The kinetochore is the macromolecular protein complex that drives chromosome segregation in eukaryotes. Its most fundamental function is to connect centromeric DNA to dynamic spindle microtubules. Studies in popular model eukaryotes have shown that centromere protein (CENP)-A is critical for DNA-binding, whereas the Ndc80 complex is essential for microtubule-binding. Given their conservation in diverse eukaryotes, it was widely believed that all eukaryotes would utilize these components to make up a core of the kinetochore. However, a recent study identified an unconventional type of kinetochore in evolutionarily distant kinetoplastid species, showing that chromosome segregation can be achieved using a distinct set of proteins. Here, I review the discovery of the two kinetochore systems and discuss how their studies contribute to a better understanding of the eukaryotic chromosome segregation machinery.

Faithful transmission of genetic information from generation to generation is essential for the survival of all organisms. In eukaryotes, chromosome replication occurs during S phase and duplicated chromosomes are physically connected by the cohesin complex [1] (Figure 1). Chromosome segregation is directed by the kinetochore, the proteinaceous structure that assembles onto the centromeric DNA and interacts with spindle microtubules during mitosis and meiosis [25]. In many species, centromeres consist of repetitive sequences that span several kilobases up to megabases [6]. Kinetochores assemble onto a small portion of these sequences and their positional information is epigenetically inherited from generation to generation [79]. Microtubules are dynamic polymers that grow and shrink by incorporating or dissociating α-/β-tubulin subunits at their tips [10]. By interacting with microtubules, kinetochores enable duplicated chromosomes to align at the center of the spindle. Accurate segregation requires sister kinetochores to make bi-oriented attachments to microtubules emanating from opposite poles. Attachment errors must be recognized and corrected prior to anaphase [11]. The spindle checkpoint monitors attachment status and delays the onset of anaphase until all chromosomes form correct bi-oriented attachments [12,13]. Once all chromosomes have achieved bi-orientation, the anaphase promoting complex (APC/C) is activated, leading to the cleavage of cohesin complexes and separation of sister chromatids [1416].

Chromosome segregation.

Figure 1.
Chromosome segregation.

Chromosome duplication occurs during S phase, linking sister chromatids by cohesion. When cells enter mitosis, a bipolar spindle is assembled and kinetochores start to interact with dynamic microtubules. Once all chromosomes achieve bi-oriented attachments, the spindle checkpoint is satisfied. Subsequent activation of APC/C leads to the destruction of cohesion, segregating sister chromatids away from each other in anaphase.

Figure 1.
Chromosome segregation.

Chromosome duplication occurs during S phase, linking sister chromatids by cohesion. When cells enter mitosis, a bipolar spindle is assembled and kinetochores start to interact with dynamic microtubules. Once all chromosomes achieve bi-oriented attachments, the spindle checkpoint is satisfied. Subsequent activation of APC/C leads to the destruction of cohesion, segregating sister chromatids away from each other in anaphase.

Close modal

Given the central role of kinetochores in co-ordinating chromosome segregation, elucidating their molecular mechanism of action is a fundamental issue in cell biology. The most basic kinetochore functions are to bind DNA and dynamic microtubules. What kind of proteins carry out these functions? How did they evolve? Identifying and characterizing the constituent proteins are clearly necessary as the first step to address these questions. Here, I review the discovery of two different kinetochore systems and discuss how their studies will contribute to a better understanding of chromosome segregation machines in eukaryotes. I place emphasis on the unconventional kinetoplastid kinetochore and refer readers to comprehensive reviews on the conventional kinetochore [3,4,1719].

Discovery

Centromeres/kinetochores were first recognized cytologically as primary constrictions on each chromosome where spindle microtubules attach [20]. Kinetochore structures were visualized by electron microscopy [2123], revealing a trilaminar structure at the periphery of the centromere. Identification of kinetochore components was made possible by the discovery of autoimmune sera from patients with scleroderma spectrum disease that stained the centromere regions of chromosomes [24]. Using the sera, the first kinetochore proteins (centromere proteins, CENP-A, B, C) were recognized [25,26]. Their cDNA clones were subsequently isolated and protein sequences revealed [2729].

Although the first kinetochore proteins were discovered in humans, many additional kinetochore proteins were subsequently identified in budding yeast owing to the power of genetics and its simple centromere structure [30]. For example, biochemical purifications of centromere DNA-binding proteins [31], genetic screens for mutants that have increased rates of chromosome loss [32,33], one-hybrid [34] and two-hybrid screens [35] all led to the identification of many kinetochore proteins. More recently, affinity purification and mass spectrometry analysis accelerated the identification [36,37].

Increased sensitivity in homology detection algorithms has allowed the identification of homologous proteins in more and more species as the number of sequenced genomes has expanded [3841]. Genome-wide RNAi screens [42,43] and proteomic analyses of isolated chromosomes [44,45] also identified kinetochore proteins. To date, more than 80 proteins have been identified that at least transiently localize to kinetochore regions, including structural kinetochore proteins (e.g. CENP-A, CCAN, KNL1, Mis12 and Ndc80), regulatory proteins [e.g. Aurora B/chromosomal passenger complex (CPC), Mps1, Plk1, PP1 and PP2A], and spindle checkpoint proteins (e.g. Bub1, Bub3, BubR1/Mad3, Mad1 and Mad2). After identifying proteins that localize at kinetochores, the next crucial step is to reveal their function. Most notably, which proteins bind DNA or microtubules?

Identification of DNA-binding kinetochore proteins

Cloning and sequencing of CENP-A cDNA revealed that it is a histone H3 variant [29,46]. This was a very exciting finding because it not only revealed that CENP-A is a DNA-binding protein but also suggested a possible mechanism for how kinetochore positions could be epigenetically inherited [47,48]. Subsequent studies have shown that CENP-A is essential for kinetochore assembly in various eukaryotes [42,49,50]. Given its importance in kinetochore biology, extensive studies have been performed to better understand the nature and function of CENP-A ([5154]).

Other proteins that can bind DNA, at least in vitro, include CENP-CMif2 [55,56], CENP-UAme1/CENP-QOkp1 [57], and CENP-T/W/S/X [5860]. Minor groove DNA-binding motifs are found in some of these proteins: the AT-hook [61] in CENP-CMif2 and CENP-UAme1/CENP-QOkp1 [56,57,62], and the SPKK motif [63] in CENP-A [64]. However, these short motifs are often poorly conserved even among closely related species, suggesting that they are probably not essential for their functions but may contribute to enhancing their DNA-binding affinity [56]. Regardless, the presence of these short motifs could be indicative of DNA-binding activities for a given protein.

Identification of microtubule-binding kinetochore proteins

Human Ndc80HEC was originally identified in a two-hybrid screen with retinoblastoma protein as bait [65], whereas its yeast homolog was identified in a spindle pole body preparation [66]. Ndc80 interacts with Nuf2, Spc24, and Spc25, which together form a rod-shaped complex [67,68]. All four subunits contain globular domains and extensive coiled-coil regions. Although depletion studies showed that the Ndc80 complex is essential for kinetochore–microtubule attachment [36,69,70], it was not clear whether the complex directly interacts with microtubules or recruits other kinetochore proteins that bind microtubules. It was an in vitro microtubule co-sedimentation assay that revealed that the former was the case [71,72]. However, it took nearly a decade since the identification of Ndc80 to assign it a function at the kinetochore. Given its now recognized importance in microtubule-binding, extensive studies have been carried out to mechanistically understand how the Ndc80 complex binds to microtubules ([7378]).

Kinetochore subcomplexes and hierarchical assembly

Another important aspect of kinetochore research is to understand its design principles. Electron microscopy studies have visualized the overall architecture of kinetochores within cells [79,80] as well as isolated kinetochores [81]. At the molecular level, kinetochores are composed of biochemically stable/isolatable subcomplexes, such as the Ndc80 subcomplex, Mis12 subcomplex, and CCAN [37,8286]. It is thought that these subcomplexes assemble in a hierarchical manner at the kinetochore [82]. CENP-A is at the base of this hierarchy and recruits CENP-C [52,87], which in turn recruits the Mis12 subcomplex [88,89] as well as CCAN components [90,91]. The KNL1, Mis12, and Ndc80 subcomplexes can form a larger complex called the KMN network that has microtubule-binding activities [71,92]. The relative positions of these subcomplexes within kinetochores have been determined by super-resolution microscopy, revealing the overall protein architecture of kinetochores [9395].

Structural features

The majority of kinetochore proteins lack obvious functional domains or similarity to other proteins at the primary sequence level. However, structural analyses have revealed common domains in proteins that do not have significant sequence similarity. For example, the RWD domain was found in Spc24, Spc25, CENP-PCtf19, CENP-OMcm21, Csm1, Mad1, and KNL1 proteins [67,68,92,96,97], calponin homology domain in Ndc80 and Nuf2 proteins [72,73], and tetratricopeptide repeat domain in Bub1, BubR1, and Mps1 proteins [98,99]. Furthermore, CENP-LIml3 has similarity to the bacterial recombination-associated protein RdgC and TATA-binding protein TBP [100,101], CENP-M folds like a small GTPase [102], the C-terminal dimerization domain of CENP-CMif2 has similarity to the bacterial transcription factor Hth-3 [56], and point centromere-specific Ndc10 has similarity to a bacterial tyrosine recombinase [103,104]. These studies highlight the importance of structural analysis in kinetochore research to reveal distant homology. They also suggest that the highly complicated kinetochores likely originated from a few protein modules aided by gene duplication and functional diversification [105].

How widely are these kinetochore proteins conserved?

Microtubules appear to be utilized for chromosome segregation in all eukaryotes studied thus far, and α-/β-tubulins are one of the most highly conserved proteins [106108]. Kinetochore proteins, in contrast, vary to a great extent in primary sequences [109]. Yet, thanks to increased sensitivity in homology search algorithms, we now know that many of the above-mentioned kinetochore proteins are conserved in various eukaryotes [40,62]. CENP-A, CENP-C, and components of the Ndc80 complex are widely found in diverse eukaryotes, suggesting that most eukaryotes utilize these kinetochore proteins to bind DNA or microtubules. However, although CENP-A was thought to be conserved in all eukaryotes, a recent study showed that CENP-A is absent in some holocentric insects [110]. Furthermore, two popular model eukaryotes (Drosophila melanogaster and Caenorhabditis elegans) appear to have lost the majority of CCAN components [111114]. Therefore, it has become clear that the repertoire of kinetochore proteins can be dramatically different among eukaryotes despite functional conservation [115].

More extreme cases were found in kinetoplastids, a group of unicellular eukaryotes defined by the presence of kinetoplast which is a large structure in the mitochondrion that contains the mitochondrial DNA [116]. When the genomes of three kinetoplastid species were sequenced (Trypanosoma brucei, Trypanosoma cruzi, and Leishmania major), none of the core structural kinetochore proteins were identified [117119]. In contrast, components of basic cell cycle machinery were readily identified, including the CDK/cyclin system, the cohesin complex, separase, the condensin complex, Aurora B kinase, Polo-like kinase, MAD2, APC/C, and proteasomes [120]. However, chromosome segregation depends on spindle microtubules [121], and electron microscopy has visualized kinetochore-like electron-dense plaques that appear to form end-on attachments to spindle microtubules in mitotic cells [122124]. Yet, nothing was known about the identity of kinetochore components in kinetoplastids. Do they have conventional kinetochore proteins that simply diverged too far in primary sequences? Or do they have previously unforeseen types of kinetochore proteins?

Discovery: personal reflections

My scientific career started in 2003 when I joined Dr Yoshinori Watanabe's group in Tokyo to study how Polo-like kinase modulates kinetochore function to promote meiosis-specific chromosome segregation in fission yeast [125]. I then moved to Dr Sue Biggins' laboratory in Seattle to study mitotic kinetochores using budding yeast. There, I achieved the isolation of native kinetochores for the first time [44,126,127] and obtained a PhD. When I was thinking about projects for my post-doctoral training, I became interested in finding out whether basic kinetochore proteins are conserved in all eukaryotes. After I learned that kinetoplastids do not have any apparent canonical kinetochore components including CENP-A [128,129], I joined the laboratory of Dr Keith Gull in Oxford in November 2010 to identify kinetochore proteins in T. brucei. There were no magical antibodies that recognized the centromere regions of its chromosomes, so how could I achieve this goal from scratch? With my expertise in protein purification, I first immunoprecipitated conserved mitotic proteins and performed mass spectrometry to identify interacting proteins [the immunoprecipitation (IP)/mass spectrometry (MS) approach]. Because spindle checkpoint components co-purify with kinetochore proteins in other eukaryotes [130], I first made a YFP fusion for MAD2, the only spindle checkpoint protein homolog identified in T. brucei. However, this protein turned out to localize at basal bodies, having no detectable signal at kinetochores [120]. Consistent with this observation, IP/MS of YFP-MAD2 only led to the identification of another protein that localized to basal bodies (B. Akiyoshi, unpublished data). I then made YFP fusions for various proteins, including AUK1 (Aurora B kinase homolog [131]), CDC20 (activator of the APC/C [132]), and XMAP215 (microtubule-associated protein [133,134]). Again, although IP/MS of these proteins worked and identified interacting proteins, this approach failed to identify any protein that localized at kinetochores (B. Akiyoshi, unpublished data).

