Вы находитесь на странице: 1из 15

Cell, Vol.

112, 407–421, February 21, 2003, Copyright 2003 by Cell Press

Centromeres and Kinetochores: Review


From Epigenetics to Mitotic
Checkpoint Signaling
Don W. Cleveland,1,* Yinghui Mao,1 elements of the mitotic checkpoint, they control cell
and Kevin F. Sullivan2 cycle advance during cell division.
1
Ludwig Institute for Cancer Research and
Department of Cellular and Molecular Medicine Defining the Locus
University of California, San Diego The centromere challenges the classic view of a genetic
9500 Gilman Drive locus. Dogma and experience suggest that a chromo-
La Jolla, California 92093 somal locus is defined by its DNA sequence, and its
2
The Scripps Research Institute function is contained with the information content pres-
10555 N. Torrey Pines Road ent in the primary sequence, e.g., recognition sites for
La Jolla, California 92037 DNA binding proteins. This simple situation is correct for
the centromere of budding yeast, where point mutations
abolish activity. In other organisms, however, including
The centromere is a chromosomal locus that ensures fission yeast, specific centromere-nucleating DNA se-
delivery of one copy of each chromosome to each quences have not been found, and centromere function
daughter at cell division. Efforts to understand the exhibits properties of an epigenetic locus, behaving as
nature and specification of the centromere have dem- a self-replicating protein complex that resides on cen-
onstrated that this central element for ensuring inheri- tromere DNA but is not determined by it.
tance is itself epigenetically determined. The kineto- Although the details differ markedly, resolution of bio-
chore, the protein complex assembled at each chemically and structurally distinct centromere domains
centromere, serves as the attachment site for spindle in many organisms has revealed common aspects of
microtubules and the site at which motors generate organization (Figure 1B), providing a sketch of a simple
forces to power chromosome movement. Unattached “universal centromere architecture.” Chromatin is the
kinetochores are also the signal generators for the key feature and the centromere domain is built on a
mitotic checkpoint, which arrests mitosis until all distinct type of nucleosome found nowhere else in the
kinetochores have correctly attached to spindle mi- genome, in which histone H3 is replaced by a divergent
crotubules, thereby representing the major cell cycle (50% identity) homolog usually referred to as CENP-A
control mechanism protecting against loss of a chro- (Smith, 2002). This core is flanked by another chromatin
mosome (aneuploidy). domain, a highly phased nucleosome array in S. cerevis-
iae and heterochromatin in other species that plays a
role in cohesion of sister centromeres (Bernard et al.,
Introduction 2001). Both domains are required for complete centro-
mere function.
Chromosome movement on the spindle during mitosis
and meiosis is powered and regulated by the centro- The Single Nucleosomal Centromere
mere, a discrete locus on each chromosome. Originally of Budding Yeast
identified as the primary constriction, this region is a The budding yeast centromere is by far the simplest
complex chromosomal substructure (Figure 1A). While and most dissected one (Figure 2). Spanning only ⵑ125
the term centromere is generally taken to refer to the bp of DNA, S. cerevisiae centromeres contain three dis-
DNA segment that confers centromere function, the cy- tinct DNA sequence elements (Figure 2A). Two of these,
tologically visible centromere is more complex . In mito- CDE I (purple, Figure 2A) and CDE III (red, Figure 2A),
sis, a proteinaceous structure, the kinetochore, assem- are conserved at each chromosome and function as
bles at the surface of the centromere and acts as the site sequence-specific elements by the criterion that single
of spindle microtubule binding. By electron microscopy, point mutations abolish their activity. The other, CDE II
the kinetochore appears as a narrow band of dense (blue in Figure 2A), is not conserved in sequence among
chromatin just at the surface of the primary constriction, the yeast chromosomes, but the approximate length
the inner kinetochore, and a laminar outer kinetochore (76–84 bp) and A⫹T rich content (90%) are maintained
domain (frequently misnamed a “plate”) that contains (reviewed in detail by Cheeseman et al., 2002b). The
many of the microtubule binding and signal transduction 125 bp centromere DNA is flanked on either side by
molecules discussed below. The monotonous hetero- a uniquely positioned (i.e., highly phased) nucleosome
chromatin that links two sister kinetochores has array rich in cohesins and thought to provide a functional
emerged as a distinct functional entity within the centro- context for centromere activity (Laloraya et al., 2000;
mere-kinetochore complex, important for cohesion and Tanaka et al., 1999).
kinetochore regulation. The boundaries of the centro- Somewhat ironically, CDE II, the unconserved DNA
mere locus therefore extend beyond the kinetochore- element, interacts with Cse4 (Stoler et al., 1995), the
forming region. In this review, we discuss how centro- homolog of CENP-A, the most highly conserved protein
meres and their kinetochores assemble, how they power motif known for the centromere (Figures 2A and 2B).
mitotic chromosome movements, and how, as signaling CENP-A/Cse4 nucleosomes are specifically associated
with centromere DNA in vivo (Meluh et al., 1998), strongly
*Correspondence: dcleveland@ucsd.edu suggesting their assembly as a nucleosomal complex at
Cell
408

Figure 1. Organization of Centromeres


(A) Overall organization of the centromere. A mitotic chromosome has been sectioned along the plane of the spindle axis, revealing the
symmetric bipolar organization of a chromosome fully engaged on the spindle. (Right) Key elements have been pseudo colored. (Violet) The
inner centromere, a heterochromatin domain that is a focus for cohesins and regulatory proteins such as Aurora B and Kin I. (Red) The inner
kinetochore, a region of distinctive chromatin composition attached to the primary constriction. (Yellow) The outer kinetochore, the site of
microtubule binding, is comprised of a diverse group of microtubule motor proteins, regulatory kinases, microtubule binding proteins, and
mitotic checkpoint proteins.
(B) Schematic illustration of centromere loci. Organization of centromeric DNA sequences from the four example organisms. (Top) Budding
yeast with a 125 bp centromere comprised of three sequence domains (pink, red, yellow). Fission yeast centromeres show an organized
structure, with a nonconserved central core (red), flanking inner repeats (pink arrows) at which the CENP-A-containing nucleosomes assemble,
and conserved outer repeats (stippled purple). The Drosophila centromere spans ⵑ400 kb (red) embedded in constitutive heterochromatin
(purple). (Bottom) Human centromeres have sizes approaching 10 Mb and are comprised of ␣-I satellite DNA (red) and a more divergent, less
regular ␣-II satellite (pink), flanked by heterochromatin (purple).

the centromere (Figure 2B). A second CDE II interacting with both core centromere and distal kinetochore or
protein, Mif2 (Meluh and Koshland, 1995), is the homolog spindle components (Ortiz et al., 1999; Wigge and Kilm-
of an essential mammalian kinetochore chromatin bind- artin, 2001; Janke et al., 2001). The Dam1p complex
ing protein CENP-C (Figure 2B). (Cheeseman et al., 2001; Janke et al., 2002) (also known
Centromere function in yeast also depends critically as DASH [Li et al., 2002]) may be the central component
on sequence specific DNA-protein interactions. CDE III of microtubule binding (Figures 2C and 2D), with its
binds an essential four subunit protein complex, Cbf3 activity regulated by the yeast Aurora B kinase (Ipl1)
(Gardner et al., 2001) (Figures 2B and 2C). CDE I binds (Tanaka et al., 2002; Biggins et al., 1999). Present in a
a nonessential bHLH DNA binding protein, Cbf1, that conserved complex with the yeast homologs of INCENP
interacts with Cbf3 components (Hemmerich et al., (Sli15) and survivin (Bir1), respectively (Bolton et al.,
2000). The geometry dictates that the Cbf1-Cbf3 interac- 2002), Ipl1-dependent phosphorylation of at least three
tion must take place across the DNA gyres wound over Dam1p components are essential for microtubule cap-
the Cse4 nucleosome, perhaps stabilizing the higher ture (Cheeseman et al., 2002a).
order complex like a protein clamp (Figure 2B). The From the model in Figure 2D, with the components
Cbf3 complex produces an extended footprint (80 bp) drawn to scale, it is immediately apparent that connect-
on centromere DNA (Russell et al., 1999; Espelin et al., ing a large microtubule (25 nm) with a small, 10 nm
1997). The resulting broad interaction surface may be nucleosomal centromere requires multiple attachments,
important for stable centromere binding or to provide a especially considering that the microtubule-kinetochore
large binding surface for more distal kinetochore pro- linkage is dynamic, with individual connections broken
teins, or both. The flanking chromatin domain functions and reformed as the microtubule grows or shrinks.
in cohesion (Megee et al., 1999; Tanaka et al., 1999).
How is the centromeric nucleoprotein complex at- Fission Yeast Centromere: Chromatin
tached stably to a single, highly dynamic (e.g., He et al., Not a Specific DNA Sequence
2000) spindle microtubule? Four key protein complexes The critical importance of the chromatin environment
comprised of ⬎20 components are central players in surrounding the core CENP-A chromatin domain has
outer kinetochore assembly and microtubule binding been most clearly elucidated in S. pombe. Each of the
(Figures 2C and 2D). The Ctf19 complex and the Ndc80 three centromeres, which range in size from 40–100 kb,
complexes appear to represent “adaptors” that interact possess a pair of repeated sequence arrays that are
Review
409

Figure 2. Molecular Details of a Simple Kinetochore


(A) The centromere DNA of S. cerevisiae is shown at top with the three sequence elements, CDE I (purple), CDE II (blue), and CDE III (red),
highlighted. Numbering corresponds to chromosome 11. Proteins that bind to the centromere DNA elements are color coded to match the
corresponding DNA binding domain and are drawn to scale assuming a globular structure. Human homologs, where known, are denoted in
italics. CBF3 contains two copies of Ctf3 and two dimers of Ndc10, but their arrangement in the complex is not known and is likely to be
more asymmetric than shown here.
(B) The nucleosomal centromere. Cse4 (blue and purple spacefill) forms a specialized nucleosome at the centromere, modeled here on the
structure of the nucleosome-containing histone H3 (Luger et al., 1997). Other histones (gray) are shown only as backbone traces. The position
of centromere DNA on the nucleosome is derived from Keith and Fitzgerald-Hayes (2000) by placing CDE I (purple) in contact with one of the
two Cse4 molecules. Most of the remaining histone core surface is in contact with CDE II (blue), while CDE III (red, behind) projects into the
internucleosomal linker DNA. The divergent N-terminal tail of Cse4 is drawn in purple. (Right) The nucleosome is rotated 180⬚ and is shown
with Cbf1 (purple), Mif2 (blue), and Cbf3 (red) placed on DNA elements CDE I, -II and -III, respectively, illustrating the relative sizes and
positions of the core centromere components.
(C) Kinetochore complexes associate with the centromere nucleoprotein core. Four key kinetochore protein complexes discussed in the text
are illustrated. Interactions, genetic or biochemical, are indicated by black arrows, while Ipl1 substrates are shown with red arrows.
(D) A model of budding yeast centromere-microtubule interaction. A microtubule (green) is shown binding to the centromere in the context
of a 30 nm fiber.

