TPX2

Last updated
TPX2
Protein TPX2 PDB 1ol5.png
Available structures
PDB Ortholog search: PDBe RCSB
Identifiers
Aliases TPX2 , C20orf1, C20orf2, DIL-2, DIL2, FLS353, GD:C20orf1, HCA519, HCTP4, REPP86, p100, microtubule nucleation factor, TPX2 microtubule nucleation factor
External IDs OMIM: 605917 MGI: 1919369 HomoloGene: 8107 GeneCards: TPX2
Orthologs
SpeciesHumanMouse
Entrez
Ensembl
UniProt
RefSeq (mRNA)

NM_012112

NM_001141975
NM_001141976
NM_001141977
NM_001141978
NM_028109

Contents

RefSeq (protein)

NP_036244

NP_001135447
NP_001135448
NP_001135449
NP_001135450
NP_082385

Location (UCSC) Chr 20: 31.74 – 31.8 Mb Chr 2: 152.69 – 152.74 Mb
PubMed search [3] [4]
Wikidata
View/Edit Human View/Edit Mouse

Targeting protein for Xklp2 is a protein that in humans is encoded by the TPX2 gene. [5] [6] [7] It is one of the many spindle assembly factors that play a key role in inducing microtubule assembly and growth during M phase.

Key domains of TPX2

TPX2 has been reported to have two NLS-containing domains that mediate its localization to microtubules; one in the amino-terminal domain, and the other in the carboxy-terminal domain. [8] [9] In addition to an NLS, the carboxy-terminal domain of TPX2 consists of tandem repeats that cover over two-thirds of the protein and are computationally predicted to consist of predominantly alpha-helical content. [10] [11] This region can be further divided into five clusters of conserved residues separated by unstructured regions: α3-7. [11] α3-6 all contain a central α-helical region that is followed by a characteristic "FKARP" motif. [11] α7 is longer and exhibits a long α-helical stretch that is computationally predicted to form a coiled coil. [11] Lastly, the final 35 amino acids of the carboxy-terminus of TPX2 are responsible for interacting with tetrameric kinesin Eg5. [12] [13]

TPX2 contains one KEN box (K-E-N) motif at amino acid 87 and three D-box (R-X-X-L) motifs at amino acids 119, 341, and 708. [14] Both motif types have been suspected to be important in regulation and degradation of TPX2 by the APC/C (see "Regulation of TPX2 in the Cell Cycle"), as typically mutations in these motifs render substrates resistant to ubiquitination by the APC/C. [15] [16] However, in vitro ubiquitination assays have shown that only the first 83 amino acids of the N-terminal region of TPX2 along with the KEN box are pertinent for recognition by Cdh1, an activator of the APC/C. [14]

Role in microtubule assembly

TPX2 has been shown in several biochemical assays to behave as a microtubule-associated protein (MAP) and co-localize with spindle microtubules during M-phase. [5] [17] [18] [9] [19] It plays a role in microtubule nucleation and is regulated by importin proteins.

TPX2 serves as a complement, depleting importin α affinity in order to allow RanGTP-induced microtubule nucleation. This has been demonstrated both in vitro in Xenopus laevis egg extracts, and with the human homologue in vivo in HeLa cells. [20] [18] TPX2 is also important in activating and recruiting Aurora A kinase, a kinase responsible for phosphorylating TPX2 and essential for cell proliferation. [9] In the presence of nuclear import factor importin α, TPX2 is bound and prevented from binding Aurora A kinase, though it is still able to bind microtubules via its amino-terminal domain. [9] This leads to inhibition of M phase microtubule nucleation. In contrast, TPX2 is freed from inhibition by displacement of importin α via RanGTP, though RanGTP is not required for free TPX2 activity, as TPX2 has been shown to induce microtubule assembly in the absence of exogenous and depletion of endogenous RanGTP. [20] This suggests that TPX2 is downstream of RanGTP activity, but whether TPX2 is directly regulated by RanGTP still remains to be determined.

The mechanism by which TPX2 promotes microtubule nucleation has yet to be determined. One proposed mechanism has been based on TPX2's role in directly suppressing tubulin subunit off-rates at the microtubule tip during microtubule assembly and disassembly, verified by fluorescence microscopy. [21] This is made possible partially by TPX2's role in sequestering free tubulin subunits and nucleating small multi-subunit tubulin complexes, which inadvertently also slows the rate of growth by decreasing the effective free tubulin concentration. [21] TPX2's stabilization of the microtubule in its polymer form therefore contributes to microtubule nucleation. Computational simulations speculate that TPX2 suppresses tubulin subunit kinetics at the microtubule tip by randomly increasing the bond stability between adjacent tubulin subunits. [21]

In addition, TPX2 has been shown to be important in chromatin-dependent spindle assembly. Even with duplicated centrosomes, TPX2 has been demonstrated to be required for formation of a stable bipolar spindle with overlapping antiparallel microtubule arrays. [18] More specifically, TPX2 contributes to microtubule branching during spindle assembly by cooperating with augmin in order to amplify microtubule mass and preserve its polarity. [22] Branching nucleation by TPX2 is observed without RanGTP, though more fan-shaped microtubule structures are formed when both RanGTP and TPX2 are present. [22] The rate of branched formation is also enhanced in the presence of both components compared to Ran alone. [22]

