Download REVIEW Centrosome replication, genomic instability and

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts
no text concepts found
Transcript
Leukemia (2002) 16, 767–775
 2002 Nature Publishing Group All rights reserved 0887-6924/02 $25.00
www.nature.com/leu
REVIEW
Centrosome replication, genomic instability and cancer
A Krämer, K Neben and AD Ho
Medizinische Klinik und Poliklinik V, Ruprecht-Karls-Universität, Heidelberg, Germany
Karyotypic alterations, including whole chromosome loss or
gain, ploidy changes, and a variety of chromosome aberrations
are common in cancer cells. If proliferating cells fail to coordinate centrosome duplication with DNA replication, this will
inevitably lead to a change in ploidy, and the formation of monopolar or multipolar spindles will generally provoke abnormal
segregation of chromosomes. Indeed, it has long been recognized that errors in the centrosome duplication cycle may be
an important cause of aneuploidy and thus contribute to cancer
formation. This view has recently received fresh impetus with
the description of supernumerary centrosomes in almost all
solid human tumors. As the primary microtubule organizing
center of most eukaryotic cells, the centrosome assures symmetry and bipolarity of the cell division process, a function that
is essential for accurate chromosome segregation. In addition,
a growing body of evidence indicates that centrosomes might
be imortant for initiating S phase and completing cytokinesis.
Centrosomes undergo duplication precisely once before cell
division. Recent reports have revealed that this process is
linked to the cell division cycle via cyclin-dependent kinase
(cdk) 2 activity that couples centriole duplication to the onset
of DNA replication at the G1/S phase transition. Alterations in
G1/S phase regulating proteins like the retinoblastoma protein,
cyclins D and E, cdk4 and 6, cdk inhibitors p16INK4A and
p15INK4B, and p53 are among the most frequent aberrations
observed in human malignancies. These alterations might not
only lead to unrestrained proliferation, but also cause karyotypic instability by uncontrolled centrosome replication. Since
several excellent reports on cell cycle regulation and cancer
have been published, this review will focus on the role of centrosomes in cell cycle progression, as well as causes and
consequences of aberrant centrosome replication in human
neoplasias.
Leukemia (2002) 16, 767–775. DOI: 10.1038/sj/leu/2402454
Keywords: centrosomes; cell cycle; cyclin-dependent kinases;
cyclins; aneuploidy; genomic instability
Genomic instability
A relationship between genomic damage and cancer development has been suggested since the beginning of the 20th century.1 The genetic alterations occurring in human tumors can
be divided into four major categories: (1) subtle sequence
changes; (2) gene amplifications or deletions; (3) chromosome
translocations; and (4) alterations in chromosome number.2
Whereas subtle sequence instabilities represented by nucleotide-excision repair-associated instability (NIN) and microsatellite instability (MIN) are rare, chromosomal instability (CIN),
involving gains and losses of whole chromosomes, is likely to
occur in most human malignancies.2,3 It is estimated that
numeric chromosomal imbalances, referred to as aneuploidy,
are the most prevalent genetic changes recorded among over
Correspondence: A Krämer, Medizinische Klinik und Poliklinik V,
Ruprecht-Karls-Universität Heidelberg, Hospitalstr. 3, 69115 Heidelberg, Germany; Fax: 49 6221 565721
Received 19 June 2001; accepted 7 January 2002
20 000 solid tumors analyzed thus far.3,4 Whether aneuploidy
is a cause or a consequence of cancer has long been debated.5
Several lines of evidence now argue in favor of aneuploidy
as a discrete chromosome mutation event that contributes to
malignant transformation and tumor progression. Most interestingly, in colorectal and endometrial cancers, there is an
inverse relationship between MIN and CIN.6 Tumors showing
microsatellite instability are diploid, whereas microsatellite
stable malignancies are usually aneuploid and exhibit
increased rates of chromosomal changes, suggesting that
MIN and CIN are alternative mechanisms of malignant
transformation.
Changes in DNA content and chromosome number have
been shown to occur already in preneoplastic lesions like oral
leukoplakia,7 early cervical neoplasias,8 and small benign
tumors of the colon,9,10 or breast11 and the deviations from a
normal karyotype increase as the tumors enlarge and eventually become malignant. Furthermore, it has recently been
demonstrated that the ploidy status in cells of oral leukoplakia
is a reliable predictor of progression into oral squamous cell
carcinoma.7 It has also been reported that an aneuploid DNA
content indicates a high risk of malignant transformation in
patients with ulcerative colitis12 and Barrett’s esophagus.13
These clinical observations are supported by in vitro studies in
which numeric chromosome aberrations are induced at early
stages of transformation.14,15 In a transgenic mouse model,
mice with a 1 Mb duplication of a portion of chromosome
11 synthenic with human chromosome 17 developed corneal
hyperplasia and thymic tumors.15 With regard to tumor progression and prognosis aneuploidy has been shown to be an
indicator of metastasis in patients with gastric carcinoma,16 as
well as papillary thyroid carcinoma17 and associated with a
poor outcome as compared to diploid neoplasms in a wide
variety of malignancies.16–19
Recent studies have revealed that specific nonrandom
numerical chromosome aberrations correlate with distinct
tumor phenotypes. Tumors from patients with hereditary papillary renal carcinomas commonly show a trisomy of chromosome 7, resulting in nonrandom duplication of a mutant MET
allele, thereby implicating this trisomy in tumorigenesis.20,21
In addition to trisomy 7, trisomy 17 in papillary tumors and
trisomy 8 in mesoblastic renal cell cancers are commonly
seen. In situ hybridization analysis of cervical intraepithelial
neoplasias has provided evidence that chromosome 1, 7, and
X aneusomies are associated with progression towards cervical carcinoma.8 Monosomy 5 or 7 seems to define a distinct
subgroup of acute myelogenous leukemia, commonly evolving from therapy-related myelodysplastic syndromes and
reflecting a poor prognosis.22 These findings support a role of
specific numeric chromosome aberrations in malignant
transformation.
Another argument in favor of a role of chromosomal instability in the process of malignant transformation is that aneuploidy seems to be a dynamic chromosome mutation event
Centrosomes and cancer
A Krämer et al
768
associated with a subsequently even further elevated rate of
chromosome instability. Lengauer and coworkers6 have
shown that in colorectal tumors with chromosomal instability
gains or losses of chromosomes occurred in excess of 10−2
per chromosome per generation. It has been estimated that
tumors of the colon, breast, pancreas and prostate may lose
an average of 25% of their alleles. In another study it was
found that the rate of chromosomal gains or losses is proportional to the degree of the pre-existing aneuploidy in the
transformed cells, hence that aneuploidy by itself further
destabilizes the karyotype, thereby altering the cellular phenotype and leading to tumor progression.23
There are many potential mitotic targets, which could cause
unequal segregation of chromosomes, among those chromosomal, spindle microtubule and centrosomal defects.5 Correct
sister chromid separation during metaphase–anaphase transition depends on the proper and timely degradation of a
group of proteins called securins that inhibit sister chromatid
separation. A human securin analogue has been identified to
be identical with the pituitary tumor transforming gene, which
is overexpressed in certain malignancies and exhibits transforming activity in vitro.24 Survivin, an inhibitor of apoptosis
protein that is selectively expressed in G2/M phase of the cell
cycle, specifically associates with microtubules of the mitotic
spindle, thereby preventing apoptosis induction at the G2/M
spindle checkpoint. Overexpression of survivin in tumor cells
may overcome this apoptotic checkpoint and favor aberrant
progression of aneuploid cells through mitosis.25 Furthermore,
in some cancers displaying chromosomal instability, the loss
of mitotic spindle checkpoint function is associated with the
mutational inactivation of genes controlling proper kinetochore-microtubule attachment.26 Normally, stable, bipolar
attachment of the kinetochores to spindle microtubules leads
to dissociation of the kinetochore-binding proteins BUB and
MAD from the kinetochore. If the kinetochore is not connected to the spindle microtubules, the BUB and MAD kinases
remain attached and arrest the cell in metaphase by delaying
activation of the cyclin destruction machinery. Cahill and
coworkers27 have found that a small fraction of human colorectal cancers display mutations in BUB proteins and that
induced overexpression of mutant hBUB1 induces chromosomal instability. Interstingly, truncations of the adenomatous
polyposis coli (APC) gene, whose protein product accumulates at the kinetochore during mitosis, is a high-affinity substrate for BUB kinases in vitro and has been shown to be
mutated in familial colon cancer and most early stages of
sporadic colorectal tumors, also lead to extensive chromosome and spindle aberrations and might contribute to
chromosomal instability in colorectal cancer28,29 Furthermore,
hMAD1 seems to be inactivated by the Tax protein of the
human T cell leukemia virus 1, leading to multinucleation and
genomic instability as well.30
Among the targets implicated in genomic instability and
malignant transformation, it has long been recognized that
errors in centrosome replication may be an important cause
of aneuploidy and might thus contribute to cancer formation.
