation88]. Additionally, a recent study showed that inhibitors against KIF2C’s depolymerization activity were sufficient to induce aneuploidy in different cancer cell lines [Citation89], highlighting KIF2Cs, function in safeguarding chromosome stability during mitosis. In line, Wagenbach et al. reported that slight changes in KIF2C’s expression pattern induces CIN, a weak overexpression as well as a scarce downregulation led in their reporter FKBP-MCAK CRISPR cell model to a significant number of cells with lagging chromosomes in telophase [Citation90]. In summary, these data illustrate how KIF2C’s increased and decreased expression in cancer cells is able to fuel tumorigenesis.

4.2. KIF2C and resisting cell death

Interestingly, MTs in recent years have been shown to be directly involved in the DNA damage response in at least three levels: (1) cytoplasmic MTs facilitate the transport of proteins involved in DNA damage repair into the nucleus, (2) MTs are able to modulate the chromatin structure and induce chromatin reorganization, (3) DNA double-strand breaks (DSBs) are mobilized by MTs, which promote the recruitment of DNA repair proteins to these sites [Citation91]. In line with this notation, a recent study indicates that KIF2C is also recruited to DNA damage sites, which was dependent on the activity of poly (ADP-ribose) polymerase 1 (PARP) and Ataxia-telangiectasia mutated (ATM) [Citation40]. Abolishing KIF2C’s activity impaired double strand break repair via both non-homologous end joining and homologous recombination [Citation40]. This, in addition to its MT destabilizing function, may explain how KIF2C is involved in resisting cell death induced by known anti-cancer drugs, such as taxanes and doxorubicin, that are known to cause DNA damage [Citation92–94]. This is in line with the high incidence of missense mutations or amplifications in the KIF2C gene in pancreatic adenocarcinoma [Citation95] or the mutation of E403K in KIF2C found in colorectal cancer patients [Citation96].

4.3. Activating invasion and metastasis

Since the process of cell migration is highly dependent on both actin filaments and MTs [Citation97], it is not surprising that multiple studies highlight the importance of KIF2C in migration, invasion and metastasis [Citation36,Citation37,Citation57,Citation98,Citation99]. Mechanistically, the first function of KIF2C during these processes is to facilitate the modulation of the cell polarization by fostering dissimilarities in MT dynamics between the leading and trailing edges of cells [Citation35]. This process is highly dependent on an Aurora A-KIF2C-RAC1 signaling pathway, in which Aurora A finely tunes KIF2C’s depolymerization activity in the growing MTs in the interphase [Citation35,Citation57]. Secondly, the centrosome repositioning is highly dependent on active MT dynamics in migrating cells. Knockdown of KIF2C was associated with defects in centrosome repositioning in retinal pigment epithelial cells (RPE-1) [Citation57], which disturbs the MT dynamics required to form a leading and trailing edge in migrating cells. Third, the orchestrated assembly and disassembly of focal adhesions requires MTs for endo- and exocytotic processes [Citation100]. Interfering with the expression of KIF2C leads to severe defects during FA turnover [Citation57] and deregulated phosphorylation events of key FA proteins including paxillin and focal adhesion kinase (FAK) [Citation36]. Fourth, the polymerization of the actin filaments and MTs is interdependent. Interfering with the plus-tip dynamics in the interphase by knockdown or overexpression of KIF2C highly reduced the polymerization of actin filaments [Citation36]. These data underscore the key roles of MAPs during cell migration, especially of KIF2C in cell motility of cancer cells.

4.4. Avoiding immune destruction

The recognition of malignant cells and the subsequent destruction of malignant cells is a key barrier to tumor formation and progression [Citation86]. As a consequence, malignant tumors form an immunosuppressive tumor microenvironment (TME) [Citation101]. In this complex process, there is evidence indicating that KIF2C could be associated with the immune response. Studies in hepatocellular carcinoma (HCC) [Citation102], glioma [Citation103,Citation104], kidney renal clear cell carcinoma (KIRC) [Citation105] and pancreatic ductal adenocarcinoma (PDAC) [Citation99] suggest that a high expression of KIF2C interferes with proper immune cell infiltration by recruiting increased numbers of tumor-associated macrophages, cancer-associated fibroblasts, myeloid-derived suppressor cells and Treg into the TME [Citation95]. The disturbed immune response was connected with decreased levels of immunomodulatory IL18 and IL1β based on RNA sequencing and protein analyses of PDAC samples [Citation99]. Further, KIF2C was shown to be involved in the complement and coagulation cascade, cytokine-cytokine receptor interaction, and IL17 signaling pathway by GO and KEGG pathway analyses of kidney renal clear cell carcinoma (KIRC) and liver hepatocellular carcinoma [Citation105], suggesting an interesting new function for KIF2C. This could be related to its role in intracellular trafficking and involvement in endocytosis, though further investigation is required [Citation36]. Additionally, KIF2C is able to negatively regulate ciliogenesis, which is a cell organelle shown to be involved in cell-cell signaling, immune cell function and immune cell metabolism [Citation106]. Deregulated degradation or overexpressed KIF2C causes an accumulation on the basal bodies in the following G1/G0 phase and reduced ciliogenesis [Citation107]. This results in broad cellular consequences such as decreased signaling of the canonical hedgehog pathway, a pathway exclusively mediated by the primary cilium, which has implications for cell fate and self-renewal [Citation107,Citation108].

4.5. Cellular senescence

Cellular senescence is known as a protective mechanism for maintaining a controlled tissue homeostasis and serves as a complementary mechanism to cellular apoptosis [Citation84]. Both senescence and apoptosis should help organisms to remove damaged, dysfunctional or unnecessary cells [Citation109]. However, several decades after the discovery of senescent cells, a multitude of studies showed that the malignancy of cancer cells is fueled by senescent cells [Citation110]. Intriguingly, KIF2C was shown to be involved in the induction of cellular senescence in human primary cells (dermal fibroblasts and umbilical vein endothelial cells) [Citation111]. Knockdown of KIF2C induced cellular senescence that was highly dependent on p53 but was not associated with the known senescence regulator p16 [Citation111]. In line with this observation, a recent study could show that KIF2C together with four other cellular senescence-related genes, is associated with poor prognosis of patients with idiopathic pulmonary fibrosis [Citation112].

In summary, these observations clearly support the concept that KIF2C is involved in at least five different hallmarks of cancer, suggesting a possible role in tumor development, progression and malignancy.

5. Molecular mechanisms involved in KIF2C’s oncogenic potential in multiple cancer entities

Numerous studies support KIF2C’s oncogenic potential showing that it is highly expressed in tumor cells from different origins and that the degree of de-differentiation of tumor cells depends on the expression level of KIF2C [Citation95,Citation113–117]. This increased expression was also found in two independent studies at the mRNA level in the entire patient sample cohort of the TCGA (The Cancer Genome Atlas) database including 33 different cancer entities [Citation95,Citation118]. KIF2C’s role in multiple vital functions and its high protein/gene expression will be discussed in various cancer entities including breast cancer, non-small lung cancer, hepatocellular carcinoma and gastric cancer, highlighting its crucial role in carcinogenesis.

5.1. Breast cancer

Breast cancer represents the best studied cancer entity correlated with KIF2C’s expression, prognostic value and involvement in therapy resistance ( and ). The first studies were conducted almost two decades ago, where Nishidate and colleagues identified KIF2C as one of 34 genes that were expressed differentially in breast tumors with lymph node metastasis compared to tumors without metastasis [Citation119]. Moreover, two in vitro studies showed that protein regulator of cytokinesis 1 (PRC1) and KIF2C formed a functional complex and that both were upregulated in several breast cancer cell lines [Citation120,Citation121]. Knockdown of either PRC1 or KIF2C inhibited proliferation of multiple breast cancer cell lines, which was associated with morphological changes especially during mitosis, in these cancer cells [Citation121]. The gene level of KIF2C was found to be correlated with poor outcomes in breast cancer patients [Citation122]. Moreover, the gene methylation of KIF2C was significantly elevated in luminal A breast cancer samples with prognostic value, together with three other cell cycle and proliferation regulators Ki-67, UBE2C (ubiquitin conjugating enzyme E2C) and HDAC4 (histone deacetylase 4) [Citation123]. A bioinformatic multi-layer interference approach of context-dependent gene networks revealed that KIF2C was a target gene for ER−/HER2− (estrogen/human epidermal growth factor receptor 2) breast cancers and its gene expression was positively regulated by E2F1 (E2F transcription factor 1), which has crucial roles in cell cycle control [Citation124,Citation125]. This was continued by three other bioinformatic approach studies [Citation126–128], which displayed that a four gene panel consisting of KIF2C, KIF1A, FAM134B (Reticulophagy Regulator 1) and ALCAM (Activated Leukocyte Cell Adhesion Molecule) had diagnostic translational potential [Citation126]. This four gene panel was further shown to have implications in several hallmarks of cancer, including enabling replicative immortality, sustaining proliferative signaling, activating invasion and metastasis, resisting cell death and deregulating cell metabolism that are connected to breast cancer heterogeneity [Citation126]. Moreover, the expression of KIF2C, PLK1 and MAD2L1 (mitotic arrest deficient 2 like 1) was correlated to reduced overall survival (OS) in invasive ductal breast carcinoma [Citation127]. In basal-like breast cancer, KIF2C is a crucial hub gene, together with nine other genes [Citation128]. These data were corroborated by a study analyzing the cohort of the cancer genome atlas [Citation129]. It was found that a high expression of KIF2C correlated indeed with poor OS and, vice versa, patients with a low expression of KIF2C had significantly more favorable OS [Citation129,Citation130]. To further support these findings, the expression of KIF2C was found to be up-regulated in all breast cancer subtypes compared to normal tissue, with the triple negative breast cancer (TNBC) subtype expressing the highest level of KIF2C. To further support these findings, the expression of KIF2C was found to be up-regulated in all breast cancer subtypes compared to normal tissue, with the triple negative breast cancer (TNBC) subtype expressing the highest level of KIF2C. Corroborating these data, five genome sequencing sets consisting of 63 TNBC samples and 169 non-TNBC samples were analyzed and five hub genes TPX2 (Targeting protein for Xklp2), CTPS1 (CTP Synthase 1), KIF2C, MELK (maternal embryonic leucine zipper kinase) and CDCA8 (cell division cycle associated 8) were again identified [Citation131]. In agreement, it was found that KIF2C expression together with three other genes was correlated with a reduced regression free survival and highly expressed KIF2C was connected to earlier relapse in TNBC patients [Citation132]. Interestingly, a screen of exclusively metastatic breast cancers revealed similar data [Citation133]. Again, KIF2C was identified among five important hub genes TPX2, CDCA8, BUB1B (Budding Uninhibited by Benzimidazoles 1 Homolog Beta), CCNA2 (cyclin A2) and KIF2C associated with a significant enhanced risk for distant metastasis [Citation133]. These data were further consolidated by seven recently published studies, which corroborated the valuable role of KIF2C as a biomarker correlated with poor prognosis for all breast cancer subtypes [Citation130,Citation131,Citation134–138] (, OS; mean HR 1.79 [1.49–2.20]). Interestingly, KIF2Chigh breast cancer tissue was significantly correlated with increased stroma cells [Citation130]. This could be associated with an increased population of cancer-associated fibroblasts, which are also related to poor prognosis of patients with luminal breast cancer subtypes [Citation139].

Figure 3. High expression of KIF2C is associated with reduced OS and DSF in BC, HCC and NSCLC. (A–C) The meta-analyses concerning KIF2C/MCAK’s importance as a prognostic marker is depicted as forest plots for OS of BC, NSCLC and HCC patients (A–C). (D) Meta-quantification of KIF2C/MCAK involvement in DSF in HCC patients. All individual studies were scored based on their patient cohort. The weighted score (ws) is visualized in a heatmap, dark blue (high ws), yellow (intermediate ws) and red (low ws). The overall mean value (MV) extracted from all studies was calculated by integrating the weighted score. Abbreviations: OS: overall survival; DFS: disease-free survival.