arranged in an inverted repeat around an unconserved mating type locus (Volpe et al., 2002; Partridge et al.,
central core sequence (Figure 1B) (Clarke, 1998). Both 2002). A key function of this heterochromatin domain
central core and inner repeat sequences are bound by is the recruitment of the fission yeast cohesin (Rad21)
the CENP-A homolog (Cnp1) and two conserved pro- (Bernard et al., 2001; Tomonaga et al., 2000), the central
teins, Mis6 and Mis12, homologs of subunits of the Ctf19 component of the protein glue that holds duplicated
complex from budding yeast (Goshima et al., 1999; Par- chromatids together.
tridge et al., 2000; Takahashi et al., 2000).
Unlike budding yeast, no unique sequence within it is Metazoans: The Mega Multisubunit Centromere
sufficient in S. pombe, with flanking sequences critical Metazoan centromere organization is in much more
for the establishment and maintenance of centromere complex than that in the yeasts (Figure 1B). Spanning
function (Partridge et al., 2000; Ngan and Clarke, 1997; 0.3–5 Mbp of DNA, human centromeres contain exten-
Ekwall et al., 1997). These are modified by methylation sive (1,500 to ⬎30,000 copies), tandemly repeated
of lysine 9 of histone H3, which in turn recruits Swi6, arrays of a 171 bp sequence element termed ␣ satellite
the fission yeast equivalent of vertebrate HP1 (Lachner (Figure 3A). Centromere function has been mapped to ␣
et al., 2001). These are characteristic biochemical mark- satellite arrays by centromeric deletions, either naturally
ers of heterochromatin, which likely require elements of occurring on the X chromosome (Schueler et al., 2001)
the RNAi system in S. pombe by analogy with Cen- or those induced by telomere insertion into the Y (Brown
homology-dependent heterochromatic silencing at the et al., 1994). Centromeric satellite DNAs are not con-
Cell
410

Figure 3. Organization of Human Centromeres


(A) DNA organization. The hierarchic organization of an ␣ satellite DNA is illustrated, with a 171 bp monomer sequence shown at top. Monomer
sequences are iterated with nucleotide sequence variations to form a higher order repeat (colored arrows), which itself is tandemly iterated
with high (⬎99%) sequence conservation to form extensive arrays of higher order repeats. At bottom is a diagram of the centromere region
of chromosome 10, illustrating the ⵑ2 Mb ␣ satellite array with surrounding pericentric satellite arrays (SAT2 and SAT3).
(B) Discontinuous distribution of centromere proteins. Hypotonic stretching of chromosomes followed by immunofluorescence with human
anti-centromere antibodies revealed a punctuate distribution of centromere antigens (Blower et al., 2002), suggesting a multiple subunit
organization at the centromere/kinetochore complex. (Right) Extended chromatin fibers from human interphase cells demonstrate discontinuous
CENP-A domains (red) interspersed with histone H3 domains (green). Images provided by Beth Sullivan (Boston University).
(C) Folding centromeric chromatin in mitotic chromosomes. The discontinuous CENP-A chromatin fiber (left) forms a single domain at the
surface of the chromosome and therefore must be folded or coiled to achieve the polarized distribution of CENP-A and H3 sites in mature
mitotic kinetochores (Sullivan et al., 2001). The coil shown illustrates only one possible folding pathway in the context of an idealized centromere
region, showing that CENP-A (red) assembles onto regular ␣-I satellite sequences (green), with ␣-II satellite (teal blue) and pericentric satellite
sequences (orange, purple) flanking this site. At right, a sinuous path of chromosomal DNA through the centromere region is proposed on
the basis of the differential localization of ␣-I and ␣-II sequences to the outer and inner centromere regions.

served in sequence among metazoans, but share: (1) S. pombe depend strongly, and those of metazoan cells
their presence in very large tandem arrays, and (2) unit primarily, on epigenetic factors rather than DNA se-
repeat lengths that tend toward multiples of the nucleo- quence for activity. Three lines of evidence support this
some repeat length (Henikoff et al., 2001). (An exception statement. First, centromere sequences are evolving at
is Drosophila, where the functionally mapped centro- an unusually high rate, coevolving with their essential
mere consists of simple [5 bp] satellite sequences inter- partner CENP-A, and show no obvious sequence con-
spersed with various transposons [Sun et al., 1997]). servation that links divergent species or even different
CENP-A nucleosomes are bound to ␣ satellite DNA chromosomes in the case of Drosophila (Henikoff et al.,
in human centromeres (Vafa and Sullivan, 1997), but 2001). Second, centromere DNA sequences by them-
these are not uniformly distributed. Stretched chromatin selves are unable to specify centromere function: stable
fibers reveal interspersed CENP-A- and histone H3-con- dicentric chromosomes have been found in which one
taining nucleosomes (Figures 3B and 3C; Blower et al., centromere has been silenced with no obvious re-
2002). These foci may represent kinetochore “subunits” arrangement of centromere DNA (Sullivan and Willard,
(for which the budding yeast represents a unitary exam- 1998). Third, acquisition of centromere function has
ple) that assemble together (Zinkowski et al., 1991) to been found on certain rearranged chromosomes lacking
form multiple binding sites for the multiple microtubules an endogenous centromere (Figure 4A). In a well studied
that attach to metazoan centromeres. case, a stable derivative of human chromosome 10 with
Primate centromeres frequently have two major ␣ sat- a deletion spanning the centromere was found to have a
ellite families adjoining each other (He et al., 1998): a functional centromere at position 10q25. Mitotic stability
highly regular ␣-I and an ␣-II that varies widely in mono-
through development and adult life in this patient dem-
mer sequences and repeat structure. CENP-A is bound
onstrated that this neocentromere functions and binds
primarily to ␣-I satellite sequences, at the surface of the
to ⬎20 kinetochore, centromere, and chromatin proteins
chromosome (Ando et al., 2002). In contrast, ␣-II satellite
(Saffery et al., 2000). This region shows no centromeric
is localized in the interior or central domain of the centro-
mere, where components such as INCENP, Aurora B, DNA or any other distinctive feature (Barry et al., 1999).
and cohesin are concentrated (Figure 3C). This two- CENP-A is present over a 460 kb domain (Lo et al., 2001),
component organization is evident throughout eukary- similar in size to the minimal Drosophila centromere (420
otes. A core domain built around CENP-A and centro- kb; Sun et al., 1997) and the smallest natural human
mere-specific chromatin binding proteins establishes centromeres (500 kb) (found on Y chromosomes, Tyler-
the kinetochore-forming component of the centromere, Smith et al., 1993). Thus, noncentromeric DNA segments
while flanking domains are enriched in proteins involved can acquire full centromere function without DNA se-
in chromatid cohesion. quence changes and be maintained at those sites indefi-
nitely—even through multiple human generations (Tyler-
The Epigenetic Centromere Smith et al., 1999).
While in budding yeast centromere DNA alone can nu- Chromatin rather than DNA may be a major determi-
cleate centromere formation de novo, centromeres in nant of centromere identity (Williams et al., 1998). During
Review
411

Figure 4. Epigenetics and DNA Sequences in Centromere Formation


(A) A human neocentromere. (Top) Human chromosome 10. (Below) Rare internal recombination has produced the two subfragments in one
patient. One (left) is a ring chromosome (r(del)10) that carries the original centromere (red). The other is a linear deletion product, mar(del)10,
that has acquired a new centromere at 10q25 (orange). The CENP-A domain of this neocentromere spans about 400 kb.
(B) Generation of neocentromeres. A Drosophila chromosome derivative, Dp1187, was mapped for centromere activity using x-irradiation to
introduce chromosomal breaks (top). In one product, an inversion placed the normally euchromatic sequences on the left arm (cyan) directly
in cis to the functional centromere. Centromeric chromatin is proposed to have spread into the now flanking euchromatic DNA. Subsequent
x-irradiation produced deletion that entirely lacks the initial centromere DNA but, nevertheless, retains substantial centromere activity (bottom)
by virtue of the centromeric chromatin that has spread into the region.
(C) Construction of a human minichromosome. A single copy of the 16mer higher order repeat of chromosome 17 ␣ satellite DNA was amplified
to generate arrays of up to 64 repeat units in a bacterial artificial chromosome vector. These were further amplified to generate fragments
up to ⵑ1 Mb in length. Synthetic ␣ satellite arrays were combined with a ␤-geo resistance marker gene (green), telomere DNA (yellow), and
human genomic DNA (gray) for transfection into HT1080 cells. Linear minichromosomes (6–10 Mb in length, 5%–10% the size of the smallest
human chromosomes) containing 17 ␣ satellite (without detectable host chromosomal DNA) were identified at a low frequency (Harrington et
al., 1997).
(D) Dissecting DNA sequence requirements by minichromosome engineering. (Top) The ␣ satellite domain of human chromosome 21, including
a highly regular ␣21-I array with a high frequency of CENP-B boxes (green ovals) and a flanking ␣21-II array that is highly irregular in sequence
organization and has a low frequency of CENP-B boxes. (Middle) Construction of large arrays containing (␣21-ICENPB⫹) or lacking (␣21-
Icenpb-) functional, 11mer CENP-B B boxes were produced in a BAC vector, amplified, and transfected into HT1080 cells. ␣21-ICENPB⫹
yielded minichromosomes in 40%–50% of transformants; ␣21-Icenpb⫺ yielded none. Amplification of a 340 bp non-␣ satellite sequence into
which CENP-B boxes were embedded (V340:CENPB⫹) yielded no minichromosomes (Ohzeki et al., 2002).

the mapping of a Drosophila centromere by X-ray- 2001). CENP-A chromatin does not form a specialized
induced chromosome breakage, a chromosome was re- domain of DNA replication (Shelby et al., 2000; Sullivan
covered that placed the centromere directly next to a and Karpen, 2001), and its loading is uncoupled from
euchromatic DNA segment (Figure 4B). A subsequent that of the conventional histones (Shelby et al., 1997;
round of x-irradiation produced stable chromosome de- Takahashi et al., 2000). CENP-A is distributed conserva-
rivatives that lacked centromere DNA but retained only tively, however, to the two daughter chromatids during
the euchromatic sequence, indicative of a centromere- S phase (Shelby et al., 2000), providing the potential for
determining chromatin structure that spread into the carrying a DNA sequence-independent heritable mark
euchromatin and which retained full centromere func- through a CENP-A directed, self-replicating chromatin
tion after subsequent X-ray-mediated severing of the assembly pathway that bears little reference to the ac-
natural centromere. tual DNA upon which it sits (for more detailed account,
The universal association of centromere function with see Sullivan et al., 2001).
a distinct histone also supports a role for chromatin in
centromere determination. CENP-A appears to be at Determinants of De Novo Centromere Formation
the root of the kinetochore assembly process and is It is clear that non-DNA elements play a major role in
required for assembly of most distal kinetochore com- centromere determination, but do DNA sequences not
ponents examined (Howman et al., 2000; Oegema et al., play a role at all? This has been answered by engineering
Cell
412

artificial mammalian chromosomes. Minichromosomes Subsequent capture by the unattached kinetochore of