The remaining part of this review will focus on causes and
consequences of aberrant centrosome duplication in human
neoplasias.
Centrosome structure and function
The centrosome was among the first organelles described by
early cell biologists at the end of the 19th century and was
Leukemia
viewed to be as important as the discovery of the nucleus.31
Centrosomes were named during this time by Theodor Boveri
on the basis of their central position in the cell. The importance of centrosomes for cell division was well recognized,
although microtubules had not yet been discovered and it was
not yet known that centrosomes are the main microtubule
organizing centers. The sea urchin egg was one of the first
systems studied by Theodor Boveri, who elaborated an outstanding wealth of information on centrosomes by using iron
hematoxylin as the primary staining technique.32 After Boveri’s remarkable discoveries, research on centrosomes stagnated and was not persued for many years. The focus changed
back again when immunofluorescence microscopy and the
availability of immunological probes opened a renaissance of
centrosome research which is now being extended to the
molecular level.
As the primary microtubule organizing center of most eukaryotic cells, the mammalian centrosome functions as the site
for the nucleation and organization of microtubules and thereby plays a fundamental role in organizing the interphase
cytoskeleton and the mitotic spindle.33,34
The centrosome consists of three major domains, the most
prominent being the centriolar domain and the pericentriolar
domain35,36 (Figure 1). The centriolar domain defines the
center of the centrosome, and dependent of the stage of the
cell cycle contains either a single centriole or a centriole pair.
The pericentriolar domain encircles the centriole as a cloud
of pericentriolar material. These domains together have an
average diameter of approximately 1 ␮m in most animal cells.
The centrosome does not end at the edge of the pericentriolar
domain. Rather, it extends out and integrates with the surrounding cytoplasm and organelles through microtubules and
microfilaments. These outer reaches of the centrosome are
called outer centrosomal domain.
Figure 1
Schematic diagram of a centriole pair. Centrioles are
arranged at right angles and connected at their proximal ends by
fibrous material. The mature mother centriole consists of nine triplet
microtubules and has appendages and satellites at its distal end,
whereas the immature daughter centriole does not. Adapted with permission from Ref. 127.
Centrosomes and cancer
A Krämer et al
The centriolar domain contains the centriole, a barrelshaped, cylindrical organelle that is found in most, but not all
animal cells. Centrioles are structurally similar to basal bodies,
which are found at the base of eukaryotic cilia and flagella.37,38 In some organisms, the same structure acts as basal
body in interphase and centriole in mitosis. During fertilization, basal bodies of sperm function as templates for the
organization of maternal material into the first centrosome of
the egg.
A mature centriole consists of nine triplet microtubules. The
outer surface of the centriole is associated with matrix material
and electron-dense satellite structures along its length, a series
of appendages at the distal end, and during replication, an
immature or daughter centriole at its proximal end.
The functions of the centrioles within the centrosome
remain unclear. Plant cells, yeast, fungi and oocytes in some
embryonic systems are known to have centrosomes without
centrioles, suggesting that at least in some cases the centrioles
are not essential for centrosome organization and function.39
However, recent results demonstrate that centrioles are
required for organizing the centrosomal components into a
single stable structure in vertebrate cells.40 Since the reproductive capacity of a centrosome depends on its centriole
content,41 and since centrosomes lacking centrioles do not
reproduce in animal cells,42 centrioles likely play an
important role in centrosome reproduction, which must be
tightly controlled to maintain genetic stability.
Surrounding the centriole, the pericentriolar material contains a large number of coiled-coil proteins and functions as
the site for microtubule polymerization.36–38 One well-characterized component of the pericentriolar material is the ␥-tubulin ring complex, a conserved, ring-shaped multiprotein complex of all microtubule organizing centers that seems to serve
as a template for microtubule polymerization from closely
related ␣- and ␤-tubulins.43 The ␥-tubulin ring complex consists of ␥-tubulin and at least five other proteins, two of those
being Spc97p and Spc98p.44 Little is known about how ␥tubulin ring complexes are recruited to, and anchored at centrosomes in mammalin cells. However, pericentrin, a large
coiled-coil protein that is known to form a dynamic reticular
lattice in the pericentriolar material, has been implicated in
the recruitment of ␥-tubulin complexes to centrosomes in vertebrate cells.45 In addition to the ␥-tubulin ring complex
around 100 proteins have been described to be contained
within the centrosome. However, it is useful to consider only
those proteins as true components of the centrosome whose
localization is microtubule- and cell cycle-independent, since
increasing evidence suggests that several proteins may use
association with the centrosome to ensure proper segregation
at mitosis. The pericentriolar material is preferentially associated with mature centrioles. The size of this domain is cell
cycle-dependent, reaching its largest dimensions at the metaphase–anaphase transition and a low point at telophase.37
The centrosome has three important functions: (1) it
nucleates the polymerization of tubulin subunits into long
microtubule polymers; (2) it organizes the nucleated microtubules into useful arrays; and (3) it duplicates onces every cell
cycle, thereby licensing cell cycle progression from G1 into S
phase, as well as exit from cytokinesis. According to a current
model, it seems likely that microtubule nucleation and anchoring are two separate activities, mediated by different classes
of molecules. After being nucleated in the pericentriolar
material by ␥-tubulin ring complexes, microtubules are
released from their nucleating sites and translocated to sites
of anchoring at the subdistal appendages of the maternal cen-
triole, which contain ninein and centriolin as anchor proteins,
but lack microtubule-nucleating proteins.38,46,47
It has long been assumed that centrosomes are necessary for
mitotic spindle assembly. However, higher plants and some
developmental systems lack classical centrosomes but still
organize normal spindles48,49 and spindle assembly by noncentrosomal pathways has recently been demonstrated even
for mammalian cells.50,51 Therefore, centrosome-mediated
spindle assembly might provide a redundant pathway to
ensure high fidelity of chromosome segregation or guarantee
equal distibution of centrosomes to daughter cells for completion of other essential cellular functions by these
organelles.
In addition to classical centrosomal functions like microtubule nucleation and spindle assembly, recent studies with vertebrate cells suggest that centrosomes are required for somatic
cell cycle progression from G1 into S phase, as well as for
completion of cytokinesis. Microsurgical or laser ablation of
centrosomes from interphase cells does not prevent mitotic
spindle assembly, but leads to either incomplete cytokinesis
and, regardless of whether or not cytokinesis was completed,
to an arrest in late G1 phase of the following cell cycle.50,51
Another study also observed that the exit from cytokinesis in
Drosophila cells is dependent on presence and post-anaphase
repositioning of centrosomes.52 These studies indicate that
centrosomes might be involved in checkpoints that monitor
completion of mitosis prior to cytokinesis and the presence of
aberrant centrosome numbers or excess DNA prior to
initiation of DNA replication.53,54 Alternatively, centrosomes
might be directly required to activate final stages of cytokinesis and DNA replication.
