Open access peer-reviewed chapter

Two Tails for Neurofibromin: A Tale of Two Microtubule-Associated Proteins

Written By

Charoula Peta, Emmanouella Tsirimonaki, Constantinos Fedonidis, Xeni Koliou, Nikos Sakellaridis and Dimitra Mangoura

Submitted: 23 March 2021 Reviewed: 05 April 2021 Published: 29 July 2021

DOI: 10.5772/intechopen.97574

From the Edited Volume

Clinical and Basic Aspects of Neurofibromatosis Type 1

Edited by Juichiro Nakayama and Yuichi Yoshida

Chapter metrics overview

366 Chapter Downloads

View Full Metrics

Abstract

Neurofibromatosis type 1, NF-1, is a common monogenic (NF1) disease, characterized by highly variable clinical presentation and high predisposition for tumors, especially those of astrocytic origin (low- to high-grade gliomas). Unfortunately, very few genotype–phenotype correlations have been possible, and the numerous identified mutations do not offer help for prognosis and patient counselling. Whole gene deletion in animals does not successfully model the disease, as NF-1 cases caused by point mutations could be differentially affected by cell type-specific alternative splice variants of NF1. In this chapter, we will discuss the differential Microtubule-Associated-Protein (MAP) properties of NLS or ΔNLS neurofibromins, produced by the alternatively splicing of exon 51, which also contains a Nuclear Localization Sequence (NLS), in the assembly of the mitotic spindle and in faithful genome transmission. We will also commend on the major theme that emerges about NLS-containing tumor suppressors that function as mitotic MAPs.

Keywords

  • NLS and ΔNLS neurofibromins
  • astrocyte
  • glioblastoma
  • astral microtubules
  • spindle
  • Microtubule-Associated-Proteins
  • chromosome segregation

1. Introduction

Neurofibromatosis type 1 (NF-1) is a common, complex multisystem cancer predisposition syndrome, with a worldwide incidence at birth of 1: 2–3000 people [1] and a documented high mortality mostly due to malignancies [2]. NF-1 is caused by autosomal, dominantly inherited or de novo (50: 50 [3]) pathogenic mutations in the NF1 gene, which encodes the large protein neurofibromin. The NF1 gene was identified 30 years ago, yet with over 3000 different mutations identified thus far [4], only very few genotype–phenotype correlations have been postulated [5, 6, 7]. Affected individuals may present with a wide range of clinical manifestations, mostly from the Central Nervous System, CNS and the Peripheral Nervous System, PNS [1], as the NF1 gene remains highly expressed there, whereas is downregulated in most other tissues in the adult. Thus, most NF-1 tumors are found in the CNS (gliomas) or the PNS (plexiform neurofibromas, malignant peripheral nerve sheath tumors, or the hallmark of the disease sub- and cutaneous neurofibromas), while there is increased risk for other cancers mostly of neural crest origin [8, 9].

In particular high-grade gliomas are more frequent in adults with NF-1, which have 5 times greater risk for glioblastoma (GBM) than the general population [10]. In addition, the great mutation rate of the NF1 gene, which has also made its cloning impossible, is now recognized in ~20% of sporadic GBM [11]. In terms of specific treatments none are available for the cancers of NF-1 patients. Many drugs, like anti-angiogenic agents [12], have shown no responses, and MEK1/2 inhibitors have been approved only for plexiform neurofibromas. GBM prognosis remains dire (~2 years) even with the use of the highly cytotoxic temozolomide, while clinicians daily struggle with decisions for affected children. Unfortunately, gliomas frequently are resistant to temozolomide therapy and the candidate mechanism, for other tumors too, is the formation of tumor microtubes [13]. These recently recognized long, highly rich in actin, dynamic membrane protrusions of astrocytoma cells form a network for multicellular communication that promotes tumor growth and invasion of the brain. Therefore, understanding the cytoskeleton associations of neurofibromin is highly important in the effort to identify new therapeutic targets.

Confirmation of causative mutation with molecular diagnosis is a difficult task, as the large NF1 gene of over 400 Kb and 57 exons has no mutational hot spots and one of the highest mutation rates in human genetic disorders, which explains the high incidence of de novo variants even within the same family. The complicated behavior of the gene is further highlighted from genetic manipulations of mice. When Nf1 along with two more tumor suppressor genes (TSGs) were targeted with CRISPR/Cas in the forebrain of E13.5 mice, aggressive tumors resembling human GBMs were produced; however, whole-genome sequencing of the induced GBMs hinted to a very variable repair of CRISPR-induced double-strand breaks, potentially locus-specific [14]. Thus, mouse genetic NF-1 models have been marginally helpful in designing prognostic tests or therapies.

Even when timely, molecular diagnosis may only rarely offer help for prognosis or consultation [5, 7] and the challenge to correlate genotype–phenotype in this disease of uncontrolled cell growth and tumorigenesis remains largely unmet. It is our opinion that the impact of the various mutations will be best appreciated, once our currently limited knowledge on the functions of the distinct neurofibromin protein domains (Figure 1) will be expanded. As we will elaborate, these domains perform critical functions, evidently through inter- and intra-molecular interactions, most notably with tubulins, all of which are altered by cell type-specific, alternative splicing events and post-translational modifications, with mutations adding a supra level of complexity that must be met.

Figure 1.

Major NF1 splicing events and neurofibromin domains with functional importance in the CNS. A, the two major alternative exons in the NF1 gene, namely 31 and 51, which produce GRDI or GRDII and NLS or ΔNLS transcripts, respectively. B, Neurofibromin domains of known functional importance: CSRD, Cysteine/Serine-Rich domain; GRD, GAP-related domain; SEC14, Yeast Sec14p-like domain; PH, pleckstrin homology domain; CTD, C-Terminal domain and NLS, Nuclear localization sequence. Amino acid numbers for all putative isoforms are based on GRDI-NLS neurofibromin (Ensembl transcript NF1–201).

Indeed, the quest for genotype–phenotype correlation is complicated by developmental stage- and tissue- or cell type-specific alternative splicing events of the NF1 gene, which is downregulated in most tissues in the adult except CNS and PNS tissues. While several alternative exons have been described, three are common in the CNS, namely 9a/9br, 31 (former 23), and 51 (former 43) [15, 16, 17]. Addition of the small alternative exon 9a/9br correlates well with neuronal differentiation and is downregulated in oligodendrogliomas [15, 18], but no specific function assignments have been made thus far. In contrast, inclusion of the other two, namely 31 and 51, does have important functional consequences, as will be discussed next.

Skipping of alternative exon 31, which corresponds to the center of the RasGAP-related domain (GRD) of neurofibromin (Figure 1A) through which neurofibromin inactivates Ras, generates two variants accordingly named as GRD type I, GRDI and, if exon 31 is included, as GRD-type II, GRDII [16]. Exon 31 is mostly skipped in CNS neurons early on, whereas it is retained as the prominent transcript in astrocytes [19, 20, 21]. Due to the central role of Ras in many cellular functions and in carcinogenesis, GRDs have received high attention. Both GRDs are functional RasGAPs [22], when overexpressed in vitro [23] and as we showed in vivo [24], albeit GRDI is a much more potent RasGAP than GRDII.