Another idea I had in mind was to carry out a YFP-tagging screen. The T. brucei genome has approximately 9000 proteins, which means that I should identify kinetochore proteins if I study all of them. This was a simple and straightforward idea, but I did not know where to start. It was before the development of PCR-based tagging methods in the laboratory [135], meaning that I had to make targeting plasmids for each gene [136]. Coincidentally, Dr Christine Clayton's group published a cell cycle transcriptome analysis of different cell cycle stages in March 2011 in PLoS ONE [137]. Impressed by the quality of data, I immediately started tagging uncharacterized genes whose transcript levels were up-regulated in either S phase or mitosis. This approach paid off. Examination of only 30 genes identified one corresponding protein that had a characteristic kinetochore localization pattern: formation of dots in the nucleus during S phase, their alignment on mitotic spindles in metaphase, and close association with spindle poles during late anaphase. This was my first ‘Eureka!’ moment and happened in June 2011. Later I named the protein KKT1 for kinetoplastid kinetochore protein 1. Using KKT1 as bait, I aimed to identify more kinetochore proteins. However, my first IP of YFP-KKT1 completely failed using the same protocol that had worked for those conserved proteins mentioned above (cryolysis using mortar and pestle). It turned out that the YFP-KKT1 signal remained in the detergent-insoluble fraction, so I tried various methods to release it. I eventually found that sonication efficiently solubilized KKT1, and immunoprecipitation from this sample led to the identification of 12 proteins as possible KKT1 interaction partners. YFP-tagging revealed typical kinetochore localization patterns for all of them, which was another moment of big excitement. By continuing the hunt and repeating the IP/MS of these newly identified proteins, I identified 19 proteins by March 2012 (KKT1–19). These proteins all had kinetochore-like localization and their RNAi-mediated knockdowns resulted in chromosome mis-segregation. However, there was a fairly valid criticism I had to face. ‘Are they really kinetochore proteins?’

In other eukaryotes, people have used various approaches to examine kinetochore localization such as immunoelectron microscopy [138], chromatin immunoprecipitation (ChIP) [139], chromosome spreads [140], and co-localization with other known kinetochore proteins [36]. I first tried to perform immunoelectron microscopy to examine whether KKT proteins localize at the kinetochore-like electron-dense plaques [124], but it turned out that the fixation condition completely killed the YFP epitope. Then, I switched my effort to a ChIP approach, although it had a different problem. That is, the resolution of previously mapped centromere positions [141] did not allow me to determine which specific DNA fragments should be examined by PCR or Southern blot. I struggled >1 year at this stage. Eventually, deep sequencing of ChIP samples revealed that KKT2 and KKT3 were specifically enriched in the centromere regions, thereby showing that these KKT proteins are bona fide kinetochore proteins. We first submitted the paper to Nature, but it was rejected without review as being uninteresting, which happened to be the same reaction to Bill Earnshaw's paper on the discovery of CENP-A, B, C [26,142]. We next submitted it to Cell and it was accepted within 1 month [143].

I recently identified one more kinetochore protein, KKT20 [144], so 20 KKT proteins have been identified to date (Figure 2). It is important to mention that there may be more kinetochore proteins still waiting to be identified. For example, proteins that weakly or very transiently interact with KKT proteins may well have been missed in the IP/MS approach. Other methods, such as BioID [145] and genome-wide GFP tagging screen, may identify the rest of kinetochore proteins (if any). Another caveat is that any gene not predicted to encode a protein in the T. brucei genome database will remain unidentified. With these caveats in mind, I decided to study these KKT proteins in depth.

Diagram of T. brucei KKT proteins.

Figure 2.
Diagram of T. brucei KKT proteins.

Identified domains and motifs are shown. Putative subcomplexes are grouped in dotted boxes. Adapted from [143].

Figure 2.
Diagram of T. brucei KKT proteins.

Identified domains and motifs are shown. Putative subcomplexes are grouped in dotted boxes. Adapted from [143].

Close modal

Sequence analysis

Initial BLAST and HMMER searches of KKT proteins identified their homologous proteins in both parasitic and free-living kinetoplastids [143]. For most KKT proteins, these searches did not reveal any significant hit outside kinetoplastids. Some KKT proteins had similarity to various proteins due to the presence of the following conserved domains (Figure 2): a BRCT (BRCA1 C terminus) domain in KKT4, an FHA (Forkhead-associated) domain in KKT13, a WD40-like domain in KKT15, and a protein kinase domain in KKT2, KKT3, KKT10, and KKT19. BRCT and FHA are typically protein–protein interaction domains and are often found in proteins involved in the DNA damage response [146]. WD40 is also a protein–protein interaction domain that is one of the most abundant domains in eukaryotic genomes [146]. T. brucei has 190 predicted eukaryotic protein kinases, and comparative kinome analysis suggested that KKT2 and KKT3, despite having characteristic features of active protein kinases, do not have a clear affiliation to any known group or family among eukaryotic kinases [147]. Although KKT10 and KKT19 have been classified as members of the CLK/Lammer subfamily in the CMGC family, they may have adapted to carry out kinetochore functions in kinetoplastids as judged by the significant differences between KKT10/KKT19 and the human or Arabidopsis CLK/Lammer kinases [143].

I have also been using various sequence analysis software to try to reveal the function and nature of KKT proteins. This includes multiple sequence alignments to determine conserved regions [148], coiled-coil predictions [149], disordered region predictions [150], motif scans [151,152], secondary structure predictions [153,154], and HHpred [155]. Based on these analyses, I recently identified a highly divergent polo box domain (DPB) in the C-terminal region of KKT2, KKT3, and KKT20 [144]. KKT20 does not have a kinase domain, but has similarity to KKT2 and KKT3 in the central domain, suggesting that these proteins likely share common ancestry and that the ancestor might be PLK. In fact, a simple BLAST search showed that the kinase domains of KKT2 and KKT3 are more similar to that of PLK than other kinases. KKT10 and KKT19 exhibit a high degree of similarity at the protein level, as do KKT17 and KKT18. These results suggest that gene duplication played an important role in the evolution of kinetoplastid kinetochores, as seen for conventional kinetochores.

Predictions of functions

Similar to conventional kinetochores, kinetoplastid kinetochores appear to consist of subcomplexes as judged by the IP/MS results and differential localization patterns [143]. On the basis of IP/MS data, I assigned several putative subcomplexes that could represent functional modules: the KKT14/KKT15 subcomplex, the KKT16/KKT17/KKT18 subcomplex, and the KKT6/KKT7/KKT8/KKT9/KKT10/KKT11/KKT12/KKT19 subcomplex (Figure 2).

Studies of conventional kinetochore proteins have shown that their functions are often manifested in localization patterns [17]. For example, DNA-binding kinetochore components (e.g. CENP-A and CENP-C) are often constitutively localized at centromeres, whereas microtubule-binding components (e.g. Ndc80 and KNL1) localize to kinetochores specifically during M phase. Although the unconventional kinetoplastid kinetochore proteins may not follow the same principle, there is no evidence to believe otherwise. In fact, the observation that components of the putative subcomplexes mentioned above largely follow the same localization pattern supports this possibility (Figure 3). In this sense, the most promising candidates for DNA-binding proteins are KKT2, KKT3, and KKT4.

Differential localization timings of KKT proteins.

Figure 3.
Differential localization timings of KKT proteins.

Adapted from [143].

Figure 3.
Differential localization timings of KKT proteins.

Adapted from [143].

Close modal

Which KKT proteins bind DNA?

KKT2 and KKT3 have putative DNA-binding motifs (AT-hook and SPKK) in some, but not all, kinetoplastids. This is reminiscent of conventional DNA-binding kinetochore proteins such as CENP-A and CENP-C (see above), supporting the possibility that KKT2 and KKT3 may bind DNA. Besides these motifs, KKT2 and KKT3 have three domains conserved among kinetoplastids: an N-terminal protein kinase domain, a central domain that has conserved Cys residues, and the C-terminal DPB. Expression of truncated proteins in vivo can determine which region(s) is important for kinetochore/centromere localization, which might correspond to their DNA-binding domains (e.g. see [55,156,157] for CENP-C studies). It will be necessary to directly test whether KKT2 and KKT3 have any DNA-binding activity in vitro using pure proteins. Isolation of native KKT proteins or subcomplexes from trypanosomes in good quantity or stoichiometry turned out to be difficult, at least in my hands. Therefore, recombinant KKT proteins (full length or truncations) will need to be expressed and purified to carry out in vitro assays.

While in vitro assays are important to dissect their DNA-binding activities, in vivo analyses will be essential to determine their relevance. RNAi-mediated knockdown showed that KKT2 is important for faithful chromosome segregation [143], and a kinome-wide RNAi screen showed that KKT3 is essential for growth [158], suggesting that these homologous proteins likely have non-redundant functions. It will be necessary to examine whether other KKT proteins fail to localize to kinetochores in the absence of KKT2 or KKT3 to determine the localization hierarchy. Although inducible RNAi-mediated knockdown experiments can be easily performed in T. brucei [159,160], other methods such as conditional knockdowns [161], conditional knockouts [162], or a protein degradation system [163] may be necessary to achieve better depletion efficiency and observe defects.

Besides putative DNA-binding domains, KKT2 and KKT3 also have a protein kinase domain classified as unique among eukaryotic kinases [147]. What are the substrates of these protein kinases? Considering the absence of centromere-specific histones in kinetoplastids [128,164], KKT2/KKT3 might phosphorylate histones specifically at the centromere and this phosphorylation mark might provide centromere identity. Or they might phosphorylate and recruit other KKT proteins onto the kinetochore. Understanding the functions of KKT2 and KKT3 will be key to elucidating the assembly of kinetoplastid kinetochores.

Which KKT proteins bind microtubules?

Unlike the DNA side, sequence analysis did not reveal any obvious domain implicated in microtubule-binding. In the light of conventional kinetochore research, it will be important to invest efforts to obtain recombinant KKT proteins and perform microtubule co-sedimentation assays. Although tubulins purified from bovine brain are commonly used in sedimentation assays, it may be important to use trypanosome tubulins to detect binding [165,166].

The primary candidates for microtubule-binding kinetochore proteins are the KKT14 and KKT15 proteins whose localization pattern resembles that of Ndc80. It will be crucial to test these and other KKT proteins in a sedimentation assay. Once microtubule-binding proteins are identified, more detailed analyses will need to be carried out to determine whether the mechanism of microtubule-binding is distinct from conventional kinetochore proteins.

Design principles and regulations

It is also important to reveal the interaction network of KKT proteins. For example, how does the KKT10 kinase localize at kinetochores from S phase and dissociate at the metaphase–anaphase transition? What happens if KKT10 does not dissociate properly? To answer these questions, it will be necessary to reveal which proteins interact with KKT10 and how their interactions are regulated. Numerous phosphorylation sites have been identified on the 20 KKT proteins in my IP/MS analyses (B. Akiyoshi, unpublished data), so it is likely that phosphorylation plays important roles in regulating interactions as in other eukaryotes [167171]. Although it is challenging to understand the interaction network of 20 proteins, this will be key to gaining insights into the architectural design of kinetoplastid kinetochores.

Aurora B is a highly conserved protein kinase found in all sequenced eukaryotes, including kinetoplastids [172]. It is the catalytic subunit of the CPC that displays dynamic localization patterns during mitosis and regulates various mitotic functions, including kinetochore regulations [173]. As in other eukaryotes, Aurora B localizes to kinetochore regions during prometaphase/metaphase in T. brucei ([131] and my unpublished data). Therefore, its kinetochore function might be conserved in kinetoplastids. In line with this possibility, KKT7 has putative-binding motifs for the PP1 phosphatase (RSVTF conserved among kinetoplastids and SILK found in Leishmania) that is known to counteract Aurora B activities [174,175].