Figure 3. High expression of KIF2C is associated with reduced OS and DSF in BC, HCC and NSCLC. (A–C) The meta-analyses concerning KIF2C/MCAK’s importance as a prognostic marker is depicted as forest plots for OS of BC, NSCLC and HCC patients (A–C). (D) Meta-quantification of KIF2C/MCAK involvement in DSF in HCC patients. All individual studies were scored based on their patient cohort. The weighted score (ws) is visualized in a heatmap, dark blue (high ws), yellow (intermediate ws) and red (low ws). The overall mean value (MV) extracted from all studies was calculated by integrating the weighted score. Abbreviations: OS: overall survival; DFS: disease-free survival.

In addition to its role as a prognostic marker, KIF2C has been reported to be linked to therapy resistance to taxanes [Citation89,Citation93,Citation94] and doxorubicin (DOX) [Citation92]. It was shown in malignant (MDA-MB-231, HeLa and HuH7) and nonmalignant (RPE-1 and CHO) cell lines that taxane resistant cells changed their MT environment and recruited more septin9_i1 and septin9_i3 from actin fibers to MTs, which led to tubulin tyrosination and long polyglutamylated chains. This increased the recruitment of MT destabilizing proteins such as KIF2C and CLIP170, resulting in increased MT dynamics and a reduced sensitivity toward MT stabilizing agents [Citation93,Citation94]. Furthermore, KIF2C was able to directly bind pyruvate kinase M2 (PKM2). This prevented the ubiquitination of PKM2 and its proteasomal degradation. The increased levels of PKM2 promoted autophagy and glycolysis, both mechanisms known to enhance DOX resistance [Citation92]. In conclusion, KIF2C increases breast cancer tumorgenicity, malignancy and is sufficient to induce chemoresistance by multiple mechanisms. As a consequence, KIF2C is regarded as a potential cancer therapy target. In fact, several studies showed that depletion or knockdown of KIF2C interferes with proliferation, migration and invasion capacity of cancer cells [Citation36,Citation37,Citation116,Citation121]. A recent analysis of breast cancer samples integrated in The Cancer Genome Atlas (TCGA) revealed that KIF2C, along with two cell cycle master regulators CCNB1 (cyclin B1) and Plk1, is a promising target for breast cancer therapy [Citation140]. In line with this notion, three microRNAs (miR-10b-5p, miR-485-5p and miR-181c) targeting KIF2C were described [Citation116,Citation117], leading to a reduced gene expression. A high expression of these three microRNAs was positively correlated with patient survival [Citation117]. Interestingly, a recent study found three potent inhibitors against the depolymerization activity of KIF2C [Citation89]. These compounds were able to reduce the clonogenic survival rate of TNBC cells and one compound was able to sensitize TNBC cells to paclitaxel [Citation89]. In summary, these reports corroborate KIF2C as a prognostic biomarker and novel compounds enable KIF2C to be a potential target for breast cancer therapy.

5.2. Hepatocellular carcinoma (HCC)

The involvement of KIF2C in HCC was extensively studied in vitro and in vivo during the last decade [Citation141–143]. These investigations could show that KIF2C modulates the proliferation, migration, metastasis and tumor growth in vitro and in a mouse model [Citation141–143]. Additionally, KIF2C affected the stemness of HCC cells by interfering with several key pathways including mTORC1 (mammalian target of rapamycin complex 1) [Citation144], Ras/MAPK (rat sarcoma/mitogen-activated protein kinase) [Citation98], Wnt/β-catenin (wingless/beta-catenin) [Citation144] and MEK/ERK (mitogen-activated protein kinase kinase/extracellular signal-regulated kinases) [Citation145]. Specifically, it was shown that KIF2C is a downstream target of Wnt/β-catenin signaling, leading to its transcriptional activation [Citation144]. In turn, KIF2C enhanced the mTORC1 activity by directly binding to TBC1D7 (Tre2-Bub2-Cdc16 domain family member 7), which interfered with the stability of the mTORC1 inhibitory TSC1-TSC2 complex [Citation144]. This is an essential step in the progression of HCC, since the mTORC1 activity was shown to stimulate cell growth, metabolic re-programming, proliferation and inhibition of apoptosis in HCC patients [Citation146]. In accordance with this, another study could show that the ANLN (anillin actin-binding protein)-KIF2C signaling axis promoted bone metastasis of HCC by activating mTORC1 [Citation147]. Moreover, knockdown of KIF2C in HCC cells gave rise to the transcriptional upregulation of the PI3K/Akt (phosphatidylinositol 3′-kinase/protein kinase B) and MAPK signaling pathways [Citation98], two key pathways in the progression of HCC [Citation148]. At a functional level, KIF2C was shown to promote cell proliferation, migration, invasion, cell cycle progression and inhibited apoptosis as well as induced the epithelial-to-mesenchymal transition (EMT), possibly by the activation of the Ras/MAPK and PI3K/Akt pathways [Citation98]. In support of these observations, knockdown of KIF2C decreased the expression of important EMT transcription factors and downstream targets, including Snail and vimentin [Citation145]. Additionally, the silencing of KIF2C was also associated with decreased p-MEK and p-ERK signals, reducing proliferation, migration and invasion capacity of the epithelial breast cancer cell line HCC1954 [Citation145]. These findings indicate KIF2C’s multifaceted roles in HCC progression, in addition to its known function as regulator for MT stability as well as dynamics to safeguard chromosome separation during mitosis [Citation17,Citation83].

Consistent with the outlined molecular mechanisms of how KIF2C fuels the malignancy of HCC cells, a large number of investigations have been able to correlate highly expressed KIF2C in HCC patients with several clinicopathological characteristics ( and ). Twelve different studies indicate that the high expression of KIF2C is associated with a significantly reduced overall survival [Citation98,Citation102,Citation141,Citation142,Citation144,Citation145,Citation149–154] (, OS; mean HR 1.84 [1.29-2.80]). Furthermore, KIF2C expression was correlated to increased tumor grading [Citation142,Citation153], staging [Citation98], RFS [Citation98,Citation150], DFS [Citation98,Citation141,Citation144,Citation151,Citation153] (, DSF; mean HR 1.81 [1.20–2.61]), differentiation of the tumor [Citation144], relapse [Citation144], and immune cell infiltration [Citation102]. Consequently, based on diverse bioinformatical approaches, KIF2C was found to be a key hub gene for progression of HCC with its broad involvement in various “Gene Ontology” (GO) and “Kyoto Encyclopedia of Genes and Genomes” (KEGG) pathways containing mitotic nuclear division, nuclear division, chromosome segregation, mitotic sister chromatid segregation, mitotic spindle, spindle pole and kinetochore [Citation152]. Additionally, KIF2C forms interaction networks with other cell cycle master regulators such as CCNB1, CDC20, CENPM (centromere protein M), TOP2A (DNA topoisomerase II α), MYBL2 (Myb-related protein B), PLK1, UBE2C and BIRC5 (survivin) [Citation155]. A recent study utilized the protein and gene expression of KIF2C and RAC1 in HCC tissue to characterize a specific HCC patient subtype, which significantly correlated with worse prognosis, clinicopathological grade, increased tumor mutation burden, higher CD8+ T cell infiltration and an altered chemotherapy drug sensitivity [Citation156]. Therefore, KIF2C has been suggested as a prognostic biomarker [Citation149,Citation152–154,Citation156–160] and as a treatment target for HCC [Citation141,Citation144,Citation155].

5.3. Non-small cell lung cancer (NSCLC)

In non-small cell lung cancer (NSCLC) tissues, the gene expression of KIF2C was found to be significantly upregulated compared to paired non-tumor samples in three independent RNA sequencing datasets [Citation161]. Moreover, a high expression of KIF2C has been associated with advanced TNM stage in NSCLC patients, especially in the most important and common type of lung adenocarcinoma, and consequently with a dramatically decreased OS ( and ) [Citation161,Citation162]. These results could be confirmed in several NSCLC cell lines and IHC samples at the protein level, which was again correlated with reduced OS, higher primary tumor staging (T stage), worsened differentiation status and lymph node metastasis [Citation163]. KIF2C knockdown inhibited proliferation, colony formation, migration and invasion of NSCLC cell lines [Citation163]. Interestingly, KIF2C may be targeted by microRNA-325-3p, microRNA-34a-5p and microRNA-186-3p observed via a luciferase reporter assay or transcription factor-miRNA-mRNA network computing [Citation163–165]. In accordance with this finding, microRNA-325-3p is considered as a tumor suppressor and its expression was decreased in NSCLC samples [Citation166]. It could be shown that the expression of KIF2C and miR-325-3p were inversely correlated in the 30 NSCLC samples [Citation163]. In line, five recently published bioinformatic studies could identify KIF2C as key signaling hub gene in the progression of lung cancer [Citation164,Citation167–170]. Three of them verified the significantly reduced OS associated with a high expression of KIF2C [Citation164,Citation168,Citation169] and all five studies suggested various hub genes, including KIF2C as prognostic biomarker and therapeutic target [Citation167–170] (, OS; mean HR 2.18 [1.73–2.84]). Interestingly, this study found multiple drugs, which either interacted with KIF2C or had the potential to regulate KIF2C [Citation164]. Additionally, KIF2Chigh tissue expression was negatively correlated with the infiltration of CD4+ and CD8+ T cells [Citation164,Citation169]. Functionally, KIF2C could be shown to promote proliferation, migration and invasion of different NSCLC cell lines and to reduce apoptosis [Citation165]. These effects were likely mediated by the activation of the Akt pathway, as overexpression of KIF2C increased the phosphorylation level of p-Akt, its downstream targets p-GSK-3β and β-catenin [Citation165], which are known for their roles in cancer migration and invasion [Citation171]. The upregulation of KIF2C is likely associated with the hypomethylation of its promotor region [Citation164]. In conclusion, these data highlight that KIF2C is an important protein with oncogenic potential for NSCLC progression involved in metastasis, differentiation, staging and OS.

5.4. Glioma

The first report of KIF2C’s association with the malignancy of gliomas was published by Bie and colleagues in 2012 [Citation172]. Their work indicated that KIF2C gene and protein expression was related to glioma grading and patient OS. Mechanistically, the expression level of KIF2C was strongly correlated with the proliferation marker Ki-67 [Citation172]. This unfavorable relationship between the expression of KIF2C and the OS of glioma patients was further reinforced by three reports showing that KIF2C was an important hub gene in the overall TCGA glioma dataset [Citation173], low-grade gliomas [Citation174] and secondary glioblastomas [Citation175]. Moreover, two recent in silico investigations strengthened the role of KIF2C as key hub gene [Citation103,Citation104], verified its association with poor survival rates in multiple glioma grades (2-4) and a disturbed immune infiltration [Citation103], indicating its diagnostic and prognostic value.

5.5. Esophageal squamous cell carcinoma (ESCC)

Based on immunohistochemical staining of malignant tissue sections, a high expression of KIF2C increased the incidence of a high pathologic tumor and poor tumor differentiation status in male patients, though not in female patients [Citation176]. Among patients with similar pathological tumor node metastasis stages, the prognosis was worse in male patients with KIF2Chigh expression. The OS and DFS were significantly shorter in male patients with high KIF2C expression compared to female patients with high or low KIF2C expression. Its high expression was linked to a significantly increased risk of a higher pathological tumor status and poorer tumor differentiation. The KIF2C expression in ESCC seemed to serve as an independent prognostic marker for male, but not for female patients. Since KIF2C is highly expressed in testis, Duan and colleagues speculate that KIF2C might require an androgen rich environment to carry out its full function, explaining the sex-biased results [Citation176].