Nevertheless, no significant rescue capacity of GRDs alone has been shown for NF-1 phenotypes, leading to the characterization of such phenotypes as Ras-independent by many researchers [25, 26, 27]. Along several such scientific efforts, the importance of other domains (Figure 1B) in the allosteric regulation of GRD has been established. Indeed, collective experimental evidence has postulated that neurofibromin domains may bind each other to form dimers [28], as well as, multiple proteins to coordinate Ras signalling ([25, 26, 27, 29, 30], reviewed in [31]). For example, in glioma cells overexpression of CSRD plus GRD -after phosphorylations by PKCε or α- binding to cortical F-actin increases and imposes a positive allosteric regulation on GRD and thus intense Ras deactivation, which is sufficient to switch the effect of EGF signalling from proliferation to differentiation [30]. While this mechanism was the first provided answer to the open question of how RasGAPs are recruited to the membrane, its clinical significance was directly postulated when large cohorts of NF-1 patients, heterozygous for nonsynonymous mutations of any of five successive amino acids in the CSRD, were found to have high, >50% predisposition to malignancies as compared to the general NF-1-affected population [57]. SEC14, also reported to halt glioma cell invasion [32], is a domain that mediates binding to phospholipids [33, 34] and, as we showed, imposes, like CSRD, a positive allosteric regulation on GRD and accelerates Ras deactivation, potent enough to switch the activation of ERK from an analogous to a digital mode [24].

Exon 51 in the CTD contains the NLS, a sequence of basic amino acid clusters required for proteins of >45 kDa to dock onto the nuclear pore complex as a cargo for nuclear import. The necessary energy expenditure and the directionality of the import is provided by a gradient of RanGDP in the cytoplasm and RanGTP in the nucleus. A similar Ran gradient, generated around duplicated chromosomes during mitosis, allows the release of NLS-containing mitotic proteins that regulate spindle assembly and congression of chromosomes [35, 36]. NF1 exon 51 may be also alternatively transcribed, producing NLS or non NLS (ΔNLS) transcripts and corresponding NLS or ΔNLS neurofibromin isoforms (Figure 1). We first identified in silico this bipartite NLS and documented experimentally that most neurofibromin molecules reside in the nucleus in neurons [37].

Later genetic analysis [17] revealed that in those fetal tissues, which will not retain high levels of NF1 expression in the adult, ΔNLS transcripts are expressed in higher percentages. In contrast, in tissues that NF1 remains high in the adult, fetal expression of ΔNLS is very low early on and increases with development. Typical examples for the former is the liver (ΔNLS constitutes 25% of total NF1 expression in the fetus and only 15% in the adult) and for the latter is the brain, where the meek expression of 1% rises by 4-fold in the adult. Thus, there is an upregulation of ΔNLS transcripts in the tissue most implicated in NF-1 pathology, that is the CNS [17].

Moreover, we recently addressed the pressing question of developmental regulation of exon 51 skipping/inclusion in CNS cell types, using chick embryo or the early postnatal mouse or rat brains. We find that expression of ΔNLS is first detected only when neurons become postmitotic with its levels rising from negligent to 10% of total NF1 in mature neurons; in astrocytes in culture, ΔNLS transcripts rise along with those for glial fibrillary acidic protein (GFAP) and reach a level of ~5% of total NF1 [19]. Thus, our studies postulate that in both neurons and astrocytes as many as four variants and neurofibromin isoforms may be expressed (Figure 1), while expression of ΔNLS transcripts and ΔNLS neurofibromins correlates with neuronal and astrocytic differentiation and underline the necessity to study the individual functions of ΔNLS and NLS neurofibromins.

The importance of NLS inclusion was totally unexplored, till we documented a few years ago that neurofibromin controls the pivotal function of chromosome congression on the mitotic spindle [38] (Figure 2) and then proved that depletion of NLS neurofibromins deregulates spindle assembly and positioning, leading to aneuploidy and increased micronuclei formation [39]. More importantly, these studies established the function of neurofibromins as MAPs. All currently known impacts of this function will be presented in more detail in the next section.

Figure 2.

Neurofibromins regulate spindle configuration and chromosome congression. A, Mitotic spindles consist of three major types of microtubules (MTs): astral MTs that radiate from the centrosomes/poles, microtubule bundles (K-fibers) to link kinetochores to poles, and interpolar bundles to separate poles, elongate the spindle, and bridge K-fibers [40]. B, SF268 glioblastoma cells, transduced with mock or NF1-specific siRNAs, are stained for neurofibromin, β-tubulin, and chromatin, as indicated. In mock siRNA-cells, neurofibromin decorates astral MTs (asterisks) and both K-fibers and interpolar MTs (arrowheads), in a symmetric spindle with properly aligned chromosomes at its equator (white arrows). Depletion of neurofibromins (siNF1) causes irregularities in the spindle geometry and chromosome congression aberrations (yellow arrows) [38]. Images are the maximal intensity projection of 0.34 μm confocal plane stacks.

In concluding this introduction, we believe that the importance of exploring novel yet fundamental questions on the functions of NLS neurofibromin isoforms is tantamount for understanding patient phenotypes and designing prognostic tools for NF-1 glioma growth and NF-1 mutation-specific therapies. Therefore, in this Chapter we will focus on novel insights on the MAP function of neurofibromins from our recent studies.

Advertisement

2. Neurofibromins as MAPs and their role in mitosis

Tubulins rapidly form highly dynamic noncovalent polymers, the microtubules, which execute essential functions for the constant yet ever-changing needs of all cells, such as function-coupled shapes, directed intracellular transport, migration, and, most importantly for the development of an organism, properly oriented cell divisions with accurate genomic transmission. For cell division, several types of MTs organize, elongate, and orient a bipolar spindle, through which chromosomes will position at the spindle equator for faithful sister chromatid separation and then segregation to the two daughter cells ([41, 42, 43] and Figure 2A).

Accordingly, multitudes of structurally different MAP proteins associate with MTs to regulate MT nucleation, polymerization, organization, bundling, and crosslinking in preparation for and completion of mitosis. The availability of mitotic MAPs is tightly regulated by coordinated transcription, as well as by cell cycle-dependent post-translational modifications, most often phosphorylations that control protein trafficking, homeostasis, and inter- or intramolecular interactions [44, 45, 46]. Aberrations in these processes may lead to aneuploidy and on to tumorigenesis, thus the ability of MAPs to alter MT dynamics is recognized for its prognostic value in cancer and as a target for cancer chemotherapies [47, 48].

Association of neurofibromin with cytosolic MTs was first established by confocal microscopy in fibroblasts and the molecule was proposed to act as a MAP, through a small segment (residues 815–834 in the CSRD) bearing in silico homology to MAPs Tau and MAP2 [49]. Since then, diverse experimental approaches, including co-immunoprecipitations, co-purifications, in vitro MT assembly, and affinity precipitations, have further documented this interaction with cytosolic [23, 24, 29, 37, 38, 50] and with mitotic MTs [38, 39].

Indeed, confocal image analysis of primary or tumor cells derived from the ectoderm or the neural crest and quadruply stained for β-tubulin, neurofibromin, F-actin and chromatin/chromosomes, shows that pools of endogenous neurofibromin colocalize with cytoplasmic MTs (e.g., rat astrocytes in Figure 3A, yellow arrows), as well as with F-actin, mainly at the cell cortex and lamellipodia ([37, 38, 39]; Figure 3B, yellow arrows). Interestingly, no association could be established with any intermediate filament in neural cells, as for example with the abundant astrocytic marker GFAP (Figure 3C), except for nuclear lamins [38].

Figure 3.

Endogenous neurofibromins colocalize with both A. cytoplasmic MTs and B. F-actin, but not with C. GFAP. Rat primary astrocytes are stained for neurofibromin, chromatin (Hoechst 33258) and A, β-tubulin, B, F-actin (phalloidin) or C, GFAP. Images are the maximal intensity projection of 0.34 μm confocal planes; yellow arrows point to co-localization of neurofibromin with cytoplasmic MTs in A and actin in B; white arrows point to nuclear neurofibromin.