The TPX2 region necessary for branching microtubule nucleation resides in its carboxy-terminal half (amino acids 319-716), [22] with TPX2 domains α5-7 as the minimal necessary requirement and domains α3-4 serving as contributors to nucleation efficiency by enabling earlier induction at faster rates. The amino-terminal half of TPX2 also increases the efficiency of the reaction. [11] TPX2 α5-7 is different from the remainder of the protein in that it contains conserved regions in its amino acid sequence that share sequence similarity with two known γ-TuRC nucleation activator motifs: SPM and γ-TuRC. [11] The SPM-like motif is found within the α5 domain, while the γTuNA-like motif is found to start in the α5 domain and stretch into the SPM-like motif. Without these two motifs, no microtubule nucleation was observed in vitro, though microtubule binding ability was maintained. [11] However, these two motifs are not the only essential ones in microtubule branching nucleation; the FKARP motifs of α5 and α6 are also essential for stimulating this process. [11] Furthermore, the α-helical region stretch of domain α7 and the C-terminal residues that interact with Eg5 are critical for microtubule branching nucleation as well. [11] While α5-7 domains are important in this process, none of them have intrinsic microtubule nucleation activity. [11]

In terms of binding to and bundling microtubules, at least any of the three domains α3-7 of TPX2 are necessary for significant binding and bundling in vitro. [11] Furthermore, it is likely that the domains cooperatively mediate microtubule binding and bundling, as successive addition or subtraction of a domain does not result in a linear change in microtubule binding and bundling capacity. [11]

Activation and reciprocation through Aurora A kinase

TPX2 recruits and activates Aurora A kinase by utilizing its short 43 amino acid long amino-terminal sequence to bind the catalytic domain of Aurora A, locking the kinase into its active conformation. [23] [24] More specifically, this interaction positions the activation segment of the kinase into a more favorable conformation for substrate binding and swings the crucial phosphothreonine residue, a target usually exposed and accessible for deactivation of Aurora A kinase by PP1, into a buried position, thereby locking Aurora A into an active conformation. [23] Notably, this recognition between TPX2 and Aurora A is analogous to that between the cAMP-dependent protein kinase (cAPK) catalytic core and its flanking region, suggesting a recurring theme in kinase regulation. [23] Activated Aurora A in turn phosphorylates TPX2, but it is still unclear how Aurora A’s phosphorylation of TPX2 affects its activities.

Role in cleavage arrest and interaction with Eg5

When four-fold TPX2 over the endogenous level was injected into a blastomere with a two-cell embryo, cleavage arrest was induced. [12] This arrest has been attributed to the amino acids 471-715 of the carboxy-terminus of the TPX2 protein, with the last 35 amino acids being absolutely necessary for inducing cleavage arrest. [12] During cytokinesis failure, cycles of DNA synthesis and mitosis continue. Notably, spindle poles fail to segregate, leading to a failure to establish a bipolar spindle, a spindle midzone, and a central spindle complex. [12] Because cleavage furrow ingression is primarily triggered by signals from the spindle midzone, [25] [26] these biological phenotypes could account for the failure of this event due to the inability to activate the spindle checkpoint. [12] Instead of a bipolar spindle, both spindle poles are in apposition, with an impairment of pushing forces generated by interpolar microtubules. [12]

The mechanistic cause behind cleavage arrest is attributed to TPX2’s ability to directly bind motor protein Eg5, which requires the last 35 amino acids of the TPX2 carboxy-terminus for its interaction. [12] When Eg5 was co-injected with TPX2 in vivo, cleavage furrow arrest was blocked and ingression was observed. This suggests that the carboxy-terminus of TPX2 regulates spindle pole movement via an Eg5-dependent mechanism. [12]

Binding with Xlp2

When bound to microtubules, TPX2 recruits a plus-end directed motor protein, Xlp2, a protein that is required in early mitosis and localizes to spindle poles, to microtubule minus ends of asters. [17] [27] [28] Like TPX2’s localization to microtubules, this recruitment is also RanGTP independent. [17] [29]

Regulation of TPX2 in the cell cycle

Monitoring TPX2 gene mRNA expression during cell cycle progression in synchronized HeLa cells revealed that TPX2 expression is high in G2/M phase, decreases dramatically upon G1 phase entry, increases upon entry into S phase, and peaks again at the next G2/M phase. [30] [14] This is correlated by results showing an increased stability of TPX2 in S-phase extracts compared to that of TPX2 in mitotic extracts, indicated by a significant increase in TPX2 half-life. [14] The drop in TPX2 is consistent with the drastic reorganization in structure and dynamics of the mitotic spindle. [31]