were initially formed de novo (Harrington et al., 1997) a microtubule from the opposite spindle pole produces
by transfecting a mixture of a large synthetic chromo- biorientation and congression to the cell center in a
some 7 ␣ satellite array, telomeric sequences, a se- rapid, discontinuous series of movements, again medi-
lectable marker, and genomic DNA fragments (for origin ated by kinetochore motors (Figure 5B, red chromatids).
and other functions) (Figure 4C). This produced stably The presence at kinetochores of active plus and minus
transmitted microchromosomes with functional centro- end motors has been demonstrated, with net direction
meres containing chromosome 7 ␣ satellite sequences of movement switchable in vitro by phosphorylation (Hy-
and typical kinetochore components, but only in a hand- man and Mitchison, 1991). In metazoans (Figure 6A),
ful of extremely rare instances. Subsequently, a yeast known kinetochore motors include the kinesin family
artificial chromosome (YAC) carrying chromosome 21 ␣ member CENP-E and cytoplasmic dynein.
satellite arrays was retrofitted with telomeres by recom-
bination and introduced into human cells, yielding a Motors in Congression and Anaphase
much enhanced frequency of microchromosomes con- A long perplexing question has been how congression
taining newly assembled functional centromeres, but is achieved. With a common set of motors at both sister
only when the YAC contained ␣-I type satellite arrays kinetochores, how is it that movement of a bioriented
(Ikeno et al., 1998). pair to the center is dominant, especially for those (like
The homogeneous, highly regular ␣-I DNA has a high the red pair in Figure 5B) that can have many more
frequency of sites for the centromere-associated DNA microtubules bound to the kinetochore first attached
binding protein CENP-B; ␣-II does not. This suggested and that are initially much closer to one pole? Laser
a role for CENP-B in centromere formation, an idea pre- ablation to disconnect the two chromatids or destroy
viously dismissed by demonstration that CENP-B null either kinetochore established that almost all of the force
mice are viable and fertile with only mild phenotypic is generated by the leading kinetochore (Khodjakov and
effects on gonad size (Hudson et al., 1998). By engi- Rieder, 1996), i.e., the one whose bound microtubules
neering two ␣-I satellite arrays such that they differed are shortening and whose motors are moving toward
only by the presence of functional CENP-B boxes (Fig- the microtubule minus ends.
ure 4D), CENP-B was shown to be necessary for de Two soluble dynein inhibitors (p50 dynamitin and a
novo centromere formation, but it functions efficiently dynein antibody) disrupt the alignment of kinetochores
only in the context of ␣-satellite DNA. Since heterochro- at metaphase (Sharp et al., 2000). Mutations in the teth-
matin formation is a highly cooperative process driven ers (Rough deal [Rod] and Zeste white 10 [ZW10]) that
by a series of mutually reinforcing reactions (Richards link dynein to kinetochores attenuate the rate of pole-
and Elgin, 2002), components such as CENP-B likely ward chromosome movement (Savoian et al., 2000), im-
function in mediating centromeric chromatin modifica- plicating dynein as a likely primary motor for con-
tion, as has been seen for the three CENP-B homologs gression.
in S. pombe (Nakagawa et al., 2002). Other motors, especially the plus end motor CENP-E,
stabilize microtubule capture at both sisters (McEwen
et al., 2001; Putkey et al., 2002) in part through a contri-
Kinetochores Power Chromosome
bution of CENP-E in maintaining attachment of kineto-
Movements in Mitosis
chores to the end of a disassembling microtubule (Lom-
Since the term “mitosis” was introduced over 100 years
billo et al., 1995). Chromosome alignment is precluded
ago by Fleming, a challenge has been to understand
by disruption of CENP-E function in vitro (Wood et al.,
how cells divide and faithfully transmit chromosomes at
1997) or in vivo (Putkey et al., 2002). A final class of
each cell division. In a typical somatic cell cycle (Figure kinetochore component, exemplified by the Kin I sub-
5A), chromosomes in prophase initiate condensation, group of the kinesin family (including XKCM1 in Xenopus
then, upon disassembly of the nuclear envelope and [Walczak et al., 1996] and MCAK in mammals [Worde-
the interphase microtubule array, the fully compacted man and Mitchison, 1995]), is a microtubule disassem-
chromosomes spill into what was the cytoplasm to pro- blase [Desai et al., 1999]. This novel action acts at and
duce prometaphase. As a nascent mitotic spindle as- between the two sister kinetochores, using the normal
sembles, a dynamic process of repetitive search by un- kinesin ATPase cycle to power removal of tubulin sub-
stable microtubules ensues for capture of chromosomes units.
at their kinetochores (Figure 1A). Initial capture is fre- Chromosomes also experience forces exerted along
quently by binding of one kinetochore of a duplicated the chromosomes arms as a result of spindle microtu-
chromosome pair along the side of a spindle microtubule bule interaction with plus-end directed microtubule mo-
(Figure 5B, violet chromatid pair), allowing rapid (up to tors bound to chromatin (chromokinesins). The first
1 ␮m/s) poleward translocation along that microtubule identified, Drosophila Nod, is required for proper align-
powered by a minus-end directed, kinetochore bound ment of meiotic chromosomes that have not undergone
microtubule motor, almost certainly cytoplasmic dynein recombination (Afshar et al., 1995). Immunodepleting
(Rieder and Alexander, 1990). This is followed by attach- another (Kid) prevents normal chromosome alignment
ment of additional microtubules (up to 30 in humans (Antonio et al., 2000; Funabiki and Murray, 2000), while
[Rieder, 1981] or seven at mouse kinetochores [Putkey antibody-induced inhibition of human Kid blocks chro-
et al., 2002]), motor action at attachment sites, and oscil- mosome oscillations (like those of the red chromatid
latory movements linked to continued growth and pair, Figure 5C), with chromosome arms atypically ex-
shrinkage of those kinetochore bound microtubules tending toward spindle poles during congression (Lev-
(Figures 5B and 5C, orange chromatid pair). esque and Compton, 2001).
Review
413

Figure 5. Chromosome Movements in Mi-


tosis
(A) Summary of the stages of mitosis. (Red)
DNA; (green) microtubules. Images provided
by Andreas Merdes (University of Edinburgh).
(B) Chromosome movements during mitosis:
(violet) one kinetochore of a duplicated chro-
matid pair becomes laterally attached to a
single microtubule and rapidly powers the
pair poleward. (Orange) The mono-oriented
chromosome acquires additional microtu-
bules and then oscillates near the pole. (Red)
After capture by the unattached kinetochore
of a microtubule from the opposite pole, the
now bioriented chromosome congresses to
the spindle equator. (Blue) At anaphase, the
sister chromatids separate and undergo pole-
ward motion.
(C) The speeds of chromosome movements
for the chromatids in (B). (Violet) Mono-ori-
ented rapid poleward translocation; (orange)
oscillatory movements prior to biorientation;
(red) bioriented congression and oscillation;
and (blue) poleward motion in anaphase.
Adapted from Skibbens et al. (1993). Repro-
duced from The Journal of Cell Biology, 1993,
vol. 122, pp. 859–875 by copyright permission
of the Rockefeller University Press.

Modeling Chromosome Congression 1998]). Considering that dynein is tethered to the under-
Among the early models to explain congression were lying kinetochore by ZW10/Rod (see model in Figure
proposals that the force generated was proportional to 6B), congression would be achieved if the affinity of
the length of the attached microtubule (Hays et al., 1982; these tethers for kinetochore sites was weakened by
Ostergren, 1951). The explanation was as perplexing as increasing microtubule occupancy at that kinetochore.
the initial problem, as no molecular mechanism was In other words, if distortion of the kinetochore occurs
apparent for how such a length-dependent force could as more microtubules bind and this decreases the
actually be generated through action at kinetochores. strength of ZW10/Rod binding, then dynein’s action
With the demonstration that ZW10 and Rod apparently would simply pull itself and ZW10/Rod off of the more
stream poleward after microtubule capture (Williams et highly occupied, trailing kinetochore, producing the
al., 1996), as do many of the kinetochore bound compo- streaming of ZW10 so prominently seen on both sides
nents of the mitotic checkpoint (Howell et al., 2000, of the sister kinetochores after congression (Williams
2001), it seems to us that a model compatible with the et al., 1996). The continuing oscillations at metaphase
key evidence is that congression is powered, at least in (Figures 5B and 5C, red chromatid pair) would simply
part, by cytoplasmic dynein asymmetrically bound to reflect stoichastic changes in microtubule number on
the leading kinetochore of a chromatid pair. the two sides, thereby switching which kinetochore
The 1-50 pN force generated during chromosome could more tightly hang onto its dynein.
movement (Alexander and Rieder, 1991; Nicklas, 1983, A cautionary note, however, is that there are multiple
1988) is within the capacity of one or a few dynein mole- contributors and multiple solutions to congression so
cules (ⵑ2–6 pN per motor molecule [Shingyoji et al., that the model put forward here is likely only be one
Cell
414

Figure 6. Kinetochore Microtubule Capture and Chromosome Congression


(A) Kinetochore microtubules are captured by motor proteins, including CENP-E and cytoplasmic dynein, the latter in associated with dynactin,
ZW10, and Rod. Kin I acts as a microtubule disassemblase at and between the two sister kinetochores. Microtubule plus-end tracking proteins
(such as EB1, APC) track microtubule “plus” ends and contribute to kinetochore microtubule stability and attachment. Plus-end directed
chromokinesins are localized on chromosome arms and are responsible for driving the chromosome arms away from the poles (the “polar
ejection” force).
(B) A model for chromosome congression in prometaphase driven by dynein-dependent minus-end directed kinetochore motility. Increasing
microtubule occupancy at the trailing kinetochore lowers ZW10/Rod affinity for that kinetochore causing dynein to pull itself and ZW10/Rod
out of the kinetochore and to stream along the kinetochore microtubule. On the leading kinetochore with fewer microtubules bound, ZW10/
Rod is tethered more tightly, and the dynein power stroke moves the chromosome toward the spindle equator.