In an earlier study with primary neurons, we found that neurofibromin, in addition to its colocalization with cytosolic MTs, localizes also in the nucleus and identified a bipartite NLS in the CTD (Figure 1B) as the probable mediator of nuclear entry [37]. Previously thought as an artifact in the skin epithelium, nuclear neurofibromin is detected with a variety of techniques, i.e., immunocytochemistry (e.g., white arrows in Figure 3), subcellular fractionations, or proteomics [28, 29, 37, 38, 39, 51, 52, 53] in all cells of an ectodermal origin.

We next provided evidence that the nuclear entry of neurofibromin is active, that is through interactions of its NLS with the Ran/importin system [38], as now shown in cancer breast cells [54]. Moreover, we established that a requirement for the cell-cycle regulated nuclear entry of neurofibromin is phosphorylation by Protein Kinase C (PKC) on Serine2808, a residue relatively close to the NLS [29, 38], which is retained in both isoforms. As neurofibromin expression patterns and nuclear regulation appeared to have all the attributes of a mitotic factor, in particular of a MAP, we next addressed the possibility in cells that regularly undergo mitosis.

Thus, we have postulated that neurofibromin primarily co-localizes with β-tubulin at all stages of spindle assembly from prophase to metaphase (e.g., Figure 2B) and then through the transformation of the spindle to a machinery for chromosome segregation and cell division, that’s is through anaphase, telophase and cytokinesis [38, 39]. Again, neurofibromin’s localization onto the spindle is apparent in all cells of ectodermal origin, with no exception. More relevant for gliomagenesis, endogenous neurofibromin in primary cortical or cerebellar astrocytes colocalizes with all three tubulin classes on microtubular structures, that is with α- and β-tubulin throughout mitosis and with γ-tubulin at the centrosomes at interphase and when duplicated for entrance to mitosis [19, 38, 39].

Experimentally, at least three neurofibromin domains have been previously identified to bind tubulins, namely GRD, SEC14, and CTD. Affinities to tubulin for the first two domains were explored for regulation of neurofibromin’s RasGAP activity and the third for baiting neurofibromin associated proteins or for nuclear import studies. Thus, GRDI-tubulin interactions lead to a partial inhibition of its cytosolic GAP activity [23] and certain patient mutations in GRD impair the ability of neurofibromin to associate with MTs [55], while the competition of tubulin with H-Ras for binding to SEC14 that we found in COS cells may provide an explanation and a mechanism for neurofibromin dissociation away from cytoplasmic microtubules [24]. As for CTD, it baits the plus-end MAP Collapsin response mediator protein 2 (CRMP2) [50], while we have shown high affinity of GFP-CTD(+NLS) to α-, β- and γ-tubulins [38].

The importance of neurofibromin as a mitotic MAP was first discovered, when we showed that siRNA-depletion of all transcripts and thus of all neurofibromin isoforms leads to severe errors in chromosome congression, with chromosomes remaining unattached or randomly away from the spindle equator even at full metaphasic spindle length (Figure 2B, yellow arrows and [38]). Typically, the unstable lateral interactions between kinetochores and microtubules, which dominate early prometaphase, lead to the reproducible arrangement of chromosomes in an equatorial ring, or torus-like distribution on the surface of the spindle [56]. This loss of the toroidal arrangement of chromosomes with neurofibromin depletion has to be the first evidence that neurofibromin may act to stabilize microtubule for chromosome congression. Consistent with this, using overexpressions of our human CTD construct but not of other domains, abnormal bypassing of mitotic arrest was rescued in the yeast, after the yeast homologs of neurofibromin Ira1 and 2 were deleted [25].

The importance of neurofibromin isoforms as major mitotic MAPs was next discovered, when we probed the individual effects of neurofibromin isoforms that differ in the sequence of the 41 amino acids encoded by exon 51, namely of ΔNLS- and NLS-neurofibromins, on mitotic spindle assembly and faithful genomic transmission [39]. These effects will be highlighted next.

Advertisement

3. NLS and ΔNLS neurofibromins are different MAPS

To further address the mechanism by which neurofibromin regulates chromosome congression and considering together that a. neurofibromin accumulates in the nucleus in a Ran-dependent manner at late S/G2 and resides on the spindle throughout mitosis [38], b. the major cellular target in NF-1 for abnormal proliferation and carcinogenesis is the astrocyte [10], and c. the higher expression of NLS- over ΔNLS -NF1 transcripts in astrocytes [19], we next evaluated separately the roles of ΔNLS- and NLS-isoforms in spindle assembly and chromosome segregation in an astrocytic cell context.

For these purposes, we have generated SF268 glioblastoma cell lines that stably express, under the control of doxycycline, shRNAs specifically designed to degrade either both GRDI- and GRDII-ΔNLS or both GRDI- and GRDII-NLS-NF1 transcripts (referred to as NLS-cells and ΔNLS-cells, respectively). This reversible genetic modification has allowed us to pose the question of possible different functions of ΔNLS- and NLS-isoforms and dissect their roles in mitosis [39].

Confocal image analysis of cells immunostained for β-tubulin and neurofibromin and co-stained for filamentous actin, along quantitation of colocalization, postulates that ΔNLS neurofibromins are absent from the nucleus [39]. Moreover, in ΔNLS-cells association of neurofibromin with F-actin is significantly limited, especially in lamellipodia, whereas NLS-neurofibromins richly decorate them along other actin structures. Association with tubulin is not significantly reduced in ΔNLS-cells, yet microtubule organizing centers (MTOCs) are discerned with difficulty, because MTs organize a dense but non-radial network. To the contrary, NLS neurofibromin colocalization with β-tubulin is significantly enhanced, although β-tubulin intensity itself is not increased [39].

This different robustness of MTOC formation among ΔNLS-cells and the parental or NLS-cells is functionally validated with cell migration after wound (scratch) assays. In astrocytes, relocation of their major MTOC, the centrosome, between the nucleus and the leading-edge during migration is well explained [57]. When positions of centrosomes and nuclei are observed in cells stained for γ-tubulin and Hoechst, respectively, confocal microscopy shows that, unlike parental and NLS-cells, it is readily apparent that centrosomes in ΔNLS-cells fail to position properly, remaining randomly oriented [39]. Hence, time-lapse video microscopy of cells during wound healing confirms that NLS-cells and the parental SF268 cells move with a directed, multicellular movement, while ΔNLS-cells, moving almost as fast, perform a palindromic motion and fail to repair the “scratch wound” (videos in [39]). Overall, this is the first time that neurofibromin is linked to astrocytic cell migration, and, at least the loss of NLS neurofibromins, to defective centrosome positioning and functional cell polarity [39].

Both types of neurofibromins retained colocalization with β-tubulin on the mitotic spindle, albeit colocalization levels with NLS-neurofibromin are, as also for cytosolic MTs, significantly raised (Figure 4, images, plots, and colocalization means; [39]). Considering together that NLS neurofibromins do not affect MT densities, whereas ΔNLS-neurofibromins inversely regulate MT densities both in the cytoplasm and the spindle [39], these data document differential MAP properties for ΔNLS-neurofibromins as compared to NLS-neurofibromins.

Figure 4.

NLS- and ΔNLS-neurofibromins have different affinities for mitotic MTs. Naïve, ΔNLS-, and NLS- SF268 cells are stained for neurofibromin and β-tubulin. The first two columns contain single focal 0.34 μm planes of a confocal stack and the third column their mergings. Scatter plots show signal intensity in each plane (Volocity®) and numbers are the colocalization means±SE; arrow indicates the statistically significant difference in NLS- versus ΔNLS- or parental cells.