Aurora B is critical for the destabilization of improper microtubule attachments by phosphorylating kinetochores. It has been proposed that the extent of phosphorylation is dependent on the distance from its substrates, thereby explaining why improper tension-less attachments can be specifically destabilized [176180]. A key element in this hypothesis is the enrichment of Aurora B at the inner centromere regions, which become physically separated from bi-oriented kinetochores due to microtubule-pulling forces. Indeed, distinct space (∼1 µm) has been observed in between metaphase sister kinetochores in various eukaryotes [181], including yeasts [182], humans [183], plants [184], Cyanidioschyzon merolae [185], and Plasmodium [186]. In contrast, there is no evidence for such space in kinetoplastids. Electron microscopy studies showed that when sister kinetochore pairs interact with microtubules from opposite poles, they have a back-to-back configuration without distinct space between the two structures in T. brucei, T. cruzi, and Leishmania [123,124,187]. The analysis of YFP-tagged KKT proteins in T. brucei supports these observations [143], so Aurora B may never be significantly separated from sister kinetochores in kinetoplastids. Understanding their Aurora B regulation might provide novel insights into the mechanism of chromosome bi-orientation. Interestingly, a recent study in budding yeast showed that its chromosomes can bi-orient without the inner centromere localization of Aurora B [188]. These observations imply that there is still a lot to learn about the molecular mechanism of chromosome segregation in eukaryotes.

The spindle checkpoint is a surveillance mechanism that ensures high fidelity chromosome segregation in eukaryotes [13,189]. Although widely conserved among diverse eukaryotes [190,191], the spindle checkpoint is dispensable for the survival of yeasts, flies, and even human cells under certain conditions [192196]. Indeed, checkpoint-independent mechanisms are also important for the regulation of the APC/C activities [197201]. Interestingly, there is no evidence of functional spindle checkpoint or feedback control in kinetoplastids. For example, disruption of mitotic spindles or inhibition of DNA replication do not noticeably delay the onset of anaphase or cytokinesis in T. brucei [120,202,203]. Nonetheless, KKT4 and KKT20 co-purified with significant amounts of APC/C components [143,144], and the BRCT domain of a human microcephalin (MCPH1) protein has been shown to interact with the APC/C, at least in vitro [204]. Therefore, KKT4 and KKT20 might directly affect the activation/inactivation of the APC/C, regulating anaphase onset in a feedforward manner.

Which organisms have unconventional kinetochores?

Kinetoplastids are a widespread group of unicellular eukaryotes found in diverse environmental conditions, including hydrothermal vents [205208]. They are defined by the presence of kinetoplast, a large structure in the mitochondrion that contains the mitochondrial DNA [116]. Kinetoplastids can be divided into two subclasses, the early diverging Prokinetoplastina and the Metakinetoplastina [209]. The latter can be further divided into three orders for the bodonid species (Eubodonida, Parabodonida, and Neobodonida) and the trypanosomatids [210]. KKT proteins are found in all sequenced bodonids (Bodo saltans and Trypanoplasma borreli) [211] and trypanosomatids (e.g. T. brucei, T. cruzi, Leishmania, and Phytomonas spp.) [117119,212]. Furthermore, at least some KKT proteins are present in an early diverging Prokinetoplastina, Perkinsela sp. [213], suggesting that KKT proteins are likely a conserved feature of all kinetoplastids (my unpublished data).

Kinetoplastids belong to Euglenozoa, which is a large and diverse group of flagellates [214]. Four lineages are currently known: Kinetoplastids, Euglenids, Diplonemids, and Symbiontids [215]. Genome sequences are available only for kinetoplastids, and it was not known whether KKT proteins are present specifically in kinetoplastids or common among Euglenozoa [143]. Interestingly, examination of Euglena gracilis transcriptome [216] revealed clear homologs of conventional kinetochore proteins, including Ndc80, Nuf2, Spc25, and several candidates for CENP-A (based on the presence of longer loop 1 [217]), as well as spindle checkpoint proteins (Mad1, Mad2, and Mad3) (my unpublished data). In contrast, I failed to identify any obvious KKT proteins in this organism. Although its genome sequence will need to confirm it, these data suggest that E. gracilis likely has conventional kinetochore proteins, not KKT proteins. If true, it may be appropriate to re-classify Euglenozoa into two separate groups (see below). It will be interesting to examine whether other Euglenozoa species (e.g. Diplonema papillatum [218220] and Calkinsia aureus [221]) utilize conventional or unconventional kinetochore proteins.

Despite the long research on chromosome segregation machinery in eukaryotes, we know little about its evolutionary origins or history. For example, what kind of segregation machinery was used in the last eukaryotic common ancestor (LECA)? The notions that microtubules are used in all eukaryotes studied thus far and that eukaryotes and tubulins likely evolved from Archaea [222224] suggest that the LECA probably used tubulin-based polymers to drive chromosome segregation. In contrast, there is little trace of prokaryotic traits in kinetochore proteins, making it difficult to reveal what kind of proteins might have been used to bridge between DNA and tubulin-based polymers in this hypothetical eukaryote.

Another difficulty stems from the uncertainty of the position of the root of the eukaryotic tree of life, meaning that we still do not have concrete views about which organisms are the earliest-branching eukaryotes [214,225]. Therefore, it remains unclear whether the unique kinetoplastid kinetochore represents an ancient feature or a derived one. However, among several competing hypotheses [226228], it has been proposed that Euglenozoa (or species within Euglenozoa) may represent the earliest-branching eukaryotes [229] based on their unique mitochondrial cytochromes c/c1 and the absence of a recognizable biogenesis apparatus for these proteins [230]. The discovery of unconventional kinetochore proteins in kinetoplastids supports this hypothesis. If kinetoplastids are among the earliest-branching eukaryotes (Figure 4A), then the LECA might have possessed the KKT-based kinetochore (Figure 4Ai). In this scenario, only kinetoplastids have retained them while other eukaryotes have lost them and invented conventional kinetochores. Alternatively, it is equally possible that LECA had conventional kinetochores that have been lost in kinetoplastids (Figure 4Aii). It is also possible that LECA had a hitherto unknown type of kinetochores (Figure 4Aiii). Furthermore, the finding that Euglena has conventional kinetochore proteins raises the possibility that kinetoplastids, not Euglena, could be the earliest-branching eukaryotes, although other interpretations are also possible (Figure 4B). Genome sequences of other Euglenozoa species might provide further insights into this question.

Hypothetical models on early eukaryotic history and origin of kinetoplastid kinetochores.

Figure 4.
Hypothetical models on early eukaryotic history and origin of kinetoplastid kinetochores.

(A) The earliest-branching kinetoplastid hypothesis. In this scenario, LECA may have had KKT kinetochores (i), conventional kinetochores (ii), or unknown type of kinetochores (iii). Conventional and KKT-based kinetochores are shown in blue and red, respectively. Note that this is a highly simplified and non-comprehensive set of possibilities of early eukaryotic history and origins of kinetoplastid kinetochores. (B) The earliest-branching Euglenozoa hypothesis based on their unique cytochrome c biogenesis machinery. In this scenario, either kinetoplastids (i) or euglenids (ii) could be the earliest-branching eukaryotes.

Figure 4.
Hypothetical models on early eukaryotic history and origin of kinetoplastid kinetochores.

(A) The earliest-branching kinetoplastid hypothesis. In this scenario, LECA may have had KKT kinetochores (i), conventional kinetochores (ii), or unknown type of kinetochores (iii). Conventional and KKT-based kinetochores are shown in blue and red, respectively. Note that this is a highly simplified and non-comprehensive set of possibilities of early eukaryotic history and origins of kinetoplastid kinetochores. (B) The earliest-branching Euglenozoa hypothesis based on their unique cytochrome c biogenesis machinery. In this scenario, either kinetoplastids (i) or euglenids (ii) could be the earliest-branching eukaryotes.

Close modal

Regardless of the evolutionary origins, understanding the nature of unconventional kinetochore proteins will likely provide important insights into the fundamental principles of eukaryotic segregation machines. Furthermore, the unique kinetochore proteins represent an ideal drug target against parasitic kinetoplastids [231,232]. Understanding the molecular functions of KKT proteins should therefore contribute to developing specific drugs. Finally, it is worth mentioning that there is no organism known so far that contains both types of kinetochores. Can they coexist in a given organism? Is the conventional kinetochore system more accurate than the unconventional one? Are there as-yet different types of kinetochores to be found? Surely a lot of exciting discoveries will be made in the next few decades in the field of chromosome segregation research.

APC/C, anaphase promoting complex/cyclosome; CENP, centromere protein; ChIP, chromatin immunoprecipitation; CPC, chromosomal passenger complex; DPB, divergent polo box domain; IP, immunoprecipitation; KKT, kinetoplastid kinetochore; LECA, last eukaryotic common ancestor; MCPH1, microcephalin; MS, mass spectrometry.

I am supported by a Sir Henry Dale Fellowship jointly funded by the Wellcome Trust and the Royal Society [grant number 098403/Z/12/Z] as well as a Wellcome-Beit Prize Fellowship [grant number 098403/Z/12/A].

I thank John Archibald and Julius Lukeš for sharing unpublished Perkinsela genome data, and Ellis O'Neill and Rob Field for the predicted protein sequence dataset of Euglena gracilis. I also thank Sue Biggins and Gabriele Marcianò for comments on the manuscript.

The Author declares that there are no competing interests associated with this manuscript.