aspect. Chromosome fragments produced by laser cut- increases dynamic instability (Tirnauer et al., 1999). Re-
ting still congress with only a single kinetochore (Khodja- moval of EB1 has demonstrated that it is important for
kov et al., 1997). Similarly, in some cells, dynein inhibi- spindle microtubule stabilization (Rogers et al., 2002).
tors apparently do not affect congression, although they During congression of a chromatid pair, EB1 is found
disrupt mitotic checkpoint signaling (Howell et al., 2001). at the ends of kinetochore bound microtubules that are
growing, but not at those that are disassembling (Tir-
nauer et al., 2002). EB1 interacts (Su et al., 1995) with
Mitotic Complexities: Nonmotor Attachments the human adenomatous polyposis coli (APC) tumor
In budding yeast where the intranuclear spindle forms suppressor protein (not to be confused with APC/C, the
prior to centromere duplication during S phase, nonmo- E3 ubiquitin ligase used in controlling anaphase onset—
tor microtubule-associated proteins (MAPs) seem to be see below). APC also binds to and stabilizes microtu-
primary in microtubule capture, while dynein plays a role bules (Zumbrunn et al., 2001), localizes to the ends of
in spindle positioning, but not chromosome movement microtubules embedded in kinetochores, forms a com-
(Yeh et al., 1995). The Dam1p complex (Figures 2C and plex with mitotic checkpoint proteins Bub1 and Bub3,
2D) interacts physically with central kinetochore pro- and is a substrate for both of the Bub1 and BubR1
teins of both the Ctf3 and Ndc80 complexes and binds kinases in vitro (Kaplan et al., 2001). One possible expla-
to microtubules directly in vitro (Cheeseman et al., 2001), nation for these observations is that EB1/APC complex
consistent with a direct role in mediating kinetochore- may be one of the nonmotor linker(s) that connect micro-
microtubule attachments. tubule attachment and the spindle checkpoint signaling
A group of microtubule plus-end binding proteins machinery on the kinetochore.
(e.g., the EB1 protein family) might also be involved in
mediating interactions between microtubules and kinet- Anaphase Movements Using Motors and Flux
ochores (Figure 6A). The yeast EB1 homolog BIM1 local- Two mechanisms for poleward chromosome movement
izes to the plus ends of cytoplasmic microtubules and in anaphase, the primary function of mitosis, are now
Review
415

known: kinetochore motors and microtubule flux. That it was an unattached kinetochore that was responsible
motors are used is hardly surprising, given their contri- came from the seminal demonstration that laser ablation
bution to congression. For motors, the most obvious of the last unattached kinetochore produces anaphase
candidate is dynein. Although not required for chromo- onset within about 15 min (Rieder et al., 1995). A kineto-
some segregation in budding yeast (Yeh et al., 1995), chore-dependent wait anaphase signal is also sug-
disruption of dynein clearly shows an anaphase chromo- gested in budding yeast: blocking centromere assembly
some segregation defect in Tetrahymena (Lee et al., (by destruction of the Cbf3 component Ndc10) elimi-
1999) and in kinetochore- and chromatid-to-pole move- nates mitotic delay in the presence of microtubule as-
ments in flies (Sharp et al., 2000). sembly inhibitors (Gardner et al., 2001).
Flux, however, is much less intuitive. Discovered by
photobleaching fluorescent microtubules (Gorbsky et Generating a Diffusible Inhibitor
al., 1987) and confirmed using fluorescence photoac- of Advance to Anaphase
tivation of tubulin assembled into kinetochore bound Genetics in yeast initially identified seven components
microtubules (Mitchison and Salmon, 1992), flux repre- of the mitotic checkpoint, Mad1–Mad3 (Mitotic arrest
sents continuous addition of tubulin subunits at kineto- defective) (Li and Murray, 1991), Bub1–Bub3 (Budding
chores, coupled to disassembly at the poles driven by uninhibited by benomyl) (Hoyt et al., 1991), and Mps1,
plus-end directed, pole bound microtubule motors pre- a kinase that is also essential for spindle pole body
sumably pulling on kinetochore microtubules and sliding duplication (Weiss and Winey, 1996). There are verte-
them poleward. Subunits are lost from the minus ends of brate homologs of all of these except Bub2. Initially
the microtubules, which must be tethered to the spindle demonstrated for Mad2 (Chen et al., 1996; Li and Be-
pole, but in a way that subunits can be disassembled. nezra, 1996), all have now been identified to bind to and
The major mechanism for chromosome movement in act at unattached kinetochores including Mad1 (Chen
anaphase in vertebrate somatic cells is motor-powered et al., 1998), Bub1 (Taylor and McKeon, 1997), Bub3
kinetochore movement coupled to microtubule disas- (Taylor et al., 1998), BubR1 (the mammalian Mad3) (Tay-
sembly at the kinetochore (Mitchison and Salmon, 1992; lor et al., 1998), and Mps1 (Abrieu et al., 2001). Additional
Walters et al., 1996). Here, poleward flux makes a rela- required contributors without yeast counterparts are
tively minor contribution with chromosome-to-pole known in metazoans. Without the kinetochore-associ-
movement three to eight times faster than flux. Yeast ated microtubule motor protein CENP-E, a binding part-
appear to lack microtubule depolymerization at poles ner of BubR1, the checkpoint cannot be established or
and poleward microtubule flux during anaphase (Malla- maintained in vitro (Abrieu et al., 2000) or in mice (Putkey
varapu et al., 1999). In Xenopus egg extracts, however, et al., 2002). Inhibition by mutation (Basto et al., 2000)
anaphase A movement occurs at rates similar to pole- or antibody injection (Chan et al., 2000) of ZW10 and Rod
ward spindle microtubule flux (Desai et al., 1998), consis- have revealed both to also be required for checkpoint
tent with flux as the predominant mechanism. Else- signaling. A final component with extensive sequence
where, the situation is controversial: fluorescent speckle homology to and overlapping function with Bub3 is Rae1
microscopy has been used to claim a dominant role (Babu et al., 2003), initially identified with an additional
for flux (Maddox et al., 2002) in syncytial Drosophila role in nuclear transport.
embryos, while other efforts have found that dynein in- Although the nature of the direct molecular interac-
hibitors disrupt chromatid-to-pole movement during tion(s) between checkpoint proteins and kinetochores
anaphase A (Savoian et al., 2000; Sharp et al., 2000). It and the interdependencies of checkpoint proteins have
seems safe to conclude that in general, both mecha- not been determined (see Musacchio and Hardwick
nisms are used to produce anaphase force at kineto- [2002] for details), the basic plan of the signaling cas-
chores. cade is established (Figure 7). Mad2 is recruited to unat-
tached kinetochores in a complex with Mad1 (Chen et
Unattached Kinetochores as Signaling Devices al., 1998). BubR1 and Bub1, both kinases, are required
for the Mitotic Checkpoint for generation and then rapid release from kinetochores
In order to assure accurate segregation, the mitotic of one or more inhibitors of Cdc20 (Fizzy in flies [Dawson
checkpoint (also known as the spindle assembly check- et al., 1993], p55 in mammals [Kallio et al., 1998] or Slp1p
point) acts to block entry into anaphase until both kineto- in fission yeast [Kim et al., 1998]). This sequesters or
chores of every duplicated chromatid pair have attached inactivates Cdc20, which is required for substrate recog-
correctly to spindle microtubules. What has emerged in nition by the anaphase promoting complex/cyclosome
the decade since its initial description is that unattached (APC/C) (substrates include securin [Zur and Brandeis,
kinetochores and/or those not under microtubule- 2001] and cyclin B [Raff et al., 2002]), whose ubiquitin-
induced tension are the central signaling elements that mediated destruction by the proteosome is required for
produce a “wait anaphase” signal. A trio of cell-biologi- anaphase onset (Cohen-Fix et al., 1996; Fang et al.,
cal experiments provided the primary insights that a 1998; Zou et al., 1999).
single, unattached kinetochore is enough to inhibit ana- The identity of the kinetochore-derived wait anaphase
phase onset. By filming mitoses, it was initially found inhibitor (or inhibitors) is not established. An activated
that anaphase ensues about 20 min after the last kineto- form of Mad2 (frequently referred to as Mad2*) has long
chore attaches to the spindle (Rieder et al., 1994) and been thought to be the inhibitory signal released. Micro-
that by repeated detachment of a meiotic chromosome injection of antibodies to Mad2 first revealed premature
from a spindle by manipulation with a microneedle de- anaphase onset and chromosome missegregation
layed anaphase indefinitely (Li and Nicklas, 1995). That (Gorbsky et al., 1998). Mad2 only binds to unattached
Cell
416

Figure 7. Activating and Silencing the Mitotic


Checkpoint
Unattached kinetochores bind a collection of
mitotic checkpoint components. These in-
clude at least four kinases [BubR1 (Chan et
al., 1999), Bub1 (Roberts et al., 1994), Mps1
(Abrieu et al., 2001), and Mapk (Takenaka et
al., 1998)]. These and the microtubule motor
CENP-E are required for rapid recruitment of
the Mad1/Mad2 complex. Mad2 cycles rap-
idly, releasing from kinetochores in a modi-
fied form or complex that sequesters Cdc20,
thus inhibiting recognition by the ubiquitin li-
gase APC/C of substrates, including securin,
whose destruction is required for advance to
anaphase. ZW10/Rod, the kinetochore tar-
geting components for cytoplasmic dynein,
are required for generation of the inhibitory
signal, possibly because they mitigate the
rapid release of the inhibitor from kineto-
chores. After microtubule attachment, ten-
sion is generated between the bioriented ki-
netochores, checkpoint signaling is silenced,
and anaphase ensues after decay of the pre-
viously made inhibitor.

kinetochores and the association at those kinetochores (Sudakin et al., 2001; Fraschini et al., 2001). This has led
is extremely dynamic, with release and rebinding to un- to the suggestion that the kinetochore contribution may
attached kinetochores within 25 s (Howell et al., 2000). modify APC/C itself to increase its affinity for MCC (Su-
Since a tetrameric form of recombinant Mad2 forms a dakin et al., 2001). While possible, this is unattractive
ternary complex with Cdc20 and APC/C in vitro, blocks as the only kinetochore-derived inhibitor because in or-
APC/C ubiquitination activity, and arrests Xenopus em- der to keep APC/C inactive, such a model would require
bryos and egg extracts in mitosis (Fang et al., 1998), it a single kinetochore to be capable of rapidly recruiting,
has been proposed as a candidate for the wait anaphase and modifying, all cellular APC/C—a daunting task.
inhibitor. However, no checkpoint-dependent Mad2 More plausible are models with built-in signal amplifica-
oligomer generation has yet been identified in vivo. tion, for example, where a very abundant component
The recruitment of Mad2 at kinetochores has often (like Mad2) is activated and released by kinetochores,
been taken as a measure of ongoing checkpoint signal thereby producing a high concentration of an inhibitor
generation (and release). However, inhibition of ZW10/ to saturate available Cdc20.
Rod yields an inactive checkpoint despite prominent Lastly, kinetochores directly attract and may modify
Mad2 binding at kinetochores (Chan et al., 2000). Dimi- the target for inhibition, Cdc20, which cycles at kineto-
nution of Hec1, the homolog of yeast Ndc80, on the chores (Kallio et al., 2002) with kinetics even faster than
other hand, yields a chronically activated checkpoint reported for Mad2 (complete turnover within about 5 s).
with no Mad2 bound to kinetochores (Martin-Lluesma How this relates to checkpoint signaling is made less
et al., 2002). Thus, it is now abundantly clear that the clear by unaltered cycling after checkpoint signal pro-
steady-state level of Mad2 bound to kinetochores is not duction is silenced by microtubule attachment.
a faithful reporter for checkpoint activation or inacti-
vation. The Checkpoint Signal Has Limited Diffusibility
Another candidate for the wait anaphase signal is While the identity(ies) and mechanism of generation at
BubR1, which directly binds Cdc20 and APC/C elements unattached kinetochores of the Cdc20-APC/C inhibi-
Cdc16, Cdc27, and APC7 (Chan et al., 1999; Wu et al., tor(s) remains imperfectly defined, the activated form
2000) and by doing so can block Cdc20 activation of must be diffusible. The primary evidence for this is sim-
APC/C for mitotic substrates. The inhibitory activity has ple logic: in order to delay anaphase so as to prevent
been argued to be BubR1 alone without a contribution loss of a single chromosome, one unattached kineto-
of Mad2 (Tang et al., 2001) or in a complex (named MCC) chore (for example, on a chromatid pair trapped by
that apparently contains stoichiometric amounts of chance behind one of the spindle poles) must be able
BubR1, Bub3, Mad2, and Cdc20 (Sudakin et al., 2001). to inhibit APC/C-dependent destruction of securin. This
An equivalent quaternary complex containing Mad3 (the requires that the lone unattached kinetochore release a
yeast BubR1), Bub3, Mad2, and Cdc20 has also been sufficiently strong (diffusible) signal to inhibit all Cdc20-
observed in budding yeast (Fraschini et al., 2001; Hard- APC/C.
wick et al., 2000). Further, the much higher APC/C inhibi- But, there are strict limits to such diffusion. By analysis
tory activity in vitro of MCC (⬎3,000-fold more potent of cells containing two distinct spindles (after cell fu-
than tetrameric Mad2) (Sudakin et al., 2001) would make sion), one or a few unattached kinetochores have been
this complex an attractive candidate for a diffusible in- found unable to prevent anaphase in an adjacent spindle
hibitory signal were it not that it (and its yeast counter- in which all kinetochores are attached (Rieder et al.,
part) is formed in a kinetochore-independent manner 1997). The limited range of the APC/C inhibition almost
Review
417