5.6. Gastric cancer

The gene expression of KIF2C was reported to be significantly higher in gastric cancer tissues compared to its expression in nonmalignant tissues [Citation95,Citation177]. Furthermore, overexpression of KIF2C resulted in significantly higher lymphatic invasion, lymph node metastasis, serosal invasion and a reduced patient survival [Citation113]. As a result, KIF2Chigh gastric cancers correlated with a poor patient diagnosis with reduced OS, DSS and PFI [Citation177] (). In vitro studies showed that overexpression of KIF2C enhanced cell migration and proliferation with a correlation between the high levels of KIF2C and multiple proliferation genes (CCN2A, CCNB1, CCNB2, CCNE1, CDK1 and CDK2) in gastric cancer tissues [Citation113,Citation177]. Another study showed that KIF2C antigen peptides could be used to lyse gastric and colon cancer cells in an HLA class I, and CD8-restricted manner [Citation178]. Moreover, knockdown of KIF2C reduced cell proliferation of gastric cancer cells (GES-1, AGS, MKN-45, NCI-N87, and SNU-1) [Citation177]. These reports suggest that KIF2C may be a clinical marker that correlates with immune cell infiltration and may prove to be an effective therapeutic target in the treatment of patients with KIF2C overexpressing gastric cancer.

5.7. Pancreatic cancer (PCa)

The first evidence that KIF2C may be involved in the progression and poor prognosis of PDAC patients was obtained using single-cell RNA-sequencing (scRNA-seq.) [Citation95]. The pan-cancer scRNA-seq. data displayed that KIF2C expression is highly enriched in CD4+ T cells, fibroblasts, NK cells, and most significantly in malignant cells in a comprehensive PDAC cohort (24 primary PDAC tumors and 11 control pancreases) [Citation95]. Interestingly, 1.8% of all patients in the TCGA cohort including PAAD (Pancreatic Adenocarcinoma) had genetic alterations in the KIF2C gene, which included predominantly missense mutations or amplifications [Citation95]. The KIF2C gene and protein expression were enhanced in PAAD. Increased levels of KIF2C expression were significantly associated with poor OS and DSF. High KIF2C was significantly associated with E2F, EGFR, MYC, TP53 and KRAS [Citation95,Citation179]. In support of this, two recent studies with pancreatic cancer (PCa) and two with PAAD found that KIF2C was highly upregulated in pancreatic cancer tissues and correlated with poor patient prognosis and survival [Citation99,Citation179–181]. Specifically, KIF2Chigh PDACs were significantly correlated with reduced OS, DFS, PFS and postoperative survival time (PST) [Citation99]. Moreover, PCa samples with a high expression of KIF2C reduced patient OS, DFI, DSS and PFI [Citation181]. These clinicopathological features in KIF2Chigh PCas is likely associated with an altered immune cell infiltration, TME, immune checkpoint activation and a MAPK signaling pathway inhibitor resistance [Citation181]. The expression levels of KIF2C were inversely linked with a reduced methylation level of its DNA [Citation181]. Additionally, upregulated KIF2C was associated with high tumor staging (IV), shorter OS and poorer differentiation [Citation99]. Functionally, the overexpression of KIF2C was shown to be associated with enhanced migration, invasion and colony formation capacity in PCa cells lines ASPC-1 and MIA-PaCa2 [Citation99]. Moreover, these cells had decreased interleukin-1β (IL1β) and IL18 levels, and upregulated CDC20. Finally, a study with a pancreatic xenograft mouse model with MIA-PaCa2 cells reported that a downregulation of KIF2C inhibited the formation of subcutaneous tumors and lung metastasis [Citation99]. In line with a deregulated cytokine secretion, KIF2C expression was also associated with a relative abundance of tumor-infiltrating lymphocytes (TILs) and CD4+ T cells in PAAD patients [Citation180]. In further support, several KIF genes including KIF2C, KIF4A, KIF11, KIF14, KIF15, KIF16B, KIF20A, KIF22 and KIF25 were core enriched in the immunologic signature of PAAD samples [Citation179]. Overall, these data strengthen the crucial role of KIF2C in progression of PCa and suggest an involvement in regulating the expression of important immunomodulatory cytokines as well as fueling the malignancy of the cancer cells.

5.8. Renal cell cancer (RCC)

In order to find out prognostic genes, a gene analysis was performed using 533 RNA samples of KIRC from the TCGA. KIF2C and two other members of the KIF-family (KIF23 and KIF4A) were found as prognostic genes in KIRC [Citation182]. The study also examined KEGG pathways that have impacted OS of patients with KIRC. These included the cell cycle, homologous recombination and the p53 signaling pathway [Citation182]. KIF2C was implicated in the outcomes of KIRC patients, although no correlation of KIF2C with any clinicopathologic features of KIRC was shown. A further study suggested that long non-coding RNAs (lncRNAs) of TTK, CENPE, KIF2C, BUB1, and RAD51AP1 (RAD51 Associated Protein 1) could act as potential biomarkers for chromophobe renal cell carcinoma progression and prognosis [Citation183]. Indeed, by analyzing the enormous TCGA and gene expression omnibus database, four other studies found that high expression of KIF2C was significantly correlated with reduced OS and DSF in patients with KIRC [Citation95,Citation105,Citation118,Citation184]. Moreover, the expression of KIF2C was significantly associated with infiltrated immune cells such as B cell, CD8+ T cell, CD4+ T cell, myeloid-derived suppressor cells, Tregs cells and non-immune cells including cancer-associated fibroblasts [Citation105]. The authors suggest that KIF2C is a prognostic biomarker linked to immunosuppression and an interesting target for immunotherapy [Citation105]. Interestingly, a recent report showed that the DNA of KIF2C is significantly hypomethylated in RCC [Citation185]. This is in line with an interesting hypothesis that altered DNA methylation preferably occurs during early carcinogenesis, which is a hallmark of ccRCC [Citation186]. These findings reveal that KIF2C might be a promising biomarker in patients with RCC, as it is frequently deregulated in this cancer entity and correlates with poor prognosis and an immunosuppressive TME.

5.9. Colorectal cancer

Similar to gastric cancer, KIF2C was found to be overexpressed in colorectal cancer tissues from 195 patients when compared to the corresponding normal tissue [Citation114,Citation115]. The overexpression was associated with increased lymph node metastasis, venous invasion, lymphatic invasion, peritoneal dissemination, and Dukes’ staging classification, leading to a poor survival of colorectal cancer patients with KIF2Chigh expression [Citation114]. This was recently corroborated by two independent studies [Citation187,Citation188]. While the first identified upregulated KIF2C as one of four key proto-oncogenes in a comprehensive transcriptomic study of colorectal cancer tissue stages I-IV, adenocarcinoma and mucinous adenocarcinoma [Citation188]. The second included KIF2C into a three gene prognostic signature together with poly (ADP-ribose) polymerase 1 binding protein (PARPBP) and kinetochore-localized astrin/SPAG5 binding protein (KNSTRN) based on a machine learning based approach [Citation187]. These signatures were able to accurately predict the clinical outcomes of stage IV CRC patients, correlated with an increased cancer stem cell phenotype in these patients [Citation187]. Another study used a computational prediction of disease-associated non-synonymous polymorphism analysis to filter out pathological mutations in a large pool of datasets [Citation96]. Their analysis found that the mutation of E403K in KIF2C, which interferes with protein conformation and stability, was associated with the development of colorectal cancer [Citation96]. These observations underline the crucial importance of KIF2C in the progression of colorectal cancer, which might be a possible treatment target by using an antigen specific immunotherapy to stimulate the immunosuppressive environment by triggering a spontaneous CD4+ T cell response of the Th1-type in colorectal cancer [Citation115].

5.10. Endometrial cancer (EC)

KIF2C was found to be highly upregulated among 344 other genes differentially expressed in endometrial cancer [Citation189]. By generating a protein-protein interaction network, 15 genes were identified with a highly prognostic potential including KIF2C [Citation189]. This is in line with two other studies, which also generated hub gene networks for EC with similar genes containing KIF2C [Citation190,Citation191]. These analyses underscore the notion that the expression of KIF2C is negatively associated with the OS of EC patients [Citation189–191] and is correlated with a poor differentiation status in uterine corpus endometrial carcinoma [Citation191]. Interestingly, on a functional level, KIF2C was shown to stimulate the proliferation, migration and invasion of two endometrial cancer cell lines (Ishikawa and RL95-2) [Citation192]. Additionally, knockdown of KIF2C in a xenograft mouse model significantly reduced the tumor growth and the expression level of KIF2C was negatively correlated to CD8+ T cell invasion [Citation192], implicating KIF2C’s role in recruiting immune cells in the TME, important in prognosis and in potential EC intervention.

5.11. Ovarian cancer (OC)

The data of KIF2C in OC are inconclusive and illustrate that a fine-tuned expression of KIF2C is highly associated with the patient outcome. In an analysis of a dataset of 396 OC samples and 54 controls, upregulated KIF2C was detected as one of 12 hub genes for OC progression in a protein interaction network, and its expression was negatively associated with poor OS of patients with epithelial ovarian cancer [Citation193]. Another bioinformatic approach found a regulatory network between the upstream transcription factors SMAD4 (SMAD Family Member 4), NFKB1 (Nuclear Factor Kappa B Subunit 1), SMAD3, TP53 (Tumor Protein P53), HNF4A (Hepatocyte Nuclear Factor 4 Alpha) and their potential downstream targets KIF2C, STAT3 (signal transducer and activator of transcription 3) and BUB1 [Citation194]. Intriguingly, the authors showed that KIF2C was downregulated in the platinum resistant ovarian cancer cell line A2780 [Citation194]. FOXM1 was proposed to be an alternative upstream regulator and a transcriptional activator of KIF2C, a transcriptional activator that binds to the KIF2Cs promotor and is associated with paclitaxel resistance, consequently, silencing FOXM1 resulted in downregulation of KIF2C [Citation195]. Ni and colleagues reported in their discussion that low expression of KIF2C was correlated with decreased survival and DSF of OC patients [Citation196]. By contrast, another study showed that KIF2C is upregulated in OC samples, together with nine other hub genes, but could not confirm a significant impact on OS [Citation197]. These studies show that the regulation of KIF2C must be fine-tuned and that its overexpression as well as downregulation are associated with the pathogenesis of OC.

5.12. Acute lymphoblastic leukemia (ALL)

While KIF2C has been considered a promising novel biomarker and prognostic marker for various solid tumors, its role in leukemia is hardly defined. One study showed that KIF2C was overexpressed in patients with relapsed acute lymphoblastic leukemia (ALL) [Citation198]. Moreover, this study displayed a correlation between the expression of KIF2C [Citation198] and its binding partner KIF18B [Citation199], together both proteins are associated with a risk of relapse in ALL [Citation198]. Mechanistically, they found that a functional KIF2C was necessary for the accurate lymphopoiesis in zebrafish embryos, which share 50% structural homology to KIF2C [Citation198]. The dependence of hematopoietic stem cells survival on high expression of KIF2C/KIF18B may be a possible pathomechanism of the occurrence of ALL relapse [Citation170].

5.13. Other cancer entities

Several single reports or dataset analyses deal with the expression of KIF2C in cervical cancer, nasopharyngeal carcinoma (NPC), laryngeal squamous cell carcinoma (LSCC), laryngeal squamous-cell cancer and bladder cancer [Citation184,Citation200–204]. In cervical cancer, the authors showed that KIF2C mutation is strongly associated with the survival rate, and that KIF2C expression was significantly upregulated in cervical cancer tissues and cervical cancer cells [Citation184]. KIF2C enhanced cell proliferation, invasion, and migration in vitro and increased tumor growth in vivo. KIF2C knockdown promotes the activation of the p53 signaling pathway. A rescue assay with KIF2C and p53 double knockdown partially reversed the inhibitory influence of KIF2C silencing on cervical cancer processes [Citation184]. The study suggested that KIF2C might be a novel therapeutic target for cervical cancer. Zuo et al. highlighted the role of KIF2C in NPC, with their data indicating that KIF2C is aberrantly overexpressed in multiple mRNA datasets of NPC [Citation200]. Additionally, the knockdown of KIF2C in HNE-1 and CNE-1 NPC cell lines led to reduced motility and migration, increased mitotic defects and an enhanced paclitaxel sensitivity, indicating the broad role of KIF2C in NPC [Citation200]. In LSCC, KIF2C was found to be a hub gene in two independent bioinformatic approaches using multiple comprehensive publicly available RNA expression datasets [Citation201,Citation202], but further investigations are required to decipher the clinicopathological impact of KIF2C’s expression in this tumor entity. Finally, a study indicated KIF2C as a key hub gene in bladder cancer. In their relatively small patient cohort, they could not correlate a high expression of KIF2C with the OS [Citation203]. In support, Yang et al. found that circular RNA circRGNEF promoted bladder cancer progression via the miR-548/KIF2C signaling axis [Citation204]. They found that the overexpression of KIF2C was associated with increased invasion, migration and proliferation in the bladder cancer cell lines T24 and UM-UC-3 [Citation204]. In sum, all of these studies point to the notion that KIF2C is associated with increased cancer malignancy in diverse cancer entities.