1
Nasmyth
,
K.
and
Haering
,
C.H.
(
2009
)
Cohesin: its roles and mechanisms
.
Annu. Rev. Genet.
43
,
525
558
doi:
2
McIntosh
,
J.R.
,
Molodtsov
,
M.I.
and
Ataullakhanov
,
F.I.
(
2012
)
Biophysics of mitosis
.
Q. Rev. Biophys.
45
,
147
207
doi:
3
Biggins
,
S.
(
2013
)
The composition, functions, and regulation of the budding yeast kinetochore
.
Genetics
194
,
817
846
doi:
4
Cheeseman
,
I.M.
(
2014
)
The kinetochore
.
Cold Spring Harb. Perspect. Biol.
6
,
a015826
doi:
5
Pesenti
,
M.E.
,
Weir
,
J.R.
and
Musacchio
,
A.
(
2016
)
Progress in the structural and functional characterization of kinetochores
.
Curr. Opin. Struct. Biol.
37
,
152
163
doi:
6
Allshire
,
R.C.
and
Karpen
,
G.H.
(
2008
)
Epigenetic regulation of centromeric chromatin: old dogs, new tricks?
Nat. Rev. Genet.
9
,
923
937
doi:
7
Black
,
B.E.
and
Cleveland
,
D.W.
(
2011
)
Epigenetic centromere propagation and the nature of CENP-A nucleosomes
.
Cell
144
,
471
479
doi:
8
Fukagawa
,
T.
and
Earnshaw
,
W.C.
(
2014
)
The centromere: chromatin foundation for the kinetochore machinery
.
Dev. Cell
30
,
496
508
doi:
9
Westhorpe
,
F.G.
and
Straight
,
A.F.
(
2015
)
The centromere: epigenetic control of chromosome segregation during mitosis
.
Cold Spring Harb. Perspect. Biol.
7
,
a015818
doi:
10
Desai
,
A.
and
Mitchison
,
T.J.
(
1997
)
Microtubule polymerization dynamics
.
Annu. Rev. Cell Dev. Biol.
13
,
83
117
doi:
11
Nicklas
,
R.B.
(
1997
)
How cells get the right chromosomes
.
Science
275
,
632
637
doi:
12
Murray
,
A.W.
(
2011
)
A brief history of error
.
Nat. Cell Biol.
13
,
1178
1182
doi:
13
Musacchio
,
A.
(
2015
)
The molecular biology of spindle assembly checkpoint signaling dynamics
.
Curr. Biol.
25
,
R1002
R1018
doi:
14
Pines
,
J.
(
2011
)
Cubism and the cell cycle: the many faces of the APC/C
.
Nat. Rev. Mol. Cell Biol.
12
,
427
438
doi:
15
Primorac
,
I.
and
Musacchio
,
A.
(
2013
)
Panta rhei: the APC/C at steady state
.
J. Cell Biol.
201
,
177
189
doi:
16
Chang
,
L.
and
Barford
,
D.
(
2014
)
Insights into the anaphase-promoting complex: a molecular machine that regulates mitosis
.
Curr. Opin. Struct. Biol.
29
,
1
9
doi:
17
Cheeseman
,
I.M.
and
Desai
,
A.
(
2008
)
Molecular architecture of the kinetochore-microtubule interface
.
Nat. Rev. Mol. Cell Biol.
9
,
33
46
doi:
18
Santaguida
,
S.
and
Musacchio
,
A.
(
2009
)
The life and miracles of kinetochores
.
EMBO J.
28
,
2511
2531
doi:
19
Verdaasdonk
,
J.S.
and
Bloom
,
K.
(
2011
)
Centromeres: unique chromatin structures that drive chromosome segregation
.
Nat. Rev. Mol. Cell Biol.
12
,
320
332
doi:
20
Flemming
,
W.
(
1882
)
Zellsubstanz, kern und zelltheilung
,
F.C.W. Vogel Leipzig
21
Luykx
,
P.
(
1965
)
The structure of the kinetochore in meiosis and mitosis in Urechis eggs
.
Exp. Cell Res.
39
,
643
657
doi:
22
Brinkley
,
B.R.
and
Stubblefield
,
E.
(
1966
)
The fine structure of the kinetochore of a mammalian cell in vitro
.
Chromosoma
19
,
28
43
doi:
23
Jokelainen
,
P.T.
(
1967
)
The ultrastructure and spatial organization of the metaphase kinetochore in mitotic rat cells
.
J. Ultrastruct. Res.
19
,
19
44
doi:
24
Moroi
,
Y.
,
Peebles
,
C.
,
Fritzler
,
M.J.
,
Steigerwald
,
J.
and
Tan
,
E.M.
(
1980
)
Autoantibody to centromere (kinetochore) in scleroderma sera
.
Proc. Natl Acad. Sci. USA
77
,
1627
1631
doi:
25
Guldner
,
H.H.
,
Lakomek
,
H.J.
and
Bautz
,
F.A.
(
1984
)
Human anti-centromere sera recognise a 19.5 kD non-histone chromosomal protein from HeLa cells
.
Clin. Exp. Immunol.
58
,
13
20
PMID:
[PubMed]
26
Earnshaw
,
W.C.
and
Rothfield
,
N.
(
1985
)
Identification of a family of human centromere proteins using autoimmune sera from patients with scleroderma
.
Chromosoma
91
,
313
321
doi:
27
Earnshaw
,
W.C.
,
Sullivan
,
K.F.
,
Machlin
,
P.S.
,
Cooke
,
C.A.
,
Kaiser
,
D.A.
,
Pollard
,
T.D.
et al
(
1987
)
Molecular cloning of cDNA for CENP-B, the major human centromere autoantigen
.
J. Cell Biol.
104
,
817
829
doi:
28
Saitoh
,
H.
,
Tomkiel
,
J.
,
Cooke
,
C.A.
,
Ratrie
,
H.
,
Maurer
,
M.
,
Rothfield
,
N.F.
et al
(
1992
)
CENP-C, an autoantigen in scleroderma, is a component of the human inner kinetochore plate
.
Cell
70
,
115
125
doi:
29
Sullivan
,
K.F.
,
Hechenberger
,
M.
and
Masri
,
K.
(
1994
)
Human CENP-A contains a histone H3 related histone fold domain that is required for targeting to the centromere
.
J. Cell Biol.
127
,
581
592
doi:
30
Clarke
,
L.
and
Carbon
,
J.
(
1980
)
Isolation of a yeast centromere and construction of functional small circular chromosomes
.
Nature
287
,
504
509
doi:
31
Lechner
,
J.
and
Carbon
,
J.
(
1991
)
A 240 kD multisubunit protein complex, CBF3, is a major component of the budding yeast centromere
.
Cell
64
,
717
725
doi:
32
Maine
,
G.T.
,
Sinha
,
P.
and
Tye
,
B.K.
(
1984
)
Mutants of S. cerevisiae defective in the maintenance of minichromosomes
.
Genetics
106
,
365
–385 PMID:
[PubMed]
33
Spencer
,
F.
,
Gerring
,
S.L.
,
Connelly
,
C.
and
Hieter
,
P.
(
1990
)
Mitotic chromosome transmission fidelity mutants in Saccharomyces cerevisiae
.
Genetics
124
,
237
–249 PMID:
[PubMed]
34
Ortiz
,
J.
,
Stemmann
,
O.
,
Rank
,
S.
and
Lechner
,
J.
(
1999
)
A putative protein complex consisting of Ctf19, Mcm21, and Okp1 represents a missing link in the budding yeast kinetochore
.
Genes Dev.
13
,
1140
1155
doi:
35
Hofmann
,
C.
,
Cheeseman
,
I.M.
,
Goode
,
B.L.
,
McDonald
,
K.L.
,
Barnes
,
G.
and
Drubin
,
D.G.
(
1998
)
Saccharomyces cerevisiae Duo1p and Dam1p, novel proteins involved in mitotic spindle function
.
J. Cell Biol.
143
,
1029
1040
doi:
36
Wigge
,
P.A.
and
Kilmartin
,
J.V.
(
2001
)
The Ndc80p complex from Saccharomyces cerevisiae contains conserved centromere components and has a function in chromosome segregation
.
J. Cell Biol.
152
,
349
360
doi:
37
Cheeseman
,
I.M.
,
Anderson
,
S.
,
Jwa
,
M.
,
Green
,
E.M.
,
Kang
,
J.-s.
,
Yates
, III,
J.R.
et al
(
2002
)
Phospho-regulation of kinetochore-microtubule attachments by the Aurora kinase Ipl1p
.
Cell
111
,
163
172
doi:
38
Buchwitz
,
B.J.
,
Ahmad
,
K.
,
Moore
,
L.L.
,
Roth
,
M.B.
and
Henikoff
,
S.
(
1999
)
Cell division: a histone-H3-like protein in C. elegans
.
Nature
401
,
547
548
doi:
39
Goshima
,
G.
,
Kiyomitsu
,
T.
,
Yoda
,
K.
and
Yanagida
,
M.
(
2003
)
Human centromere chromatin protein hMis12, essential for equal segregation, is independent of CENP-A loading pathway
.
J. Cell Biol.
160
,
25
39
doi:
40
Meraldi
,
P.
,
McAinsh
,
A.D.
,
Rheinbay
,
E.
and
Sorger
,
P.K.
(
2006
)
Phylogenetic and structural analysis of centromeric DNA and kinetochore proteins
.
Genome Biol.
7
,
R23
doi:
41
Schleiffer
,
A.
,
Maier
,
M.
,
Litos
,
G.
,
Lampert
,
F.
,
Hornung
,
P.
,
Mechtler
,
K.
et al
(
2012
)
CENP-T proteins are conserved centromere receptors of the Ndc80 complex
.
Nat. Cell Biol.
14
,
604
613
doi:
42
Oegema
,
K.
,
Desai
,
A.
,
Rybina
,
S.
,
Kirkham
,
M.
and
Hyman
,
A.A.
(
2001
)
Functional analysis of kinetochore assembly in Caenorhabditis elegans
.
J. Cell Biol.
153
,
1209
1226
doi:
43
Goshima
,
G.
,
Wollman
,
R.
,
Goodwin
,
S.S.
,
Zhang
,
N.
,
Scholey
,
J.M.
,
Vale
,
R.D.
et al
(
2007
)
Genes required for mitotic spindle assembly in Drosophila S2 cells
.
Science
316
,
417
421
doi:
44
Akiyoshi
,
B.
,
Nelson
,
C.R.
,
Ranish
,
J.A.
and
Biggins
,
S.
(
2009
)
Quantitative proteomic analysis of purified yeast kinetochores identifies a PP1 regulatory subunit
.
Genes Dev.
23
,
2887
2899
doi:
45
Ohta
,
S.
,
Bukowski-Wills
,
J.-C.
,
Sanchez-Pulido
,
L.
,
Alves Fde
,
L.
,
Wood
,
L.
,
Chen
,
Z.A.
et al
(
2010
)
The protein composition of mitotic chromosomes determined using multiclassifier combinatorial proteomics
.
Cell
142
,
810
821
doi:
46
Palmer
,
D.K.
,
O'Day
,
K.
,
Trong
,
H.L.
,
Charbonneau
,
H.
and
Margolis
,
R.L.
(
1991
)
Purification of the centromere-specific protein CENP-A and demonstration that it is a distinctive histone
.
Proc. Natl Acad. Sci. USA
88
,
3734
3738
doi:
47
Earnshaw
,
W.C.
and
Migeon
,
B.R.
(
1985
)
Three related centromere proteins are absent from the inactive centromere of a stable isodicentric chromosome
.
Chromosoma
92
,
290
296
doi:
48
Voullaire
,
L.E.
,
Slater
,
H.R.
,
Petrovic
,
V.
and
Choo
,
K.H.
(
1993
)
A functional marker centromere with no detectable alpha-satellite, satellite III, or CENP-B protein: activation of a latent centromere?
Am. J. Hum. Genet.
52
,
1153
–1163 PMID:
[PubMed]
49
Collins
,
K.A.
,
Castillo
,
A.R.
,
Tatsutani
,
S.Y.
and
Biggins
,
S.
(
2005
)
De novo kinetochore assembly requires the centromeric histone H3 variant
.
Mol. Biol. Cell
16
,
5649
5660
doi:
50
Liu
,
S.-T.
,
Rattner
,
J.B.
,
Jablonski
,
S.A.
and
Yen
,
T.J.
(
2006
)
Mapping the assembly pathways that specify formation of the trilaminar kinetochore plates in human cells
.
J. Cell Biol.
175
,
41
53
doi:
51
Tachiwana
,
H.
,
Kagawa
,
W.
,
Shiga
,
T.
,
Osakabe
,
A.
,
Miya
,
Y.
,
Saito
,
K.
et al
(
2011
)
Crystal structure of the human centromeric nucleosome containing CENP-A
.
Nature
476
,
232
235
doi:
52
Kato
,
H.
,
Jiang
,
J.
,
Zhou
,
B.-R.
,
Rozendaal
,
M.
,
Feng
,
H.
,
Ghirlando
,
R.
et al
(
2013
)
A conserved mechanism for centromeric nucleosome recognition by centromere protein CENP-C
.
Science
340
,
1110
1113
doi:
53
Falk
,
S.J.
,
Guo
,
L.Y.
,
Sekulic
,
N.
,
Smoak
,
E.M.
,
Mani
,
T.
,
Logsdon
,
G.A.
et al
(
2015
)