Examinations of colocalization with γ-tubulin on the duplicated centrosome, show a 25% decrease for NLS-neurofibromins, while ΔNLS neurofibromins show no statistical differences on this aspect. Yet, centrosomes in ΔNLS cells have larger volumes (1.8x), indicating that NLS neurofibromins may help form a more efficient mitotic centrosome, in terms of future spindle assembly [38, 39]. In human cells, the centrosome is the major MTOC for spindle MT assembly and duplicated centrosomes serve as poles to orient the spindle. More specifically, γ-tubulin and its several associated proteins form a large ring complex (γ-TuRC) that serves to nucleate highly dynamic MTs from the spindle, K-fibers from kinetochores, and interpolar bundles that elongate the spindle [43, 58, 59]. Because γ-TuRC is dispensable for this purpose on occasion [60], the role of MAP-dependent regulation in the nucleation to centrosomes, whether increasing [61] or inhibiting nucleation [62], has been highlighted.

The differential properties of the NLS and ΔNLS isoforms as MAPs on centrosome size are further highlighted by the effects on MT nucleation prior to nuclear envelope breakdown, when the relatively sparse microtubule formations of interphase cells transfigure into a bipolar spindle. Indeed, our studies have documented that the following parameters are greatly affected with depletion of NLS neurofibromins [39]:

3.1 Astral MT formation and spindle positioning

A most striking difference is the abnormal astral MT growth patterns with loss of NLS neurofibromins, as the average length of astral MTs in NLS-cells grows to 5.9 ± 0.33 μm over 4.0 ± 0.19 μm in parentals (p < 0.0001), while is robustly diminished in ΔNLS-cells (Figures 4 and 5, asterisks; [39]). Proper astral formation is required for spindle position and aberrations of astral MTs correlate well with spindle misorientation [41], and we too find that loss of astral MTs with loss of NLS-neurofibromin leads to statistically differential positioning of the spindle by several degrees [39]. A number of diverse families of proteins impact the timely nucleation and maintenance of astral MTs, yet few may cause loss of astral MTs. Notably, functional ablation (phosphoablating mutants) of End-binding protein 2 (EB2), an MT plus-end MAP that binds to MT lattices in a phosphorylation-dependent manner during mitosis, leads to a marked delay in anaphase onset and abnormalities in chromosome congression [63]. Multifunction proteins ALIX and RACK1 acting through regulation of other MAPs [64] or motor proteins [65], are also essential for proper astral MT elongation, spindle orientation and chromosome segregation. Whether neurofibromins regulate in addition the actions of other MAPs remain to be investigated; it is clear, however, that NLS neurofibromins are essential for astral MTs formation.

Figure 5.

Spindle configuration and chromosome congression are regulated by neurofibromin isoforms. A, Immunofluorescence confocal images of cells at metaphase stained for β-tubulin and chromatin of SF268, NLS, and ΔNLS-SF268 cells. B, 3D reconstructions (IMARIS) of the same images for better viewing show that parental and NLS-cells have a rich array of astral MTs (asterisks) and properly congressed chromosomes (white arrows). In contrast, ΔNLS cells lack astral MTs (asterisks) and display abnormal chromosome congression (yellow arrows).

3.2 Spindle length

Both parental and NLS-cells with fully developed metaphasic spindles have normally congressed chromosomes at the spindle equator (Figure 5, white arrows), albeit the latter have significantly shorter spindles (pole to pole distance x_ = 8.91 ± 0.22 μm versus 10.8 ± 0.2 μm; p < 0.0001 [39]). In stark contrast, cells expressing only ΔNLS-neurofibromins have very poorly aligned chromosomes at the equator (Figure 5, yellow arrows), although their metaphasic spindle length is significantly longer (x_ = 11.5 ± 0.15 μm; p < 0.01;). In over 50% of ΔNLS cells, the majority of chromosomes from a wide diffused ring and altogether lack the typical tight alignment seen at metaphase (Figure 5, yellow arrows).

3.3 Spindle geometry

When confocal z-planes of β-tubulin and Hoechst fluorescence signals are reconstructed in three-dimensions, the dramatically different spindle geometries, formed in the absence of NLS-neurofibromins, become readily apparent (Figure 5B). The anastral spindles of ΔNLS cells feature large hollows by the equator and chromosomes forming queues on some prominent thick K-fibers, while over half of the cells have unaligned chromosomes, or a 4-fold increase compared to control and NLS-cells [39]. In interpreting this geometry, we assume that thicker MT formations may result from upregulation of the augmin-mediated, local amplification of MTs, as augmin targets γ-TuRCs to nucleate preexisting MTs [66]; in parallel, bridging (Figure 2A) MTs [40] delay to develop, hence the spindle equator is almost devoid of tubulin signals. As these metaphasic patterns [39] strongly resemble those typically seen at prometaphase when unstable interactions of MTs dominate [56], the important role of neurofibromins as MAPs for mitosis is further highlighted.

3.4 Duration of mitotic phases

Abnormal positioning of the spindle often associates with altered times spent at mitotic stages. Quantification of the mitotic stage distribution for each cell type by flow cytometry validates this prediction, since loss of NLS-neurofibromin elicits an almost 50% increase in time spent at metaphase [39]. In contrast, NLS cells have significantly lower percentages in prophase and metaphase over parentals, which, combined with higher percentages in cytokinesis, reflected an overall acceleration through metaphase. Considered together, these results document for the first time that neurofibromin actively participates in the progression of mitosis. Moreover, these data further support the notion that NLS- and ΔNLS-neurofibromins may exert opposing effects during aster formation and spindle assembly, as, in parental cells, these two parameters appear to be the arithmetic sum of the results obtained with each isoform type [39].

3.5 Chromosome segregation fidelity

In parental (Figures 2 and 5, white arrows) and NLS-neurofibromin expressing cells (Figure 5, white arrows), chromosomes move in a coordinated manner towards the opposed poles and chromosome compaction is readily seen. In cells expressing only ΔNLS-neurofibromins, these parameters are again inversely regulated, namely, despite the prolonged time spent at metaphase, a significant >40% increase in cells with chromosome segregation errors mainly lagging chromosomes is documented (Figure 5, yellow arrows); similar delays in chromosome compaction in ΔNLS-neurofibromin cells are recorded in telophase [39]. The described effects on spindle assembly and chromosome segregation perturbations are readily traced in the high frequency of micronuclei, and a 5-fold increase in the numbers of cells with micronuclei within 10 mitotic cycles [39]. Micronuclei may facilitate rapid karyotype evolution, as their few chromosomes, unprotected from DNA damage, often undergo chromothripsis and chromoanasynthesis and then get incorporated into the genome of the host cell within just 1–2 mitoses [67].

Summarizing, these data establish for the first time that NLS- and ΔNLS-neurofibromins actively participate in the formation of mitotic asters and spindles, and efficient, error-free chromosome congression, possibly by exerting opposing effects. The question then rises about the possible mechanisms that would explain their different interactions with tubulins and microtubules. Drawing from immunoprecipitations studies with various antibodies, whereby different amounts of endogenously expressed neurofibromin is recovered from ΔNLS- or NLS-cell lysates [39], we have to reasonably presume that inclusion, or not, of the 41 amino acids of exon 51 may alter the conformation of the molecule. Numerous examples exist when one to few residues change the functional properties of a protein by imposing changes on protein conformation and post-translational modifications. Thus, an expected differential conformation of the ΔNLS or NLS neurofibromins would impact both its known intramolecular and intermolecular interactions.

In support of this argument the affinity of NLS- is higher for β- and lower for γ-tubulin when compared to ΔNLS-neurofibromins. Moreover, revisiting the question of MAP domains in the primary sequence of neurofibromin, we have identified, at higher percentages of similarity than the previously suggested [49], three other small Tau-like motifs [39], one of which corresponds to codons apposed to 50–51 or 50–52 exon junctions and could be affected by the inclusion or skipping of exon 51.