769
Centrosome replication
In animal cells, the interphase centrosome reproduces or
duplicates only once per cell cycle, thereby ensuring a strictly
bipolar mitotic spindle axis.55 Duplication of the single centrosome is initiated at the G1/S transition and completed before
mitosis, where the duplicated centrosomes play a role in
organizing the poles of the mitotic spindle. The centrosomes
are segregated at mitosis such that each of the two cells
resulting from division receives only one. Duplication of the
centrosome is semiconservative: the paired centriole splits and
a new centriole forms in association with each, creating two
centrosomes56 (Figure 2). Centrioles replicate during S phase
concurrent with DNA replication. During this time a small
procentriole bud forms adjacent to the proximal wall of each
parenting centriole, which then gradually elongates.57 By G2,
when the cell possesses a 4N DNA content, it also possesses
four centrioles arranged into two pairs referred to as diplosomes. Within each diplosome the mother and daughter centrioles are orthogonally oriented so that a line through the long
axis of the daughter points to the wall of the mother. This
relationship is then maintained through the ensuing mitosis.
During the next cell cycle, a slow process of maturation of
the daughter centriole takes place which transforms it into a
fully differentiated centriole apparently similar in all respects
to the mother centriole. In parallel, the centrosome matrix
changes in size and structure, displaying a varying number of
satellite structures, the significance of which is not clear.58,59
Recent studies indicate that many regulatory pathways
might control centrosome duplication. The most likely candidates for initiating centrosome replication during G1 phase are
extracellular signals. It seems reasonable to assume that the
Leukemia
Centrosomes and cancer
A Krämer et al
770
Figure 2
Centrosome reproduction cycle. Centrioles are represented as barrels. Grey shading represents the mature mother centriole (M). The immature daughter centriole (D) is displayed in white.
Adapted with permission from Ref. 36.
environmental signals responsible for activating centrosome
doubling would vary from cell to cell depending upon their
phenotypes. However, to date, the only extracellular ligand
that has been shown to affect centrosome doubling is epidermal growth factor (EGF) which has the capacity to activate
the entire repertoire of events necessary for centrosome replication in many cell types.60 The cellular pathways that are
activated following receptor stimulation remain to be defined.
However, recent studies suggest that the ultimate intracellular
target are cyclin/cdk2 complexes.61–64 The major difference
between studies utilizing embryonic cells and somatic cells is
the demonstration that cyclin E is the cdk2 partner in embryonic cells,61,62 while cyclin A is the cdk2 activating subunit
in somatic cells.63,64 This may simply indicate differences
between cell cycle regulation in embryonic and somatic cells.
For example, in embryonic cells, cyclin E is responsible for
activating DNA synthesis, whereas cyclin A fulfills that role in
somatic cell types.65 In addition, in embryonic cells cyclin E
levels do not oscillate during the cell cycle,66 an observation
that fits nicely with the demonstration that rounds of centrosome replication will occur in the absence of protein synthesis
in embryonic cells.67 In addition to cdk2 activation, centrosome duplication in somatic cells seems to be strictly dependent on phosphorylation of pRb and subsequent transcriptional activity of E2F64 (Figure 3).
One of the centrosomal substrates of cdk2 activity seems to
be nucleophosmin, a protein that associates with the centrosome during mitosis and thereby prevents centriole splitting
and duplication until late G1 phase, when cdk2 activity rises
and leads to phosphorylation and subsequent dissociation or
degradation of nucleophosmin from the centrosome.68
Another downstream phosphorylation target of cdk2 is the
mouse orthologue of yeast Mps-1p, whose kinase activity is
required for centrosome duplication.69 Phosphorylation of
Mps-1p by cdk2 leads to stabilization of Mps-1p protein levels
at the G1/S boundary, ensuring once more that centrosome
duplication is coordinated with S phase entry.
Apart from G1/S phase regulatory molecules, additional proteins appear to be involved in the decision to replicate centrosomes by somatic cells. Prominent among those are
Leukemia
Figure 3
Regulation of the G1/S phase transition with subsequent
DNA replication and centrosome duplication in somatic cells. Phosphorylation of the retinoblastoma protein (Rb) by cyclin D-cdk4/6 and
cyclin E-cdk2 leads to transcriptional activation of E2F followed by
transactivation of the cyclin A promoter and other E2F target genes.
The activity of cdks is negatively regulated by cdk inhibitors like
p16INK4A, p21CIP1 and p27KIP1. Both DNA replication and centrosome
duplication are regulated through the pRb pathway. E2F transcription
factors are required for both processes. In addition to E2F, cyclin A
may be the preferred partner of cdk2 for centrosome duplication,
whereas cyclin E may be primarily responsible for promoting the G1/S
transition and initiating DNA replication.
NEK2/cNAP1, p53/p21KIP1, ZYG-1, Aurora kinases, and the
ubiquitin-mediated proteolysis pathway.
NEK2 is a centrosomal protein kinase that phosphorylates
cNAP1, a part of the structure linking parental centrioles in a
cell cycle-regulated manner, thereby inducing the splitting of
centrioles as the initial event of centrosome duplication in late
G1 phase.70,71 Mutational inactivation of the p53/p21CIP1 pathway,72,73 or STK15/BTAK/Aurora 2,74,75 a member of the Aurora family of protein kinases has been shown to induce multiple rounds of centrosome replication within a single cell
cycle, suggesting a role for these molecules in the regulation
of centrosome duplication. In Caenorhabditis elegans, a novel
kinase called ZYG-1 has been demonstrated to be necessary
for centrosome duplication.76 ZYG-1 mutant embryos failed
to form a bipolar spindle, even though cell cycle progression
appeared normal, resulting in the assembly of a monopolar
spindle at mitosis. Furthermore, it has been observed that
ubiquitin-mediated proteolysis is required for centrosome
duplication. Inhibition of the proteasome in Xenopus extracts
inhibited centrosome replication and centriole splitting, arguing in favor of the hypothesis that proteolysis is involved in
the degradation of connections between centrioles early in the
centrosome duplication process.77 On the other hand, contrasting data in mice and Drosophila have shown that knockouts of SCF ubiquitin ligase components result in centrosome
overreplication, suggesting that proteolysis prevents uncontrolled centrosome duplication.78,79 Additional studies are
needed to clarify the precise role of the proteins in the centrosome replication process.
Centrosome aberrations in human malignancies
Centrosomal abnormalities have been detected in several disease processes including neurodegeneration,80,81 autoimmune
Centrosomes and cancer
A Krämer et al
disorders,82 and viral or bacterial infections.83–85 They may
also play a role in aging.86 Most frequently, however, centrosome abnormalities are observed in cancer cells.
Karyotypic alterations, including whole chromosome loss or
gain, ploidy changes, and a variety of chromosome aberrations are common in cancer cells.3,87 If proliferating cells
fail to coordinate centrosome duplication with DNA replication, this will inevitably lead to a change in ploidy, and the
formation of monopolar or multipolar spindles will generally
provoke abnormal segregation of chromosomes. Indeed, it has
long been recognized that errors in the centrosome duplication cycle may be an important cause of aneuploidy and
thus contribute to cancer formation.1 This view has recently
received fresh impetus with the description of supernumerary
centrosomes in many solid human tumors including brain,88,89
breast,89–92 lung,89 colon,89 prostate,89 pancreas,93 bile duct,94
and head and neck.95 Hematological malignancies like nonHodgkin’s lymphomas and acute leukemias also display centrosome aberrations at high frequencies96,97 (Figure 4). Most
centrosome defects fall into three groups: aberrant microtubule nucleation, abnormal centrosome duplication, and a failure to correctly separate during mitosis. Centrosomes of tumor
cells display diverse structural alterations, including (1) an
increase in centrosome number and volume; (2) accumulation
of excess pericentriolar material; (3) supernumerary centrioles;
and (4) inappropriate phosphorylation of centrosome proteins.89,90 In addition, tumor centrosomes show functional
abnormalities characterized by increased microtubule
nucleation activity.90,98 There are at least two functional
consequences of centrosome defects that may play important
roles during neoplastic transformation and tumor progression.