CENP-C reshapes and stabilizes CENP-A nucleosomes at the centromere
.
Science
348
,
699
703
doi:
54
Westhorpe
,
F.G.
,
Fuller
,
C.J.
and
Straight
,
A.F.
(
2015
)
A cell-free CENP-A assembly system defines the chromatin requirements for centromere maintenance
.
J. Cell Biol.
209
,
789
801
doi:
55
Sugimoto
,
K.
,
Yata
,
H.
,
Muro
,
Y.
and
Himeno
,
M.
(
1994
)
Human centromere protein C (CENP-C) is a DNA-binding protein which possesses a novel DNA-binding motif
.
J. Biochem.
116
,
877
881
PMID:
[PubMed]
56
Cohen
,
R.L.
,
Espelin
,
C.W.
,
De Wulf
,
P.
,
Sorger
,
P.K.
,
Harrison
,
S.C.
and
Simons
,
K.T.
(
2008
)
Structural and functional dissection of Mif2p, a conserved DNA-binding kinetochore protein
.
Mol. Biol. Cell
19
,
4480
4491
doi:
57
Hornung
,
P.
,
Troc
,
P.
,
Malvezzi
,
F.
,
Maier
,
M.
,
Demianova
,
Z.
,
Zimniak
,
T.
et al
(
2014
)
A cooperative mechanism drives budding yeast kinetochore assembly downstream of CENP-A
.
J. Cell Biol.
206
,
509
524
doi:
58
Hori
,
T.
,
Amano
,
M.
,
Suzuki
,
A.
,
Backer
,
C.B.
,
Welburn
,
J.P.
,
Dong
,
Y.
et al
(
2008
)
CCAN makes multiple contacts with centromeric DNA to provide distinct pathways to the outer kinetochore
.
Cell
135
,
1039
1052
doi:
59
Nishino
,
T.
,
Takeuchi
,
K.
,
Gascoigne
,
K.E.
,
Suzuki
,
A.
,
Hori
,
T.
,
Oyama
,
T.
et al
(
2012
)
CENP-T-W-S-X forms a unique centromeric chromatin structure with a histone-like fold
.
Cell
148
,
487
501
doi:
60
Takeuchi
,
K.
,
Nishino
,
T.
,
Mayanagi
,
K.
,
Horikoshi
,
N.
,
Osakabe
,
A.
,
Tachiwana
,
H.
et al
(
2014
)
The centromeric nucleosome-like CENP-T-W-S-X complex induces positive supercoils into DNA
.
Nucleic Acids Res.
42
,
1644
1655
doi:
61
Aravind
,
L.
and
Landsman
,
D.
(
1998
)
AT-hook motifs identified in a wide variety of DNA-binding proteins
.
Nucleic Acids Res.
26
,
4413
4421
doi:
62
Westermann
,
S.
and
Schleiffer
,
A.
(
2013
)
Family matters: structural and functional conservation of centromere-associated proteins from yeast to humans
.
Trends Cell Biol.
23
,
260
269
doi:
63
Suzuki
,
M.
(
1989
)
SPKK, a new nucleic acid-binding unit of protein found in histone
.
EMBO J.
8
,
797
804
PMID:
[PubMed]
64
Malik
,
H.S.
,
Vermaak
,
D.
and
Henikoff
,
S.
(
2002
)
Recurrent evolution of DNA-binding motifs in the Drosophila centromeric histone
.
Proc. Natl Acad. Sci. USA
99
,
1449
1454
doi:
65
Chen
,
Y.
,
Riley
,
D.J.
,
Chen
,
P.L.
and
Lee
,
W.H.
(
1997
)
HEC, a novel nuclear protein rich in leucine heptad repeats specifically involved in mitosis
.
Mol. Cell. Biol.
17
,
6049
6056
doi:
66
Wigge
,
P.A.
,
Jensen
,
O.N.
,
Holmes
,
S.
,
Souès
,
S.
,
Mann
,
M.
and
Kilmartin
,
J.V.
(
1998
)
Analysis of the Saccharomyces spindle pole by matrix-assisted laser desorption/ionization (MALDI) mass spectrometry
.
J. Cell Biol.
141
,
967
977
doi:
67
Ciferri
,
C.
,
De Luca
,
J.
,
Monzani
,
S.
,
Ferrari
,
K.J.
,
Ristic
,
D.
,
Wyman
,
C.
et al
(
2005
)
Architecture of the human ndc80-hec1 complex, a critical constituent of the outer kinetochore
.
J. Biol. Chem.
280
,
29088
29095
doi:
68
Wei
,
R.R.
,
Sorger
,
P.K.
and
Harrison
,
S.C.
(
2005
)
Molecular organization of the Ndc80 complex, an essential kinetochore component
.
Proc. Natl Acad. Sci. USA
102
,
5363
5367
doi:
69
McCleland
,
M.L.
,
Kallio
,
M.J.
,
Barrett-Wilt
,
G.A.
,
Kestner
,
C.A.
,
Shabanowitz
,
J.
,
Hunt
,
D.F.
et al
(
2004
)
The vertebrate Ndc80 complex contains Spc24 and Spc25 homologs, which are required to establish and maintain kinetochore-microtubule attachment
.
Curr. Biol.
14
,
131
137
doi:
70
DeLuca
,
J.G.
,
Dong
,
Y.
,
Hergert
,
P.
,
Strauss
,
J.
,
Hickey
,
J.M.
,
Salmon
,
E.D.
et al
(
2005
)
Hec1 and nuf2 are core components of the kinetochore outer plate essential for organizing microtubule attachment sites
.
Mol. Biol. Cell
16
,
519
531
doi:
71
Cheeseman
,
I.M.
,
Chappie
,
J.S.
,
Wilson-Kubalek
,
E.M.
and
Desai
,
A.
(
2006
)
The conserved KMN network constitutes the core microtubule-binding site of the kinetochore
.
Cell
127
,
983
997
doi:
72
Wei
,
R.R.
,
Al-Bassam
,
J.
and
Harrison
,
S.C.
(
2007
)
The Ndc80/HEC1 complex is a contact point for kinetochore-microtubule attachment
.
Nat. Struct. Mol. Biol.
14
,
54
59
doi:
73
Ciferri
,
C.
,
Pasqualato
,
S.
,
Screpanti
,
E.
,
Varetti
,
G.
,
Santaguida
,
S.
,
Dos Reis
,
G.
et al
(
2008
)
Implications for kinetochore-microtubule attachment from the structure of an engineered Ndc80 complex
.
Cell
133
,
427
439
doi:
74
Powers
,
A.F.
,
Franck
,
A.D.
,
Gestaut
,
D.R.
,
Cooper
,
J.
,
Gracyzk
,
B.
,
Wei
,
R.R.
et al
(
2009
)
The Ndc80 kinetochore complex forms load-bearing attachments to dynamic microtubule tips via biased diffusion
.
Cell
136
,
865
875
doi:
75
Alushin
,
G.M.
,
Ramey
,
V.H.
,
Pasqualato
,
S.
,
Ball
,
D.A.
,
Grigorieff
,
N.
,
Musacchio
,
A.
et al
(
2010
)
The Ndc80 kinetochore complex forms oligomeric arrays along microtubules
.
Nature
467
,
805
810
doi:
76
Sarangapani
,
K.K.
,
Akiyoshi
,
B.
,
Duggan
,
N.M.
,
Biggins
,
S.
and
Asbury
,
C.L.
(
2013
)
Phosphoregulation promotes release of kinetochores from dynamic microtubules via multiple mechanisms
.
Proc. Natl Acad. Sci. USA
110
,
7282
–7287 PMID:
[PubMed]
77
Ji
,
Z.
,
Gao
,
H.
and
Yu
,
H.
(
2015
)
Kinetochore attachment sensed by competitive Mps1 and microtubule binding to Ndc80C
.
Science
348
,
1260
1264
doi:
78
Hiruma
,
Y.
,
Sacristan
,
C.
,
Pachis
,
S.T.
,
Adamopoulos
,
A.
,
Kuijt
,
T.
,
Ubbink
,
M.
et al
(
2015
)
Competition between MPS1 and microtubules at kinetochores regulates spindle checkpoint signaling
.
Science
348
,
1264
1267
doi:
79
Dong
,
Y.
,
Vanden Beldt
,
K.J.
,
Meng
,
X.
,
Khodjakov
,
A.
and
McEwen
,
B.F.
(
2007
)
The outer plate in vertebrate kinetochores is a flexible network with multiple microtubule interactions
.
Nat. Cell Biol.
9
,
516
522
doi:
80
McIntosh
,
J.R.
,
Grishchuk
,
E.L.
,
Morphew
,
M.K.
,
Efremov
,
A.K.
,
Zhudenkov
,
K.
,
Volkov
,
V.A.
et al
(
2008
)
Fibrils connect microtubule tips with kinetochores: a mechanism to couple tubulin dynamics to chromosome motion
.
Cell
135
,
322
333
doi:
81
Gonen
,
S.
,
Akiyoshi
,
B.
,
Iadanza
,
M.G.
,
Shi
,
D.
,
Duggan
,
N.
,
Biggins
,
S.
et al
(
2012
)
The structure of purified kinetochores reveals multiple microtubule-attachment sites
.
Nat. Struct. Mol. Biol.
19
,
925
929
doi:
82
De Wulf
,
P.
,
McAinsh
,
A.D.
and
Sorger
,
P.K.
(
2003
)
Hierarchical assembly of the budding yeast kinetochore from multiple subcomplexes
.
Genes Dev.
17
,
2902
2921
doi:
83
Pinsky
,
B.A.
,
Tatsutani
,
S.Y.
,
Collins
,
K.A.
and
Biggins
,
S.
(
2003
)
An Mtw1 complex promotes kinetochore biorientation that is monitored by the Ipl1/Aurora protein kinase
.
Dev. Cell
5
,
735
745
doi:
84
Obuse
,
C.
,
Iwasaki
,
O.
,
Kiyomitsu
,
T.
,
Goshima
,
G.
,
Toyoda
,
Y.
and
Yanagida
,
M.
(
2004
)
A conserved Mis12 centromere complex is linked to heterochromatic HP1 and outer kinetochore protein Zwint-1
.
Nat. Cell Biol.
6
,
1135
1141
doi:
85
Foltz
,
D.R.
,
Jansen
,
L.E.T.
,
Black
,
B.E.
,
Bailey
,
A.O.
,
Yates
,
J.R.
and
Cleveland
,
D.W.
(
2006
)
The human CENP-A centromeric nucleosome-associated complex
.
Nat. Cell Biol.
8
,
458
469
doi:
86
Okada
,
M.
,
Cheeseman
,
I.M.
,
Hori
,
T.
,
Okawa
,
K.
,
McLeod
,
I.X.
,
Yates
, III,
J.R.
et al
(
2006
)
The CENP-H-I complex is required for the efficient incorporation of newly synthesized CENP-A into centromeres
.
Nat. Cell Biol.
8
,
446
457
doi:
87
Carroll
,
C.W.
,
Milks
,
K.J.
and
Straight
,
A.F.
(
2010
)
Dual recognition of CENP-A nucleosomes is required for centromere assembly
.
J. Cell Biol.
189
,
1143
1155
doi:
88
Przewloka
,
M.R.
,
Venkei
,
Z.
,
Bolanos-Garcia
,
V.M.
,
Debski
,
J.
,
Dadlez
,
M.
and
Glover
,
D.M.
(
2011
)
CENP-C is a structural platform for kinetochore assembly
.
Curr. Biol.
21
,
399
405
doi:
89
Screpanti
,
E.
,
De Antoni
,
A.
,
Alushin
,
G.M.
,
Petrovic
,
A.
,
Melis
,
T.
,
Nogales
,
E.
et al
(
2011
)
Direct binding of Cenp-C to the Mis12 complex joins the inner and outer kinetochore
.
Curr. Biol.
21
,
391
398
doi:
90
Klare
,
K.
,
Weir
,
J.R.
,
Basilico
,
F.
,
Zimniak
,
T.
,
Massimiliano
,
L.
,
Ludwigs
,
N.
et al
(
2015
)
CENP-C is a blueprint for constitutive centromere-associated network assembly within human kinetochores
.
J. Cell Biol.
210
,
11
22
doi:
91
McKinley
,
K.L.
,
Sekulic
,
N.
,
Guo
,
L.Y.
,
Tsinman
,
T.
,
Black
,
B.E.
and
Cheeseman
,
I.M.
(
2015
)
The CENP-L-N complex forms a critical node in an integrated meshwork of interactions at the centromere-kinetochore interface
.
Mol. Cell
60
,
886
898
doi:
92
Petrovic
,
A.
,
Mosalaganti
,
S.
,
Keller
,
J.
,
Mattiuzzo
,
M.
,
Overlack
,
K.
,
Krenn
,
V.
et al
(
2014
)
Modular assembly of RWD domains on the Mis12 complex underlies outer kinetochore organization
.
Mol. Cell
53
,
591
605
doi:
93
Joglekar
,
A.P.
,
Bloom
,
K.
and
Salmon
,
E.D.
(
2009
)
In vivo protein architecture of the eukaryotic kinetochore with nanometer scale accuracy
.
Curr. Biol.
19
,
694
699
doi:
94
Wan
,
X.
,
O'Quinn
,
R.P.
,
Pierce
,
H.L.
,
Joglekar
,
A.P.
,
Gall
,
W.E.
,
DeLuca
,
J.G.
et al
(
2009
)
Protein architecture of the human kinetochore microtubule attachment site
.
Cell
137
,
672
684
doi:
95
Magidson
,
V.
,
He
,
J.
,
Ault
,
J.G.
,
O'Connell
,
C.B.
,
Yang
,
N.
,
Tikhonenko
,
I.
et al
(
2016
)
Unattached kinetochores rather than intrakinetochore tension arrest mitosis in taxol-treated cells