Our results show that the direction of the ΔNLS or NLS knockdown effects is most often opposite and suggest that the two functionally interact when both present. Whether this occurs through the formation of a dimer, if neurofibromins form dimers [28, 68] in eukaryotic cells at normal neurofibromin concentrations, is an intriguing question. Indeed, how the NLS and ΔNLS conformations may affect dimer formation is an interesting experimental goal.

Given the higher abundance of NLS transcripts irrespectively of GRD type that we observe in neurons and astrocytes [19], it is not possible to have only NLS-ΔNLS heterodimers. It is, however, possible to have homodimers only, or dimerization to be driven by properties that GRDI or GRDII attain. If any of the latter are entertained in eukaryotic cells, then an additional level of regulation is to be expected. At any rate, loss of the amino acids and the NLS encoded by exon 51 suffices to produce a different MAP. By the same token, the expression of two closely related isoforms yet with differential effects on MT structures further suggests that an extra layer of regulation on MT dynamics is thereby served by neurofibromins.

Advertisement

4. Conclusions: NLS-containing, tumor suppressor MAPS

A major theme that emerges from our studies and studies by others is that several MAPs have been described as tumor suppressors and correspondingly several proteins, identified as such, are found to function as mitotic MAPs. Another typical characteristic of such tumor suppressors is the presence of an NLS in their amino acid sequence, which regulates both their timely nuclear import in preparation of mitosis and their release during spindle assembly. All currently known tumor suppressor proteins with such attributes are listed in Table 1.

ProteinGeneFunctions served
APCAPCMT stabilization, astral MT formation, compaction of mitotic chromatin, chromosome segregation [69, 70, 71]
ATIP3MTUS1MT stability, cell polarity and migration, centrosome number, metaphase spindle length [72, 73, 74]
BRCA1BRCA1MT nucleation, spindle assembly, centromeric cohesion [75, 76, 77]
CYLDCYLDMT polymerization and stability, especially of astral MTs; spindle positioning [78, 79]
DAB2IPDAB2IPMT stability [80]
DLC-2/STARD13STARD13MT stability, spindle positioning, chromosome segregation [81]
DLG1/ SAP97SAP97MTs polarization, centrosome positioning [57, 82]
FEZ1/LZTS1FEZ1/LZTS1MT assembly, chromosome segregation [83, 84]
FHITFHITMT assembly, spindle disassembly [85, 86]
NAV3NAV3MT stability [87]
NeurofibrominNF1MT polymerization, cell migration, astral MT formation, spindle assembly and positioning, chromosome segregation [19, 24, 29, 37, 38, 39]
NF2/MerlinNF2MT polymerization; actin cytoskeleton organization, signalling scaffolding at the membrane [88, 89, 90]
p53TP53Clustering of centrosomes, pole formation [91, 92, 93]
PTENPTENCentrosome and spindle pole motility, spindle assembly, chromosome segregation [45, 94]
RASSF1ARASSF1AMT stability, spindle organization, chromosome segregation [95, 96, 97]
RB1RB1MT dynamics, centrosome number and condensation, chromosome segregation [98, 99, 100]
VHLVHLMT stability, spindle positioning [101, 102]
WT1WT1Chromosomal segregation, mitotic checkpoint [103]

Table 1.

Tumor Suppressors with a functional NLS and established roles as MAPs.

Perturbations of spindle assembly and chromosome segregation, when tumor suppressors that act as mitotic MAPs are lost or mutated, is a first step to aneuploidy. Given the usually compromised ability for DNA repair and the increased replication stress in these genetic backgrounds, the resulting aneuploidy may additionally feed chromosomal instability (CIN) and thus rapid evolution of karyotypes with clonal expansion advantages and tumorigenesis [104, 105]. Hence, the study of the regulation of NLS-containing tumor suppressors must receive high attention in the collective effort of understanding their mechanism of action and for developing better prognostic and possibly therapeutic approaches.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