First, aberrant centrosome duplication may adversely affect
maintenance of cell polarity in interphase cells because cytoplasmic architecture and directional vesicular trafficking may
be disorganized in a cell with multiple centrosomes. Second,
centrosome defects may increase the incidence of multipolar
mitoses that lead to chromosomal segregation abnormalities
Figure 4
Indirect immunofluorescence staining of normal bone
marrow (a) and acute myelogenous leukemia (b), mantle cell lymphoma (c) and Hodgkin’s lymphoma (d). Cells were immunostained
with an antibody to pericentrin, followed by a FITC-conjugated secondary antibody. In part adapted from Refs 96 and 97.
and aneuploidy. While normal-appearing metaphase plates
with bipolar spindles are typical in malignant tumors, aberrant
mitotic figures are not infrequent. A careful serial section electron microscopic study in which individual spindle microtubules were traced to the poles of breast cancer cells with
multipolar spindles revealed that particular spindle poles may
possess one, two, supernumerary, or no centrioles at all.91 In
some cases, individual tumor bipolar spindles were observed
to have microtubule foci containing supernumerary centrioles.
Thus, multiple centrosomes containing supernumerary centrioles can either lead to multipolar mitosis or through coalescence, may result in bipolar spindles. When a bipolar spindle
is produced, equal partition of sister chromatids is the anticipated outcome. On the other hand, multipolar or monopolar
spindles will result in aneuploid or polyploid daughter cells
with cell death being the most common outcome for aneuploid daughter cells that lack essential chromosomes. However, aneuploid cells with an appropriate complement of
chromosomes or parts thereof do survive and, in rare cases,
may acquire comparative growth advantage.99 Survival and
perpetuation of the transformed clone, however, requires an
additional step – the resumption of mitotic stability through
the assembly of a bipolar, not multipolar, spindle. Current
data support coalescence of multiple centrosomes into two
spindle poles as a mechanism for regulating the number of
functional centrosomes in tumor cells.91,98,100
The mechanistic significance of centrosome aberrations for
the induction of genomic instability has been a matter of
intensive debate. On one hand, it has been proposed that
abnormal centrosome numbers may directly cause mitotic
defects that lead to genomic instability. According to this
model, centrosome abnormalities emerge early during neoplastic progression. Alternatively, centrosome abnormalities
may simply reflect genomic instability and accumulate in parallel with other cellular abnormalities. These two models have
different implications for the diagnostic and prognostic value
of centrosome abnormalities in human tumors. Importantly,
it has been demonstrated recently that expression of human
papilloma virus (HPV) 16 E7 oncoprotein, which binds to and
destabilizes the Rb tumor suppressor protein, is associated
with an abnormal synthesis of centrioles and induces abnormal centrosome numbers and genomic instability early during
neoplastic progression in primary human epithelial
cells.101,102 Consistent with these observations, preinvasive as
well as invasive HPV-associated genital lesions contain abnormal centrosome numbers that are associated with mitotic
abnormalities and numeric chromosomal alterations.101 Similarly, the development of pancreatic cancer in transgenic mice
expressing the simian virus 40 tumor antigen, which inactivates Rb and p53 proteins, is characterized by supernumerary
centrioles and the sequential appearance of tetraploid and
then multiple aneuploid cell populations.103
As a second line of evidence, centrosome defects have been
found in a fraction of preinvasive cancers of the prostate and
breast, as well as in some adenomas of the pancreas.93,104 In
an orthotopic implantation model of human pancreatic cancer
cells into pancreata of nude mice centrosome abnormalities
were detected in only a small fraction of parental cells in culture, whereas the frequency markedly increased in cells isolated from the pancreatic xenografts.105 Abnormal centrosome
numbers were found at even higher frequencies in metastatic
foci, as compared to the pancreatic xenografts. In addition,
centrosome abnormalities occur at a higher frequency in
advanced stages as compared to early stages of different
malignancies.94,95,104 In mantle cell lymphoma, the frequency
771
Leukemia
Centrosomes and cancer
A Krämer et al
772
of centrosome aberrations correlates with tetraploidization of
clinically aggressive blastoid variants.96 Furthermore, centrosome abnormalities were detectable at a significantly higher
frequency in aggressive diffuse large B cell lymphomas as
compared to indolent follicular lymphomas, suggesting that
abnormal centrosomes may contribute to the aquisition of an
increasing karyotypic instability typically seen in aggressive
lymphomas.
A comparison of mismatch repair deficient diploid vs mismatch repair proficient aneuploid colorectal carcinoma cell
lines revealed the exclusive occurrence of centrosome amplification and instability in all of the aneuploid tumor cell lines
analyzed.106 All diploid tumors contained centrosomes that
were functionally and structurally indistinguishable from those
in normal human fibroblasts.
Taken together, these observations support the notion that
the integrity of the centrosome plays a central role in the
development of genomic instability.
Molecular scenarios for the origin of centrosome
aberrations
The detailed mechanisms by which centrosome aberrations
develop in malignant tumors are largely unknown. Because
the centrosome is the microtubule organizing center of
interphase cells, and because this organelle must duplicate
and separate in G1 phase ultimately to become the poles of
the mitotic spindle, several lines of evidence point to a G1
phase cell cycle checkpoint that ensures proper centriole
duplication and separation.61–64,107 Accordingly, defects of
G1/S phase regulatory proteins should not only lead to unrestrained proliferation, but also cause karyotypic instability by
uncontrolled centrosome replication.
A physiological example of centrosome hyperreplication
supporting this view are endomitotic cell cycles of megakaryocytes. In order to become polyploid, these cells replicate
their DNA without karyokinesis and cytokinesis. Megakaryocytes undergo a normal cell cycle progression with G1, S and
G2 phases, nuclear membrane breakdown, sister chromatid
separation, and chromosome pole to pole movement.108
However, late mitosis stages like anaphase B, telophase and
cytokinesis are skipped, resulting in cells that contain a single
polylobulated typical nucleus with a 2xN ploidy. Microscopic
examination shows that a large number of centrosomes is
present in polyploid megakaryocytes. These findings have
been explained by the sustained expression of the G1 cyclins
D3, A and E, accompanied by pRb hyperphosphorylation at
least in megakaryocytic cell lines.109,110 Most interestingly,
careful analysis of multipolar endomitotic spindles revealed
an asymmetrical, aneuploid distribution of chromosomes
between the spindle poles and, subsequently, the different
lobes of interphase megakaryocyte nuclei despite an intact
metaphase/anaphase checkpoint.111 These findings suggest
that centrosome hyperreplication during G1/S phase by upregulated function of cdk2 indeed leads to aberrant chromosome distribution during mitosis.
Recently, the multiple-centrosome phenotype has been
recapitulated in budding yeast by disrupting the activity of
specific cdk complexes.112 In budding yeast, nine different
cyclins which associate with a single cdk, p34Cdc28, can be
divided into three groups – three G1 cyclins, two S phase Bcyclins, and four mitotic B-cyclins. Analogous to the findings
in eukaryotic cells, spindle pole body (SPB) duplication is
initiated by G1 cyclins. For SPB reduplication, however, S
Leukemia
phase B-cyclin activity seems to be necessary, whereas mitotic
B-cyclins inhibit SPB reduplication to ensure the dependence
of SPB reduplication on the completion of mitosis and subsequent down-regulation of mitotic cyclin/cdk activity. Similarly, in Xenopus, centrosomes were found to reduplicate in
the continuous presence of cyclin E/cdk2 activity when the
mitotic cyclin A/cdk1 and cyclin B/cdk1 activities were
absent.50 Xenopus extracts depleted for cyclin E/cdk2 could
duplicate, but not reduplicate centrosomes, indicating that an
activity analogous to yeast G1 cyclins – most likely cyclin
D/cdk4 – may be responsible for the initial duplication of the
centrosome in Xenopus.