.
J. Cell Biol.
212
,
307
319
doi:
96
Schmitzberger
,
F.
and
Harrison
,
S.C.
(
2012
)
RWD domain: a recurring module in kinetochore architecture shown by a Ctf19–Mcm21 complex structure
.
EMBO Rep.
13
,
216
222
doi:
97
Kim
,
S.
,
Sun
,
H.
,
Tomchick
,
D.R.
,
Yu
,
H.
and
Luo
,
X.
(
2012
)
Structure of human Mad1 C-terminal domain reveals its involvement in kinetochore targeting
.
Proc. Natl Acad. Sci. USA
109
,
6549
6554
doi:
98
Bolanos-Garcia
,
V.M.
,
Kiyomitsu
,
T.
,
D'Arcy
,
S.
,
Chirgadze
,
D.Y.
,
Grossmann
,
J.G.
,
Matak-Vinkovic
,
D.
et al
(
2009
)
The crystal structure of the N-terminal region of BUB1 provides insight into the mechanism of BUB1 recruitment to kinetochores
.
Structure
17
,
105
116
doi:
99
Nijenhuis
,
W.
,
von Castelmur
,
E.
,
Littler
,
D.
,
De Marco
,
V.
,
Tromer
,
E.
,
Vleugel
,
M.
et al
(
2013
)
A TPR domain-containing N-terminal module of MPS1 is required for its kinetochore localization by Aurora B
.
J. Cell Biol.
201
,
217
231
doi:
100
Guo
,
Q.
,
Tao
,
Y.
,
Liu
,
H.
,
Teng
,
M.
and
Li
,
X.
(
2013
)
Structural insights into the role of the Chl4-Iml3 complex in kinetochore assembly
.
Acta Crystallogr. D Biol. Crystallogr.
69
,
2412
2419
doi:
101
Hinshaw
,
S.M.
and
Harrison
,
S.C.
(
2013
)
An Iml3-Chl4 heterodimer links the core centromere to factors required for accurate chromosome segregation
.
Cell Rep.
5
,
29
36
doi:
102
Basilico
,
F.
,
Maffini
,
S.
,
Weir
,
J.R.
,
Prumbaum
,
D.
,
Rojas
,
A.M.
,
Zimniak
,
T.
et al
(
2014
)
The pseudo GTPase CENP-M drives human kinetochore assembly
.
eLife
3
,
e02978
doi:
103
Cho
,
U.-S.
and
Harrison
,
S.C.
(
2012
)
Ndc10 is a platform for inner kinetochore assembly in budding yeast
.
Nat. Struct. Mol. Biol.
19
,
48
55
doi:
104
Perriches
,
T.
and
Singleton
,
M.R.
(
2012
)
Structure of yeast kinetochore Ndc10 DNA-binding domain reveals unexpected evolutionary relationship to tyrosine recombinases
.
J. Biol. Chem.
287
,
5173
5179
doi:
105
Ohno
,
S.
(
1970
)
Evolution by gene duplication
,
Springer-Verlag
,
New York
106
Wickstead
,
B.
and
Gull
,
K.
(
2011
)
The evolution of the cytoskeleton
.
J. Cell Biol.
194
,
513
525
doi:
107
Drechsler
,
H.
and
McAinsh
,
A.D.
(
2012
)
Exotic mitotic mechanisms
.
Open Biol.
2
,
120140
doi:
108
Findeisen
,
P.
,
Mühlhausen
,
S.
,
Dempewolf
,
S.
,
Hertzog
,
J.
,
Zietlow
,
A.
,
Carlomagno
,
T.
et al
(
2014
)
Six subgroups and extensive recent duplications characterize the evolution of the eukaryotic tubulin protein family
.
Genome Biol. Evol.
6
,
2274
2288
doi:
109
Henikoff
,
S.
,
Ahmad
,
K.
and
Malik
,
H.S.
(
2001
)
The centromere paradox: stable inheritance with rapidly evolving DNA
.
Science
293
,
1098
1102
doi:
110
Drinnenberg
,
I.A.
,
deYoung
,
D.
,
Henikoff
,
S.
and
Malik
,
H.S.
(
2014
)
Recurrent loss of CenH3 is associated with independent transitions to holocentricity in insects
.
eLife
3
,
e03676
doi:
111
Cheeseman
,
I.M.
,
Niessen
,
S.
,
Anderson
,
S.
,
Hyndman
,
F.
,
Yates
, III,
J.R.
,
Oegema
,
K.
et al
(
2004
)
A conserved protein network controls assembly of the outer kinetochore and its ability to sustain tension
.
Genes Dev.
18
,
2255
2268
doi:
112
Przewloka
,
M.R.
,
Zhang
,
W.
,
Costa
,
P.
,
Archambault
,
V.
,
D'Avino
,
P.P.
,
Lilley
,
K.S.
et al
(
2007
)
Molecular analysis of core kinetochore composition and assembly in Drosophila melanogaster
.
PLoS ONE
2
,
e478
doi:
113
Liu
,
Y.
,
Petrovic
,
A.
,
Rombaut
,
P.
,
Mosalaganti
,
S.
,
Keller
,
J.
,
Raunser
,
S.
et al
(
2016
)
Insights from the reconstitution of the divergent outer kinetochore of Drosophila melanogaster
.
Open Biol.
6
,
150236
doi:
114
Richter
,
M.M.
,
Poznanski
,
J.
,
Zdziarska
,
A.
,
Czarnocki-Cieciura
,
M.
,
Lipinszki
,
Z.
,
Dadlez
,
M.
et al
(
2016
)
Network of protein interactions within the Drosophila inner kinetochore
.
Open Biol.
6
,
150238
doi:
115
Drinnenberg
,
I.A.
,
Henikoff
,
S.
and
Malik
,
H.S.
(
2016
)
Evolutionary turnover of kinetochore proteins: a Ship of Theseus?
Trends Cell Biol.
26
,
498
510
doi:
116
Vickerman
,
K.
(
1962
)
The mechanism of cyclical development in trypanosomes of the Trypanosoma brucei sub-group: an hypothesis based on ultrastructural observations
.
Trans. R. Soc. Trop. Med. Hyg.
56
,
487
495
doi:
117
Berriman
,
M.
,
Ghedin
,
E.
,
Hertz-Fowler
,
C.
,
Blandin
,
G.
,
Renauld
,
H.
,
Bartholomeu
,
D.C.
et al
(
2005
)
The genome of the African trypanosome Trypanosoma brucei
.
Science
309
,
416
422
doi:
118
El-Sayed
,
N.M.
,
Myler
,
P.J.
,
Bartholomeu
,
D.C.
,
Nilsson
,
D.
,
Aggarwal
,
G.
,
Tran
,
A.-N.
et al
(
2005
)
The genome sequence of Trypanosoma cruzi, etiologic agent of Chagas disease
.
Science
309
,
409
415
doi:
119
Ivens
,
A.C.
,
Peacock
,
C.S.
,
Worthey
,
E.A.
,
Murphy
,
L.
,
Aggarwal
,
G.
,
Berriman
,
M.
et al
(
2005
)
The genome of the kinetoplastid parasite, Leishmania major
.
Science
309
,
436
442
doi:
120
Akiyoshi
,
B.
and
Gull
,
K.
(
2013
)
Evolutionary cell biology of chromosome segregation: insights from trypanosomes
.
Open Biol.
3
,
130023
doi:
121
Ersfeld
,
K.
and
Gull
,
K.
(
1997
)
Partitioning of large and minichromosomes in Trypanosoma brucei
.
Science
276
,
611
614
doi:
122
Vickerman
,
K.
and
Preston
,
T.M.
(
1970
)
Spindle microtubules in the dividing nuclei of trypanosomes
.
J. Cell. Sci.
6
,
365
383
PMID:
[PubMed]
123
Solari
,
A.J.
(
1980
)
The 3-dimensional fine structure of the mitotic spindle in Trypanosoma cruzi
.
Chromosoma
78
,
239
255
doi:
124
Ogbadoyi
,
E.
,
Ersfeld
,
K.
,
Robinson
,
D.
,
Sherwin
,
T.
and
Gull
,
K.
(
2000
)
Architecture of the Trypanosoma brucei nucleus during interphase and mitosis
.
Chromosoma
108
,
501
513
doi:
125
Kim
,
J.
,
Ishiguro
,
K.
,
Nambu
,
A.
,
Akiyoshi
,
B.
,
Yokobayashi
,
S.
,
Kagami
,
A.
et al
(
2015
)
Meikin is a conserved regulator of meiosis-I-specific kinetochore function
.
Nature
517
,
466
471
doi:
126
Akiyoshi
,
B.
,
Sarangapani
,
K.K.
,
Powers
,
A.F.
,
Nelson
,
C.R.
,
Reichow
,
S.L.
,
Arellano-Santoyo
,
H.
et al
(
2010
)
Tension directly stabilizes reconstituted kinetochore-microtubule attachments
.
Nature
468
,
576
579
doi:
127
Akiyoshi
,
B.
and
Biggins
,
S.
(
2012
)
Reconstituting the kinetochore–microtubule interface: what, why, and how
.
Chromosoma
121
,
235
250
doi:
128
Lowell
,
J.E.
and
Cross
,
G.A.M.
(
2004
)
A variant histone H3 is enriched at telomeres in Trypanosoma brucei
.
J. Cell Sci.
117
,
5937
5947
doi:
129
Talbert
,
P.B.
,
Bayes
,
J.J.
and
Henikoff
,
S.
(
2009
) Evolution of centromeres and kinetochores: a two-part fugue. In
The kinetochore
(
De Wulf
,
P.
and
Earnshaw
,
W.C.
, eds
), pp.
1
37
,
Springer
,
New York
130
Kiyomitsu
,
T.
,
Obuse
,
C.
and
Yanagida
,
M.
(
2007
)
Human Blinkin/AF15q14 is required for chromosome alignment and the mitotic checkpoint through direct interaction with Bub1 and BubR1
.
Dev. Cell
13
,
663
676
doi:
131
Li
,
Z.
,
Lee
,
J.H.
,
Chu
,
F.
,
Burlingame
,
A.L.
,
Günzl
,
A.
and
Wang
,
C.C.
(
2008
)
Identification of a novel chromosomal passenger complex and its unique localization during cytokinesis in Trypanosoma brucei
.
PLoS ONE
3
,
e2354
doi:
132
Herzog
,
F.
,
Primorac
,
I.
,
Dube
,
P.
,
Lenart
,
P.
,
Sander
,
B.
,
Mechtler
,
K.
et al
(
2009
)
Structure of the anaphase-promoting complex/cyclosome interacting with a mitotic checkpoint complex
.
Science
323
,
1477
1481
doi:
133
Hsu
,
K.-S.
and
Toda
,
T.
(
2011
)
Ndc80 internal loop interacts with Dis1/TOG to ensure proper kinetochore-spindle attachment in fission yeast
.
Curr. Biol.
21
,
214
220
doi:
134
Miller
,
M.P.
,
Asbury
,
C.L.
and
Biggins
,
S.
(
2016
)
A TOG protein confers tension sensitivity to kinetochore-microtubule attachments
.
Cell
165
,
1428
1439
doi:
135
Dean
,
S.
,
Sunter
,
J.
,
Wheeler
,
R.J.
,
Hodkinson
,
I.
,
Gluenz
,
E.
and
Gull
,
K.
(
2015
)
A toolkit enabling efficient, scalable and reproducible gene tagging in trypanosomatids
.
Open Biol.
5
,
140197
doi:
136
Kelly
,
S.
,
Reed
,
J.
,
Kramer
,
S.
,
Ellis
,
L.
,
Webb
,
H.
,
Sunter
,
J.
et al
(
2007
)
Functional genomics in Trypanosoma brucei: a collection of vectors for the expression of tagged proteins from endogenous and ectopic gene loci
.
Mol. Biochem. Parasitol.
154
,
103
109
doi:
137
Archer
,
S.K.
,
Inchaustegui
,
D.
,
Queiroz
,
R.
and
Clayton
,
C.
(
2011
)
The cell cycle regulated transcriptome of Trypanosoma brucei
.
PLoS ONE
6
,
e18425
doi:
138
Brenner
,
S.
,
Pepper
,
D.
,
Berns
,
M.W.
,
Tan
,
E.
and
Brinkley
,
B.R.
(
1981
)
Kinetochore structure, duplication, and distribution in mammalian cells: analysis by human autoantibodies from scleroderma patients
.
J. Cell Biol.
91
,
95
102
doi:
139
Meluh
,
P.B.
and
Koshland
,
D.
(
1997
)
Budding yeast centromere composition and assembly as revealed by in vivo cross-linking
.
Genes Dev.
11
,
3401
3412
doi:
140
Hayashi
,
A.
,
Ogawa
,
H.
,
Kohno
,
K.
,
Gasser
,
S.M.
and
Hiraoka
,
Y.
(
1998
)
Meiotic behaviours of chromosomes and microtubules in budding yeast: relocalization of centromeres and telomeres during meiotic prophase
.
Genes Cells
3
,
587
601
doi:
141
Obado
,
S.O.
,
Bot
,
C.
,
Nilsson
,
D.
,
Andersson
,
B.
and
Kelly
,
J.M.
(
2007
)
Repetitive DNA is associated with centromeric domains in Trypanosoma brucei but not Trypanosoma cruzi
.
Genome Biol.
8
,
R37
doi:
142
Earnshaw