certainly reflects the limited signal strength produced (Gorbsky et al., 1998), is insensitive to inactivation with
by a single unattached kinetochore, coupled to the short Mad2 antibody injection (Skoufias et al., 2001).
natural half-life of the inhibitor(s), whose destruction is This has lead to the proposal that there are two
complete within about 15 min after silencing production branches of checkpoint signaling and silencing. One, in
(Rieder et al., 1995). In addition, after release and local which the signal released is an activated, inhibitory form
diffusion the inhibitor may be concentrated by trafficking of Mad2, is silenced by microtubule attachment, while
to spindle poles by cytoplasmic dynein (see below), the other, presumably involving conversion of BubR1
where APC/C is enriched (Tugendreich et al., 1995). This into a Cdc20 inhibitor, is silenced by tension. This is not
is compatible with the finding that many checkpoint unappealing, but is at best a very murky argument, since
components are rapidly trafficked poleward by dynein attachment and tension are intimately interrelated. It has
after release (Howell et al., 2001). been long known that tension stabilizes attachment (Ault
and Nicklas, 1989). CENP-E elimination, for example,
Silencing the Checkpoint Signal: yields loss of kinetochore tension and fewer (McEwen
Attachment and Tension et al., 2001; Putkey et al., 2002), less stably bound micro-
There is a continuing controversy as to whether the tubules. Moreover, BubR1 (Mad3), Bub1, or Mad2 can-
mitotic checkpoint is silenced by microtubule attach- not suppress the Cdc20-APC/C activity independently
ment (Rieder et al., 1995) or by the tension exerted be- of the others. Thus, it seems to us that rather than sepa-
tween bioriented kinetochore pairs after attachment (Li rate parts of checkpoint silencing, attachment and ten-
and Nicklas, 1995) and whether activities of subsets of sion really represent two halves of the same coin, inte-
the known components are selectively silenced by one grally interrelated.
or the other (Skoufias et al., 2001; Zhou et al., 2002).
McIntosh formally proposed the tension model under
Mitotic Checkpoint Defects Promote Aneuploidy
which the mechanical tension generated by poleward-
and Tumorigenesis
directed forces acting on both kinetochores of a bi-
The mitotic checkpoint in budding yeast, where the in-
oriented chromatid pair turns off checkpoint signaling
tranuclear spindle forms quickly after centromeres are
(McIntosh, 1991). Compelling evidence for a tension re-
duplicated in S, is a real checkpoint activated only to
quirement initially emerged using praying mantid sper-
arrest mitosis in the relatively rare instances when at-
matocytes in meiosis I, which have a Y chromosome
tachment is delayed. In most other organisms, it is not
and two genetically different X chromosomes. Although
a checkpoint at all. Rather, it is an essential cell cycle
the Y is supposed to pair with both X chromosomes,
control pathway activated at every mitosis/meiosis im-
occasionally it pairs only with one, thus yielding a mono-
mediately upon nuclear envelope disassembly. Loss of
oriented X. When this occurs, anaphase onset is delayed
Mad2, Bub3, or Rae1 in mice is lethal early, with cells
by up to 9 hr but can be triggered to initiate almost
accumulating mitotic errors and undergoing apoptosis
immediately by application of mechanical tension ap-
by embryonic day 5 or 6 (Dobles et al., 2000; Kalitsis et
plied across the mono-oriented kinetochore by use of
al., 2000; Babu et al., 2003). Similarly, in Drosophila,
a force-calibrated microneedle (Li and Nicklas, 1995).
loss of Bub1 causes chromosome missegregation and
This is a general finding in meiosis: eliminating tension
lethality (Basu et al., 1999). Microinjection of antibodies
between homologous chromosomes by preventing re-
to Mad2 yields premature anaphase onset and chromo-
combination delayed anaphase onset in yeast, presum-
some missegregation (Gorbsky et al., 1998). Haplo-
ably from checkpoint activation. This delay was elimi-
insufficiency in Mad2 provokes late onset, self-limiting
nated by genetically allowing sister kinetochores to
lung tumors (Michel et al., 2001). Reduction in Bub3
inappropriately separate during meiosis I, thereby
or Rae1 generates aneuploidy in vitro and a sharply
allowing each homolog to biorient in the absence of
increased susceptibility to chemically induced tumori-
recombination and restore tension across the sisters
genesis (Babu et al., 2003). Thus, the primary mission of
(Shonn et al., 2000). Similarly, in cells that enter mitosis
the checkpoint is to prevent such errors in chromosome
without a prior round of DNA replication, the unrepli-
segregation, a hallmark of human tumor progression
cated chromatids attach to spindle microtubules, but
(Hartwell and Kastan, 1994).
no tension can be developed and the mitotic checkpoint
is chronically activated (Stern and Murray, 2001).
This is not the whole story, however. In the initial Conclusions
demonstration that kinetochores are the signaling ele- It is now clear that the centromere and its associated
ments for the checkpoint, destruction of the last unat- kinetochore are much more than simple attachment
tached kinetochore eliminated checkpoint signaling de- sites for spindle microtubules. Mammalian centromeres
spite lack of tension on the kinetochore of the sister are much more complex than initially imagined, repre-
chromatid (Rieder et al., 1995). Similarly, in maize, satis- senting repeated assemblies of the simple, one nucleo-
fying the checkpoint requires tension in meiosis, but in some centromeres of budding yeast. Central to genetic
mitosis, attachment is sufficient (Yu et al., 1999). Further- inheritance, in almost all examples known they are them-
more, in PtK1 cells, loss of Mad2 recruitment to kineto- selves determined not by DNA sequence, but by one or
chores depends on microtubule attachment, not tension more epigenetic elements. They are active components
(Waters et al., 1998). Similarly, very low doses of the in microtubule capture, stabilization, and in powering
microtubule inhibitor vinblastine produce loss of tension chromosome movements essential to proper segrega-
and kinetochore bound Mad2 and a checkpoint arrest tion. More than that, they are the signaling elements for
that, unlike the case after inhibition of spindle assembly controlling cell cycle advance through mitosis.
Cell
418