In line with these observations, both inactivation of pRb and
overexpression of cyclin E have been shown to induce abnormal centrosome synthesis and genomic instability.101,113,114
Overexpression of cyclin A, E2F-2, and E2F-3 also leads to
multiple rounds of centrosome duplication.64 Also, overexpression of HPV-16 E6 and E7 oncoproteins, which inactivate p53 and pRb, has been shown to induce centrosome
aberrations with subsequent genomic instability.101,102,115 On
the other hand, overexpression of p16INK4A, p21CIP1 and
p27KIP1 is able to block centrosome duplication.63,64,116
Mutational inactivation of p53 has also been described to
initiate multiple rounds of centrosome replication within a single cell cycle.72,114,117,118 This effect seems to be mediated at
least in part through the p53/p21CIP1 axis, since induced
expression of the cdk2 inhibitor p21CIP1 in p53−/− cells partially restores centrosome duplication control, suggesting that
p21CIP1 is one of multiple effector pathways of the p53mediated regulation of the centrosome duplication cycle.73 In
line with these reports, reduced p21CIP1 expression as well as
overexpression of the p53 inhibitor mdm2 results in centrosome overduplication, gross nuclear abnormalities and polyploidy in human cells.106,117–119 In addition, centrosome aberrations and chromosome instability resulting from mutational
inactivation of the tumor suppressors BRCA1 and BRCA2
might also be mediated by interaction and inactivation of the
p53/p21CIP1 axis.120–124 Interestingly, in prolonged cell culture,
primary epithelial cells from p53-null mice display centrosome hyperamplification and numerical chromosome instability at high frequency in early to mid passage cells.118 Both,
heterogeneity in the number of chromosomes and centrosome
amplification diminish in late passage cells, giving rise to distinct subpopulations of cells which might have aquired chromosome compositions that promise an optimal growth in a
given environment.
Defects which lead to centrosome amplification, aneuploidy and transformation in vitro without aparent association to
G1/S phase regulatory events include overexpression of centrosome-associated kinases, such as breast tumor amplified
kinase (STK15/BTAK/Aurora 2).74 This protein is found to be
overexpressed in multiple human tumor types.125 Overexpression of the centrosomal protein pericentrin in prostate
cell lines has also been shown to lead to severe centrosome
and spindle defects, cellular disorganization, enhanced soft
agar growth and genomic instability.104 Highly significant
increases in chromosome missegregation associated with centrosome fragmentation have also been described in cell lines
lacking the mammalian DNA damage repair genes XRCC2
and XRCC3.126
In summary, centrosomes are structurally and numerically
abnormal in the vast majority of human malignancies. Observation of centrosome aberrations in preinvasive cancers, correlation of the extent of centrosome abnormalities with the
extent of chromosome instability, and centrosome aberrations
Centrosomes and cancer
A Krämer et al
preceding other manifestations of genomic instability in
experimental systems all argue in favor of a role of centrosome
abnormalities in the development of karyotype instability in
human cancers. Since centrosome replication has been found
to be ultimately controlled by pRb and cdk2 activity, the commonly observed abrogation of the pRb pathway in human
tumors will not only facilitate progression towards DNA replication, but may also deregulate the centrosome duplication
cycle and thereby lead to genomic instability. However, a
direct link between centrosome aberrations and cancer has
still to be established. A better understanding of how centrosomes contribute to cell cycle regulation and aneuploidy may
help to define novel therapeutic strategies to suppress genomic instability at early stages of malignant transformation.
References
1 Boveri T. Zur Frage der Entstehung maligner Tumoren. FischerVerlag: Jena, 1914.
2 Lengauer C, Kinzler KW, Vogelstein B. Genetic instabilities in
human cancers. Nature 1998; 396: 643–649.
3 Mitelman F, Johansson B, Mertens F. Catalog of Chromosome
Aberrations in Cancer, Vol. 2. Wiley-Liss: New York, 1994.
4 Heim S, Mitelman F. Cancer Cytogenetics. Wiley-Liss: New
York, 1995.
5 Sen S. Aneuploidy and cancer. Curr Opin Oncol 2000; 12:
82–88.
6 Lengauer C, Kinzler KW, Vogelstein B. Genetic instability in colorectal cancers. Nature 1997; 386: 623–627.
7 Sudbo J, Kildal W, Risberg B, Koppang HS, Danielsen HE, Reith
A. DNA content as a prognostic marker in patients with oral leukoplakia. New Engl J Med 2001; 344: 1270–1278.
8 Bulten J, Poddighe PJ, Robben JC, Gemmink JH, de Wilde PC,
Hanselaar GAGJM. Interphase cytogenetic analysis of cervical
intraepithelial neoplasia. Am J Pathol 1998; 152: 495–503.
9 Bardi G, Parada LA, Bomme L, Pandis N, Willen R, Johansson
B, Jeppsson B, Beroukas K, Heim S, Mitelman F. Cytogenetic
comparisons of synchronous carcinomas and polyps in patients
with colorectal cancer. Br J Cancer 1997; 76: 765–769.
10 Bomme L, Bardi G, Pandis N, Fenger C, Kronborg O, Heim S.
Cytogenetic analysis of colorectal adenomas: karyotypic comparisons of synchronous tumors. Cancer Genet Cytogenet 1998;
106: 66–71.
11 Silverstein MJ. Ductal Carcinoma In Situ of the Breast. Williams &
Wilkins: Baltimore, 1997.
12 Lindberg JO, Stenling RB, Rutegard JN. DNA aneuploidy as a
marker of permalignancy in surveillance of patients with ulcerative colitis. Br J Surg 1999; 86: 947–950.
13 Barrett MT, Sanchez CA, Prevo LJ, Wong DJ, Galipeau PC, Paulson TG, Rabinovitch PS, Reid BJ. Evolution of neoplastic cell lineages in Barrett oesophagus. Nat Genet 1999; 22: 106–109.
14 Li R, Yerganian G, Duesberg P, Krämer A, Willer A, Rausch C,
Hehlmann R. Aneuploidy correlated 100% with chemical transformation of Chinese hamster cells. Proc Natl Acad Sci USA
1997; 94: 14506–14511.
15 Liu P, Zhang H, McLellan A, Vogel H, Bradley A. Embryonic
lethality and tumorigenesis caused by segmental aneuploidy on
mouse chromosome 11. Genetics 1998; 150: 1155–1168.
16 Sasaki O, Kido K, Nagahama S. DNA ploidy, Ki-67 and p53 as
indicators of lymph node metastasis in early gastric carcinoma.
Anal Quant Cytol Histol 1999; 21: 85–88.
17 Sturgis CD, Caraway NP, Johnston DA, Sherman SI, Kidd L, Katz
RL. Image analysis of papillary thyroid carcinoma fine-needle
aspirates: significant association between aneuploidy and death
from disease. Cancer 1999; 87: 155–160.
18 Abad M, Ciudad J, Rincon MR, Silva I, Paz-Bouza JI, Lopez A,
Alonso AG, Bullon A, Orfao A. DNA aneuploidy by flow cytometry is an independent prognostic factor in gastric cancer. Anal
Cell Pathol 1998; 16: 223–231.
19 Magennis DP. Nuclear DNA in histological and cytological
specimens: measurement and prognostic significance. Br J
Biomed Sci 1997; 54: 140–148.
20 Flechter JA. Renal and bladder cancers. In: Wolman SR, Sell S,
Totowa NJ (eds). Human Cytogenetic Cancer Markers. Humana
Press: Totowa, 1997, pp 169–202.
21 Zhuang Z, Park WS, Pack S, Schmidt L, Vortmeyer AO, Pak E,
Pham T, Weil RJ, Candidus S, Lubensky IA, Linehan WM, Zbar
B, Weirich G. Trisomy 7-harbouring non-random duplication of
the mutant MET allele in hereditary papillary renal carcinomas.