Overall, TPX2 has been shown through in vivo experiments to be regulated by the APC/CCdh1 pathway. [14] The instability and drop in TPX2 at mitotic exit is dependent on both the anaphase-promoting complex/cyclosome (APC/C) and an ubiquitin ligase integral in mitotic progression, along with APC/C's activator protein, Cdh1. [14] [32] This is a result of TPX2 being bound directly by Cdh1, and not Cdc20 or any other substrate of APC/CCdh1, and designated for degradation by APC/C. [14] Moreover, the Cdh1-TPX2 binding interaction produces the TPX2 stability seen during mitosis up until mitotic exit: The amino-terminal region of Cdh1 (amino acids 1-125) can act as a dominant negative mutant when expressed in mammalian cells, stabilizing APC/CCdh1 substrates such as TPX2 by competitive binding. [14]

Role in the nucleus

When the cell is in interphase, because of its ability to bind to importin α and β, TPX2 has been found localized in the nucleus. [5] [17] This has been proposed to be a physical mechanism by which proteins that operate in M phase are inactivated in interphase. TPX2 during M-phase accumulates at the poles of spindles in a “dynein-dynactin-dependent way.” [17] [9] The mechanism of this localization currently remains unclear, but it is not RanGTP dependent despite its downfield position from RanGTP activity, as TPX2 in Xenopus laevis egg extracts have been shown to accumulate at the center of microtubule asters (after the addition of centrosomes, taxol, or DMSO) and bind to pure microtubules in the presence of importins. [19]

Though nuclear import of TPX2 is thought to sequester TPX2 away from cytoplasmic tubulin in order to solely prevent premature spindle assembly, [33] [34] roles of nuclear TPX2 have recently been discovered. One of these roles is with the DNA damage response, where depletion of TPX2 in cells leads to a transient increase in γ-H2AX (the phosphorylated form of H2AX, the form that serves as a marker of DNA damage response amplification) levels in cells treated with ionizing radiation, [35] and overexpression of TPX2 leads to a decrease in the number of ionizing radiation-induced MDC1 foci and γ-H2AX levels. [35] This is supported by the discovery of TPX2 accumulation at DNA double strand breaks and association with the machinery of DNA damage response that controls the amplification of γ-H2AX. [35] However, the exact molecular mechanisms by which TPX2 impacts the ionizing radiation-dependent γ-H2AX levels still remains to be discovered.  Note that TPX2’s function in the DNA damage response is independent of its mitotic function, and is therefore independent of apoptosis.

When no ionizing radiation is present, TPX2 readily associates with the chromatin. [36] Interestingly, overexpression of TPX2 in these conditions produces abnormal DAPI staining patterns, where DAPI staining is more structured and compartmentalized than the typical uniformly-distributed DAPI staining in wild type cells. [36] Moreover, when TPX2 levels were depleted in unirradiated cells, no significant changes in γ-H2AX levels were found, [35] but the levels of H4K16ac, the acetylated form of H4K16 (a histone post-translationally modified during DNA damage response), decreased. [36] This decrease is unaffected by ionizing radiation, yet correlates with the decrease in γ-H2AX under such conditions. A result of this decrease is a defect in BP531 (p53 binding protein 1) recruitment to chromosomal breaks, [36] as recruitment is dependent on the acetylation status of H4K16. [37] As with TPX2 with regards to its impact on ionizing radiation-dependent γ-H2AX levels, the molecular mechanism by which TPX2 affects the acetylation status of H4K16 remains to be discovered.

Relevance in cancer

Because of its integral role in microtubule assembly and therefore mitosis, TPX2 is found to be overexpressed in different types of human cancers including hepatocellular carcinoma (HCC), [30] medullary thyroid cancer, [38] bladder carcinoma, [39] and estrogen receptor-positive metastatic breast cancer [40] and contributes to tumor growth and metastasis. [30] In HCC, TPX2 has been shown to be positively correlated with poor prognosis, metastasis, and recurrence. [41] [42] [43] Studies on TPX2 in HCC have also showed that TPX2 promotes tumoriogenesis and liver cancer cell growth by increasing tumor spheroid and diminishing cell growth inhibition, demonstrated by knocking out endogenous expression of TPX2 using TPX2 si-RNA. [30]

As a result, TPX2 has recently been a topic of interest for learning more about the relationship between mitotic errors and tumorigenesis, along with novel cancer therapies. So far, research on depleting TPX2 via TPX2 si-RNA in HCC cells in vitro has shown significant effects in diminishing cell motility and invasion (i.e. metastasis), along with diminishing proteins involved in the G1 to S phase transition. [30] Similar results have been shown with TPX2 depletion in esophageal cancer EC9706 cells, leading to reduced cancer cell growth and invasion ability, [44] and in cervical [45] and pancreatic cancer [46] with regards to reduced tumor growth using TPX2 si-RNA transfection.

In liver cancer cells, TPX2 depletion has been linked to increased genomic instability, resulting in multinucleation and DNA damage. [30] While many tumor cells in general accumulate mutations in genomic instability that enable them to have a growth advantage in tumor promotion and transformation, [47] high chromosomal instability can act as a tumor-suppressing mechanism by leading to cell death. [48] [49] Therefore, the significant aneuploidy and genomic instability at mitotic division via TPX2 depletion can serve as a potential therapeutic target for cancer patients by eliminating highly proliferating cells.