6. Conclusion

In this systematic review, we have highlighted the current data concerning KIF2C’s molecular working mechanisms and its potential involvement in diverse signaling pathways. Further, its association with at least five hallmarks of cancer was emphasized, explaining its oncogenic potential. In line, we have reviewed its expression in various cancer entities and discussed its potential clinical significance, including a meta-analysis correlating KIF2C’s expression with the OS of BC, NSCLC and HCC patients. The data show that the gene expression of KIF2C is significantly increased in at least 33 different cancer types compared to benign tissue [Citation95,Citation118]. Upregulated KIF2C is involved in multiple cellular activities, including proliferation [Citation17], migration and invasion [Citation36], DNA repair and therapy resistance [Citation92] as well as cell-cell and cell-ECM signaling by modulating the primary cilium [Citation39]. The data also suggest that KIF2Chigh gene or protein expression in cancer tissues could be a diagnostic marker for a number of cancer entities and a prognostic marker for patient OS and PSF. This highlights that a precisely regulated expression and activity of KIF2C is required to safeguard its vital cellular functions. Its deregulation is either induced directly by an altered gene expression, changed DNA methylation, abnormal mRNA regulation by for example lncRNAs or miRNAs, or indirectly by post-translational modifications via altered kinases such as Aurora A, Aurora B, Plk1 or Cdk1.

7. Study limitations and outlook

Although a lot of work has been done in recent years, most of the summarized data involved multi-omics approaches and computational science, which were heavily based on whole genomic transcriptomics. More detailed studies including proteomics are needed to solidify the significance of KIF2C as a diagnostic/prognostic marker for cancer patients, as clearly proposed for breast cancer, NSCLC and HCC (). Moreover, clinical trials including KIF2C expression data or the use of highlighted gene signatures for the design of treatment schedules for multiple cancer entities would largely expand the significance of KIF2C’s role in the outcome of patients. In addition, it is of importance to investigate whether the increase of KIF2C with its oncogenic potential is a driving factor in carcinogenesis because of its role in chromosome instability, or whether it is a pathological consequence of the unrestricted cell cycle of cancer cells.

Recent progress in establishing patient-derived tumor organoids (PDO) and stroma organoids in combination with CRISPR technologies provide novel tools to study the molecular regulatory network of KIF2C deregulation in elaborated in vivo models [Citation205,Citation206]. Moreover, PDOs enable us to recapitulate KIF2C’s involvement in therapy resistance toward known MT-interfering agents such as taxanes. It would be of great interest to find out if PDOs from taxane resistant patients have deregulated MT depolymerization machinery, compared to sensitive patients. Furthermore, the advances in conditional knockout mice confer the possibility to study KIF2C’s role during malignant tumor as well as organ formation and development, as shown recently for the nervous system in conditional knockout KIF2Cflox/flox/NestinCre mice [Citation207]. These models can be combined with specific inhibitory small molecule compounds against KIF2C [Citation89] or activating/inhibiting modifications on KIF2C introduced by the precise CRISPR technology, which make it possible to explore the individual roles of KIF2C during these vital cellular processes. The novel omics and scRNA-seq. technologies [Citation208] will help to decipher in which cell population KIF2C is enriched within the TME and which pathways are involved in this process, as displayed by recent studies for KIF2C in PDAC [Citation95,Citation99]. Finally, it will be of great interest, if KIF2C is indeed able to modulate the immune response in the TME as suggested in some cancer entities and how this is facilitated by a protein mainly known for its MT destabilizing function. These modern tools and techniques will help to confirm that KIF2C can serve as a diagnostic and prognostic marker for patients with various cancers.

Authors’ contributions

NNK, AR, and HM conceived the manuscript. AR designed the search strategy. AR, NNK and HM performed the article screening and data extraction. AR and NNK wrote the initial manuscript. JY and HM modified the manuscript. CS, FL and JY did critical reading. All authors read and approved the final manuscript.

Abbreviations
ALCAM=

Activated Leukocyte Cell adhesion Molecule

ALL=

acute lymphoblastic leukemia

APC=

anaphase-promoting complex

ANLN=

anillin actin-binding protein

ATM=

Ataxia-telangiectasia mutated

BUB1B=

Budding Uninhibited by Benzimidazoles 1 Homolog Beta

CDCA8=

cell division cycle associated 8

CENPM=

centromere protein M

CIN=

chromosomal instability

ccRCC=

clear cell renal cell carcinoma

CTPS1=

CTP Synthase 1

CCNA2=

cyclin A2

CCNB1=

cyclin B1

Cdk1=

cyclin-dependent kinase 1

CLASP=

cytoplasmic linker-associated protein

DSF=

disease-free survival

TOP2A=

DNA topoisomerase II α

MYBL2=

Myb-related protein B

DSBs=

double-strand breaks

DOX=

doxorubicin

E2F1=

E2F transcription factor 1

EMT=

epithelial-to-mesenchymal transition

ER=

estrogen

ECV=

extracellular environment

ECM=

extracellular matrix

ERK=

extracellular signal-regulated kinase

FAK=

focal adhesion kinase

FA=

focal adhesion

HCC=

hepatocellular carcinoma

HDAC4=

histone deacetylase 4

HNF4A=

Hepatocyte Nuclear Factor 4 Alpha

HER2=

human epidermal growth factor receptor 2

IL=

interleukin

KIRC=

kidney renal clear cell carcinoma

KNSTRN=

kinetochore-localized astrin/SPAG5 binding protein

K-Ras=

Kirsten rat sarcoma virus

LSCC=

laryngeal squamous cell carcinoma

mTORC1=

mammalian target of rapamycin complex 1

MELK=

maternal embryonic leucine zipper kinase

miRNAs=

micro RNAs

MAPs=

microtubule associated proteins

TIP150=

tracking protein of 150 KDa

MTOCs=

microtubule organizing centers

MT=

microtubule

MAPK=

mitogen-activated protein kinase

MAD2L1=

mitotic arrest deficient 2 like 1

NPC=

nasopharyngeal carcinoma

NSCLC=

non-small cell lung cancer

lncRNAs=

long non-coding RNAs

NFKB1=

Nuclear Factor Kappa B Subunit 1

NuMa=

nuclear mitotic apparatus

NuSAP=

nucleolar spindle-associated protein

OS=

overall survival

PAK1=

p21-activated kinase 1

PCa=

pancreatic cancer

PDAC=

pancreatic ductal adenocarcinoma

PDOs=

patient-derived tumor organoids

Plk1=

Polo-like kinase 1

PARPBP=

poly (ADP-ribose) polymerase 1 binding protein

PARP=

poly (ADP-ribose) polymerase 1

PFS=

progression-free survival

PST=

postoperative survival time

PRC1=

protein regulator of cytokinesis 1

PKM2=

pyruvate kinase M2

PRISMA=

preferred reporting items for Systematic Reviews and Meta-Analyses

RAC1=

Rac Family Small GTPase 1

RAD51AP1=

RAD51 Associated Protein 1

RCC=

renal cell cancer

RITA=

RBP-J interacting and tubulin-associated protein

FAM134B=

Reticulophagy Regulator 1

RPE-1=

retinal pigment epithelial cells

scRNA-seq=

single-cell RNA-sequencing

STAT3=

signal transducer and activator of transcription 3

SMAD4=

SMAD Family Member 4

BIRC5=

survivin

TBC1D7=

Tre2-Bub2-Cdc16 domain family member 7

TCGA=

The Cancer Genome Atlas

TNBC=

triple negative breast cancer

TPX2=

Targeting protein for Xklp2

TBC1D7=

Tre2-Bub2-Cdc16 domain family member 7

TME=

tumor microenvironment

TP53=

Tumor Protein P53

T stage=

tumor staging

TILs=

tumor-infiltrating lymphocytes

UBE2C=

ubiquitin conjugating enzyme E2C

WS=

weighted score

Wnt=

wingless

XMAP215=

Xenopus microtubule-associated protein

Supplemental material

Acknowledgments

We thank Dr. S.C. Hoock for the critical reading of this manuscript.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Data availability statement

All data generated or analyzed during this study are included in this article and its supplementary information.

Correction Statement

This article was originally published with errors, which have now been corrected in the online version. Please see Correction (http://dx.doi.org/10.1080/10408363.2024.2327873).

Additional information

Funding

This project was funded by the “Deutsche Forschungsgemein­schaft“ [DFG, German Research Foundation, project number 413992926, 406036362 and 438690235].