Nat Genet 1998; 20: 66–69.
22 Löwenberg B, Downing JR, Burnett A. Acute myeloid leukemia.
N Engl J Med 1999; 341: 1051–1062.
23 Duesberg P, Rausch C, Rasnick D, Hehlmann R. Genetic instability of cancer cells is proportional to their degree of aneuploidy.
Proc Natl Acad Sci USA 1998; 95: 13692–13697.
24 Zou H, McGarry TJ, Bernal T, Kirschner MW. Identification of a
vertebrate sister-chromatid separation inhibitor involved in transformation and tumorigenesis. Science 1999; 285: 418–422.
25 Li F, Ambrosini G, Chu EY, Plescia J, Tognin S, Marchisio PC,
Altieri DC. Control of apoptosis and mitotic spindle checkpoint
by survivin. Nature 1998; 396: 580–584.
26 Orr-Weaver TL, Weinberg RA. A checkpoint on the road to cancer. Nature 1998; 392: 223–224.
27 Cahill DP, Lengauer C, Yu J, Riggins GJ, Willson JKV, Markowitz
SD, Kinzler KW, Vogelstein B. Mutations of mitotic checkpoint
genes in human cancers. Nature 1998; 392: 300–303.
28 Kaplan KB, Burds AA, Swedlow JR, Bekir SS, Sorger PK, Nathke
IS. A role for adenomatous polyposis coli protein in chromosome
segregation. Nat Cell Biol 2001; 3: 429–432.
29 Fodde R, Kuipers J, Rosenberg C, Smits R, Kielman M, Gaspar
C, van Es JH, Breukel C, Wiegant J, Giles RH, Clevers H.
Mutations in the APC tumour suppressor gene cause chromosomal instability. Nat Cell Biol 2001; 3: 433–438.
30 Jin DY, Spencer F, Jeang KT. Human T cell leukemia virus type
1 oncoprotein Tax targets the human mitotic checkpoint protein
MAD1. Cell 1998; 33: 81–91.
31 Flemming W. Studien über die Entwicklungsgeschichte der
Najaden. Sitzungber Akad Wissensch Wien 1875; 71: 81–147.
32 Boveri T. Zellen-Studien: Über die Natur der Centrosomen.
Fischer: Jena, 1900.
33 Brinkley BR. Microtubule organizing centers. Ann Rev Cell Biol
1985; 1: 145–172.
34 Tassin AM, Bornens M. Centrosome structure and microtubule
nucleation in animal cells. Biol Cell 1999; 91: 343–354.
35 Vorobjev IA, Nadezhdina ES. The centrosome and its role in the
organization of microtubules. Int Rev Cytol 1987; 106: 227–293.
36 Mack GJ, Ou Y, Rattner JB. Integrating centrosome structure with
protein composition and function in animal cells. Microsc Res
Tech 2000; 49: 409–419.
37 Urbani L, Stearns T. The centrosome. Curr Biol 1999; 9:
R315–R317.
38 Doxsey S. Re-evaluating centrosome function. Nature Rev 2001;
2: 688–698.
39 Karsenti E, Maro B. Centrosomes and the spatial distribution of
microtubules in animal cells. TIBS 1986; 11: 460–463.
40 Bobinnec Y, Khodjakov A, MirLM, Rieder CL, Edde B, Bornens
M. Centriole disassembly in vivo and its effect on centrosome
structure and function in vertebrate cells. J Cell Biol 1998; 143:
1575–1589.
41 Sluder G, Rieder CL. Centriole number and the reproductive
capacity of spindle poles. J Cell Biol 1985; 100: 887–896.
42 Sluder G, Miller FJ, Rieder CL. Reproductive capacity of sea
urchin centrosomes without centrioles. Cell Motil Cytoskelet
1989; 13: 264–273.
43 Moritz M, Braunfeld MB, Sedat JW, Alberts B, Agard DA.
Microtubule nucleation by ␥-tubulin-containing rings in the centrosome. Nature 1995; 378: 638–640.
44 Murphy SM, Urbani L, Stearns T. The mammalian ␥-tubulin complex contains homologues of the yeast spindle pole body components spc97p and spc98p. J Cell Biol 1998; 141: 663–674.
45 Dictenberg JB, Zimmerman W, Sparks CA, Young A, Vidair C,
Zheng Y, Carrington W, Fay FS, Doxsey SJ. Pericentrin and
gamma-tubulin form a protein complex and are organized into
a novel lattice at the centrosome. J Cell Biol 1998; 141: 163–174.
46 Mogensen MM, Malik A, Piel M, Bouckson-Castaing V, Bornens
M. Microtubule minus-end anchorage at centrosomal and non-
773
Leukemia
Centrosomes and cancer
A Krämer et al
774
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
Leukemia
centrosomal sites: the role of ninein. J Cell Sci 2000; 113:
3013–3023.
Piel M, Meyer P, Khodjakov A, Rieder CL, Bornens M. The
respective contributions of the mother and daughter centrioles to
centrosome activity and behavior in vertebrate cells. J Cell Biol
2000; 149: 317–330.
Compton DA. Spindle assembly in animal cells. Ann Rev
Biochem 2000; 69: 95–114.
Megraw TL, Kao LR, Kaufman TC. Zygotic development without
functional mitotic centrosomes. Curr Biol 2001; 11: 116–120.
Hinchcliffe EH, Miller FJ, Cham M, Khodjakov A, Sluder G.
Requirement of a centrosomal activity for cell cycle progression
through G1 into S phase. Science 2001; 291: 1547–1550.
Khodjakow A, Cole RW, Oakley BR, Rieder CL. Centrosomeindependent mitotic spindle formation in vertebrates. Curr Biol
2000; 10: 59–67.
Piel M, Nordberg J, Euteneuer U, Bornens M. Centrosome-dependent exit of cytokinesis in animal cells. Science 2001; 291:
1550–1553.
Doxsey SJ. Centrosomes as command centres for cellular control.
Nature Cell Biol 2001; 3: E105–E108.
Andreassen PR, Lohez OD, Lacroix FB, Margolis RL. Tetraploid
state induces p53-dependent arrest of nontransformed mammalian cell in G1. Mol Biol Cell 2001; 12: 1315–1328.
Kellogg DR, Moritz M, Alberts BM. The centrosome and cellular
organization. Annu Rev Biochem 1994; 63: 639–674.
Kochanski RS, Borisy GG. Mode of centriole duplication and distribution. J Cell Biol 1990; 110: 1599–1605.
Kuriyama R, Borisy GG. Centriole cycle in Chinese hamster
ovary cells as determined by whole-mountelectron microscopy.
J Cell Biol 1981; 91: 814–821.
Rieder CL, Borisy GG. The centrosome cycle in PTK2 cells: asymmetric distribution and structural changes in the pericentriolar
material. Biol Cell 1982; 44: 117–132.
Vorobjev IA, Chentsov YS. Centrioles in the cell cycle. I. Epithelial cells. J Cell Biol 1982; 98: 938–949.
Balczon R, Bao L, Zimmer WE, Brown K, Zinkowski RP, Brinkley
BR. Dissociation of centrosome replication events from DNA
synthesis and mitotic division in hydroxyurea arrested Chinese
hamster ovary cells. J Cell Biol 1995; 130: 105–115.
Hinchcliffe EH, Li C, Thompson EA, Maller JL, Sluder G. Requirement of cdk2-cyclin E activity for repeated centrosome reproduction in Xenopus egg extracts. Science 1999; 283: 851–854.
Lacey KR, Jackson PK, Stearns T. Cyclin-dependent kinase control of centrosome duplication. Proc Natl Acad Sci USA 1999;
96: 2817–2822.
Matsumoto Y, Hayashi K, Nishida E. Cyclin-dependent kinase 2
(Cdk2) is required for centrosome duplication in mammalian
cells. Curr Biol 1999; 9: 429–432.