,
W.C.
(
2015
)
Discovering centromere proteins: from cold White hands to the A, B, C of CENPs
.
Nat. Rev. Mol. Cell Biol.
16
,
443
449
doi:
143
Akiyoshi
,
B.
and
Gull
,
K.
(
2014
)
Discovery of unconventional kinetochores in kinetoplastids
.
Cell
156
,
1247
1258
doi:
144
Nerusheva
,
O.O.
and
Akiyoshi
,
B.
(
2016
)
Divergent polo box domains underpin the unique kinetoplastid kinetochore
.
Open Biol.
6
,
150206
doi:
145
Morriswood
,
B.
,
Havlicek
,
K.
,
Demmel
,
L.
,
Yavuz
,
S.
,
Sealey-Cardona
,
M.
,
Vidilaseris
,
K.
et al
(
2013
)
Novel bilobe components in Trypanosoma brucei identified using proximity-dependent biotinylation
.
Eukaryotic Cell
12
,
356
367
doi:
146
Reinhardt
,
H.C.
and
Yaffe
,
M.B.
(
2013
)
Phospho-Ser/Thr-binding domains: navigating the cell cycle and DNA damage response
.
Nat. Rev. Mol. Cell Biol.
14
,
563
580
doi:
147
Parsons
,
M.
,
Worthey
,
E.A.
,
Ward
,
P.N.
and
Mottram
,
J.C.
(
2005
)
Comparative analysis of the kinomes of three pathogenic trypanosomatids: Leishmania major, Trypanosoma brucei and Trypanosoma cruzi
.
BMC Genomics
6
,
127
doi:
148
Katoh
,
K.
and
Standley
,
D.M.
(
2013
)
MAFFT multiple sequence alignment software version 7: improvements in performance and usability
.
Mol. Biol. Evol.
30
,
772
780
doi:
149
Lupas
,
A.
,
Van Dyke
,
M.
and
Stock
,
J.
(
1991
)
Predicting coiled coils from protein sequences
.
Science
252
,
1162
1164
doi:
150
Bryson
,
K.
,
McGuffin
,
L.J.
,
Marsden
,
R.L.
,
Ward
,
J.J.
,
Sodhi
,
J.S.
and
Jones
,
D.T.
(
2005
)
Protein structure prediction servers at University College London
.
Nucleic Acids Res.
33
,
W36
W38
doi:
151
Bailey
,
T.L.
,
Boden
,
M.
,
Buske
,
F.A.
,
Frith
,
M.
,
Grant
,
C.E.
,
Clementi
,
L.
et al
(
2009
)
MEME suite: tools for motif discovery and searching
.
Nucl. Acids Res.
doi:
152
Finn
,
R.D.
,
Coggill
,
P.
,
Eberhardt
,
R.Y.
,
Eddy
,
S.R.
,
Mistry
,
J.
,
Mitchell
,
A.L.
et al
(
2016
)
The Pfam protein families database: towards a more sustainable future
.
Nucleic Acids Res.
44
,
D279
D285
doi:
153
Buchan
,
D.W.A.
,
Minneci
,
F.
,
Nugent
,
T.C.O.
,
Bryson
,
K.
and
Jones
,
D.T.
(
2013
)
Scalable web services for the PSIPRED Protein Analysis Workbench
.
Nucleic Acids Res.
41
,
W349
W357
doi:
154
Drozdetskiy
,
A.
,
Cole
,
C.
,
Procter
,
J.
and
Barton
,
G.J.
(
2015
)
JPred4: a protein secondary structure prediction server
.
Nucl. Acids Res.
43
,
W389
W394
doi:
155
Söding
,
J.
,
Biegert
,
A.
and
Lupas
,
A.N.
(
2005
)
The HHpred interactive server for protein homology detection and structure prediction
.
Nucleic Acids Res.
33
,
W244
W248
doi:
156
Yang
,
C.H.
,
Tomkiel
,
J.
,
Saitoh
,
H.
,
Johnson
,
D.H.
and
Earnshaw
,
W.C.
(
1996
)
Identification of overlapping DNA-binding and centromere-targeting domains in the human kinetochore protein CENP-C
.
Mol. Cell. Biol.
16
,
3576
3586
doi:
157
Milks
,
K.J.
,
Moree
,
B.
and
Straight
,
A.F.
(
2009
)
Dissection of CENP-C-directed centromere and kinetochore assembly
.
Mol. Biol. Cell
20
,
4246
4255
doi:
158
Jones
,
N.G.
,
Thomas
,
E.B.
,
Brown
,
E.
,
Dickens
,
N.J.
,
Hammarton
,
T.C.
and
Mottram
,
J.C.
(
2014
)
Regulators of Trypanosoma brucei cell cycle progression and differentiation identified using a kinome-wide RNAi screen
.
PLoS Pathog.
10
,
e1003886
doi:
159
Ngô
,
H.
,
Tschudi
,
C.
,
Gull
,
K.
and
Ullu
,
E.
(
1998
)
Double-stranded RNA induces mRNA degradation in Trypanosoma brucei
.
Proc. Natl Acad. Sci. USA
95
,
14687
14692
doi:
160
Wang
,
Z.
,
Morris
,
J.C.
,
Drew
,
M.E.
and
Englund
,
P.T.
(
2000
)
Inhibition of Trypanosoma brucei gene expression by RNA interference using an integratable vector with opposing T7 promoters
.
J. Biol. Chem.
275
,
40174
40179
doi:
161
Merritt
,
C.
and
Stuart
,
K.
(
2013
)
Identification of essential and non-essential protein kinases by a fusion PCR method for efficient production of transgenic Trypanosoma brucei
.
Mol. Biochem. Parasitol.
190
,
44
49
doi:
162
Kim
,
H.-S.
,
Li
,
Z.
,
Boothroyd
,
C.
and
Cross
,
G.A.M.
(
2013
)
Strategies to construct null and conditional null Trypanosoma brucei mutants using Cre-recombinase and loxP
.
Mol. Biochem. Parasitol.
191
,
16
19
doi:
163
Nishimura
,
K.
,
Fukagawa
,
T.
,
Takisawa
,
H.
,
Kakimoto
,
T.
and
Kanemaki
,
M.
(
2009
)
An auxin-based degron system for the rapid depletion of proteins in nonplant cells
.
Nat. Meth.
6
,
917
922
doi:
164
Siegel
,
T.N.
,
Hekstra
,
D.R.
,
Kemp
,
L.E.
,
Figueiredo
,
L.M.
,
Lowell
,
J.E.
,
Fenyo
,
D.
et al
(
2009
)
Four histone variants mark the boundaries of polycistronic transcription units in Trypanosoma brucei
.
Genes Dev.
23
,
1063
1076
doi:
165
MacRae
,
T.H.
and
Gull
,
K.
(
1990
)
Purification and assembly in vitro of tubulin from Trypanosoma brucei brucei
.
Biochem. J.
265
,
87
93
doi:
166
Widlund
,
P.O.
,
Podolski
,
M.
,
Reber
,
S.
,
Alper
,
J.
,
Storch
,
M.
,
Hyman
,
A.A.
et al
(
2012
)
One-step purification of assembly-competent tubulin from diverse eukaryotic sources
.
Mol. Biol. Cell
23
,
4393
4401
doi:
167
Gascoigne
,
K.E.
,
Takeuchi
,
K.
,
Suzuki
,
A.
,
Hori
,
T.
,
Fukagawa
,
T.
and
Cheeseman
,
I.M.
(
2011
)
Induced ectopic kinetochore assembly bypasses the requirement for CENP-A nucleosomes
.
Cell
145
,
410
422
doi:
168
Akiyoshi
,
B.
,
Nelson
,
C.R.
and
Biggins
,
S.
(
2013
)
The Aurora B kinase promotes inner and outer kinetochore interactions in budding yeast
.
Genetics
194
,
785
789
doi:
169
Gascoigne
,
K.E.
and
Cheeseman
,
I.M.
(
2013
)
CDK-dependent phosphorylation and nuclear exclusion coordinately control kinetochore assembly state
.
J. Cell Biol.
201
,
23
32
doi:
170
Kim
,
S.
and
Yu
,
H.
(
2015
)
Multiple assembly mechanisms anchor the KMN spindle checkpoint platform at human mitotic kinetochores
.
J. Cell Biol.
208
,
181
196
doi:
171
Rago
,
F.
,
Gascoigne
,
K.E.
and
Cheeseman
,
I.M.
(
2015
)
Distinct organization and regulation of the outer kinetochore KMN network downstream of CENP-C and CENP-T
.
Curr. Biol.
25
,
671
677
doi:
172
Hochegger
,
H.
,
Hégarat
,
N.
and
Pereira-Leal
,
J.B.
(
2013
)
Aurora at the pole and equator: overlapping functions of Aurora kinases in the mitotic spindle
.
Open Biol.
3
,
120185
doi:
173
Carmena
,
M.
,
Wheelock
,
M.
,
Funabiki
,
H.
and
Earnshaw
,
W.C.
(
2012
)
The chromosomal passenger complex (CPC): from easy rider to the godfather of mitosis
.
Nat. Rev. Mol. Cell Biol.
13
,
789
803
doi:
174
Francisco
,
L.
,
Wang
,
W.
and
Chan
,
C.S.
(
1994
)
Type 1 protein phosphatase acts in opposition to Ipl1 protein kinase in regulating yeast chromosome segregation
.
Mol. Cell. Biol.
14
,
4731
4740
doi:
175
Liu
,
D.
,
Vleugel
,
M.
,
Backer
,
C.B.
,
Hori
,
T.
,
Fukagawa
,
T.
,
Cheeseman
,
I.M.
et al
(
2010
)
Regulated targeting of protein phosphatase 1 to the outer kinetochore by KNL1 opposes Aurora B kinase
.
J. Cell Biol.
188
,
809
820
doi:
176
Tanaka
,
T.U.
,
Rachidi
,
N.
,
Janke
,
C.
,
Pereira
,
G.
,
Galova
,
M.
,
Schiebel
,
E.
et al
(
2002
)
Evidence that the Ipl1-Sli15 (Aurora kinase-INCENP) complex promotes chromosome bi-orientation by altering kinetochore-spindle pole connections
.
Cell
108
,
317
329
doi:
177
Andrews
,
P.D.
,
Ovechkina
,
Y.
,
Morrice
,
N.
,
Wagenbach
,
M.
,
Duncan
,
K.
,
Wordeman
,
L.
et al
(
2004
)
Aurora B regulates MCAK at the mitotic centromere
.
Dev. Cell
6
,
253
268
doi:
178
Watanabe
,
Y.
(
2010
)
Temporal and spatial regulation of targeting aurora B to the inner centromere
.
Cold Spring Harb. Symp. Quant. Biol.
75
,
419
423
doi:
179
Lampson
,
M.A.
and
Cheeseman
,
I.M.
(
2011
)
Sensing centromere tension: Aurora B and the regulation of kinetochore function
.
Trends Cell Biol.
21
,
133
140
doi:
180
Zaytsev
,
A.V.
,
Segura-Peña
,
D.
,
Godzi
,
M.
,
Calderon
,
A.
,
Ballister
,
E.R.
,
Stamatov
,
R.
et al
(
2016
)
Bistability of a coupled Aurora B kinase-phosphatase system in cell division
.
eLife
5
,
e10644
doi:
181
Bloom
,
K.S.
(
2014
)
Centromeric heterochromatin: the primordial segregation machine
.
Annu. Rev. Genet.
48
,
457
484
doi:
182
Goshima
,
G.
and
Yanagida
,
M.
(
2000
)
Establishing biorientation occurs with precocious separation of the sister kinetochores, but not the arms, in the early spindle of budding yeast
.
Cell
100
,
619
633
doi:
183
Cooke
,
C.A.
,
Heck
,
M.M.
and
Earnshaw
,
W.C.
(
1987
)
The inner centromere protein (INCENP) antigens: movement from inner centromere to midbody during mitosis
.
J. Cell Biol.
105
,
2053
2067
doi:
184
Dawe
,
R.K.
,
Reed
,
L.M.
,
Yu
,
H.G.
,
Muszynski
,
M.G.
and
Hiatt
,
E.N.
(
1999
)
A maize homolog of mammalian CENPC is a constitutive component of the inner kinetochore
.
Plant Cell Online
11
,
1227
1238
doi:
185
Fujiwara
,
T.
,
Tanaka
,
K.
,
Kuroiwa
,
T.
and
Hirano
,
T.
(
2013
)
Spatiotemporal dynamics of condensins I and II: evolutionary insights from the primitive red alga Cyanidioschyzon merolae
.
Mol. Biol. Cell
24
,
2515
2527
doi:
186
Prensier
,
G.
and
Slomianny
,
C.
(
1986
)
The karyotype of Plasmodium falciparum determined by ultrastructural serial sectioning and 3D reconstruction
.
J. Parasitol.
72
,
731
736
doi:
187
Ureña
,
F.
(
1986
)
Three-dimensional reconstructions of the mitotic spindle and dense plaques in three species of Leishmania
.
Z. Parasitenkd.
72
,
299
306
doi:
188
Campbell
,
C.S.
and
Desai
,
A.
(
2013
)