References

  • Gudimchuk NB, McIntosh JR. Regulation of microtubule dynamics, mechanics and function through the growing tip. Nat Rev Mol Cell Biol. 2021;22(12):777–795. doi:10.1038/s41580-021-00399-x.
  • Saunders WS. Mitotic spindle pole separation. Trends Cell Biol. 1993;3(12):432–437. doi:10.1016/0962-8924(93)90032-v.
  • Brouhard GJ, Rice LM. Microtubule dynamics: an interplay of biochemistry and mechanics. Nat Rev Mol Cell Biol. 2018;19(7):451–463. doi:10.1038/s41580-018-0009-y.
  • Wu J, Akhmanova A. Microtubule-organizing centers. Annu Rev Cell Dev Biol. 2017;33(1):51–75. doi:10.1146/annurev-cellbio-100616-060615.
  • Vineethakumari C, Lüders J. Microtubule anchoring: attaching dynamic polymers to cellular structures [mini review]. Front Cell Dev Biol. 2022;10:867870. doi:10.3389/fcell.2022.867870.
  • Michaels TC, Feng S, Liang H, et al. Mechanics and kinetics of dynamic instability. Elife. 2020;9:e54077. doi:10.7554/eLife.54077.
  • Voter WA, Erickson HP. The kinetics of microtubule assembly. Evidence for a two-stage nucleation mechanism. J Biol Chem. 1984;259(16):10430–10438. doi:10.1016/S0021-9258(18)90982-8.
  • Cleary JM, Hancock WO. Molecular mechanisms underlying microtubule growth dynamics. Curr Biol. 2021;31(10):R560–R573. doi:10.1016/j.cub.2021.02.035.
  • Duellberg C, Cade NI, Holmes D, et al. The size of the EB cap determines instantaneous microtubule stability. Elife. 2016;5:e13470. doi:10.7554/eLife.13470.
  • Horio T, Murata T. The role of dynamic instability in microtubule organization. Front Plant Sci. 2014;5:511. doi:10.3389/fpls.2014.00511.
  • Gottschalk AC, Hefti MM. The evolution of microtubule associated proteins – a reference proteomic perspective. BMC Genomics. 2022;23(1):266. doi:10.1186/s12864-022-08502-y.
  • Ramkumar A, Jong BY, Ori-McKenney KM. ReMAPping the microtubule landscape: how phosphorylation dictates the activities of microtubule-associated proteins. Dev Dyn. 2018;247(1):138–155. doi:10.1002/dvdy.24599.
  • Barbier P, Zejneli O, Martinho M, et al. Role of Tau as a microtubule-associated protein: structural and functional aspects. Front Aging Neurosci. 2019;11:204. doi:10.3389/fnagi.2019.00204.
  • Lindeboom JJ, Nakamura M, Saltini M, et al. CLASP stabilization of plus ends created by severing promotes microtubule creation and reorientation. J Cell Biol. 2018;218(1):190–205. doi:10.1083/jcb.201805047.
  • Steinhäuser K, Klöble P, Kreis NN, et al. Deficiency of RITA results in multiple mitotic defects by affecting microtubule dynamics. Oncogene. 2017;36(15):2146–2159. 2017/04/01 doi:10.1038/onc.2016.372.
  • Wang W, Cantos-Fernandes S, Lv Y, et al. Insight into microtubule disassembly by kinesin-13s from the structure of Kif2C bound to tubulin. Nat Commu. 2017;8(1):70.
  • Ritter A, Kreis NN, Louwen F, et al. Molecular insight into the regulation and function of MCAK. Crit Rev Biochem Mol Biol. 2015;51(4):228–245. doi:10.1080/10409238.2016.1178705.
  • Moher D, Liberati A, Tetzlaff J, et al. Preferred reporting items for systematic reviews and meta-analyses: the PRISMA statement. PLoS Med. 2009;6(7):e1000097. doi:10.1371/journal.pmed.1000097.
  • Lawrence CJ, Dawe RK, Christie KR, et al. A standardized kinesin nomenclature. J Cell Biol. 2004;167(1):19–22. doi:10.1083/jcb.200408113.
  • Ali I, Yang WC. The functions of kinesin and kinesin-related proteins in eukaryotes. Cell Adh Migr. 2020;14(1):139–152. doi:10.1080/19336918.2020.1810939.
  • Fan R, Lai K-O. Understanding how kinesin motor proteins regulate postsynaptic function in neuron. FEBS J. 2022;289(8):2128–2144. doi:10.1111/febs.16285.
  • Chen Y, Hancock WO. Kinesin-5 is a microtubule polymerase. Nat Commun. 2015;6(1):8160. doi:10.1038/ncomms9160.
  • Desai A, Verma S, Mitchison TJ, et al. Kin I kinesins are microtubule-destabilizing enzymes. Cell. 1999;96(1):69–78. doi:10.1016/s0092-8674(00)80960-5.
  • Ems-McClung SC, Walczak CE. Kinesin-13s in mitosis: key players in the spatial and temporal organization of spindle microtubules. Semin Cell Dev Biol. 2010;21(3):276–282. doi:10.1016/j.semcdb.2010.01.016.
  • Trofimova D, Paydar M, Zara A, et al. Ternary complex of Kif2A-bound tandem tubulin heterodimers represents a kinesin-13-mediated microtubule depolymerization reaction intermediate. Nature Commun. 2018;9(1):2628.
  • Cooper JR, Wagenbach M, Asbury CL, et al. Catalysis of the microtubule on-rate is the major parameter regulating the depolymerase activity of MCAK. Nat Struct Mol Biol. 2010;17(1):77–82. 2010/01/01 doi:10.1038/nsmb.1728.
  • Wordeman L, Mitchison TJ. Identification and partial characterization of mitotic centromere-associated kinesin, a kinesin-related protein that associates with centromeres during mitosis. J Cell Biol. 1995;128(1–2):95–104. doi:10.1083/jcb.128.1.95.
  • Schmieder S, Nagai M, Orlando RA, et al. Podocalyxin activates RhoA and induces actin reorganization through NHERF1 and Ezrin in MDCK cells. J Am Soc Nephrol. 2004;15(9):2289–2298. doi:10.1097/01.ASN.0000135968.49899.E8.
  • Sanhaji M, Ritter A, Belsham HR, et al. Polo-like kinase 1 regulates the stability of the mitotic centromere-associated kinesin in mitosis. Oncotarget. 2014;5(10):3130–3144. doi:10.18632/oncotarget.1861.
  • Andrews PD, Ovechkina Y, Morrice N, et al. Aurora B regulates MCAK at the mitotic centromere. Dev Cell. 2004;6(2):253–268. doi:10.1016/s1534-5807(04)00025-5.
  • Honnappa S, Gouveia SM, Weisbrich A, et al. An EB1-binding motif acts as a microtubule tip localization signal. Cell. 2009;138(2):366–376. doi:10.1016/j.cell.2009.04.065.
  • Tanenbaum ME, Medema RH, Akhmanova A. Regulation of localization and activity of the microtubule depolymerase MCAK. Bioarchitecture. 2011;1(2):80–87. doi:10.4161/bioa.1.2.15807.
  • Moore AT, Rankin KE, von Dassow G, et al. MCAK associates with the tips of polymerizing microtubules. J Cell Biol. 2005;169(3):391–397. doi:10.1083/jcb.200411089.
  • Lee T, Langford KJ, Askham JM, et al. MCAK associates with EB1. Oncogene. 200827(17):2494–2500. doi:10.1038/sj.onc.1210867.
  • Braun A, Dang K, Buslig F, et al. Rac1 and Aurora a regulate MCAK to polarize microtubule growth in migrating endothelial cells. J Cell Biol. 2014;206(1):97–112. doi:10.1083/jcb.201401063.
  • Moon HH, Kreis NN, Friemel A, et al. Mitotic centromere-associated kinesin (MCAK/KIF2C) regulates cell migration and invasion by modulating microtubule dynamics and focal adhesion turnover. Cancers (Basel). 2021;13(22):5673. doi:10.3390/cancers13225673.
  • Ritter A, Sanhaji M, Friemel A, et al. Functional analysis of phosphorylation of the mitotic centromere-associated kinesin by Aurora B kinase in human tumor cells. Cell Cycle. 2015;14(23):3755–3767. doi:10.1080/15384101.2015.1068481.
  • Eichenlaub-Ritter U. Microtubule dynamics and tumor invasion involving MCAK. Cell Cycle. 2015;14(21):3353–3353. doi:10.1080/15384101.2015.1093813.
  • Schweiggert J, Habeck G, Hess S, et al. SCF(Fbxw5) targets kinesin-13 proteins to facilitate ciliogenesis. EMBO J. 2021;40(18):e107735. doi:10.15252/embj.2021107735.
  • Zhu S, Paydar M, Wang F, et al. Kinesin Kif2C in regulation of DNA double strand break dynamics and repair. Elife. 2020;9:e51636. doi:10.7554/eLife.53402.
  • Ganem NJ, Compton DA. The KinI kinesin Kif2a is required for bipolar spindle assembly through a functional relationship with MCAK. J Cell Biol. 2004;166(4):473–478. doi:10.1083/jcb.200404012.
  • Rogers GC, Rogers SL, Schwimmer TA, et al. Two mitotic kinesins cooperate to drive sister chromatid separation during anaphase. Nature. 2004;427(6972):364–370. doi:10.1038/nature02256.
  • Ganem NJ, Upton K, Compton DA. Efficient mitosis in human cells lacking poleward microtubule flux. Curr Biol. 2005;15(20):1827–1832. doi:10.1016/j.cub.2005.08.065.
  • Yi ZY, Ma XS, Liang QX, et al. Kif2a regulates spindle organization and cell cycle progression in meiotic oocytes. Sci Rep. 2016;6(1):38574. doi:10.1038/srep38574.
  • Miyamoto T, Hosoba K, Ochiai H, et al. The microtubule-depolymerizing activity of a mitotic kinesin protein KIF2A drives primary cilia disassembly coupled with cell proliferation. Cell Rep. 2015;10(5):664–673. doi:10.1016/j.celrep.2015.01.003.
  • Akkaya C, Atak D, Kamacioglu A, et al. Roles of developmentally regulated KIF2A alternative isoforms in cortical neuron migration and differentiation. Development. 2021;148(4):dev192674. doi:10.1242/dev.192674.
  • Miki H, Setou M, Kaneshiro K, et al. All kinesin superfamily protein, KIF, genes in mouse and human. Proc Natl Acad Sci U S A. 2001;98(13):7004–7011. doi:10.1073/pnas.111145398.
  • Hood EA, Kettenbach AN, Gerber SA, et al. Plk1 regulates the kinesin-13 protein Kif2b to promote faithful chromosome segregation. Mol Biol Cell. 2012;23(12):2264–2274. doi:10.1091/mbc.E11-12-1013.
  • Wu C, Orozco C, Boyer J, et al. BioGPS: an extensible and customizable portal for querying and organizing gene annotation resources. Genome Biol. 2009;10(11):R130. doi:10.1186/gb-2009-10-11-r130.
  • Su AI, Wiltshire T, Batalov S, et al. A gene atlas of the mouse and human protein-encoding transcriptomes. Proc Natl Acad Sci U S A. 2004;101(16):6062–6067. doi:10.1073/pnas.0400782101.
  • Wang G, Hu H-B, Chang Y, et al. Rab7 regulates primary cilia disassembly through cilia excision. J Cell Biol. 2019;218(12):4030–4041. doi:10.1083/jcb.201811136.
  • Mashima Y, Nohira H, Sugihara H, et al. KIF24 depletion induces clustering of supernumerary centrosomes in PDAC cells. Life Sci Alliance. 2022;5(11):e202201470. doi:10.26508/lsa.202201470.
  • Kobayashi T, Tsang WY, Li J, et al. Centriolar kinesin Kif24 interacts with CP110 to remodel microtubules and regulate ciliogenesis. Cell. 2011;145(6):914–925. doi:10.1016/j.cell.2011.04.028.
  • Delgehyr N, Rangone H, Fu J, et al. Klp10A, a microtubule-depolymerizing kinesin-13, cooperates with CP110 to control Drosophila centriole length. Curr Biol. 2012;22(6):502–509. doi:10.1016/j.cub.2012.01.046.
  • Burns KM, Wagenbach M, Wordeman L, et al. Nucleotide exchange in dimeric MCAK induces longitudinal and lateral stress at microtubule ends to support depolymerization. Structure. 2014;22(8):1173–1183. doi:10.1016/j.str.2014.06.010.
  • Burns KM, Sarpe V, Wagenbach M, et al. HX-MS2 for high performance conformational analysis of complex protein states. Protein Sci. 2015;24(8):1313–1324. doi:10.1002/pro.2707.
  • Zong H, Hazelbaker M, Moe C, et al. Spatial regulation of MCAK promotes cell polarization and focal adhesion turnover to drive robust cell migration. Mol Biol Cell. 2021;32(7):590–604. doi:10.1091/mbc.E20-05-0301.