Meraldi P, Lukas J, Fry AM, Bartek J, Nigg EA. Centrosome duplication in mammalian somatic cells requires E2F and Cdk2-Cyclin
A. Nature Cell Biol 1999; 1: 88–93.
Hartley RS, Sible JC, Lewellyn, Maller JL. A role for cyclin E/cdk2
in the timing of the midblastula transition in Xenopus embryos.
Dev Biol 1997; 188: 312–321.
Knoblich JA, Sauer K, Jones L, Richardson H, Saint R, Lehner CF.
Cyclin E controls S phase progression and its down-regulation
during Drosophila embryogenesis is required for the arrest of cell
proliferation. Cell 1994; 77: 107–120.
Sluder G, Miller FJ, Cole R, Rieder CL. Protein synthesis and the
cell cycle: centrosome reproduction in sea urchin eggs is not
under translational control. J Cell Biol 1990; 110: 2025–2032.
Okuda M, Horn HF, Tarapore P, Tokuyama Y, Smulian AG, Chan
P-K, Knudsen ES, Hofmann IA, Snyder JD, Bove KE, Fukasawa K.
Nucleophosmin/B23 is a target of CDK2/cyclin E in centrosome
duplication. Cell 2000; 103: 127–140.
Fisk HA, Winey M. The mouse mps1p-like kinase regulates centrosome duplication. Cell 2001; 106: 95–104.
Fry AM, Meraldi P, Nigg EA. A centrosomal function for the
human Nek2 protein kinase, a member of the NIMA family of
cell cylce regulators. EMBO J 1998; 17: 470–481.
Mayor T, Stierhof YD, Tanaka K, Fry AM, Nigg EA.The centrosomal protein C-Nap1 is required for cell cycle-regulated centrosome cohesion. J Cell Biol 2000; 151: 837–846.
Fukasawa K, Choi T, Kuriyama R, Rulong S, Van de Woude GF.
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
Abnormal centrosome amplification in the absence of p53.
Science 1996; 271: 1744–1747.
Tarapore P, Horn HF, Tokuyama Y, Fukasawa K. Direct regulation of the centrosome duplication cycle by the p53p21Waf1/Cip1 pathway. Oncogene 2001; 20: 3173–3184.
Zhou H, Kuang J, Zhong L, Kuo WL, Gray JW, Sahin A, Brinkley
BR, Sen S. Tumour amplified kinase STK15/BTAK induces centrosome amplification, aneuploidy and transformation. Nature
Genet 1998; 20: 189–193.
Nigg EA. Mitotic kinases as regulators of cell division and its
checkpoints. Nat Rev Mol Cell Biol 2001; 2: 21–21.
O’Connell KF, Caron C, Kopish KR, Hurd DD, Kemphues KJ, Li
Y, White JG. The C. elegans zyg-1 gene encodes a regulator of
centrosome duplication with distinct maternal and paternal roles
in the embryo. Cell 2001; 105: 547–558.
Freed E, Lacey KR, Huie P, Lyapina SA, Deshaies RJ, Stearns T,
Jackson PK. Components of an SCF ubiquitin ligase localize to
the centrosome and regulate the centrosome duplication cycle.
Genes Dev 1999; 13: 2242–2257.
Wojcik EJ, Glover DM, Hays TS. The SCF ubiquitin ligase protein
slimb regulates centrosome duplication in Drosophila. Curr Biol
2000; 10: 1131–1134.
Nakayama K, Nagahama H, Minamishima YA, Matsumoto M,
Nakamichi I, Kitagawa K, Shirane M, Tsunematsu R, Tsukiyama
T, Ishida N, Kitagawa M, Nakayama K, Hatakeyama S. Targeted
disruption of Skp2 results in accumulation of cyclin E and
p27(Kip1), polyploidy and centrosome overduplication. EMBO J
2000; 19: 2069–2081.
Li J, Xu M, Zhou H, Ma J, Potter H. Alzheimer presenilins in the
nuclear membrane, interphase kinetochores, and centrosomes
suggest a role in chromosome segregation. Cell 1997; 90:
917–927.
Sathasivam K, Woodman B, Mahal A, Bertaux F, Wanker EE,
Shima DT, Bates GP. Centrosome disorganization in fibroblast
cultures derived from R6/2 Huntington’s disease (HD) transgenic
mice and HD patients. Hum Mol Genet 2001; 10: 2425–2435.
Gavanescu I, Vazquez-Abad D, McCauley J, Senecal JL, Doxsey
S. Centrosome proteins: a major class of autoantigens in scleroderma. J Clin Immunol 1999; 19: 166–171.
Ploubidou A, Moreau V, Ashman K, Reckmann I, Gonzalez C,
Way M. Vaccinia virus infection disrupts microtubule organization and centrosome function. EMBO J 2000; 19: 3932–3944.
Minemoto Y, Shimura M, Ishizaka Y, Masamune Y, Yamashita
K. Multiple centrosome formation induced by the expression of
Vpr gene of human immunodeficiency virus. Biochem Biophys
Res Commun 1999; 258: 379–384.
Cimolai N, Mah D, Roland E. Anticentriolar autoantibodies in
children with central nervous system manifestations of Mycoplasma pneumoniae infection. J Neurol Neurosurg Psychiatry
1994; 57: 638–639.
Schatten H, Chakrabarti A, Hedrick J. Centrosome and microtubule instability in aging Drosophila cells. J Cell Biochem 1999;
74: 229–241.
Vogelstein B, Fearon ER, Kern SE, Hamilton SR, Preisinger AC,
Nakamura Y, White R. Allelotype of colorectal carcinomas.
Science 1989; 244: 207–211.
Weber RG, Bridger JM, Benner A, Weisenberger D, Ehemann V,
Reifenberger G, Lichter P. Centrosome amplification as a possible mechanism for numerical chromosome aberrations incerebral primitive neuroectodermal tumors with TP53 mutations.
Cytogenet Cell Genet 1998; 83: 266–269.
Pihan GA, Purohit A, Wallace J, Knecht H, Woda B, Quesenberry
P, Doxsey SJ. Centrosome defects and genetic instability in malignant tumors. Cancer Res 1998; 58: 3974–3985.
Lingle WL, Lutz WH, Ingle JN, Maihle NJ, Salisbury JL. Centrosome hypertrophy in human breast tumors: implications for genomic stability and cell polarity. Proc Natl Acad Sci USA 1998;
95: 2950–2955.
Lingle WL, Salisbury JL. Altered centrosome structure is associated with abnormal mitoses in human breast tumors. Am J Pathol
1999; 155: 1941–1951.
Schneeweiss A, Sinn H-P, Ehemann V, Khbeis T, Neben K,
Lichter P, Krause U, Ho AD, Bastert G, Krämer A. Centrosome
aberrations in primary invasive breast cancer are associated with
Centrosomes and cancer
A Krämer et al
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
nodal status and hormone receptor expression (submitted for
publication).
Sato N, Mizumoto K, Nakamura M, Nakamura K, Kusumoto M,
Niiyama H, Ogawa T, Tanaka M. Centrosome abnormalities in
pancreatic ductal carcinoma. Clin Cancer Res 1999; 5: 963–970.
Kuo K-K, Sato N, Mizumoto K, Maehara N, Yonemasu H, Ker CG, Sheen P-C, Tanaka M. Centrosome abnormalities in human
carcinomas of the gallbladder and intrahepatic and extrahepatic
bile ducts. Hepatology 2000; 31: 59–64.
Gustafson LM, Gleich LL, Fukasawa K, Chadwell J, Miller MA,
Stambrook PJ, Gluckmann JL. Centrosome hyperamplification in
head and neck squamous cell carcinoma: a potential phenotypic
marker of tumor aggressiveness. Laryngoscope 2000; 110:
1798–1801.