Advertisement

Nomenclature

References

  1. 1. Uusitalo E, Leppavirta J, Koffert A, Suominen S, Vahtera J, Vahlberg T, et al. Incidence and mortality of neurofibromatosis: a total population study in Finland. The Journal of investigative dermatology. 2015;135(3):904-6
  2. 2. Evans DG, O'Hara C, Wilding A, Ingham SL, Howard E, Dawson J, et al. Mortality in neurofibromatosis 1: in North West England: an assessment of actuarial survival in a region of the UK since 1989. European journal of human genetics : EJHG. 2011;19(11):1187-91
  3. 3. McKeever K, Shepherd CW, Crawford H, Morrison PJ. An epidemiological, clinical and genetic survey of neurofibromatosis type 1 in children under sixteen years of age. The Ulster medical journal. 2008;77(3):160-3
  4. 4. Stenson PD, Mort M, Ball EV, Chapman M, Evans K, Azevedo L, et al. The Human Gene Mutation Database (HGMD((R))): optimizing its use in a clinical diagnostic or research setting. Human genetics. 2020;139(10):1197-207
  5. 5. Koczkowska M, Chen Y, Callens T, Gomes A, Sharp A, Johnson S, et al. Genotype-Phenotype Correlation in NF1: Evidence for a More Severe Phenotype Associated with Missense Mutations Affecting NF1 Codons 844-848. American journal of human genetics. 2018;102(1):69-87
  6. 6. Upadhyaya M, Huson SM, Davies M, Thomas N, Chuzhanova N, Giovannini S, et al. An absence of cutaneous neurofibromas associated with a 3-bp inframe deletion in exon 17 of the NF1 gene (c.2970-2972 delAAT): evidence of a clinically significant NF1 genotype-phenotype correlation. American journal of human genetics. 2007;80(1):140-51
  7. 7. Xu M, Xiong H, Han Y, Li C, Mai S, Huang Z, et al. Identification of Mutation Regions on NF1 Responsible for High- and Low-Risk Development of Optic Pathway Glioma in Neurofibromatosis Type I. Frontiers in genetics. 2018;9:270
  8. 8. Holzel M, Huang S, Koster J, Ora I, Lakeman A, Caron H, et al. NF1 is a tumor suppressor in neuroblastoma that determines retinoic acid response and disease outcome. Cell. 2010;142(2):218-29
  9. 9. Opocher G, Conton P, Schiavi F, Macino B, Mantero F. Pheochromocytoma in von Hippel-Lindau disease and neurofibromatosis type 1. Familial cancer. 2005;4(1):13-6
  10. 10. D'Angelo F, Ceccarelli M, Tala, Garofano L, Zhang J, Frattini V, et al. The molecular landscape of glioma in patients with Neurofibromatosis 1. Nature medicine. 2019;25(1):176-87
  11. 11. Cancer Genome Atlas Research N. Comprehensive genomic characterization defines human glioblastoma genes and core pathways. Nature. 2008;455(7216):1061-8
  12. 12. Gururangan S, Fangusaro J, Poussaint TY, McLendon RE, Onar-Thomas A, Wu S, et al. Efficacy of bevacizumab plus irinotecan in children with recurrent low-grade gliomas--a Pediatric Brain Tumor Consortium study. Neuro-oncology. 2014;16(2):310-7
  13. 13. Weil S, Osswald M, Solecki G, Grosch J, Jung E, Lemke D, et al. Tumor microtubes convey resistance to surgical lesions and chemotherapy in gliomas. Neuro-oncology. 2017;19(10):1316-26
  14. 14. Zuckermann M, Hovestadt V, Knobbe-Thomsen CB, Zapatka M, Northcott PA, Schramm K, et al. Somatic CRISPR/Cas9-mediated tumour suppressor disruption enables versatile brain tumour modelling. Nature communications. 2015;6:7391
  15. 15. Danglot G, Regnier V, Fauvet D, Vassal G, Kujas M, Bernheim A. Neurofibromatosis 1 (NF1) mRNAs expressed in the central nervous system are differentially spliced in the 5' part of the gene. Human molecular genetics. 1995;4(5):915-20
  16. 16. Suzuki Y, Suzuki H, Kayama T, Yoshimoto T, Shibahara S. Brain tumors predominantly express the neurofibromatosis type 1 gene transcripts containing the 63 base insert in the region coding for GTPase activating protein-related domain. Biochemical and biophysical research communications. 1991;181(3):955-61
  17. 17. Vandenbroucke I, Vandesompele J, De Paepe A, Messiaen L. Quantification of NF1 transcripts reveals novel highly expressed splice variants. FEBS letters. 2002;522(1-3):71-6
  18. 18. Gutmann DH, Zhang Y, Hirbe A. Developmental regulation of a neuron-specific neurofibromatosis 1 isoform. Annals of neurology. 1999;46(5):777-82
  19. 19. Peta C, Tsirimonaki E, Fedonidis C, Koliou X, Mangoura D. Depletion of the developmentally-regulated, NLS-containing, major neurofibromin isoform imposes mitotic aberrations in mouse astrocytes. 2021
  20. 20. Karouzaki S, Theofilopoulos S, Mangoura D. Differential properties of neurofibromin GRD I and II in developing neurons. Journal of neurochemistry. 2011;118, Sup.1, p192
  21. 21. Barron VA, Lou H. Alternative splicing of the neurofibromatosis type I pre-mRNA. Bioscience reports. 2012;32(2):131-8
  22. 22. Andersen LB, Ballester R, Marchuk DA, Chang E, Gutmann DH, Saulino AM, et al. A conserved alternative splice in the von Recklinghausen neurofibromatosis (NF1) gene produces two neurofibromin isoforms, both of which have GTPase-activating protein activity. Molecular and cellular biology. 1993;13(1):487-95
  23. 23. Bollag G, McCormick F, Clark R. Characterization of full-length neurofibromin: tubulin inhibits Ras GAP activity. The EMBO journal. 1993;12(5):1923-7
  24. 24. Leondaritis G, Mangoura D, editors. Regulation of neurofibromin’s GAP-related domain RasGAP activity by a Sec14-homology domain-dependent allosteric switch. Trans 61th EEBMB; 2010
  25. 25. Luo G, Kim J, Song K. The C-terminal domains of human neurofibromin and its budding yeast homologs Ira1 and Ira2 regulate the metaphase to anaphase transition. Cell cycle. 2014;13(17):2780-9
  26. 26. Ozawa T, Araki N, Yunoue S, Tokuo H, Feng L, Patrakitkomjorn S, et al. The neurofibromatosis type 1 gene product neurofibromin enhances cell motility by regulating actin filament dynamics via the Rho-ROCK-LIMK2-cofilin pathway. The Journal of biological chemistry. 2005;280(47):39524-33
  27. 27. Starinsky-Elbaz S, Faigenbloom L, Friedman E, Stein R, Kloog Y. The pre-GAP-related domain of neurofibromin regulates cell migration through the LIM kinase/cofilin pathway. Molecular and cellular neurosciences. 2009;42(4):278-87
  28. 28. Sherekar M, Han SW, Ghirlando R, Messing S, Drew M, Rabara D, et al. Biochemical and structural analyses reveal that the tumor suppressor neurofibromin (NF1) forms a high-affinity dimer. The Journal of biological chemistry. 2020;295(4):1105-19
  29. 29. Leondaritis G, Petrikkos L, Mangoura D. Regulation of the Ras-GTPase activating protein neurofibromin by C-tail phosphorylation: implications for protein kinase C/Ras/extracellular signal-regulated kinase 1/2 pathway signaling and neuronal differentiation. Journal of neurochemistry. 2009;109(2):573-83
  30. 30. Mangoura D, Sun Y, Li C, Singh D, Gutmann DH, Flores A, et al. Phosphorylation of neurofibromin by PKC is a possible molecular switch in EGF receptor signaling in neural cells. Oncogene. 2006;25(5):735-45
  31. 31. Scheffzek K, Shivalingaiah G. Ras-Specific GTPase-Activating Proteins-Structures, Mechanisms, and Interactions. Cold Spring Harbor perspectives in medicine. 2019;9(3)
  32. 32. Fadhlullah SFB, Halim NBA, Yeo JYT, Ho RLY, Um P, Ang BT, et al. Pathogenic mutations in neurofibromin identifies a leucine-rich domain regulating glioma cell invasiveness. Oncogene. 2019;38(27):5367-80
  33. 33. D'Angelo I, Welti S, Bonneau F, Scheffzek K. A novel bipartite phospholipid-binding module in the neurofibromatosis type 1 protein. EMBO reports. 2006;7(2):174-9
  34. 34. Welti S, Fraterman S, D'Angelo I, Wilm M, Scheffzek K. The sec14 homology module of neurofibromin binds cellular glycerophospholipids: mass spectrometry and structure of a lipid complex. Journal of molecular biology. 2007;366(2):551-62