  • Ogawa T, Nitta R, Okada Y, et al. A common mechanism for microtubule destabilizers-M type kinesins stabilize curling of the protofilament using the class-specific neck and loops. Cell. 2004;116(4):591–602. doi:10.1016/s0092-8674(04)00129-1.
  • Wang W, Shen T, Guerois R, et al. New insights into the coupling between microtubule depolymerization and ATP hydrolysis by kinesin-13 protein Kif2C. J Biol Chem. 2015;290(30):18721–18731. doi:10.1074/jbc.M115.646919.
  • Talapatra SK, Harker B, Welburn JP. The C-terminal region of the motor protein MCAK controls its structure and activity through a conformational switch. Elife. 2015;4:e06421. doi:10.7554/eLife.06421.
  • Maney T, Wagenbach M, Wordeman L. Molecular dissection of the microtubule depolymerizing activity of mitotic centromere-associated kinesin. J Biol Chem. 2001;276(37):34753–34758. doi:10.1074/jbc.M106626200.
  • Hertzer KM, Ems-McClung SC, Kline-Smith SL, et al. Full-length dimeric MCAK is a more efficient microtubule depolymerase than minimal domain monomeric MCAK. Mol Biol Cell. 2006;17(2):700–710. doi:10.1091/mbc.e05-08-0821.
  • Vale RD, Fletterick RJ. The design plan of kinesin motors. Annu Rev Cell Dev Biol. 1997;13(1):745–777. doi:10.1146/annurev.cellbio.13.1.745.
  • Ovechkina Y, Wagenbach M, Wordeman L. K-loop insertion restores microtubule depolymerizing activity of a "neckless" MCAK mutant. J Cell Biol. 2002;159(4):557–562. doi:10.1083/jcb.200205089.
  • Xia P, Zhou J, Song X, et al. Aurora a orchestrates entosis by regulating a dynamic MCAK–TIP150 interaction. J Mol Cell Biol. 2014;6(3):240–254. doi:10.1093/jmcb/mju016.
  • Pakala SB, Nair VS, Reddy SD, et al. Signaling-dependent phosphorylation of mitotic centromere-associated kinesin regulates microtubule depolymerization and its centrosomal localization. J Biol Chem. 2012;287(48):40560–40569. doi:10.1074/jbc.M112.399576.
  • Lan W, Zhang X, Kline-Smith SL, et al. Aurora B phosphorylates centromeric MCAK and regulates its localization and microtubule depolymerization activity. Curr Biol. 2004;14(4):273–286. doi:10.1016/S0960-9822(04)00064-8.
  • Ohi R, Sapra T, Howard J, et al. Differentiation of cytoplasmic and meiotic spindle assembly MCAK functions by Aurora B-dependent phosphorylation. Mol Biol Cell. 2004;15(6):2895–2906. doi:10.1091/mbc.e04-02-0082.
  • Ems-McClung SC, Hainline SG, Devare J, et al. Aurora B inhibits MCAK activity through a phosphoconformational switch that reduces microtubule association. Curr Biol. 2013;23(24):2491–2499. doi:10.1016/j.cub.2013.10.054.
  • Zhang X, Ems-McClung SC, Walczak CE. Aurora a phosphorylates MCAK to control ran-dependent spindle bipolarity. Mol Biol Cell. 2008;19(7):2752–2765. doi:10.1091/mbc.e08-02-0198.
  • Zong H, Carnes SK, Moe C, et al. The far C-terminus of MCAK regulates its conformation and spindle pole focusing. Mol Biol Cell. 2016;27(9):1451–1464. doi:10.1091/mbc.E15-10-0699.
  • Ritter A, Sanhaji M, Steinhäuser K, et al. The activity regulation of the mitotic centromere-associated kinesin by Polo-like kinase 1. Oncotarget. 2015;6(9):6641–6655. doi:10.18632/oncotarget.2843.
  • Sanhaji M, Friel CT, Kreis NN, et al. Functional and spatial regulation of mitotic centromere-associated kinesin by cyclin-dependent kinase 1. Mol Cell Biol. 2010;30(11):2594–2607. doi:10.1128/MCB.00098-10.
  • Belsham HR, Friel CT. A Cdk1 phosphomimic mutant of MCAK impairs microtubule end recognition. PeerJ. 2017;5:e4034. doi:10.7717/peerj.4034.
  • Jiang K, Wang J, Liu J, et al. TIP150 interacts with and targets MCAK at the microtubule plus ends. EMBO Rep. 2009;10(8):857–865. doi:10.1038/embor.2009.94.
  • Li C, Zhang Y, Yang Q, et al. NuSAP modulates the dynamics of kinetochore microtubules by attenuating MCAK depolymerisation activity. Sci Rep. 2016;6(1):18773. doi:10.1038/srep18773.
  • Seetharaman S, Etienne-Manneville S. Cytoskeletal crosstalk in cell migration. Trends Cell Biol. 2020;30(9):720–735. doi:10.1016/j.tcb.2020.06.004.
  • Garcin C, Straube A. Microtubules in cell migration. Essays Biochem. 201963(5):509–520. doi:10.1042/EBC20190016.
  • Burridge K. Focal adhesions: a personal perspective on a half century of progress. FEBS J. 2017;284(20):3355–3361. doi:10.1111/febs.14195.
  • Seetharaman S, Etienne-Manneville S. Microtubules at focal adhesions - a double-edged sword. J Cell Sci. 2019;132(19):jcs232843.
  • D'Angelo L, Myer NM, Myers KA. MCAK-mediated regulation of endothelial cell microtubule dynamics is mechanosensitive to myosin-II contractility. Mol Biol Cell. 2017;28(9):1223–1237. doi:10.1091/mbc.E16-05-0306.
  • Zaganjor E, Osborne JK, Weil LM, et al. Ras regulates kinesin 13 family members to control cell migration pathways in transformed human bronchial epithelial cells. Oncogene. 2014;33(47):5457–5466. doi:10.1038/onc.2013.486.
  • Sanhaji M, Friel CT, Wordeman L, et al. Mitotic centromere-associated kinesin (MCAK): a potential cancer drug target. Oncotarget. 2011;2(12):935–947. doi:10.18632/oncotarget.416.
  • Hanahan D. Hallmarks of cancer: new dimensions. Cancer Discov. 2022;12(1):31–46. doi:10.1158/2159-8290.CD-21-1059.
  • Hanahan D, Weinberg RA. The hallmarks of cancer. Cell. 2000;100(1):57–70. doi:10.1016/s0092-8674(00)81683-9.
  • Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. Cell. 2011;144(5):646–674. doi:10.1016/j.cell.2011.02.013.
  • Maiato H, Silva S. Double-checking chromosome segregation. J Cell Biol. 2023;222(5):e202301106. doi:10.1083/jcb.202301106.
  • Schubert M, Hong C, Jilderda LJ, et al. Cancer tolerance to chromosomal instability is driven by Stat1 inactivation in vivo. bioRxiv. 2021. doi:10.1101/2021.12.03.471107.
  • Smith JC, Husted S, Pilrose J, et al. MCAK inhibitors induce aneuploidy in triple-negative breast cancer models. Cancers (Basel). 2023;15(13):3309. doi:10.3390/cancers15133309.
  • Wagenbach M, Vicente JJ, Ovechkina Y, et al. Functional characterization of MCAK/Kif2C cancer mutations using high-throughput microscopic analysis. Mol Biol Cell. 2020;31(7):580–588. doi:10.1091/mbc.E19-09-0503.
  • Kim JM. Molecular link between DNA damage response and microtubule dynamics. Int J Mol Sci. 2022;23(13):6986. doi:10.3390/ijms23136986.
  • Jiang CF, Xie YX, Qian YC, et al. TBX15/miR-152/KIF2C pathway regulates breast cancer doxorubicin resistance via promoting PKM2 ubiquitination. Cancer Cell Int. 2021;21(1):542. doi:10.1186/s12935-021-02235-w.
  • Froidevaux-Klipfel L, Targa B, Cantaloube I, et al. Septin cooperation with tubulin polyglutamylation contributes to cancer cell adaptation to taxanes. Oncotarget. 2015;6(34):36063–36080. doi:10.18632/oncotarget.5373.
  • Targa B, Klipfel L, Cantaloube I, et al. Septin filament coalignment with microtubules depends on SEPT9_i1 and tubulin polyglutamylation, and is an early feature of acquired cell resistance to paclitaxel. Cell Death Dis. 2019;10(2):54. doi:10.1038/s41419-019-1318-6.
  • Zhang B, Liu P, Li Y, et al. Multi-omics analysis of kinesin family member 2C in human tumors: novel prognostic biomarker and tumor microenvironment regulator. Am J Cancer Res. 2022;12(11):4954–4976.
  • Kumar A, Rajendran V, Sethumadhavan R, et al. Evidence of colorectal cancer-associated mutation in MCAK: a computational report. Cell Biochem Biophys. 2013 ;67(3):837–851. 2013/12/01 doi:10.1007/s12013-013-9572-1.
  • Dogterom M, Koenderink GH. Actin–microtubule crosstalk in cell biology. Nat Rev Mol Cell Biol. 2019;20(1):38–54. doi:10.1038/s41580-018-0067-1.
  • Mo S, Fang D, Zhao S, et al. Down regulated oncogene KIF2C inhibits growth, invasion, and metastasis of hepatocellular carcinoma through the Ras/MAPK signaling pathway and epithelial-to-mesenchymal transition. Ann Transl Med. 2022;10(3):151–151. doi:10.21037/atm-21-6240.
  • Huang X, Zhao F, Wu Q, et al. KIF2C facilitates tumor growth and metastasis in pancreatic ductal adenocarcinoma. Cancers (Basel). 2023;15(5):1502. doi:10.3390/cancers15051502.
  • Nolte MA, Nolte-‘t Hoen ENM, Margadant C. Integrins control vesicular trafficking; new tricks for old dogs. Trends Biochem Sci. 2021;46(2):124–137. doi:10.1016/j.tibs.2020.09.001.
  • Vinay DS, Ryan EP, Pawelec G, et al. Immune evasion in cancer: mechanistic basis and therapeutic strategies. Semin Cancer Biol. 2015;35 Suppl:S185–S198. Dec doi:10.1016/j.semcancer.2015.03.004.
  • Li R, Cui X, Sun W, et al. ASF1B, as an independent prognostic biomarker, correlates with immune infiltrates in hepatocellular carcinoma. Comb Chem High Throughput Screen. 2023;26(7):1311–1323. doi:10.2174/1386207325666220820112111.
  • Tu B, Xiang H, Li M, et al. In silico analysis of the correlation of KIF2C with prognosis and immune infiltration in glioma. Comput Math Methods Med. 2022;2022:6320828–6320822. doi:10.1155/2022/6320828.
  • Chen S. Glioma subtypes based on the activity changes of immunologic and hallmark gene sets in cancer. Front Endocrinol (Lausanne). 2022;13:879233. doi:10.3389/fendo.2022.879233.
  • Zhang X, Li Y, Hu P, et al. KIF2C is a biomarker correlated with prognosis and immunosuppressive microenvironment in human tumors. Front Genet. 2022;13:891408. doi:10.3389/fgene.2022.891408.
  • Picon-Galindo E, Latz E, Wachten D. Primary cilia and their effects on immune cell functions and metabolism: a model. Trends Immunol. 2022;43(5):366–378. 2022/05/01/doi:10.1016/j.it.2022.03.001.
  • Anvarian Z, Mykytyn K, Mukhopadhyay S, et al. Cellular signalling by primary cilia in development, organ function and disease. Nat Rev Nephrol. 2019;15(4):199–219. 2019/04/01 doi:10.1038/s41581-019-0116-9.
  • Carballo GB, Honorato JR, de Lopes GPF, et al. A highlight on Sonic hedgehog pathway. Cell Commun Signal. 2018;16(1):11. doi:10.1186/s12964-018-0220-7.
  • Kumari R, Jat P. Mechanisms of cellular senescence: cell cycle arrest and senescence associated secretory phenotype. Front Cell Dev Biol. 2021;9:645593. doi:10.3389/fcell.2021.645593.
  • Schmitt CA, Wang B, Demaria M. Senescence and cancer—role and therapeutic opportunities. Nat Rev Clin Oncol. 2022;19(10):619–636. 2022/10/01 doi:10.1038/s41571-022-00668-4.
  • Gwon M-R, Cho JH, Kim J-R. Mitotic centromere-associated kinase (MCAK/Kif2C) regulates cellular senescence in human primary cells through a p53-dependent pathway. FEBS Lett. 2012;586(23):4148–4156. doi:10.1016/j.febslet.2012.10.012.