Related Research Articles

<span class="mw-page-title-main">Microtubule</span> Polymer of tubulin that forms part of the cytoskeleton

Microtubules are polymers of tubulin that form part of the cytoskeleton and provide structure and shape to eukaryotic cells. Microtubules can be as long as 50 micrometres, as wide as 23 to 27 nm and have an inner diameter between 11 and 15 nm. They are formed by the polymerization of a dimer of two globular proteins, alpha and beta tubulin into protofilaments that can then associate laterally to form a hollow tube, the microtubule. The most common form of a microtubule consists of 13 protofilaments in the tubular arrangement.

<span class="mw-page-title-main">Nuclear pore</span>

A nuclear pore is a part of a large complex of proteins, known as a nuclear pore complex that spans the nuclear envelope, which is the double membrane surrounding the eukaryotic cell nucleus. There are approximately 1,000 nuclear pore complexes (NPCs) in the nuclear envelope of a vertebrate cell, but this number varies depending on cell type and the stage in the life cycle. The human nuclear pore complex (hNPC) is a 110 megadalton (MDa) structure. The proteins that make up the nuclear pore complex are known as nucleoporins; each NPC contains at least 456 individual protein molecules and is composed of 34 distinct nucleoporin proteins. About half of the nucleoporins typically contain solenoid protein domains—either an alpha solenoid or a beta-propeller fold, or in some cases both as separate structural domains. The other half show structural characteristics typical of "natively unfolded" or intrinsically disordered proteins, i.e. they are highly flexible proteins that lack ordered tertiary structure. These disordered proteins are the FG nucleoporins, so called because their amino-acid sequence contains many phenylalanine–glycine repeats.

<span class="mw-page-title-main">Spindle apparatus</span> Feature of biological cell structure

In cell biology, the spindle apparatus refers to the cytoskeletal structure of eukaryotic cells that forms during cell division to separate sister chromatids between daughter cells. It is referred to as the mitotic spindle during mitosis, a process that produces genetically identical daughter cells, or the meiotic spindle during meiosis, a process that produces gametes with half the number of chromosomes of the parent cell.

<span class="mw-page-title-main">Telophase</span> Final stage of a cell division for eukaryotic cells both in mitosis and meiosis

Telophase is the final stage in both meiosis and mitosis in a eukaryotic cell. During telophase, the effects of prophase and prometaphase are reversed. As chromosomes reach the cell poles, a nuclear envelope is re-assembled around each set of chromatids, the nucleoli reappear, and chromosomes begin to decondense back into the expanded chromatin that is present during interphase. The mitotic spindle is disassembled and remaining spindle microtubules are depolymerized. Telophase accounts for approximately 2% of the cell cycle's duration.

<span class="mw-page-title-main">Tubulin</span> Superfamily of proteins that make up microtubules

Tubulin in molecular biology can refer either to the tubulin protein superfamily of globular proteins, or one of the member proteins of that superfamily. α- and β-tubulins polymerize into microtubules, a major component of the eukaryotic cytoskeleton. Microtubules function in many essential cellular processes, including mitosis. Tubulin-binding drugs kill cancerous cells by inhibiting microtubule dynamics, which are required for DNA segregation and therefore cell division.

<span class="mw-page-title-main">Spindle checkpoint</span> Cell cycle checkpoint

The spindle checkpoint, also known as the metaphase-to-anaphase transition, the spindle assembly checkpoint (SAC), the metaphase checkpoint, or the mitotic checkpoint, is a cell cycle checkpoint during mitosis or meiosis that prevents the separation of the duplicated chromosomes (anaphase) until each chromosome is properly attached to the spindle. To achieve proper segregation, the two kinetochores on the sister chromatids must be attached to opposite spindle poles. Only this pattern of attachment will ensure that each daughter cell receives one copy of the chromosome. The defining biochemical feature of this checkpoint is the stimulation of the anaphase-promoting complex by M-phase cyclin-CDK complexes, which in turn causes the proteolytic destruction of cyclins and proteins that hold the sister chromatids together.

<span class="mw-page-title-main">Kinetochore</span> Protein complex that allows microtubules to attach to chromosomes during cell division

A kinetochore is a disc-shaped protein structure associated with duplicated chromatids in eukaryotic cells where the spindle fibers attach during cell division to pull sister chromatids apart. The kinetochore assembles on the centromere and links the chromosome to microtubule polymers from the mitotic spindle during mitosis and meiosis. The term kinetochore was first used in a footnote in a 1934 Cytology book by Lester W. Sharp and commonly accepted in 1936. Sharp's footnote reads: "The convenient term kinetochore has been suggested to the author by J. A. Moore", likely referring to John Alexander Moore who had joined Columbia University as a freshman in 1932.

Aurora kinases are serine/threonine kinases that are essential for cell proliferation. They are phosphotransferase enzymes that help the dividing cell dispense its genetic materials to its daughter cells. More specifically, Aurora kinases play a crucial role in cellular division by controlling chromatid segregation. Defects in this segregation can cause genetic instability, a condition which is highly associated with tumorigenesis. The first aurora kinases were identified in Drosophila melanogaster, where mutations led to failure of centrosome separation with the monopolar spindles reminiscent of the North Pole, suggesting the name aurora.