  35. 35. Kalab P, Pralle A, Isacoff EY, Heald R, Weis K. Analysis of a RanGTP-regulated gradient in mitotic somatic cells. Nature. 2006;440(7084):697-701
  36. 36. Oh D, Yu CH, Needleman DJ. Spatial organization of the Ran pathway by microtubules in mitosis. Proceedings of the National Academy of Sciences of the United States of America. 2016;113(31):8729-34
  37. 37. Li C, Cheng Y, Gutmann DA, Mangoura D. Differential localization of the neurofibromatosis 1 (NF1) gene product, neurofibromin, with the F-actin or microtubule cytoskeleton during differentiation of telencephalic neurons. Brain research Developmental brain research. 2001;130(2):231-48
  38. 38. Koliou X, Fedonidis C, Kalpachidou T, Mangoura D. Nuclear import mechanism of neurofibromin for localization on the spindle and function in chromosome congression. Journal of neurochemistry. 2016;136(1):78-91
  39. 39. Peta C, Tsirimonaki E, Samouil D, Georgiadou K, Mangoura D. Nuclear Isoforms of Neurofibromin Are Required for Proper Spindle Organization and Chromosome Segregation. Cells. 2020;9(11)
  40. 40. Kajtez J, Solomatina A, Novak M, Polak B, Vukusic K, Rudiger J, et al. Overlap microtubules link sister k-fibres and balance the forces on bi-oriented kinetochores. Nature communications. 2016;7:10298
  41. 41. di Pietro F, Echard A, Morin X. Regulation of mitotic spindle orientation: an integrated view. EMBO reports. 2016;17(8):1106-30
  42. 42. Maiato H, Gomes AM, Sousa F, Barisic M. Mechanisms of Chromosome Congression during Mitosis. Biology. 2017;6(1)
  43. 43. Meraldi P. Centrosomes in spindle organization and chromosome segregation: a mechanistic view. Chromosome research : an international journal on the molecular, supramolecular and evolutionary aspects of chromosome biology. 2016;24(1):19-34
  44. 44. Cuijpers SAG, Vertegaal ACO. Guiding Mitotic Progression by Crosstalk between Post-translational Modifications. Trends in biochemical sciences. 2018;43(4):251-68
  45. 45. He J, Zhang Z, Ouyang M, Yang F, Hao H, Lamb KL, et al. PTEN regulates EG5 to control spindle architecture and chromosome congression during mitosis. Nature communications. 2016;7:12355
  46. 46. Reed SI. Ratchets and clocks: the cell cycle, ubiquitylation and protein turnover. Nature reviews Molecular cell biology. 2003;4(11):855-64
  47. 47. Bhat KM, Setaluri V. Microtubule-associated proteins as targets in cancer chemotherapy. Clinical cancer research : an official journal of the American Association for Cancer Research. 2007;13(10):2849-54
  48. 48. Schroeder C, Grell J, Hube-Magg C, Kluth M, Lang D, Simon R, et al. Aberrant expression of the microtubule-associated protein tau is an independent prognostic feature in prostate cancer. BMC cancer. 2019;19(1):193
  49. 49. Gregory PE, Gutmann DH, Mitchell A, Park S, Boguski M, Jacks T, et al. Neurofibromatosis type 1 gene product (neurofibromin) associates with microtubules. Somatic cell and molecular genetics. 1993;19(3):265-74
  50. 50. Patrakitkomjorn S, Kobayashi D, Morikawa T, Wilson MM, Tsubota N, Irie A, et al. Neurofibromatosis type 1 (NF1) tumor suppressor, neurofibromin, regulates the neuronal differentiation of PC12 cells via its associating protein, CRMP-2. The Journal of biological chemistry. 2008;283(14):9399-413
  51. 51. Daub H, Olsen JV, Bairlein M, Gnad F, Oppermann FS, Korner R, et al. Kinase-selective enrichment enables quantitative phosphoproteomics of the kinome across the cell cycle. Molecular cell. 2008;31(3):438-48
  52. 52. Godin F, Villette S, Vallee B, Doudeau M, Morisset-Lopez S, Ardourel M, et al. A fraction of neurofibromin interacts with PML bodies in the nucleus of the CCF astrocytoma cell line. Biochemical and biophysical research communications. 2012;418(4):689-94
  53. 53. Nousiainen M, Sillje HH, Sauer G, Nigg EA, Korner R. Phosphoproteome analysis of the human mitotic spindle. Proceedings of the National Academy of Sciences of the United States of America. 2006;103(14):5391-6
  54. 54. Zheng ZY, Anurag M, Lei JT, Cao J, Singh P, Peng J, et al. Neurofibromin Is an Estrogen Receptor-alpha Transcriptional Co-repressor in Breast Cancer. Cancer cell. 2020;37(3):387-402 e7
  55. 55. Xu H, Gutmann DH. Mutations in the GAP-related domain impair the ability of neurofibromin to associate with microtubules. Brain research. 1997;759(1):149-52
  56. 56. Magidson V, O'Connell CB, Loncarek J, Paul R, Mogilner A, Khodjakov A. The spatial arrangement of chromosomes during prometaphase facilitates spindle assembly. Cell. 2011;146(4):555-67
  57. 57. Etienne-Manneville S, Hall A. Cdc42 regulates GSK-3beta and adenomatous polyposis coli to control cell polarity. Nature. 2003;421(6924):753-6
  58. 58. Maiato H, Rieder CL, Khodjakov A. Kinetochore-driven formation of kinetochore fibers contributes to spindle assembly during animal mitosis. The Journal of cell biology. 2004;167(5):831-40
  59. 59. Meunier S, Vernos I. K-fibre minus ends are stabilized by a RanGTP-dependent mechanism essential for functional spindle assembly. Nature cell biology. 2011;13(12):1406-14
  60. 60. Consolati T, Locke J, Roostalu J, Chen ZA, Gannon J, Asthana J, et al. Microtubule Nucleation Properties of Single Human gammaTuRCs Explained by Their Cryo-EM Structure. Developmental cell. 2020;53(5):603-17 e8
  61. 61. Gard DL, Kirschner MW. A microtubule-associated protein from Xenopus eggs that specifically promotes assembly at the plus-end. The Journal of cell biology. 1987;105(5):2203-15
  62. 62. Ringhoff DN, Cassimeris L. Stathmin regulates centrosomal nucleation of microtubules and tubulin dimer/polymer partitioning. Molecular biology of the cell. 2009;20(15):3451-8
  63. 63. Iimori M, Watanabe S, Kiyonari S, Matsuoka K, Sakasai R, Saeki H, et al. Phosphorylation of EB2 by Aurora B and CDK1 ensures mitotic progression and genome stability. Nature communications. 2016;7:11117
  64. 64. Malerod L, Le Borgne R, Lie-Jensen A, Eikenes AH, Brech A, Liestol K, et al. Centrosomal ALIX regulates mitotic spindle orientation by modulating astral microtubule dynamics. The EMBO journal. 2018;37(13)
  65. 65. Ai E, Poole DS, Skop AR. RACK-1 directs dynactin-dependent RAB-11 endosomal recycling during mitosis in Caenorhabditis elegans. Molecular biology of the cell. 2009;20(6):1629-38
  66. 66. Song JG, King MR, Zhang R, Kadzik RS, Thawani A, Petry S. Mechanism of how augmin directly targets the gamma-tubulin ring complex to microtubules. The Journal of cell biology. 2018;217(7):2417-28
  67. 67. Liu S, Kwon M, Mannino M, Yang N, Renda F, Khodjakov A, et al. Nuclear envelope assembly defects link mitotic errors to chromothripsis. Nature. 2018;561(7724):551-5
  68. 68. Lupton CJ, Bayly-Jones C, D’Andrea L, Huang C, Schittenhelm RB, Venugopal H, et al. The cryo-EM structure of the neurofibromin dimer reveals the molecular basis for von Recklinghausen disease. bioRxiv. 2021
  69. 69. Dikovskaya D, Khoudoli G, Newton IP, Chadha GS, Klotz D, Visvanathan A, et al. The adenomatous polyposis coli protein contributes to normal compaction of mitotic chromatin. PloS one. 2012;7(6):e38102
  70. 70. Green RA, Wollman R, Kaplan KB. APC and EB1 function together in mitosis to regulate spindle dynamics and chromosome alignment. Molecular biology of the cell. 2005;16(10):4609-22
  71. 71. Slep KC, Rogers SL, Elliott SL, Ohkura H, Kolodziej PA, Vale RD. Structural determinants for EB1-mediated recruitment of APC and spectraplakins to the microtubule plus end. The Journal of cell biology. 2005;168(4):587-98