Krämer A, Schweizer S, Neben K, Lichter P, Krause U, Giesecke
C, Kalla J, Müller-Hermelink HK, Ho AD, Ott G. Centrosome
aberrations as a possible mechanism for chromosomal instability
in non-Hodgkin’s lymphoma (submitted for publication).
Neben K, Giesecke C, Lichter P, Ho AD, Krämer A. Centrosome
aberrations in acute myeloid leukemia are correlated to cytogenetic risk classification (in preparation).
Salisbury JL, Lingle WL, White RA, Cordes LEM, Barrett S.
Microtubule nucleating capacity of centrosomes in tissue sections. J Histochem Cytochem 1999; 47: 1265–1273.
Salisbury JL, Whitehead CM, Lingle WL, Barrett S. Centrosomes
and cancer. Biol Cell 1999; 91: 451–460.
Brinkley BR. Managing the centrosome numbers game: from
chaos to stability in cancer cell division. Trends Cell Biol 2001;
11: 18–21.
Duensing S, Duensing A, Crum CP, Münger K. Human papillomavirus type 16 E7 oncoprotein-induced abnormal centrosome
synthesis is an early event in the evolving malignant phenotype.
Cancer Res 2001; 61: 2356–2360.
Duensing S, Lee LY, Duensing A, Basile J, Piboonniyom S, Gonzalez S, Crum CP, Münger K. The human papillomavirus type
16 E6 and E7 oncoproteins cooperate to induce mitotic defects
and genomic instability by uncoupling centrosome duplication
from the cell division cycle. Proc Natl Acad Sci USA 2000; 97:
10002–10007.
Levine DS, Sanchez CA, Rabinovitch PS, Reid BJ. Formation of
the tetraploid intermediate is associated with the development
of cells with more than four centrioles in the elastase-simian virus
40 tumor antigen transgenic mouse model of pancreatic cancer.
Proc Natl Acad Sci USA 1991; 88: 6427–6431.
Pihan G, Purohit A, Wallace J, Malhotra R, Liotta L, Doxsey SJ.
Centrosome defects can account for cellular and genetic changes
that characterize prostate cancer progression. Cancer Res 2001;
61: 2212–2219.
Shono M, Sato N, Mizumoto K, Maehara N, Nakamura M, Nagai
E, Tanaka M. Stepwise progression of centrosome defects associated with local tumor growth and metastatic process of human
pancreatic carcinoma cells transplanted orthotopically into nude
mice. Lab Invest 2001; 81: 945–952.
Ghadimi BM, Sackett DL, Difilippantonio MJ, Schröck E, Neumann T, Jauho A, Auer G, Ried T. Centrosome amplification and
instability occurs exclusively in aneuploid, but not in diploid colorectal cancer cell lines, and correlates with numerical chromosomal aberrations. Genes Chromosomes Cancer 2000; 27:
183–190.
Mantel CR, Braun SE, Lee Y, Kim Y-J, Broxmeyer HE. The
interphase microtubule damage checkpoint defines an S-phase
commitment point and does not require p21waf-1. Blood 2001;
97: 1505–1507.
Vitrat N, Cohen-Solal K, Pique C, Le Couedic JP, Norol F, Larsen
AK, Katz A, Vainchenker W, Debili N. Endomitosis of human
megakaryocytes are due to abortive mitosis. Blood 1998; 91:
3711–3723.
109 Garcı́a P, Calés C. Endoreplication in megakaryoblastic cell lines
is accompanied by sustained expression of G1/S cyclins and
downregulation of cdc25C. Oncogene 1996; 13: 695–703.
110 Zimmet JM, Ladd D, Jackson CW, Stenberg PE, Ravid K. A role
for cyclin D3 in the endomitotic cell cycle. Mol Cell Biol 1997;
17: 7248–7259.
111 Roy L, Coulin P, Vitrat N, Hellio R, Debili N, Weinstein J,
Bernheim A, Vainchenker W. Asymmetrical segregation of chromosomes with a normal metaphase/anaphase checkpoint in
polyploid megakaryocytes. Blood 2001; 97: 2238–2247.
112 Haase SB, Winey M, Reed SI. Multi-step control of spindle pole
body duplication by cyclin-dependent kinase. Nature Cell Biol
2001; 3: 38–42.
113 Spruck CH, Won K-A, Reed SI. Deregulated cyclin E induces
chromosome instability. Nature 1999; 401: 297–300.
114 Mussman JG, Horn HF, Carroll PE, Okuda M, Tarapore P, Donehower LA, Fukasawa K. Synergistic induction of centrosome hyperamplification by loss of p53 and cyclin E overexpression.
Oncogene 2000; 19: 1635–1646.
115 Duensing S, Duensing A, Flores ER, Do A, Lambert PF, Münger
K. Centrosome abnormalities and genomic instability by episomal expression of human papillomavirus type 16 in raft cultures
of human keratinocytes. J Virol 2001; 75: 7712–7716.
116 Baccini V, Roy L, Vitrat N, Chagraoui H, Sabri S, Le Couedic JP, Debili N, Wendling F, Vainchenker W. Role of p21Cip1/Waf1 in
cell-cycle exit of endomitotic megakaryocytes. Blood 2001; 98:
3274–3282.
117 Carroll PE, Okuda M, Horn HF, Biddinger P, Stambrook PJ, Gleich LL, Li Y-Q, Tarapore P, Kukasawa K. Centrosome hyperamplification in human cancer: chromosome instability induced by
p53 mutation and/or Mdm2 overexpression. Oncogene 1999;
18: 1935–1944.
118 Chiba S, Okuda M, Mussman JG, Fukasawa K. Genomic convergence and suppression of centrosome hyperamplification in primary p53−/− cells in prolonged culture. Exp Cell Res 2000; 258:
310–321.
119 Mantel C, Braun SE, Reid S, Henegariu O, Liu L, Hangoc G,
Broxmeyer HE. p21cip-1/waf-1 deficiency causes deformed nuclear
architecture, centriole overduplication, polyploidy, and relaxed
microtubule damage checkpoints in human hematopoietic cells.
Blood 1999; 93: 1390–1398.
120 Xu X, Weaver Z, Linke SP, Li C, Gotay J, Wang X-W, Harris CC,
Ried T, Deng C-X. Centrosome amplification and a defective G2M cell cycle checkpoint induce genetic instability in BRCA1
exon 11 isoform-deficient cells. Molec Cell 1999; 3: 389–395.
121 Tutt A, Gabriel A, Bertwistle D, Connor F, Paterson H, Peacock
J, Ross G, Ashworth A. Absence of Brca2 causes genome instability by chromosome breakage and loss associated with centrosome amplification. Curr Biol 1999; 9: 1107–1110.
122 Brodie S, Deng C. BRCA1-associated tumorigenesis: what have
we learned from knockout mice? Trends Genet 2001; 17: S18–
S22.
123 Savelyeva L, Claas A, Matzner I, Schlag P, Hofmann V, Scherneck S, Weber B, Schwab M. Constitutional genomic instability
with inversions, duplications, and amplifications in 9p23–24 in
BRCA2 mutation carriers. Cancer Res 2001; 61: 5179–5185.
124 Hsu L-C, Doan TP, White RL. Identification of a ␥-tubulin-binding domain in BRCA1. Cancer Res 2001; 61: 7713–7718.
125 Tanaka T, Kimura M, Matsunaga K, Fukada D, Mori H, Okano
Y. Centrosomal kinase AIK1 is overexpressed in invasive ductal
carcinoma of the breast. Cancer Res 1999; 59: 2041–2044.
126 Griffin CS, Simpson PJ, Wilson CR, Thacker J. Mammalian
recombination-repair genes XRCC2 and XRCC3 promote correct
chromosome segregation. Nat Cell Biol 2000; 2: 757–761.
127 Lange BM, Gull K. Structure and function of the centriole in animal cells: progress and questions. Trends Cell Biol 1996; 6:
348–352.
775
Leukemia