References Cheeseman, I.M., Anderson, S., Jwa, M., Green, E.M., Kang, J.,
Yates, J.R., Chan, C.S., Drubin, D.G., and Barnes, G. (2002a). Phos-
Abrieu, A., Kahana, J.A., Wood, K.W., and Cleveland, D.W. (2000). pho-regulation of kinetochore-microtubule attachments by the au-
CENP-E as an essential component of the mitotic checkpoint in rora kinase ipl1p. Cell 111, 163–172.
vitro. Cell 102, 817–826. Cheeseman, I.M., Drubin, D.G., and Barnes, G. (2002b). Simple cen-
Abrieu, A., Magnaghi-Jaulin, L., Kahana, J.A., Peter, M., Castro, A., tromere, complex kinetochore: linking spindle microtubules and
Vigneron, S., Lorca, T., Cleveland, D.W., and Labbe, J. (2001). Mps1 centromeric DNA in budding yeast. J. Cell Biol. 157, 199–203.
is a kinetochore-associated kinase essential for the vertebrate mi- Chen, R.H., Waters, J.C., Salmon, E.D., and Murray, A.W. (1996).
totic checkpoint. Cell 106, 83–93. Association of spindle assembly checkpoint component XMAD2
Afshar, K., Barton, N.R., Hawley, R.S., and Goldstein, L.S. (1995). with unattached kinetochores. Science 274, 242–246.
DNA binding and meiotic chromosomal localization of the Drosoph- Chen, R.H., Shevchenko, A., Mann, M., and Murray, A.W. (1998).
ila nod kinesin-like protein. Cell 81, 129–138. Spindle checkpoint protein Xmad1 recruits Xmad2 to unattached
Alexander, S.P., and Rieder, C.L. (1991). Chromosome motion during kinetochores. J. Cell Biol. 143, 283–295.
attachment to the vertebrate spindle: initial saltatory-like behavior Clarke, L. (1998). Centromeres: proteins, protein complexes, and
of chromosomes and quantitative analysis of force production by repeated domains at centromeres of simple eukaryotes. Curr. Opin.
nascent kinetochore fibers. J. Cell Biol. 113, 805–815. Genet. Dev. 8, 212–218.
Ando, S., Yang, H., Nozaki, N., Okazaki, T., and Yoda, K. (2002). Cohen-Fix, O., Peters, J.M., Kirschner, M.W., and Koshland, D.
CENP-A, -B, and -C chromatin complex that contains the I-type (1996). Anaphase initiation in Saccharomyces cerevisiae is con-
alpha-satellite array constitutes the prekinetochore in HeLa cells. trolled by the APC-dependent degradation of the anaphase inhibitor
Mol. Cell. Biol. 22, 2229–2241. Pds1p. Genes Dev. 10, 3081–3093.
Antonio, C., Ferby, I., Wilhelm, H., Jones, M., Karsenti, E., Nebreda, Dawson, I.A., Roth, S., Akam, M., and Artavanis-Tsakonas, S. (1993).
A.R., and Vernos, I. (2000). Xkid, a chromokinesin required for chro- Mutations of the fizzy locus cause metaphase arrest in Drosophila
mosome alignment on the metaphase plate. Cell 102, 425–435. melanogaster embryos. Development 117, 359–376.
Ault, J.G., and Nicklas, R.B. (1989). Tension, microtubule re- Desai, A., Maddox, P.S., Mitchison, T.J., and Salmon, E.D. (1998).
arrangements, and the proper distribution of chromosomes in mito- Anaphase A chromosome movement and poleward spindle microtu-
sis. Chromosoma 98, 33–39. bule flux occur at similar rates in Xenopus extract spindles. J. Cell
Babu, J.R., Jeganathan, K.B., Baker, D.J., Wu, X., Kang-Decker, N., Biol. 141, 703–713.
and van Deursen, J.M. (2003). Rae1 is an essential mitotic check- Desai, A., Verma, S., Mitchison, T.J., and Walczak, C.E. (1999). Kin
point regulator that cooperates with Bub3 to prevent chromosome I kinesins are microtubule-destabilizing enzymes. Cell 96, 69–78.
missegregation. J. Cell Biol. 160, in press. Published online January Dobles, M., Liberal, V., Scott, M.L., Benezra, R., and Sorger, P.K.
27, 2003. 10.1083/jbc.200211048. (2000). Chromosome missegregation and apoptosis in mice lacking
Basu, J., Bousbaa, H., Logarinho, E., Li, Z., Williams, B.C., Lopes, C., the mitotic checkpoint Mad2. Cell 101, 635–645.
Sunkel, C.E., and Goldberg, M.L. (1999). Mutations in the essential Ekwall, K., Olsson, T., Turner, B.M., Cranston, G., and Allshire, R.C.
spindle checkpoint bub1 cause chromosome missegregation and (1997). Transient inhibition of histone deacetylation alters the struc-
fail to block apoptosis in Drosophila. J. Cell Biol. 46, 13–28. tural and functional imprint at fission yeast centromeres. Cell 91,
Barry, A.E., Howman, E.V., Cancilla, M.R., Saffery, R., and Choo, 1021–1032.
K.H. (1999). Sequence analysis of an 80 kb human neocentromere. Espelin, C.W., Kaplan, K.B., and Sorger, P.K. (1997). Probing the
Hum. Mol. Genet. 8, 217–227. architecture of a simple kinetochore using DNA-protein crosslinking.
Basto, R., Gomes, R., and Karess, R. (2000). Rough Deal and ZW10 J. Cell Biol. 139, 1383–1396.
are required for the Metaphase checkpoint in Drosophila. Nat. Cell Fang, G., Yu, H., and Kirschner, M.W. (1998). The checkpoint protein
Biol. 2, 939–943. MAD2 and the mitotic regulator CDC20 form a ternary complex with
Bernard, P., Maure, J.F., Partridge, J.F., Genier, S., Javerzat, J.P., the anaphase-promoting complex to control anaphase initiation.
and Allshire, R.C. (2001). Requirement of heterochromatin for cohe- Genes Dev. 12, 1871–1883.
sion at centromeres. Science 294, 2539–2542. Fraschini, R., Beretta, A., Sironi, L., Musacchio, A., Lucchini, G., and
Biggins, S., Severin, F.F., Bhalla, N., Sassoon, I., Hyman, A.A., and Piatti, S. (2001). Bub3 interaction with Mad2, Mad3 and Cdc20 is
Murray, A.W. (1999). The conserved protein kinase Ipl1 regulates mediated by WD40 repeats and does not require intact kineto-
microtubule binding to kinetochores in budding yease. Genes Dev. chores. EMBO J. 20, 6648–6659.
13, 532–544. Funabiki, H., and Murray, A.W. (2000). The Xenopus chromokinesin
Blower, M.D., Sullivan, B.A., and Karpen, G.H. (2002). Conserved Xkid is essential for metaphase chromosome alignment and must
organization of centromeric chromatin in flies and humans. Dev. be degraded to allow anaphase chromosome movement. Cell 102,
Cell 2, 319–330. 411–424.
Bolton, M.A., Lan, W., Powers, S.E., McCleland, M.L., Kuang, J., Gardner, G.D., Poddar, A., Yellman, C., Tavormina, P.A., Monte-
and Stukenberg, P.T. (2002). Aurora B kinase exists in a complex agudo, M.C., and Burke, D.J. (2001). The spindle checkpoint of the
with survivin and INCENP and its kinase activity is stimulated by yeast Saccharomyces cerevisiae requires kinetochore function and
survivin binding and phosphorylation. Mol. Biol. Cell 13, 3064–3077. maps to the CBF3 domain. Genetics 157, 1493–1502.
Brown, K.E., Barnett, M.A., Burgtorf, C., Shaw, P., Buckle, V.J., and Gorbsky, G.J., Sammak, P.J., and Borisy, G.G. (1987). Chromosomes
Brown, W.R. (1994). Dissecting the centromere of the human Y chro- move poleward in anaphase along stationary microtubules that co-
mosome with cloned telomeric DNA. Hum. Mol. Genet. 3, 1227–1237. ordinately disassemble from their kinetochore ends. J. Cell Biol.
Chan, G.K., Jablonski, S.A., Sudakin, V., Hittle, J.C., and Yen, T.J. 104, 9–18.
(1999). Human BUBR1 is a mitotic checkpoint kinase that monitors Gorbsky, G.J., Chen, R.H., and Murray, A.W. (1998). Microinjection
CENP-E functions at kinetochores and binds the cyclosome/APC. of antibody to Mad2 protein into mammalian cells in mitosis induces
J. Cell Biol. 146, 941–954. premature anaphase. J. Cell Biol. 141, 1193–1205.
Chan, G.K., Jablonski, S.A., Starr, D.A., Goldberg, M.L., and Yen, Goshima, G., Saitoh, S., and Yanagida, M. (1999). Proper metaphase
T.J. (2000). Human Zw10 and ROD are mitotic checkpoint proteins spindle length is determined by centromere proteins Mis12 and
that bind to kinetochores. Nat. Cell Biol. 2, 944–947. Mis6 required for faithful chromosome segregation. Genes Dev. 13,
Cheeseman, I.M., Brew, C., Wolyniak, M., Desai, A., Anderson, S., 1664–1677.
Muster, N., Yates, J.R., Huffaker, T.C., Drubin, D.G., and Barnes, G. Hardwick, K.G., Johnston, R.C., Smith, D.L., and Murray, A.W. (2000).
(2001). Implication of a novel multiprotein Dam1p complex in outer MAD3 encodes a novel component of the spindle checkpoint which
kinetochore function. J. Cell Biol. 155, 1137–1145. interacts with Bub3p, Cdc20p, and Mad2p. J. Cell Biol. 148, 871–882.
Review
419

Harrington, J.A., Van Bokkelen, G., Mays, R.W., Gustashaw, K., and ochores and centrosomes in mammalian cells. J. Cell Biol. 158,
Willard, H.F. (1997). Formation of de novo centromeres and con- 841–847.
struction of first-generation human artificial microchromosomes. Kaplan, K.B., Burds, A.A., Swedlow, J.R., Bekir, S.S., Sorger, P.K.,
Nat. Genet. 15, 345–355. and Nathke, I.S. (2001). A role for the Adenomatous Polyposis Coli
Hartwell, L.H., and Kastan, M.B. (1994). Cell cycle control and can- protein in chromosome segregation. Nat. Cell Biol. 3, 429–432.
cer. Science 266, 1821–1828. Keith, K.C., and Fitzgerald-Hayes, M. (2000). CSE4 genetically inter-
Hays, T.S., Wise, D., and Salmon, E.D. (1982). Traction force on a acts with the Saccharomyces cerevisiae centromere DNA elements
kinetochore at metaphase acts as a linear function of kinetochore CDE I and CDE II but not CDE III. Implications for the path of the
fiber length. J. Cell Biol. 93, 374–389. centromere dna around a cse4p variant nucleosome. Genetics 156,
He, D., Zeng, C., Woods, K., Zhong, L., Turner, D., Busch, R.K., 973–981.
Brinkley, B.R., and Busch, H. (1998). CENP-G: a new centromeric Khodjakov, A., and Rieder, C.L. (1996). Kinetochores moving away
protein that is associated with the alpha-1 satellite DNA subfamily. from their associated pole do not exert a significant pushing force
Chromosoma 107, 189–197. on the chromosome. J. Cell Biol. 135, 315–327.
He, X., Asthana, S., and Sorger, P.K. (2000). Transient sister chroma- Khodjakov, A., Cole, R.W., McEwen, B.F., Buttle, K.F., and Rieder,
tid separation and elastic deformation of chromosomes during mito- C.L. (1997). Chromosome fragments possessing only one kineto-
sis in budding yeast. Cell 101, 763–775. chore can congress to the spindle equator. J. Cell Biol. 136, 229–240.
Hemmerich, P., Stoyan, T., Wieland, G., Koch, M., Lechner, J., and Kim, S.H., Lin, D.P., Matsumoto, S., Kitazono, A., and Matsumoto,
Diekmann, S. (2000). Interaction of yeast kinetochore proteins with T. (1998). Fission yeast Slp1: an effector of the Mad2-dependent
centromere-protein/transcription factor Cbf1. Proc. Natl. Acad. Sci. spindle checkpoint. Science 279, 1045–1047.
USA 97, 12583–12588. Lachner, M., O’Carroll, D., Rea, S., Mechtler, K., and Jenuwein, T.
Henikoff, S., Ahmad, K., and Malik, H.S. (2001). The centromere (2001). Methylation of histone H3 lysine 9 creates a binding site for
paradox: stable inheritance with rapidly evolving DNA. Science 293, HP1 proteins. Nature 410, 116–120.
1098–1102. Laloraya, S., Guacci, V., and Koshland, D. (2000). Chromosomal
Howell, B.J., Hoffman, D.B., Fang, G., Murray, A.W., and Salmon, addresses of the cohesin component Mcd1p. J. Cell Biol. 151, 1047–
E.D. (2000). Visualization of mad2 dynamics at kinetochores, along 1056.
spindles fibers, and at spindle poles in living cells. J. Cell Biol. 150, Lee, S., Wisniewski, J.C., Dentler, W.L., and Asai, D.J. (1999). Gene
1233–1250. knockouts reveal separate functions for two cytoplasmic dyneins
Howell, B.J., McEwen, B.F., Canman, J.C., Hoffman, D.B., Farrar, in Tetrahymena thermophila. Mol. Biol. Cell 10, 771–784.
E.M., Rieder, C.L., and Salmon, E.D. (2001). Cytoplasmic dynein/ Levesque, A.A., and Compton, D.A. (2001). The chromokinesin Kid
dynactin drives kinetochore protein transport to the spindle poles is necessary for chromosome arm orientation and oscillation, but
and has a role in mitotic spindle checkpoint inactivation. J. Cell Biol. not congression, on mitotic spindles. J. Cell Biol. 154, 1135–1146.
155, 1159–1172.
Li, R., and Murray, A.W. (1991). Feedback control of mitosis in bud-
Howman, E.V., Fowler, K.J., Newson, A.J., Redward, S., MacDonald, ding yeast. Cell 66, 519–531.
A.C., Kalitsis, P., and Choo, K.H. (2000). Early disruption of centro-
meric chromatin organization in centromere protein A (Cenpa) null Li, X., and Nicklas, R.B. (1995). Mitotic forces control a cell-cycle
mice. Proc. Natl. Acad. Sci. USA 97, 1148–1153. checkpoint. Nature 373, 630–632.