Tension sensing by Aurora B kinase is independent of survivin-based centromere localization
.
Nature
497
,
118
121
doi:
189
London
,
N.
and
Biggins
,
S.
(
2014
)
Signalling dynamics in the spindle checkpoint response
.
Nat. Rev. Mol. Cell Biol.
15
,
736
747
doi:
190
Vleugel
,
M.
,
Hoogendoorn
,
E.
,
Snel
,
B.
and
Kops
,
G.J.P.L.
(
2012
)
Evolution and function of the mitotic checkpoint
.
Dev. Cell
23
,
239
250
doi:
191
Vicente
,
J.-J.
and
Cande
,
W.Z.
(
2014
)
Mad2, Bub3, and Mps1 regulate chromosome segregation and mitotic synchrony in Giardia intestinalis, a binucleate protist lacking an anaphase-promoting complex
.
Mol. Biol. Cell
25
,
2774
2787
doi:
192
Li
,
R.
and
Murray
,
A.W.
(
1991
)
Feedback control of mitosis in budding yeast
.
Cell
66
,
519
531
doi:
193
Hoyt
,
M.A.
,
Totis
,
L.
and
Roberts
,
B.T.
(
1991
)
S. cerevisiae genes required for cell cycle arrest in response to loss of microtubule function
.
Cell
66
,
507
517
doi:
194
He
,
X.
,
Patterson
,
T.E.
and
Sazer
,
S.
(
1997
)
The Schizosaccharomyces pombe spindle checkpoint protein mad2p blocks anaphase and genetically interacts with the anaphase-promoting complex
.
Proc. Natl Acad. Sci. USA
94
,
7965
7970
doi:
195
Buffin
,
E.
,
Emre
,
D.
and
Karess
,
R.E.
(
2007
)
Flies without a spindle checkpoint
.
Nat. Cell Biol.
9
,
565
572
doi:
196
Wild
,
T.
,
Larsen
,
M.S.Y.
,
Narita
,
T.
,
Schou
,
J.
,
Nilsson
,
J.
and
Choudhary
,
C.
(
2016
)
The spindle assembly checkpoint is not essential for viability of human cells with genetically lowered APC/C activity
.
Cell Rep.
14
,
1829
1840
doi:
197
Craney
,
A.
,
Kelly
,
A.
,
Jia
,
L.
,
Fedrigo
,
I.
,
Yu
,
H.
and
Rape
,
M.
(
2016
)
Control of APC/C-dependent ubiquitin chain elongation by reversible phosphorylation
.
Proc. Natl Acad. Sci. USA
113
,
1540
1545
doi:
198
Fujimitsu
,
K.
,
Grimaldi
,
M.
and
Yamano
,
H.
(
2016
)
Cyclin dependent kinase 1-dependent activation of APC/C ubiquitin ligase
.
Science
doi:
199
Jia
,
L.
,
Li
,
B.
and
Yu
,
H.
(
2016
)
The Bub1-Plk1 kinase complex promotes spindle checkpoint signalling through Cdc20 phosphorylation
.
Nat. Commun.
7
,
10818
doi:
200
Qiao
,
R.
,
Weissmann
,
F.
,
Yamaguchi
,
M.
,
Brown
,
N.G.
,
VanderLinden
,
R.
,
Imre
,
R.
et al
(
2016
)
Mechanism of APC/CCDC20 activation by mitotic phosphorylation
.
Proc. Natl Acad. Sci. USA
113
,
E2570
E2578
doi:
201
Zhang
,
S.
,
Chang
,
L.
,
Alfieri
,
C.
,
Zhang
,
Z.
,
Yang
,
J.
,
Maslen
,
S.
et al
(
2016
)
Molecular mechanism of APC/C activation by mitotic phosphorylation
.
Nature
533
,
260
264
doi:
202
Ploubidou
,
A.
,
Robinson
,
D.R.
,
Docherty
,
R.C.
,
Ogbadoyi
,
E.O.
and
Gull
,
K.
(
1999
)
Evidence for novel cell cycle checkpoints in trypanosomes: kinetoplast segregation and cytokinesis in the absence of mitosis
.
J. Cell. Sci.
112
(
Pt 24
),
4641
–4650 PMID:
[PubMed]
203
Gluenz
,
E.
,
Sharma
,
R.
,
Carrington
,
M.
and
Gull
,
K.
(
2008
)
Functional characterization of cohesin subunit SCC1 in Trypanosoma brucei and dissection of mutant phenotypes in two life cycle stages
.
Mol. Microbiol.
69
,
666
680
doi:
204
Singh
,
N.
,
Wiltshire
,
T.D.
,
Thompson
,
J.R.
,
Mer
,
G.
and
Couch
,
F.J.
(
2012
)
Molecular basis for the association of microcephalin (MCPH1) protein with the cell division cycle protein 27 (Cdc27) subunit of the anaphase-promoting complex
.
J. Biol. Chem.
287
,
2854
2862
doi:
205
Atkins
,
M.S.
,
Teske
,
A.P.
and
Anderson
,
O.R.
(
2000
)
A survey of flagellate diversity at four deep-sea hydrothermal vents in the Eastern Pacific Ocean using structural and molecular approaches
.
J. Eukaryot. Microbiol.
47
,
400
411
doi:
206
López-García
,
P.
,
Philippe
,
H.
,
Gail
,
F.
and
Moreira
,
D.
(
2003
)
Autochthonous eukaryotic diversity in hydrothermal sediment and experimental microcolonizers at the Mid-Atlantic Ridge
.
Proc. Natl Acad. Sci. USA
100
,
697
702
doi:
207
Flegontov
,
P.
,
Votýpka
,
J.
,
Skalický
,
T.
,
Logacheva
,
M.D.
,
Penin
,
A.A.
,
Tanifuji
,
G.
et al
(
2013
)
Paratrypanosoma is a novel early-branching trypanosomatid
.
Curr. Biol.
23
,
1787
1793
doi:
208
d'Avila-Levy
,
C.M.
,
Boucinha
,
C.
,
Kostygov
,
A.
,
Santos
,
H.L.C.
,
Morelli
,
K.A.
,
Grybchuk-Ieremenko
,
A.
et al
(
2015
)
Exploring the environmental diversity of kinetoplastid flagellates in the high-throughput DNA sequencing era
.
Mem. Inst. Oswaldo Cruz
110
,
956
965
doi:
209
Moreira
,
D.
,
López-García
,
P.
and
Vickerman
,
K.
(
2004
)
An updated view of kinetoplastid phylogeny using environmental sequences and a closer outgroup: proposal for a new classification of the class Kinetoplastea
.
Int. J. Syst. Evol. Microbiol.
54
,
1861
1875
doi:
210
Simpson
,
A.G.B.
,
Stevens
,
J.R.
and
Lukeš
,
J.
(
2006
)
The evolution and diversity of kinetoplastid flagellates
.
Trends Parasitol.
22
,
168
174
doi:
211
Jackson
,
A.P.
,
Otto
,
T.D.
,
Aslett
,
M.
,
Armstrong
,
S.D.
,
Bringaud
,
F.
,
Schlacht
,
A.
et al
(
2016
)
Kinetoplastid phylogenomics reveals the evolutionary innovations associated with the origins of parasitism
.
Curr. Biol.
26
,
161
172
doi:
212
Porcel
,
B.M.
,
Denoeud
,
F.
,
Opperdoes
,
F.
,
Noel
,
B.
,
Madoui
,
M.-A.
,
Hammarton
,
T.C.
et al
(
2014
)
The streamlined genome of Phytomonas spp. relative to human pathogenic kinetoplastids reveals a parasite tailored for plants
.
PLoS Genet.
10
,
e1004007
doi:
213
Tanifuji
,
G.
,
Kim
,
E.
,
Onodera
,
N.T.
,
Gibeault
,
R.
,
Dlutek
,
M.
,
Cawthorn
,
R.J.
et al
(
2011
)
Genomic characterization of Neoparamoeba pemaquidensis (Amoebozoa) and its kinetoplastid endosymbiont
.
Eukaryotic Cell
10
,
1143
1146
doi:
214
Walker
,
G.
,
Dorrell
,
R.G.
,
Schlacht
,
A.
and
Dacks
,
J.B.
(
2011
)
Eukaryotic systematics: a user's guide for cell biologists and parasitologists
.
Parasitology
138
,
1638
1663
doi:
215
Adl
,
S.M.
,
Simpson
,
A.G.B.
,
Lane
,
C.E.
,
Lukeš
,
J.
,
Bass
,
D.
,
Bowser
,
S.S.
et al
(
2012
)
The revised classification of eukaryotes
.
J. Eukaryot. Microbiol.
59
,
429
514
doi:
216
O'Neill
,
E.C.
,
Trick
,
M.
,
Hill
,
L.
,
Rejzek
,
M.
,
Dusi
,
R.G.
,
Hamilton
,
C.J.
et al
(
2015
)
The transcriptome of Euglena gracilis reveals unexpected metabolic capabilities for carbohydrate and natural product biochemistry
.
Mol. Biosyst.
11
,
2808
2820
doi:
217
Malik
,
H.S.
and
Henikoff
,
S.
(
2003
)
Phylogenomics of the nucleosome
.
Nat. Struct. Biol.
10
,
882
891
doi:
218
Marande
,
W.
,
Lukes
,
J.
and
Burger
,
G.
(
2005
)
Unique mitochondrial genome structure in diplonemids, the sister group of kinetoplastids
.
Eukaryotic Cell
4
,
1137
1146
doi:
219
Lukeš
,
J.
,
Flegontova
,
O.
and
Horák
,
A.
(
2015
)
Diplonemids
.
Curr. Biol.
25
,
R702
R704
doi:
220
de Vargas
,
C.
,
Audic
,
S.
,
Henry
,
N.
,
Decelle
,
J.
,
Mahé
,
F.
,
Logares
,
R.
et al
(
2015
)
Eukaryotic plankton diversity in the sunlit ocean
.
Science
348
,
1261605
doi:
221
Yubuki
,
N.
,
Edgcomb
,
V.P.
,
Bernhard
,
J.M.
and
Leander
,
B.S.
(
2009
)
Ultrastructure and molecular phylogeny of Calkinsia aureus: cellular identity of a novel clade of deep-sea euglenozoans with epibiotic bacteria
.
BMC Microbiol.
9
,
16
doi:
222
Yutin
,
N.
and
Koonin
,
E.V.
(
2012
)
Archaeal origin of tubulin
.
Biol. Direct
7
,
10
doi:
223
Williams
,
T.A.
,
Foster
,
P.G.
,
Cox
,
C.J.
and
Embley
,
T.M.
(
2013
)
An archaeal origin of eukaryotes supports only two primary domains of life
.
Nature
504
,
231
236
doi:
224
Spang
,
A.
,
Saw
,
J.H.
,
Jørgensen
,
S.L.
,
Zaremba-Niedzwiedzka
,
K.
,
Martijn
,
J.
,
Lind
,
A.E.
et al
(
2015
)
Complex archaea that bridge the gap between prokaryotes and eukaryotes
.
Nature
521
,
173
179
doi:
225
Embley
,
T.M.
and
Martin
,
W.
(
2006
)
Eukaryotic evolution, changes and challenges
.
Nature
440
,
623
630
doi:
226
Stechmann
,
A.
and
Cavalier-Smith
,
T.
(
2002
)
Rooting the eukaryote tree by using a derived gene fusion
.
Science
297
,
89
91
doi:
227
Rogozin
,
I.B.
,
Basu
,
M.K.
,
Csürös
,
M.
and
Koonin
,
E.V.
(
2009
)
Analysis of rare genomic changes does not support the unikont-bikont phylogeny and suggests cyanobacterial symbiosis as the point of primary radiation of eukaryotes
.
Genome Biol. Evol.
1
,
99
113
doi:
228
Katz
,
L.A.
,
Grant
,
J.R.
,
Parfrey
,
L.W.
and
Burleigh
,
J.G.
(
2012
)
Turning the crown upside down: gene tree parsimony roots the eukaryotic tree of life
.
Syst. Biol.
61
,
653
660
doi:
229
Cavalier-Smith
,
T.
(
2010
)
Kingdoms Protozoa and Chromista and the eozoan root of the eukaryotic tree
.
Biol. Lett.
6
,
342
345
doi:
230
Allen
,
J.W.A.
,
Jackson
,
A.P.
,
Rigden
,
D.J.
,
Willis
,
A.C.
,
Ferguson
,
S.J.
and
Ginger
,
M.L.
(
2008
)
Order within a mosaic distribution of mitochondrial c-type cytochrome biogenesis systems?
FEBS J.
275
,
2385
2402
doi:
231
Nishino
,
M.
,
Choy
,
J.W.
,
Gushwa
,
N.N.
,
Oses-Prieto
,
J.A.
,
Koupparis
,
K.
,
Burlingame
,
A.L.
et al
(
2013
)
Hypothemycin, a fungal natural product, identifies therapeutic targets in Trypanosoma brucei
.
elife
2
,
e00712
doi:
232
Merritt
,
C.
,
Silva
,
L.E.
,
Tanner
,
A.L.
,
Stuart
,
K.
and
Pollastri
,
M.P.
(
2014
)
Kinases as druggable targets in trypanosomatid protozoan parasites
.
Chem. Rev.
114
,
11280
11304
doi:

Author notes

*

Bungo Akiyoshi was awarded the Biochemical Society's 2016 Early Career Research Award (Cells); this article is based on the Award Lecture.

This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).