  • Yang C, Han Z, Zhan W, et al. Predictive investigation of idiopathic pulmonary fibrosis subtypes based on cellular senescence-related genes for disease treatment and management. Front Genet. 2023;14:1157258. doi:10.3389/fgene.2023.1157258.
  • Nakamura Y, Tanaka F, Haraguchi N, et al. Clinicopathological and biological significance of mitotic centromere-associated kinesin overexpression in human gastric cancer. Br J Cancer. 2007;97(4):543–549. Aug 20 doi:10.1038/sj.bjc.6603905.
  • Ishikawa K, Kamohara Y, Tanaka F, et al. Mitotic centromere-associated kinesin is a novel marker for prognosis and lymph node metastasis in colorectal cancer. Br J Cancer. 2008;98(11):1824–1829. Jun 3 doi:10.1038/sj.bjc.6604379.
  • Gnjatic S, Cao Y, Reichelt U, et al. NY-CO-58/KIF2C is overexpressed in a variety of solid tumors and induces frequent T cell responses in patients with colorectal cancer. Int J Cancer. 2010;127(2):381–393. Jul 15 doi:10.1002/ijc.25058.
  • Wang J, Yan Y, Zhang Z, et al. Role of miR-10b-5p in the prognosis of breast cancer. PeerJ. 2019;7:e7728. doi:10.7717/peerj.7728.
  • Lu H, Wang C, Xue L, et al. Human mitotic centromere-associated kinesin is targeted by MicroRNA 485-5p/181c and prognosticates poor survivability of breast cancer. J Oncol. 2019;2019:2316237–2316213. doi:10.1155/2019/2316237.
  • Qitong X, Yiming G, Feng X, et al. A Pan-Cancer Analysis of the KIF Family Gene and their Association with Prognosis, Tumor Microenvironment, and Therapeutic Targets [PREPRINT (Version 1)]. Research Square. 2021.
  • Nishidate T, Katagiri T, Lin M-L, et al. Genome-wide gene-expression profiles of breast-cancer cells purified with laser microbeam microdissection: identification of genes associated with progression and metastasis. Int J Oncol. 2004;25(4):797–819.
  • Shimo A, Nishidate T, Ohta T, et al. Elevated expression of protein regulator of cytokinesis 1, involved in the growth of breast cancer cells. Cancer Sci. 2007;98(2):174–181. doi:10.1111/j.1349-7006.2006.00381.x.
  • Shimo A, Tanikawa C, Nishidate T, et al. Involvement of kinesin family member 2C/mitotic centromere-associated kinesin overexpression in mammary carcinogenesis. Cancer Sci. 2008;99(1):62–70. doi:10.1111/j.1349-7006.2007.00635.x.
  • Sadi AM, Wang DY, Youngson BJ, et al. Clinical relevance of DNA microarray analyses using archival formalin-fixed paraffin-embedded breast cancer specimens. BMC Cancer. 2011;11(1):253:1–13. doi:10.1186/1471-2407-11-253.
  • Kamalakaran S, Varadan V, Giercksky Russnes HE, et al. DNA methylation patterns in luminal breast cancers differ from non-luminal subtypes and can identify relapse risk independent of other clinical variables. Mol Oncol. 2011;5(1):77–92. doi:10.1016/j.molonc.2010.11.002.
  • Wu M, Liu L, Hijazi H, et al. A multi-layer inference approach to reconstruct condition-specific genes and their regulation. Bioinformatics. 2013;29(12):1541–1552. doi:10.1093/bioinformatics/btt186.
  • Xie D, Pei Q, Li J, et al. Emerging role of E2F family in cancer stem cells. Front Oncol. 2021;11:723137. doi:10.3389/fonc.2021.723137.
  • Dai X, Hua T, Hong T. Integrated diagnostic network construction reveals a 4-gene panel and 5 cancer hallmarks driving breast cancer heterogeneity. Sci Rep. 2017;7(1):6827. doi:10.1038/s41598-017-07189-6.
  • Li C, Luo L, Wei S, et al. Identification of the potential crucial genes in invasive ductal carcinoma using bioinformatics analysis. Oncotarget. 2018;9(6):6800–6813. doi:10.18632/oncotarget.23239.
  • Yang K, Gao J, Luo M. Identification of key pathways and hub genes in basal-like breast cancer using bioinformatics analysis. Onco Targets Ther. 2019;12:1319–1331. doi:10.2147/OTT.S158619.
  • Song X, Zhang T, Wang X, et al. Distinct diagnostic and prognostic values of Kinesin family member genes expression in patients with breast cancer. Med Sci Monit. 2018;24:9442–9464. doi:10.12659/MSM.913401.
  • Liu S, Ye Z, Xue VW, et al. KIF2C is a prognostic biomarker associated with immune cell infiltration in breast cancer. BMC Cancer. 2023;23(1):307. doi:10.1186/s12885-023-10788-4.
  • Cao W, Jiang Y, Ji X, et al. Identification of novel prognostic genes of triple-negative breast cancer using meta-analysis and weighted gene co-expressed network analysis. Ann Transl Med. 2021;9(3):205–205. doi:10.21037/atm-20-5989.
  • Suo HD, Tao Z, Zhang L, et al. Coexpression network analysis of genes related to the characteristics of tumor stemness in triple-negative breast cancer. Biomed Res Int. 2020;2020:7575862–7575814. doi:10.1155/2020/7575862.
  • Cai Y, Mei J, Xiao Z, et al. Identification of five hub genes as monitoring biomarkers for breast cancer metastasis in silico. Hereditas. 2019;156(1):20. doi:10.1186/s41065-019-0096-6.
  • Chen G, Yu M, Cao J, et al. Identification of candidate biomarkers correlated with poor prognosis of breast cancer based on bioinformatics analysis. Bioengineered. 2021;12(1):5149–5161. doi:10.1080/21655979.2021.1960775.
  • Li TF, Zeng HJ, Shan Z, et al. Overexpression of kinesin superfamily members as prognostic biomarkers of breast cancer. Cancer Cell Int. 2020;20(1):123. doi:10.1186/s12935-020-01191-1.
  • Paizula X, Mutailipu D, Xu W, et al. Identification of biomarkers related to tumorigenesis and prognosis in breast cancer. Gland Surg. 2022;11(9):1472–1488. doi:10.21037/gs-22-449.
  • Smith JC, Husted S, Pilrose J, et al. MCAK inhibitors induce aneuploidy in triple negative breast cancer models. Cancers. 2023;15:3309. doi:10.1101/2023.05.31.543118.
  • Zhu Q, Zhang L, Sadiq FM, et al. Integrative bioinformatics and RNA sequencing based methodology results in the exploration of breast invasive carcinoma biomarkers. Am J Transl Res. 2023;15(5):3067–3091.
  • Xu A, Xu X-N, Luo Z, et al. Identification of prognostic cancer-associated fibroblast markers in luminal breast cancer using weighted gene co-expression network analysis [original research]. Front Oncol. 2023;13:1191660. doi:10.3389/fonc.2023.1191660.
  • Fang L, Liu Q, Cui H, et al. Bioinformatics analysis highlight differentially expressed CCNB1 and PLK1 genes as potential anti-breast cancer drug targets and prognostic markers. Genes (Basel). 2022;13(4):654. doi:10.3390/genes13040654.
  • Gao Z, Jia H, Yu F, et al. KIF2C promotes the proliferation of hepatocellular carcinoma cells in vitro and in vivo. Exp Ther Med. 2021;22(4):1094–1094. doi:10.3892/etm.2021.10528.
  • Chen J, Li S, Zhou S, et al. Kinesin superfamily protein expression and its association with progression and prognosis in hepatocellular carcinoma. J Cancer Res Ther. 2017;13(4):651–659. doi:10.4103/jcrt.JCRT_491_17.
  • Zhao Z, Mu H, Feng S, et al. Identification of biomarkers associated with hepatocellular carcinoma stem cell characteristics based on co-expression network analysis of transcriptome data and stemness index. Crit Rev Eukaryot Gene Expr. 2022;32(2):47–60. doi:10.1615/CritRevEukaryotGeneExpr.2021039692.
  • Wei S, Dai M, Zhang C, et al. KIF2C: a novel link between Wnt/beta-catenin and mTORC1 signaling in the pathogenesis of hepatocellular carcinoma. Protein Cell. 12(10):788–809. doi:10.1007/s13238-020-00766-y.
  • Ding Q, Jiang C, Zhou Y, et al. Kinesin family member 2C promotes hepatocellular carcinoma growth and metastasis via activating MEK/ERK pathway. Biosci Biotechnol Biochem. 2021;85(11):2241–2249. doi:10.1093/bbb/zbab154.
  • Ferrín G, Guerrero M, Amado V, et al. Activation of mTOR signaling pathway in hepatocellular carcinoma. Int J Mol Sci. 2020;21(4):1266.
  • Zheng H, Cheng ZJ, Liang B, et al. N(6)-methyladenosine modification of ANLN enhances hepatocellular carcinoma bone metastasis. Int J Biol Sci. 2023;19(4):1009–1023. doi:10.7150/ijbs.73570.
  • Sun EJ, Wankell M, Palamuthusingam P, et al. Targeting the PI3K/Akt/mTOR pathway in hepatocellular carcinoma. Biomedicines. 2021;9(11):1639. doi:10.3390/biomedicines9111639.
  • Gao Z, Chen J, Zhou Y, et al. A novel metabolism-related gene signature for predicting the prognosis of HBV-infected hepatocellular carcinoma. J Oncol. 2022;2022:2391265. doi:10.1155/2022/2391265.
  • Ji Y, Yin Y, Zhang W. Integrated bioinformatic analysis identifies networks and promising biomarkers for hepatitis B Virus-Related hepatocellular carcinoma. Int J Genomics. 2020;2020:2061024–2061018. doi:10.1155/2020/2061024.
  • Chen Y, Qian H, He X, et al. Screening of the key genes for the progression of liver cirrhosis to hepatocellular carcinoma based on bioinformatics. J Oncol. 2022;2022:2515513–2515517. doi:10.1155/2022/2515513.
  • Huang R, Liu J, Li H, et al. Identification of hub genes and their correlation with immune infiltration cells in hepatocellular carcinoma based on GEO and TCGA databases. Front Genet. 2021;12:647353. doi:10.3389/fgene.2021.647353.
  • Li X, Huang W, Huang W, et al. Kinesin family members KIF2C/4A/10/11/14/18B/20A/23 predict poor prognosis and promote cell proliferation in hepatocellular carcinoma. Am J Transl Res. 2020;12(5):1614–1639.
  • Wang D, Liu J, Liu S, et al. Identification of crucial genes associated with immune cell infiltration in hepatocellular carcinoma by weighted gene Co-expression network analysis. Front Genet. 2020;11:342. doi:10.3389/fgene.2020.00342.
  • Guo C, Tang Y, Yang Z, et al. Hallmark-guided subtypes of hepatocellular carcinoma for the identification of immune-related gene classifiers in the prediction of prognosis, treatment efficacy, and drug candidates. Front Immunol. 2022;13:958161. doi:10.3389/fimmu.2022.958161.
  • Liu ZY, Li YH, Li BW, et al. Development and validation of a vesicle-mediated transport-associated gene signature for predicting prognosis and immune therapy response in hepatocellular carcinoma. J Cancer Res Clin Oncol. 2023;149(14):13211–13230. Jul 22. doi:10.1007/s00432-023-05079-1.
  • Liu Z, Qi Y, Wang H, et al. Risk model of hepatocellular carcinoma based on cuproptosis-related genes. Front Genet. 2022;13:1000652. doi:10.3389/fgene.2022.1000652.
  • Zhang J, Lou W. A key mRNA-miRNA-lncRNA competing endogenous RNA triple Sub-network linked to diagnosis and prognosis of hepatocellular carcinoma. Front Oncol. 2020;10:340. doi:10.3389/fonc.2020.00340.
  • Ding F, Li J, Zhang Y, et al. Identifying a novel endoplasmic reticulum-related prognostic model for hepatocellular carcinomas. Oxid Med Cell Longev. 2022;2022:8248355–8248324. doi:10.1155/2022/8248355.
  • Mishra D, Mishra A, Rai SN, et al. Identification of prognostic biomarkers for suppressing tumorigenesis and metastasis of hepatocellular carcinoma through transcriptome analysis. Diagnostics. 2023;13(5):965. doi:10.3390/diagnostics13050965.