<span class="mw-page-title-main">Stathmin</span> Protein in Eukaryotes

Stathmin, also known as metablastin and oncoprotein 18 is a protein that in humans is encoded by the STMN1 gene.

<span class="mw-page-title-main">Ran (protein)</span> GTPase functioning in nuclear transport

Ran also known as GTP-binding nuclear protein Ran is a protein that in humans is encoded by the RAN gene. Ran is a small 25 kDa protein that is involved in transport into and out of the cell nucleus during interphase and also involved in mitosis. It is a member of the Ras superfamily.

<span class="mw-page-title-main">Aurora kinase A</span> Protein-coding gene in the species Homo sapiens

Aurora kinase A also known as serine/threonine-protein kinase 6 is an enzyme that in humans is encoded by the AURKA gene.

In cell biology, microtubule nucleation is the event that initiates de novo formation of microtubules (MTs). These filaments of the cytoskeleton typically form through polymerization of α- and β-tubulin dimers, the basic building blocks of the microtubule, which initially interact to nucleate a seed from which the filament elongates.

<span class="mw-page-title-main">Aurora kinase B</span> Protein

Aurora kinase B is a protein that functions in the attachment of the mitotic spindle to the centromere.

<span class="mw-page-title-main">PCNT</span> Protein-coding gene in the species Homo sapiens

Pericentrin (kendrin), also known as PCNT and pericentrin-B (PCNTB), is a protein which in humans is encoded by the PCNT gene on chromosome 21. This protein localizes to the centrosome and recruits proteins to the pericentriolar matrix (PCM) to ensure proper centrosome and mitotic spindle formation, and thus, uninterrupted cell cycle progression. This gene is implicated in many diseases and disorders, including congenital disorders such as microcephalic osteodysplastic primordial dwarfism type II (MOPDII) and Seckel syndrome.

<span class="mw-page-title-main">NEDD1</span> Protein-coding gene in the species Homo sapiens

Neural precursor cell expressed, developmentally down-regulated 1, also known as Nedd1, is a human gene and encodes the protein NEDD1.

<span class="mw-page-title-main">Anthony A. Hyman</span> British biologist

Anthony Arie Hyman FRS is a British scientist and director at the Max Planck Institute of Molecular Cell Biology and Genetics.

<span class="mw-page-title-main">Centrosome cycle</span> Centrioles are nine triplets microtubules

Centrosomes are the major microtubule organizing centers (MTOC) in mammalian cells. Failure of centrosome regulation can cause mistakes in chromosome segregation and is associated with aneuploidy. A centrosome is composed of two orthogonal cylindrical protein assemblies, called centrioles, which are surrounded by a protein dense amorphous cloud of pericentriolar material (PCM). The PCM is essential for nucleation and organization of microtubules. The centrosome cycle is important to ensure that daughter cells receive a centrosome after cell division. As the cell cycle progresses, the centrosome undergoes a series of morphological and functional changes. Initiation of the centrosome cycle occurs early in the cell cycle in order to have two centrosomes by the time mitosis occurs.

<span class="mw-page-title-main">Kinesin-like protein KIF11</span> Protein-coding gene in the species Homo sapiens

Kinesin-like protein KIF11 is a molecular motor protein that is essential in mitosis. In humans it is coded for by the gene KIF11. Kinesin-like protein KIF11 is a member of the kinesin superfamily, which are nanomotors that move along microtubule tracks in the cell. Named from studies in the early days of discovery, it is also known as Kinesin-5, or as BimC, Eg5 or N-2, based on the founding members of this kinesin family.

<span class="mw-page-title-main">KIF15</span> Protein-coding gene in the species Homo sapiens

Kinesin family member 15 is a protein that in humans is encoded by the KIF15 gene.

<span class="mw-page-title-main">Rebecca Heald</span> American cell and developmental biologist

Rebecca W. Heald is an American professor of cell and developmental biology. She is currently a Professor in the Department of Molecular & Cell Biology at the University of California, Berkeley. In May 2019, she was elected to the National Academy of Sciences. She has published over 90 research articles in peer reviewed journals.