Hoyt, M.A., Totis, L., and Roberts, B.T. (1991). S. cerevisiae genes Li, Y., and Benezra, R. (1996). Identification of a human mitotic
required for cell cycle arrest in response to loss of microtubule checkpoint gene: hsMAD2. Science 274, 246–248.
function. Cell 66, 507–517. Li, Y., Bachant, J., Alcasabas, A.A., Wang, Y., Qin, J., and Elledge,
Hudson, D.F., Fowler, K.J., Earle, E., Saffery, R., Kalitsis, P., Trowell, S.J. (2002). The mitotic spindle is required for loading of the DASH
H., Hill, J., Wreford, N.G., de Kretser, D.M., Cancilla, M.R., et al. complex onto the kinetochore. Genes Dev. 16, 183–197.
(1998). Centromere protein B null mice are mitotically and meiotically Lo, A.W., Magliano, D.J., Sibson, M.C., Kalitsis, P., Craig, J.M., and
normal but have lower body and testis weights. J. Cell Biol. 141, Choo, K.H. (2001). A novel chromatin immunoprecipitation and array
309–319. (CIA) analysis identifies a 460-kb CENP-A-binding neocentromere
Hyman, A.A., and Mitchison, T.J. (1991). Two different microtubule- DNA. Genome Res. 11, 448–457.
based motor activities with opposite polarities in kinetochores. Na- Lombillo, V.A., Nislow, C., Yen, T.J., Gelfand, V.I., and McIntosh,
ture 351, 206–211. J.R. (1995). Antibodies to the kinesin motor domain and CENP-E
Ikeno, M., Grimes, B., Okazaki, T., Nakano, M., Saitoh, K., Hoshino, inhibit microtubule depolymerization-dependent motion of chromo-
H., McGill, N.I., Cooke, H., and Masumoto, H. (1998). Construction somes in vitro. J. Cell Biol. 128, 107–115.
of YAC-based mammalian artificial chromosomes. Nat. Biotechnol. Luger, K., Mader, A.W., Richmond, R.K., Sargent, D.F., and Rich-
16, 431–439. mond, T.J. (1997). Crystal structure of the nucleosome core particle
Janke, C., Ortiz, J., Lechner, J., Shevchenko, A., Shevchenko, A., at 2.8 Å resolution. Nature 389, 251–260.
Magiera, M.M., Schramm, C., and Schiebel, E. (2001). The budding Maddox, P., Desai, A., Oegema, K., Mitchison, T.J., and Salmon,
yeast proteins Spc24p and Spc25p interact with Ndc80p and Nuf2p E.D. (2002). Poleward microtubule flux is a major component of
at the kinetochore and are important for kinetochore clustering and spindle dynamics and anaphase a in mitotic Drosophila embryos.
checkpoint control. EMBO J. 20, 777–791. Curr. Biol. 12, 1670–1674.
Janke, C., Ortiz, J., Tanaka, T.U., Lechner, J., and Schiebel, E. (2002). Mallavarapu, A., Sawin, K., and Mitchison, T. (1999). A switch in
Four new subunits of the Dam1-Duo1 complex reveal novel func- microtubule dynamics at the onset of anaphase B in the mitotic
tions in sister kinetochore biorientation. EMBO J. 21, 181–193. spindle of Schizosaccharomyces pombe. Curr. Biol. 9, 1423–1426.
Kalitsis, P., Earle, E., Fowler, K.J., and Choo, K.H. (2000). Bub3 Martin-Lluesma, S., Stucke, V.M., and Nigg, E.A. (2002). Role of
gene disruption in mice reveals essential mitotic spindle checkpoint hec1 in spindle checkpoint signaling and kinetochore recruitment
function during early embryogenesis. Genes Dev. 14, 2277–2282. of mad1/mad2. Science 297, 2267–2270.
Kallio, M., Weinstein, J., Daum, J.R., Burke, D.J., and Gorbsky, G.J. McEwen, B.F., Chan, G.K.T., Zubrowski, B., Savoian, M.S., Sauer,
(1998). Mammalian p55CDC mediates association of the spindle M.T., and Yen, T.J. (2001). CENP-E is essential for reliable bioriented
checkpoint protein Mad2 with the cyclosome/anaphase-promoting spindle attachment, but chromosome alignment can be achieved
complex, and is involved in regulating anaphase onset and late via redundant mechanisms in mammalian cells. Mol. Biol. Cell 12,
mitotic events. J. Cell Biol. 141, 1393–1406. 2776–2789.
Kallio, M.J., Beardmore, V.A., Weinstein, J., and Gorbsky, G.J. McIntosh, J.R. (1991). Structural and mechanical control of mitotic
(2002). Rapid microtubule-independent dynamics of Cdc20 at kinet- progression. Cold Spring Harb. Symp. Quant. Biol. 56, 613–619.
Cell
420

Megee, P.C., Mistrot, C., Guacci, V., and Koshland, D. (1999). The monitors sister kinetochore attachment to the spindle. J. Cell Biol.
centromeric sister chromatid cohesion site directs Mcd1p binding 127, 1301–1310.
to adjacent sequences. Mol. Cell 4, 445–450. Rieder, C.L., Cole, R.W., Khodjakov, A., and Sluder, G. (1995). The
Meluh, P.B., and Koshland, D. (1995). Evidence that the MIF2 gene checkpoint delaying anaphase in response to chromosome mono-
of Saccharomyces cerevisiae encodes a centromere protein with orientation is mediated by an inhibitory signal produced by unat-
homology to the mammalian centromere protein CENP-C. Mol. Biol. tached kinetochores. J. Cell Biol. 130, 941–948.
Cell 6, 793–807. Rieder, C.L., Khodjakov, A., Paliulis, L.V., Fortier, T.M., Cole, R.W.,
Meluh, P.B., Yang, P., Glowczewski, L., Koshland, D., and Smith, and Sluder, G. (1997). Mitosis in vertebrate somatic cells with two
M.M. (1998). Cse4p is a component of the core centromere of Sac- spindles: implications for the metaphase/anaphase transition
charomyces cerevisiae. Cell 94, 607–613. checkpoint and cleavage. Proc. Natl. Acad. Sci. USA 94, 5107–5112.
Michel, L.S., Liberal, V., Chatterjee, A., Kirchwegger, R., Pasche, B., Roberts, B.T., Farr, K.A., and Hoyt, M.A. (1994). The Saccharomyces
Gerald, W., Dobles, M., Sorger, P.K., Murty, V.V., and Benezra, R. cerevisiae checkpoint gene BUB1 encodes a novel protein kinase.
(2001). Mad2 haplo-insufficiency causes premature anaphase and Mol. Cell. Biol. 14, 8282–8291.
chromosome instability in mammalian cells. Nature 409, 355–359.
Rogers, S.L., Rogers, G.C., Sharp, D.J., and Vale, R.D. (2002). Dro-
Mitchison, T.J., and Salmon, E.D. (1992). Poleward kinetochore fiber sophila EB1 is important for proper assembly, dynamics, and posi-
movement occurs during both metaphase and anaphase-A in newt tioning of the mitotic spindle. J. Cell Biol. 158, 873–884.
lung cell mitosis. J. Cell Biol. 119, 569–582.
Russell, I.D., Grancell, A.S., and Sorger, P.K. (1999). The unstable
Musacchio, A., and Hardwick, K.G. (2002). The spindle checkpoint: F-box protein p58-Ctf13 forms the structural core of the CBF3 kinet-
structural insights into dynamic signalling. Nat. Rev. Mol. Cell Biol. ochore complex. J. Cell Biol. 145, 933–950.
3, 731–741.
Saffery, R., Irvine, D.V., Griffiths, B., Kalitsis, P., Wordeman, L., and
Nakagawa, H., Lee, J.K., Hurwitz, J., Allshire, R.C., Nakayama, J., Choo, K.H. (2000). Human centromeres and neocentromeres show
Grewal, S.I., Tanaka, K., and Murakami, Y. (2002). Fission yeast identical distribution patterns of ⬎20 functionally important kineto-
CENP-B homologs nucleate centromeric heterochromatin by pro- chore-associated proteins. Hum. Mol. Genet. 9, 175–185.
moting heterochromatin-specific histone tail modifications. Genes
Savoian, M.S., Goldberg, M.L., and Rieder, C.L. (2000). The rate of
Dev. 16, 1766–1778.
chromosome poleward motion is attenuated in Drosophila zw10 and
Ngan, V.K., and Clarke, L. (1997). The centromere enhancer mediates rod mutants. Nat. Cell Biol. 2, 948–952.
centromere activation in Schizosaccharomyces pombe. Mol. Cell.
Schueler, M.G., Higgins, A.W., Rudd, M.K., Gustashaw, K., and Wil-
Biol. 17, 3305–3314.
lard, H.F. (2001). Genomic and genetic definition of a functional
Nicklas, R.B. (1983). Measurements of the force produced by the human centromere. Science 294, 109–115.
mitotic spindle in anaphase. J. Cell Biol. 97, 542–548.
Sharp, D.J., Rogers, G.C., and Scholey, J.M. (2000). Cytoplasmic
Nicklas, R.B. (1988). The forces that move chromosomes in mitosis.
dynein is required for poleward chromosome movement during mi-
Annu. Rev. Biophys. Biophys. Chem. 17, 431–449.
tosis in Drosophila embryos. Nat. Cell Biol. 2, 922–930.
Oegema, K., Desai, A., Rybina, S., Kirkham, M., and Hyman, A.A.
Shelby, R.D., Vafa, O., and Sullivan, K.F. (1997). Assembly of CENP-A
(2001). Functional analysis of kinetochore assembly in Caenorhab-
into centromeric chromatin requires a cooperative array of nucleo-
ditis elegans. J. Cell Biol. 153, 1209–1226.
somal DNA contact sites. J. Cell Biol. 136, 501–513.
Ohzeki, J., Nakano, M., Okada, T., and Masumoto, H. (2002). CENP-B
Shelby, R.D., Monier, K., and Sullivan, K.F. (2000). Chromatin assem-
box is required for de novo centromere chromatin assembly on
bly at kinetochores is uncoupled from DNA replication. J. Cell Biol.
human alphoid DNA. J. Cell Biol. 159, 765–775.
151, 1113–1118.
Ortiz, J., Stemmann, O., Rank, S., and Lechner, J. (1999). A putative
Shingyoji, C., Higuchi, H., Yoshimura, M., Katayama, E., and Yanag-
protein complex consisting of Ctf19, Mcm21, and Okp1 represents
ida, T. (1998). Dynein arms are oscillating force generators. Nature
a missing link in the budding yeast kinetochore. Genes Dev. 13,
393, 711–714.
1140–1155.
Shonn, M.A., McCarroll, R., and Murray, A.W. (2000). Requirement
Ostergren, G. (1951). The mechanism of co-orientation in bivalents
of the spindle checkpoint for proper chromosome segregation in
and multivalents. Hereditas 37, 85–156.
budding yeast meiosis. Science 289, 300–303.
Partridge, J.F., Borgstrom, B., and Allshire, R.C. (2000). Distinct
protein interaction domains and protein spreading in a complex Skibbens, R.V., Skeen, V.P., and Salmon, E.D. (1993). Directional
centromere. Genes Dev. 14, 783–791. instability of kinetochore motility during chromosome congression
and segregation in mitotic newt lung cells: a push-pull mechanism.
Partridge, J., Scott, K., Bannister, A., Kouzarides, T., and Allshire,
J. Cell Biol. 122, 859–875.
R. (2002). cis-acting DNA from fission yeast centromeres mediates
histone H3 methylation and recruitment of silencing factors and Skoufias, D.A., Andreassen, P.R., Lacroix, F.B., Wilson, L., and Mar-
cohesin to an ectopic site. Curr. Biol. 12, 1652–1660. golis, R.L. (2001). Mammalian mad2 and bub1/bubR1 recognize dis-
tinct spindle-attachment and kinetochore-tension checkpoints.
Putkey, F.R., Cramer, T., Morphew, M.K., Silk, A.D., Johnson, R.S.,
Proc. Natl. Acad. Sci. USA 98, 4492–4497.
McIntosh, J.R., and Cleveland, D.W. (2002). Unstable kinetochore-
microtubule capture and chromosomal instability following deletion Smith, M.M. (2002). Centromeres and variant histones: what, where,
of CENP-E. Dev. Cell 3, 351–365. when and why? Curr. Opin. Cell Biol. 14, 279–285.
Raff, J.W., Jeffers, K., and Huang, J.Y. (2002). The roles of Fzy/ Stern, B.M., and Murray, A.W. (2001). Lack of tension at kinetochores
Cdc20 and Fzr/Cdh1 in regulating the destruction of cyclin B in activates the spindle checkpoint in budding yeast. Curr. Biol. 11,
space and time. J. Cell Biol. 157, 1139–1149. 1462–1467.
Richards, E.J., and Elgin, S.C. (2002). Epigenetic codes for hetero- Stoler, S., Keith, K.C., Curnick, K.E., and Fitzgerald-Hayes, M. (1995).
chromatin formation and silencing: rounding up the usual suspects. A mutation in CSE4, an essential gene encoding a novel chromatin-
Cell 108, 489–500. associated protein in yeast, causes chromosome nondisjunction
and cell cycle arrest at mitosis. Genes Dev. 9, 573–586.
Rieder, C.L. (1981). The structure of the cold-stable kinetochore
fiber in metaphase PtK1 cells. Chromosoma 84, 145–158. Su, L.K., Burrell, M., Hill, D.E., Gyuris, J., Brent, R., Wiltshire, R.,
Rieder, C.L., and Alexander, S.P. (1990). Kinetochores are trans- Trent, J., Vogelstein, B., and Kinzler, K.W. (1995). APC binds to the
ported poleward along a single astral microtubule during chromo- novel protein EB1. Cancer Res. 55, 2972–2977.
some attachment to the spindle in newt lung cells. J. Cell Biol. 110, Sudakin, V., Chan, G.K., and Yen, T.J. (2001). Checkpoint inhibition
81–95. of the APC/C in HeLa cells is mediated by a complex of BUBR1,
Rieder, C.L., Schultz, A., Cole, R., and Sluder, G. (1994). Anaphase BUB3, CDC20, and MAD2. J. Cell Biol. 154, 925–936.
onset in vertebrate somatic cells is controlled by a checkpoint that Sullivan, B., and Karpen, G. (2001). Centromere identity in Drosophila
Review
421