  • Song YJ, Tan J, Gao XH, et al. Integrated analysis reveals key genes with prognostic value in lung adenocarcinoma. Cancer Manag Res. 2018;10:6097–6108. doi:10.2147/CMAR.S168636.
  • Bai Y, Xiong L, Zhu M, et al. Co-expression network analysis identified KIF2C in association with progression and prognosis in lung adenocarcinoma. Cancer Biomark. 2019;24(3):371–382. doi:10.3233/CBM-181512.
  • Gan H, Lin L, Hu N, et al. KIF2C exerts an oncogenic role in nonsmall cell lung cancer and is negatively regulated by miR-325-3p. Cell Biochem Funct. 2019;37(6):424–431. Aug doi:10.1002/cbf.3420.
  • Abdel-Maksoud MA, Hassan F, Mubarik U, et al. An in-silico approach leads to explore six genes as a molecular signatures of lung adenocarcinoma. Am J Cancer Res. 2023;13(3):727–757.
  • Guo J, Zhang W, Sun L, et al. KIF2C accelerates the development of non-small cell lung cancer and is suppressed by miR-186-3p via the AKT-GSK3β-β-catenin pathway. Sci Rep. 2023;13(1):7288. doi:10.1038/s41598-023-30073-5.
  • Yao S, Zhao T, Jin H. Expression of MicroRNA-325-3p and its potential functions by targeting HMGB1 in non-small cell lung cancer. Biomed Pharmacother. 2015;70:72–79. doi:10.1016/j.biopha.2015.01.013.
  • Meng F, Zhang L, Ren Y, et al. Transcriptome analysis reveals key signature genes involved in the oncogenesis of lung cancer. Cancer Biomark. 2020;29(4):475–482. doi:10.3233/CBM-200110.
  • Zhao D, Mu HJ, Shi HB, et al. Identification of therapeutic targets and mechanisms of tumorigenesis in non-small cell lung cancer using multiple-microarray analysis. Medicine (Baltimore). 2020;99(44):e22815. doi:10.1097/MD.0000000000022815.
  • Chen B, Xie X, Lan F, et al. Identification of prognostic markers by weighted gene co-expression network analysis in non-small cell lung cancer. Bioengineered. 2021;12(1):4924–4935. doi:10.1080/21655979.2021.1960764.
  • Mushtaq A, Singh P, Tabassum G, et al. Unravelling hub genes as potential therapeutic targets in lung cancer using integrated transcriptomic meta-analysis and in silico approach. J Biomol Struct Dynam. 2022; 1:1–14.
  • Duda P, Akula SM, Abrams SL, et al. Targeting GSK3 and associated signaling pathways involved in cancer. Cells. 2020;9(5):1110. doi:10.3390/cells9051110.
  • Bie L, Zhao G, Wang YP, et al. Kinesin family member 2C (KIF2C/MCAK) is a novel marker for prognosis in human gliomas. Clin Neurol Neurosurg. 2012;114(4):356–360. May doi:10.1016/j.clineuro.2011.11.005.
  • Zhou J, Guo H, Liu L, et al. Construction of co-expression modules related to survival by WGCNA and identification of potential prognostic biomarkers in glioblastoma. J Cell Mol Med. 2021;25(3):1633–1644. doi:10.1111/jcmm.16264.
  • Qi C, Lei L, Hu J, et al. Serine incorporator 2 (SERINC2) expression predicts an unfavorable prognosis of Low-Grade Glioma (LGG): evidence from bioinformatics analysis. J Mol Neurosci. 2020;70(10):1521–1532. doi:10.1007/s12031-020-01620-w.
  • Zhao L, Zhang J, Liu Z, et al. Identification of biomarkers for the transition from low-grade glioma to secondary glioblastoma by an integrated bioinformatic analysis. Am J Transl Res. 2020;12(4):1222–1238.
  • Duan H, Zhang X, Wang FX, et al. KIF-2C expression is correlated with poor prognosis of operable esophageal squamous cell carcinoma male patients. Oncotarget. 2016;7(49):80493–80507. doi:10.18632/oncotarget.11492.
  • Li S, Ma Y, Wu C, et al. Knockdown of kinesin family member 2C restricts cell proliferation and induces cell cycle arrest in gastric cancer. Histol Histopathol. 2022;30:18556.
  • Kawamoto M, Tanaka F, Mimori K, et al. Identification of HLA-A*0201/-a*2402-restricted CTL epitope-peptides derived from a novel cancer/testis antigen, MCAK, and induction of a specific antitumor immune response. Oncol Rep. 2011;25(2):469–476. doi:10.3892/or.2010.1101.
  • Yang Y, Gao L, Weng NN, et al. Identification of novel molecular therapeutic targets and their potential prognostic biomarkers among kinesin superfamily of proteins in pancreatic ductal adenocarcinoma. Front Oncol. 2021;11:708900. doi:10.3389/fonc.2021.708900.
  • Xiong J, Wu R, He A, et al. Comprehensive analysis of the effects of KIF2C on prognosis, biological functions and immune infiltration in PAAD. Tissue Cell. 2022;78:101900. doi:10.1016/j.tice.2022.101900.
  • Zhang P, Gao H, Ye C, et al. Large-scale transcriptome data analysis identifies KIF2C as a potential therapeutic target associated with immune infiltration in prostate cancer. Front Immunol. 2022 ;13:905259. doi:10.3389/fimmu.2022.905259.
  • Gu Y, Lu L, Wu L, et al. Identification of prognostic genes in kidney renal clear cell carcinoma by RNA‑seq data analysis. Mol Med Rep. 2017;15(4):1661–1667. doi:10.3892/mmr.2017.6194.
  • Wu G, Xia P, Yan S, et al. Identification of unique long non-coding RNAs as putative biomarkers for chromophobe renal cell carcinoma. Per Med. 2021;18(1):9–19. doi:10.2217/pme-2020-0020.
  • Yang J, Wu Z, Yang L, et al. Characterization of kinesin family member 2C as a proto-oncogene in cervical cancer. Front Pharmacol. 2021;12:785981. doi:10.3389/fphar.2021.785981.
  • Wan B, Yang Y, Zhang Z. Identification of differentially methylated genes associated with clear cell renal cell carcinoma and their prognostic values. J Environ Public Health. 2023;2023:8405945–8405910. doi:10.1155/2023/8405945.
  • Wang J, Zhang W, Hou W, et al. Molecular characterization, tumor microenvironment association, and drug susceptibility of DNA methylation-driven genes in renal cell carcinoma. Front Cell Dev Biol. 2022;10:837919. doi:10.3389/fcell.2022.837919.
  • Wei R, Quan J, Li S, et al. Integrative analysis of biomarkers through machine learning identifies stemness features in colorectal cancer. Front Cell Dev Biol. 2021;9:724860. doi:10.3389/fcell.2021.724860.
  • Wang X, Hu S, Ji W, et al. Identification of genes associated with clinicopathological features of colorectal cancer. J Int Med Res. 2020;48(4):300060520912139. Apr doi:10.1177/0300060520912139.
  • Zhang W, Gao L, Wang C, et al. Combining bioinformatics and experiments to identify and verify key genes with prognostic values in endometrial carcinoma. J Cancer. 2020;11(3):716–732. doi:10.7150/jca.35854.
  • Fu X, Cheng S, Wang W, et al. TCGA dataset screening for genes implicated in endometrial cancer using RNA-seq profiling. Cancer Genet. 2021;254-255:40–47. Jun doi:10.1016/j.cancergen.2021.01.011.
  • Yuan Y, Chen Z, Cai X, et al. Identification of hub genes correlated with poor prognosis for patients with uterine corpus endometrial carcinoma by integrated bioinformatics analysis and experimental validation. Front Oncol. 2021;11:766947. doi:10.3389/fonc.2021.766947.
  • An L, Zhang J, Feng D, et al. KIF2C is a novel prognostic biomarker and correlated with immune infiltration in endometrial cancer. Stem Cells Int. 2021;2021:1434856–1434813. doi:10.1155/2021/1434856.
  • Li W, Liu Z, Liang B, et al. Identification of core genes in ovarian cancer by an integrative meta-analysis. J Ovarian Res. 2018;11(1):94. doi:10.1186/s13048-018-0467-z.
  • Wang W, Zhang W, Hu Y. Identification of keygenes, miRNAs and miRNA-mRNA regulatory pathways for chemotherapy resistance in ovarian cancer. PeerJ. 2021;9:e12353. doi:10.7717/peerj.12353.
  • Zhao F, Siu MK, Jiang L, et al. Overexpression of forkhead box protein M1 (FOXM1) in ovarian cancer correlates with poor patient survival and contributes to paclitaxel resistance. PLoS One. 2014;9(11):e113478. doi:10.1371/journal.pone.0113478.
  • Ni L, Chen Y, Yang J, et al. Bioinformatic analysis of key pathways and genes shared between endometriosis and ovarian cancer. Arch Gynecol Obstet. 2022;305(5):1329–1342. May doi:10.1007/s00404-021-06285-3.
  • Dong C, Tian X, He F, et al. Integrative analysis of key candidate genes and signaling pathways in ovarian cancer by bioinformatics. J Ovarian Res. 2021;14(1):92. doi:10.1186/s13048-021-00837-6.
  • Oh CK, Kang JW, Lee Y, et al. Role of kif2c, a gene related to ALL relapse, in embryonic hematopoiesis in zebrafish. Int J Mol Sci. 2020;21(9):3127.
  • Tanenbaum ME, Macurek L, van der Vaart B, et al. A complex of Kif18b and MCAK promotes microtubule depolymerization and is negatively regulated by Aurora kinases. Curr Biol. 2011;21(16):1356–1365. doi:10.1016/j.cub.2011.07.017.
  • Zuo X, Meng P, Bao Y, et al. Cell cycle dysregulation with overexpression of KIF2C/MCAK is a critical event in nasopharyngeal carcinoma. Genes Dis. 2021;10(1):212–227. doi:10.1016/j.gendis.2021.05.003.
  • Mo BY, Li GS, Huang SN, et al. The underlying molecular mechanism and identification of transcription factor markers for laryngeal squamous cell carcinoma. Bioengineered. 2021;12(1):208–224. doi:10.1080/21655979.2020.1862527.
  • Chen W, Liao L, Lai H, et al. Identification of core biomarkers associated with pathogenesis and prognostic outcomes of laryngeal squamous-cell cancer using bioinformatics analysis. Eur Arch Otorhinolaryngol. 2020;277(5):1397–1408. doi:10.1007/s00405-020-05856-5.
  • Zheng W, Zhao Y, Wang T, et al. Identification of hub genes associated with bladder cancer using bioinformatic analyses. Transl Cancer Res. 2022;11(5):1330–1343. doi:10.21037/tcr-22-1004.
  • Yang C, Li Q, Chen X, et al. Circular RNA circRGNEF promotes bladder cancer progression via miR-548/KIF2C axis regulation. Aging (Albany NY). 2020;12(8):6865–6879. doi:10.18632/aging.103047.
  • Wang J, Chen C, Wang L, et al. Patient-derived tumor organoids: new progress and opportunities to facilitate precision cancer immunotherapy. Front Oncol. 2022;12:872531. doi:10.3389/fonc.2022.872531.
  • Wang S-W, Gao C, Zheng Y-M, et al. Current applications and future perspective of CRISPR/Cas9 gene editing in cancer. Mol Cancer. 2022;21(1):57. doi:10.1186/s12943-022-01518-8.
  • Zheng R, Du Y, Wang X, et al. KIF2C regulates synaptic plasticity and cognition in mice through dynamic microtubule depolymerization. Elife. 2022;11:e72483. doi:10.7554/eLife.72483.
  • Jovic D, Liang X, Zeng H, et al. Single-cell RNA sequencing technologies and applications: a brief overview. Clin Transl Med. 2022;12(3):e694. doi:10.1002/ctm2.694.
  • Jayanthi V, Das AB et al: Grade-specific diagnostic and prognostic biomarkers in breast cancer. Genomics 2020, 112(1):388–396.
  • Zheng L, Li L et al: Six Novel Biomarkers for Diagnosis and Prognosis of Esophageal squamous cell carcinoma: validated by scRNA-seq and qPCR. Journal of Cancer 2021, 12(3):899–911.