References

  1. 1 2 3 GRCh38: Ensembl release 89: ENSG00000088325 - Ensembl, May 2017
  2. 1 2 3 GRCm38: Ensembl release 89: ENSMUSG00000027469 - Ensembl, May 2017
  3. "Human PubMed Reference:". National Center for Biotechnology Information, U.S. National Library of Medicine.
  4. "Mouse PubMed Reference:". National Center for Biotechnology Information, U.S. National Library of Medicine.
  5. 1 2 3 Heidebrecht HJ, Buck F, Steinmann J, Sprenger R, Wacker HH, Parwaresch R (July 1997). "p100: a novel proliferation-associated nuclear protein specifically restricted to cell cycle phases S, G2, and M". Blood. 90 (1): 226–33. doi: 10.1182/blood.V90.1.226 . PMID   9207457.
  6. Zhang Y, Heidebrecht H, Rott A, Schlegelberger B, Parwaresch R (Jul 1999). "Assignment of human proliferation associated p100 gene (C20orf1) to human chromosome band 20q11.2 by in situ hybridization". Cytogenetics and Cell Genetics. 84 (3–4): 182–3. doi:10.1159/000015251. PMID   10393424. S2CID   8203545.
  7. "Entrez Gene: TPX2 TPX2, microtubule-associated, homolog (Xenopus laevis)".
  8. Vos JW, Pieuchot L, Evrard JL, Janski N, Bergdoll M, de Ronde D, Perez LH, Sardon T, Vernos I, Schmit AC (October 2008). "The plant TPX2 protein regulates prospindle assembly before nuclear envelope breakdown". The Plant Cell. 20 (10): 2783–97. doi:10.1105/tpc.107.056796. PMC   2590745 . PMID   18941054.
  9. 1 2 3 4 5 Trieselmann N, Armstrong S, Rauw J, Wilde A (December 2003). "Ran modulates spindle assembly by regulating a subset of TPX2 and Kid activities including Aurora A activation". Journal of Cell Science. 116 (Pt 23): 4791–8. doi: 10.1242/jcs.00798 . PMID   14600264.
  10. Sanchez-Pulido L, Perez L, Kuhn S, Vernos I, Andrade-Navarro MA (October 2016). "The C-terminal domain of TPX2 is made of alpha-helical tandem repeats". BMC Structural Biology. 16 (1): 17. doi:10.1186/s12900-016-0070-8. PMC   5080731 . PMID   27782824.
  11. 1 2 3 4 5 6 7 8 9 10 11 12 Alfaro-Aco R, Thawani A, Petry S (April 2017). "Structural analysis of the role of TPX2 in branching microtubule nucleation". The Journal of Cell Biology. 216 (4): 983–997. doi:10.1083/jcb.201607060. PMC   5379942 . PMID   28264915.
  12. 1 2 3 4 5 6 7 8 Eckerdt F, Eyers PA, Lewellyn AL, Prigent C, Maller JL (April 2008). "Spindle pole regulation by a discrete Eg5-interacting domain in TPX2". Current Biology. 18 (7): 519–25. doi:10.1016/j.cub.2008.02.077. PMC   2408861 . PMID   18372177.
  13. Ma N, Tulu US, Ferenz NP, Fagerstrom C, Wilde A, Wadsworth P (March 2010). "Poleward transport of TPX2 in the mammalian mitotic spindle requires dynein, Eg5, and microtubule flux". Molecular Biology of the Cell. 21 (6): 979–88. doi:10.1091/mbc.e09-07-0601. PMC   2836978 . PMID   20110350.
  14. 1 2 3 4 5 6 7 8 Stewart S, Fang G (December 2005). "Anaphase-promoting complex/cyclosome controls the stability of TPX2 during mitotic exit". Molecular and Cellular Biology. 25 (23): 10516–27. doi:10.1128/MCB.25.23.10516-10527.2005. PMC   1291225 . PMID   16287863.
  15. Pfleger CM, Kirschner MW (March 2000). "The KEN box: an APC recognition signal distinct from the D box targeted by Cdh1". Genes & Development. 14 (6): 655–65. doi:10.1101/gad.14.6.655. PMC   316466 . PMID   10733526.
  16. Pfleger CM, Lee E, Kirschner MW (September 2001). "Substrate recognition by the Cdc20 and Cdh1 components of the anaphase-promoting complex". Genes & Development. 15 (18): 2396–407. doi:10.1101/gad.918201. PMC   312782 . PMID   11562349.
  17. 1 2 3 4 5 Wittmann T, Wilm M, Karsenti E, Vernos I (June 2000). "TPX2, A novel xenopus MAP involved in spindle pole organization". The Journal of Cell Biology. 149 (7): 1405–18. doi:10.1083/jcb.149.7.1405. PMC   2175143 . PMID   10871281.
  18. 1 2 3 Gruss OJ, Wittmann M, Yokoyama H, Pepperkok R, Kufer T, Silljé H, Karsenti E, Mattaj IW, Vernos I (November 2002). "Chromosome-induced microtubule assembly mediated by TPX2 is required for spindle formation in HeLa cells". Nature Cell Biology. 4 (11): 871–9. doi:10.1038/ncb870. PMID   12389033. S2CID   20151781.