is not determined in vivo by replication timing. J. Cell Biol. 154, Waters, J.C., Chen, R.H., Murray, A.W., and Salmon, E.D. (1998).
683–690. Localization of Mad2 to kinetochores depends on microtubule at-
Sullivan, B.A., and Willard, H.F. (1998). Stable dicentric X chromo- tachment, not tension. J. Cell Biol. 141, 1181–1191.
somes with two functional centromeres. Nat. Genet. 20, 227–228. Weiss, E., and Winey, M. (1996). The Saccharomyces cerevisiae
Sullivan, B.A., Blower, M.D., and Karpen, G.H. (2001). Determining spindle pole body duplication gene MPS1 is part of a mitotic check-
centromere identity: cyclical stories and forking paths. Nat. Rev. point. J. Cell Biol. 132, 111–123.
Genet. 2, 584–596. Wigge, P.A., and Kilmartin, J.V. (2001). The Ndc80p complex from
Sun, X., Wahlstrom, J., and Karpen, G.H. (1997). Molecular structure Saccharomyces cerevisiae contains conserved centromere compo-
of a functional Drosophila centromere. Cell 91, 1007–1019. nents and has a function in chromosome segregation. J. Cell Biol.
152, 349–360.
Takahashi, K., Chen, E.S., and Yanagida, M. (2000). Requirement of
Mis6 centromere connector for localizing a CENP-A-like protein in Williams, B.C., Gatti, M., and Goldberg, M.L. (1996). Bipolar spindle
fission yeast. Science 288, 2215–2219. attachments affect redistributions of ZW10, a Drosophila centro-
mere/kinetochore component required for accurate chromosome
Takenaka, K., Moriguchi, T., and Nishida, E. (1998). Activation of the
segregation. J. Cell Biol. 134, 1127–1140.
protein kinase p38 in the spindle assembly checkpoint and mitotic
arrest. Science 280, 599–602. Williams, B.C., Murphy, T.D., Goldberg, M.L., and Karpen, G.H.
(1998). Neocentromere activity of structurally acentric mini-chromo-
Tanaka, T., Cosma, M.P., Wirth, K., and Nasmyth, K. (1999). Identifi-
somes in Drosophila. Nat. Genet. 18, 30–37.
cation of cohesin association sites at centromeres and along chro-
mosome arms. Cell 98, 847–858. Wood, K.W., Sakowicz, R., Goldstein, L.S., and Cleveland, D.W.
(1997). CENP-E is a plus end-directed kinetochore motor required
Tanaka, T.U., Rachidi, N., Janke, C., Pereira, G., Galova, M., Schie-
for metaphase chromosome alignment. Cell 91, 357–366.
bel, E., Stark, M.J., and Nasmyth, K. (2002). Evidence that the Ipl1-
Sli15 (Aurora kinase-INCENP) complex promotes chromosome bi- Wordeman, L., and Mitchison, T.J. (1995). Identification and partial
orientation by altering kinetochore-spindle pole connections. Cell characterization of mitotic centromere-associated kinesin, a
108, 317–329. kinesin-related protein that associates with centromeres during mi-
tosis. J. Cell Biol. 128, 95–104.
Tang, Z., Bharadwaj, R., Li, B., and Yu, H. (2001). Mad2-independent
inhibition of APCcdc20 by the mitotic checkpoint protein Bubr1. Dev. Wu, H., Lan, Z., Li, W., Wu, S., Weinstein, J., Sakamoto, K.M., and
Cell 1, 227–237. Dai, W. (2000). p55CDC/hCDC20 is associated with BUBR1 and may
be a downstream target of the spindle checkpoint kinase. Oncogene
Taylor, S.S., and McKeon, F. (1997). Kinetochore localization of mu-
19, 4557–4562.
rine Bub1 is required for normal mitotic timing and checkpoint re-
sponse to spindle damage. Cell 89, 727–735. Yeh, E., Skibbens, R.V., Cheng, J.W., Salmon, E.D., and Bloom, K.
(1995). Spindle dynamics and cell cycle regulation of dynein in the
Taylor, S.S., Ha, E., and McKeon, F. (1998). The human homologue
budding yeast, Saccharomyces cerevisiae. J. Cell Biol. 130,
of Bub3 is required for kinetochore localization of Bub1 and a Mad3/
687–700.
Bub1-related protein kinase. J. Cell Biol. 142, 1–11.
Yu, H.G., Muszynski, M.G., and Kelly Dawe, R. (1999). The maize
Tirnauer, J.S., O’Toole, E., Berrueta, L., Bierer, B.E., and Pellman,
homologue of the cell cycle checkpoint protein MAD2 reveals kinet-
D. (1999). Yeast Bim1p promotes the G1-specific dynamics of micro-
ochore substructure and contrasting mitotic and meiotic localization
tubules. J. Cell Biol. 145, 993–1007.
patterns. J. Cell Biol. 145, 425–435.
Tirnauer, J.S., Canman, J.C., Salmon, E.D., and Mitchison, T.J.
Zhou, J., Panda, D., Landen, J.W., Wilson, L., and Joshi, H.C. (2002).
(2002). EB1 targets to kinetochores with attached, polymerizing mi-
Minor alteration of microtubule dynamics causes loss of tension
crotubules. Mol. Biol. Cell 13, 4308–4316.
across kinetochore pairs and activates the spindle checkpoint. J.
Tomonaga, T., Nagao, K., Kawasaki, Y., Furuya, K., Murakami, A., Biol. Chem. 277, 17200–17208.
Morishita, J., Yuasa, T., Sutani, T., Kearsey, S.E., Uhlmann, F., et al.
Zinkowski, R.P., Meyne, J., and Brinkley, B.R. (1991). The centro-
(2000). Characterization of fission yeast cohesin: essential anaphase
mere-kinetochore complex: a repeat subunit model. J. Cell Biol.
proteolysis of rad21 phosphorylated in the S phase. Genes Dev. 14,
113, 1091–1110.
2757–2770.
Zou, H., McGarry, T.J., Bernal, T., and Kirschner, M.W. (1999). Identi-
Tugendreich, S., Tomkiel, J., Earnshaw, W., and Hieter, P. (1995).
fication of a vertebrate sister-chromatid separation inhibitor in-
CDC27Hs colocalizes with CDC16Hs to the centrosome and mitotic
volved in transformation and tumorigenesis. Science 285, 418–422.
spindle and is essential for the metaphase to anaphase transition.
Cell 81, 261–268. Zumbrunn, J., Inoshita, K., Hyman, A.A., and Nathke, I.S. (2001).
Binding of the Adenomatous Polyposis Coli protein to microtubules
Tyler-Smith, C., Oakey, R.J., Larin, Z., Fisher, R.B., Crocker, M.,
increases microtubules stability and is regulated by GSK3b phos-
Affara, N.A., Ferguson-Smith, M.A., Muenke, M., Zuffardi, O., and
phorylation. Curr. Biol. 11, 44–49.
Jobling, M.A. (1993). Localization of DNA sequences required for
Zur, A., and Brandeis, M. (2001). Securin degradation is mediated
human centromere function through an analysis of rearranged Y
chromosomes. Nat. Genet. 5, 368–375. by fzy and fzr, and is required for complete chromatid separation
but not for cytokinesis. EMBO J. 20, 792–801.
Tyler-Smith, C., Gimelli, G., Giglio, S., Floridia, G., Pandya, A., Terzoli,
G., Warburton, P.E., Earnshaw, W.C., and Zuffardi, O. (1999). Trans-
mission of a fully functional human neocentromere through three
generations. Am. J. Hum. Genet. 64, 1440–1444.
Vafa, O., and Sullivan, K.F. (1997). Chromatin containing CENP-A
and alpha-satellite DNA is a major component of the inner kineto-
chore plate. Curr. Biol. 7, 897–900.
Volpe, T.A., Kidner, C., Hall, I.M., Teng, G., Grewal, S.I., and Martiens-
sen, R.A. (2002). Regulation of heterochromatic silencing and his-
tone H3 lysine-9 methylation by RNAi. Science 297, 1833–1837.
Walczak, C.E., Mitchison, T.J., and Desai, A. (1996). XKCM1: a Xeno-
pus kinesin-related protein that regulates microtubule dynamics
during mitotic spindle assembly. Cell 84, 37–47.
Walters, J.C., Mitchison, T.J., Rieder, C.L., and Salmon, E.D. (1996).
The kinetochore microtubule minus-end disassembly associated
with poleward flux produces a force that can do work. Mol. Biol.
Cell 7, 1547–1558.

Вам также может понравиться