  72. 72. Molina A, Velot L, Ghouinem L, Abdelkarim M, Bouchet BP, Luissint AC, et al. ATIP3, a novel prognostic marker of breast cancer patient survival, limits cancer cell migration and slows metastatic progression by regulating microtubule dynamics. Cancer research. 2013;73(9):2905-15
  73. 73. Nehlig A, Seiler C, Steblyanko Y, Dingli F, Arras G, Loew D, et al. Reciprocal regulation of Aurora kinase A and ATIP3 in the control of metaphase spindle length. Cellular and molecular life sciences : CMLS. 2021;78(4):1765-79
  74. 74. Rodrigues-Ferreira S, Nehlig A, Moindjie H, Monchecourt C, Seiler C, Marangoni E, et al. Improving breast cancer sensitivity to paclitaxel by increasing aneuploidy. Proceedings of the National Academy of Sciences of the United States of America. 2019;116(47):23691-7
  75. 75. Di Paolo A, Racca C, Calsou P, Larminat F. Loss of BRCA1 impairs centromeric cohesion and triggers chromosomal instability. FASEB journal : official publication of the Federation of American Societies for Experimental Biology. 2014;28(12):5250-61
  76. 76. Joukov V, Groen AC, Prokhorova T, Gerson R, White E, Rodriguez A, et al. The BRCA1/BARD1 heterodimer modulates ran-dependent mitotic spindle assembly. Cell. 2006;127(3):539-52
  77. 77. Sung M, Giannakakou P. BRCA1 regulates microtubule dynamics and taxane-induced apoptotic cell signaling. Oncogene. 2014;33(11):1418-28
  78. 78. Gao J, Huo L, Sun X, Liu M, Li D, Dong JT, et al. The tumor suppressor CYLD regulates microtubule dynamics and plays a role in cell migration. The Journal of biological chemistry. 2008;283(14):8802-9
  79. 79. Yang Y, Liu M, Li D, Ran J, Gao J, Suo S, et al. CYLD regulates spindle orientation by stabilizing astral microtubules and promoting dishevelled-NuMA-dynein/dynactin complex formation. Proceedings of the National Academy of Sciences of the United States of America. 2014;111(6):2158-63
  80. 80. Yu L, Shang ZF, Abdisalaam S, Lee KJ, Gupta A, Hsieh JT, et al. Tumor suppressor protein DAB2IP participates in chromosomal stability maintenance through activating spindle assembly checkpoint and stabilizing kinetochore-microtubule attachments. Nucleic acids research. 2016;44(18):8842-54
  81. 81. Vitiello E, Ferreira JG, Maiato H, Balda MS, Matter K. The tumour suppressor DLC2 ensures mitotic fidelity by coordinating spindle positioning and cell-cell adhesion. Nature communications. 2014;5:5826
  82. 82. Etienne-Manneville S, Manneville JB, Nicholls S, Ferenczi MA, Hall A. Cdc42 and Par6-PKCzeta regulate the spatially localized association of Dlg1 and APC to control cell polarization. The Journal of cell biology. 2005;170(6):895-901
  83. 83. Ishii H, Vecchione A, Murakumo Y, Baldassarre G, Numata S, Trapasso F, et al. FEZ1/LZTS1 gene at 8p22 suppresses cancer cell growth and regulates mitosis. Proceedings of the National Academy of Sciences of the United States of America. 2001;98(18):10374-9
  84. 84. Vecchione A, Baldassarre G, Ishii H, Nicoloso MS, Belletti B, Petrocca F, et al. Fez1/Lzts1 absence impairs Cdk1/Cdc25C interaction during mitosis and predisposes mice to cancer development. Cancer cell. 2007;11(3):275-89
  85. 85. Chaudhuri AR, Khan IA, Prasad V, Robinson AK, Luduena RF, Barnes LD. The tumor suppressor protein Fhit. A novel interaction with tubulin. The Journal of biological chemistry. 1999;274(34):24378-82
  86. 86. Vizeacoumar FJ, van Dyk N, F SV, Cheung V, Li J, Sydorskyy Y, et al. Integrating high-throughput genetic interaction mapping and high-content screening to explore yeast spindle morphogenesis. The Journal of cell biology. 2010;188(1):69-81
  87. 87. Cohen-Dvashi H, Ben-Chetrit N, Russell R, Carvalho S, Lauriola M, Nisani S, et al. Navigator-3, a modulator of cell migration, may act as a suppressor of breast cancer progression. EMBO molecular medicine. 2015;7(3):299-314
  88. 88. den Bakker MA, Tascilar M, Riegman PH, Hekman AC, Boersma W, Janssen PJ, et al. Neurofibromatosis type 2 protein co-localizes with elements of the cytoskeleton. The American journal of pathology. 1995;147(5):1339-49
  89. 89. Smole Z, Thoma CR, Applegate KT, Duda M, Gutbrodt KL, Danuser G, et al. Tumor suppressor NF2/Merlin is a microtubule stabilizer. Cancer research. 2014;74(1):353-62
  90. 90. Yin F, Yu J, Zheng Y, Chen Q , Zhang N, Pan D. Spatial organization of Hippo signaling at the plasma membrane mediated by the tumor suppressor Merlin/NF2. Cell. 2013;154(6):1342-55
  91. 91. Contadini C, Monteonofrio L, Virdia I, Prodosmo A, Valente D, Chessa L, et al. p53 mitotic centrosome localization preserves centrosome integrity and works as sensor for the mitotic surveillance pathway. Cell death & disease. 2019;10(11):850
  92. 92. Giannakakou P, Sackett DL, Ward Y, Webster KR, Blagosklonny MV, Fojo T. p53 is associated with cellular microtubules and is transported to the nucleus by dynein. Nature cell biology. 2000;2(10):709-17
  93. 93. Morris VB, Brammall J, Noble J, Reddel R. p53 localizes to the centrosomes and spindles of mitotic cells in the embryonic chick epiblast, human cell lines, and a human primary culture: An immunofluorescence study. Experimental cell research. 2000;256(1):122-30
  94. 94. van Ree JH, Nam HJ, Jeganathan KB, Kanakkanthara A, van Deursen JM. Pten regulates spindle pole movement through Dlg1-mediated recruitment of Eg5 to centrosomes. Nature cell biology. 2016;18(7):814-21
  95. 95. Dallol A, Hesson LB, Matallanas D, Cooper WN, O'Neill E, Maher ER, et al. RAN GTPase is a RASSF1A effector involved in controlling microtubule organization. Current biology : CB. 2009;19(14):1227-32
  96. 96. Jeon HJ, Oh JS. RASSF1A Regulates Spindle Organization by Modulating Tubulin Acetylation via SIRT2 and HDAC6 in Mouse Oocytes. Frontiers in cell and developmental biology. 2020;8:601972
  97. 97. Rong R, Jin W, Zhang J, Sheikh MS, Huang Y. Tumor suppressor RASSF1A is a microtubule-binding protein that stabilizes microtubules and induces G2/M arrest. Oncogene. 2004;23(50):8216-30
  98. 98. Coschi CH, Martens AL, Ritchie K, Francis SM, Chakrabarti S, Berube NG, et al. Mitotic chromosome condensation mediated by the retinoblastoma protein is tumor-suppressive. Genes & development. 2010;24(13):1351-63
  99. 99. Iovino F, Lentini L, Amato A, Di Leonardo A. RB acute loss induces centrosome amplification and aneuploidy in murine primary fibroblasts. Molecular cancer. 2006;5:38
  100. 100. Lyu J, Yang EJ, Zhang B, Wu C, Pardeshi L, Shi C, et al. Synthetic lethality of RB1 and aurora A is driven by stathmin-mediated disruption of microtubule dynamics. Nature communications. 2020;11(1):5105
  101. 101. Hell MP, Duda M, Weber TC, Moch H, Krek W. Tumor suppressor VHL functions in the control of mitotic fidelity. Cancer research. 2014;74(9):2422-31
  102. 102. Thoma CR, Toso A, Gutbrodt KL, Reggi SP, Frew IJ, Schraml P, et al. VHL loss causes spindle misorientation and chromosome instability. Nature cell biology. 2009;11(8):994-1001
  103. 103. Shandilya J, Roberts SG. A role of WT1 in cell division and genomic stability. Cell cycle. 2015;14(9):1358-64
  104. 104. Burrell RA, McClelland SE, Endesfelder D, Groth P, Weller MC, Shaikh N, et al. Replication stress links structural and numerical cancer chromosomal instability. Nature. 2013;494(7438):492-6
  105. 105. Gorgoulis VG, Vassiliou LV, Karakaidos P, Zacharatos P, Kotsinas A, Liloglou T, et al. Activation of the DNA damage checkpoint and genomic instability in human precancerous lesions. Nature. 2005;434(7035):907-13

Written By

Charoula Peta, Emmanouella Tsirimonaki, Constantinos Fedonidis, Xeni Koliou, Nikos Sakellaridis and Dimitra Mangoura

Submitted: 23 March 2021 Reviewed: 05 April 2021 Published: 29 July 2021