  19. 1 2 Gruss OJ, Vernos I (September 2004). "The mechanism of spindle assembly: functions of Ran and its target TPX2". The Journal of Cell Biology. 166 (7): 949–55. doi:10.1083/jcb.200312112. PMC   2172015 . PMID   15452138.
  20. 1 2 Gruss OJ, Carazo-Salas RE, Schatz CA, Guarguaglini G, Kast J, Wilm M, Le Bot N, Vernos I, Karsenti E, Mattaj IW (January 2001). "Ran induces spindle assembly by reversing the inhibitory effect of importin alpha on TPX2 activity". Cell. 104 (1): 83–93. doi: 10.1016/S0092-8674(01)00193-3 . PMID   11163242. S2CID   18948501.
  21. 1 2 3 Reid TA, Schuster BM, Mann BJ, Balchand SK, Plooster M, McClellan M, Coombes CE, Wadsworth P, Gardner MK (April 2016). "Suppression of microtubule assembly kinetics by the mitotic protein TPX2". Journal of Cell Science. 129 (7): 1319–28. doi:10.1242/jcs.178806. PMC   4852719 . PMID   26869224.
  22. 1 2 3 4 Petry S, Groen AC, Ishihara K, Mitchison TJ, Vale RD (February 2013). "Branching microtubule nucleation in Xenopus egg extracts mediated by augmin and TPX2". Cell. 152 (4): 768–77. doi:10.1016/j.cell.2012.12.044. PMC   3680348 . PMID   23415226.
  23. 1 2 3 Bayliss R, Sardon T, Vernos I, Conti E (October 2003). "Structural basis of Aurora-A activation by TPX2 at the mitotic spindle". Molecular Cell. 12 (4): 851–62. doi: 10.1016/S1097-2765(03)00392-7 . PMID   14580337.
  24. Eyers PA, Maller JL (March 2004). "Regulation of Xenopus Aurora A activation by TPX2". The Journal of Biological Chemistry. 279 (10): 9008–15. doi: 10.1074/jbc.M312424200 . PMID   14701852.
  25. Bringmann H, Hyman AA (August 2005). "A cytokinesis furrow is positioned by two consecutive signals". Nature. 436 (7051): 731–4. Bibcode:2005Natur.436..731B. doi:10.1038/nature03823. hdl: 11858/00-001M-0000-002C-9BBA-6 . PMID   16079852. S2CID   4337162.
  26. D'Avino PP, Savoian MS, Glover DM (April 2005). "Cleavage furrow formation and ingression during animal cytokinesis: a microtubule legacy". Journal of Cell Science. 118 (Pt 8): 1549–58. doi: 10.1242/jcs.02335 . PMID   15811947.
  27. Wittmann T, Boleti H, Antony C, Karsenti E, Vernos I (November 1998). "Localization of the kinesin-like protein Xklp2 to spindle poles requires a leucine zipper, a microtubule-associated protein, and dynein". The Journal of Cell Biology. 143 (3): 673–85. doi:10.1083/jcb.143.3.673. PMC   2148133 . PMID   9813089.
  28. Boleti H, Karsenti E, Vernos I (January 1996). "Xklp2, a novel Xenopus centrosomal kinesin-like protein required for centrosome separation during mitosis". Cell. 84 (1): 49–59. doi: 10.1016/S0092-8674(00)80992-7 . PMID   8548825. S2CID   17297180.
  29. Garrett S, Auer K, Compton DA, Kapoor TM (December 2002). "hTPX2 is required for normal spindle morphology and centrosome integrity during vertebrate cell division". Current Biology. 12 (23): 2055–9. doi: 10.1016/S0960-9822(02)01277-0 . PMID   12477396. S2CID   17084335.
  30. 1 2 3 4 5 6 Hsu CW, Chen YC, Su HH, Huang GJ, Shu CW, Wu TT, Pan HW (2017). "Targeting TPX2 Suppresses the Tumorigenesis of Hepatocellular Carcinoma Cells Resulting in Arrested Mitotic Phase Progression and Increased Genomic Instability". Journal of Cancer. 8 (8): 1378–1394. doi:10.7150/jca.17478. PMC   5479243 . PMID   28638452.
  31. Higuchi T, Uhlmann F (January 2005). "Stabilization of microtubule dynamics at anaphase onset promotes chromosome segregation". Nature. 433 (7022): 171–6. Bibcode:2005Natur.433..171H. doi:10.1038/nature03240. PMC   2586334 . PMID   15650742.
  32. Harper JW, Burton JL, Solomon MJ (September 2002). "The anaphase-promoting complex: it's not just for mitosis any more". Genes & Development. 16 (17): 2179–206. doi: 10.1101/gad.1013102 . PMID   12208841.
  33. Kahana JA, Cleveland DW (March 2001). "Cell cycle. Some importin news about spindle assembly". Science. 291 (5509): 1718–9. doi:10.1126/science.1059765. PMID   11253198. S2CID   82177829.
  34. Schatz CA, Santarella R, Hoenger A, Karsenti E, Mattaj IW, Gruss OJ, Carazo-Salas RE (May 2003). "Importin alpha-regulated nucleation of microtubules by TPX2". The EMBO Journal. 22 (9): 2060–70. doi:10.1093/emboj/cdg195. PMC   156067 . PMID   12727873.
  35. 1 2 3 4 Neumayer G, Helfricht A, Shim SY, Le HT, Lundin C, Belzil C, Chansard M, Yu Y, Lees-Miller SP, Gruss OJ, van Attikum H, Helleday T, Nguyen MD (December 2012). "Targeting protein for xenopus kinesin-like protein 2 (TPX2) regulates γ-histone 2AX (γ-H2AX) levels upon ionizing radiation". The Journal of Biological Chemistry. 287 (50): 42206–22. doi: 10.1074/jbc.M112.385674 . PMC   3516765 . PMID   23045526.
  36. 1 2 3 4 Neumayer G, Nguyen MD (2014-11-03). "TPX2 impacts acetylation of histone H4 at lysine 16: implications for DNA damage response". PLOS ONE. 9 (11): e110994. Bibcode:2014PLoSO...9k0994N. doi: 10.1371/journal.pone.0110994 . PMC   4217740 . PMID   25365214.
  37. van Attikum H, Gasser SM (May 2009). "Crosstalk between histone modifications during the DNA damage response". Trends in Cell Biology. 19 (5): 207–17. doi:10.1016/j.tcb.2009.03.001. PMID   19342239.
  38. Yang X, Liu G, Xiao H, Yu F, Xiang X, Lu Y, Li W, Liu X, Li S, Shi Y (July 2014). "TPX2 overexpression in medullary thyroid carcinoma mediates TT cell proliferation". Pathology & Oncology Research. 20 (3): 641–8. doi:10.1007/s12253-014-9743-4. PMID   24488334. S2CID   15649618.
  39. Yan L, Li S, Xu C, Zhao X, Hao B, Li H, Qiao B (December 2013). "Target protein for Xklp2 (TPX2), a microtubule-related protein, contributes to malignant phenotype in bladder carcinoma". Tumour Biology. 34 (6): 4089–100. doi:10.1007/s13277-013-1000-z. PMID   23873098. S2CID   15850641.
  40. Geiger TR, Ha NH, Faraji F, Michael HT, Rodriguez L, Walker RC, Green JE, Simpson RM, Hunter KW (2014). "Functional analysis of prognostic gene expression network genes in metastatic breast cancer models". PLOS ONE. 9 (11): e111813. Bibcode:2014PLoSO...9k1813G. doi: 10.1371/journal.pone.0111813 . PMC   4219783 . PMID   25368990.
  41. Liang B, Jia C, Huang Y, He H, Li J, Liao H, Liu X, Liu X, Bai X, Yang D (August 2015). "TPX2 Level Correlates with Hepatocellular Carcinoma Cell Proliferation, Apoptosis, and EMT". Digestive Diseases and Sciences. 60 (8): 2360–72. doi:10.1007/s10620-015-3730-9. PMID   26025609. S2CID   22794481.
  42. Liu Q, Yang P, Tu K, Zhang H, Zheng X, Yao Y, Liu Q (August 2014). "TPX2 knockdown suppressed hepatocellular carcinoma cell invasion via inactivating AKT signaling and inhibiting MMP2 and MMP9 expression". Chinese Journal of Cancer Research. 26 (4): 410–7. doi:10.3978/j.issn.1000-9604.2014.08.01. PMC   4153921 . PMID   25232213.
  43. Huang Y, Guo W, Kan H (October 2014). "TPX2 is a prognostic marker and contributes to growth and metastasis of human hepatocellular carcinoma". International Journal of Molecular Sciences. 15 (10): 18148–61. doi: 10.3390/ijms151018148 . PMC   4227208 . PMID   25302620.
  44. Liu HC, Zhang GH, Liu YH, Wang P, Ma JF, Su LS, Li SL, Zhang L, Liu JW (September 2014). "TPX2 siRNA regulates growth and invasion of esophageal cancer cells". Biomedicine & Pharmacotherapy. 68 (7): 833–9. doi:10.1016/j.biopha.2014.08.008. PMID   25239289.
  45. Jiang P, Shen K, Wang X, Song H, Yue Y, Liu T (June 2014). "TPX2 regulates tumor growth in human cervical carcinoma cells". Molecular Medicine Reports. 9 (6): 2347–51. doi: 10.3892/mmr.2014.2106 . PMID   24718984.
  46. Miwa T, Kokuryo T, Yokoyama Y, Yamaguchi J, Nagino M (July 2015). "Therapeutic potential of targeting protein for Xklp2 silencing for pancreatic cancer". Cancer Medicine. 4 (7): 1091–100. doi:10.1002/cam4.453. PMC   4529347 . PMID   25914189.
  47. Pan HW, Ou YH, Peng SY, Liu SH, Lai PL, Lee PH, Sheu JC, Chen CL, Hsu HC (July 2003). "Overexpression of osteopontin is associated with intrahepatic metastasis, early recurrence, and poorer prognosis of surgically resected hepatocellular carcinoma". Cancer. 98 (1): 119–27. doi: 10.1002/cncr.11487 . PMID   12833464. S2CID   20254647.
  48. McGranahan N, Burrell RA, Endesfelder D, Novelli MR, Swanton C (June 2012). "Cancer chromosomal instability: therapeutic and diagnostic challenges". EMBO Reports. 13 (6): 528–38. doi:10.1038/embor.2012.61. PMC   3367245 . PMID   22595889.
  49. Holland AJ, Cleveland DW (June 2012). "Losing balance: the origin and impact of aneuploidy in cancer". EMBO Reports. 13 (6): 501–14. doi:10.1038/embor.2012.55. PMC   3367240 . PMID   22565320.

Further reading