The Cell Cycle




Abstract


The cell cycle is a process in which a cell first duplicates its genetic information and then divides to give rise to two daughter cells. To avoid mis-segregation of the genetic materials and incorporation of errors in the genome, the cell cycle is carefully monitored. This is achieved by positioning throughout the cell cycle a number of checkpoints designed to detect DNA damage, misaligned chromosomes, and incompletely replicated DNA. Checkpoints are mediated by cyclin-dependent kinase inhibitors (CKIs) in interphase and mitosis. On the other hand, progression through the cell cycle requires the combination of cyclins and cyclin-dependent kinases (CDKs), together of which regulate activities of target proteins by phosphorylation. The proper regulation of the cell cycle is critically important for cell proliferation as mutations in genes involved in cell cycle control often lead to cancer. This chapter reviews the components of the cell cycle and its control, checkpoint mechanisms, and provide examples of pathological consequences of dysregulation of the cell cycle machinery. This information is crucial for understanding the mechanisms regulating growth, proliferation, and differentiation of cells in the gastrointestinal tract.




Keywords

Cyclins, Cyclin-dependent kinases (CDKs), CDK inhibitors (CKIs), Checkpoint, Cancer, DNA damage repair, Mitosis, RB, p53, FBXW7, Ubiquitination, Growth factors, Mitogens, Spindle assembly checkpoint (SAC)

 




Acknowledgment


This work was in part supported by grants from the National Institutes of Health (DK052230, DK093680, CA084197, and CA172113).





Components of the Cell Cycle


A typical eukaryotic cell cycle contains several distinct phases, which progress in an orderly fashion—a phase cannot commence without completion of the previous one. The four phases of the cell cycle are G 1 (G for gap), S (synthesis), G 2 , and M (mitosis) phases ( Fig. 8.1 ). The G 1 , S, and G 2 phases collectively make up the interphase. The DNA content of a cell in the G 1 phase is 2N (N is the number of chromosomes), also known as diploid, whereas the DNA content of a cell in the G 2 phase is 4N (tetraploid). The DNA content of a cell in the S phase varies between 2N and 4N, depending on the stage of replication. The M phase is in turn comprised of two processes: mitosis, in which the cell’s chromosomes are equally divided between the two daughter cells, and cytokinesis (or cell division), in which the cytoplasm of the cell divides in half to form two distinct daughter cells. Typically, the amount of time required for a single-cell cycle in actively proliferating human cells in culture is 24 hours. Of these, the M phase takes approximately 1 hour to complete and interphase takes up the remaining 23 hours.




Fig. 8.1


The phases of the cell cycle. The cell cycle begins in the G 1 phase of a diploid cell (DNA content = 2N; N is the number of chromosomes). After DNA replication is completed in the S phase, the cell enters the G 2 phase and has twice the amount of the DNA (4N) of the starting cell. This is followed by mitosis (M) and cell division, which leads to the formation of two diploid daughter cells. Cells in either mitosis or cell division (also called cytokinesis) are in the M phase, whereas those in the other three phases (G 1 , S, and G 2 ) are in the interphase. The time in which a cell spends in each phase varies among the cell type and is not drawn to scale.


In addition to the four phases of the cell cycle listed above, one phase that lies outside the cell cycle is called the G 0 (0 for zero) phase. Cells in this phase are in the resting phase, which is often the result of their leaving the G 1 phase of the cell cycle. Typically, a cell enters G 0 phase if the environment is not conducive for the progression of the cell cycle, as in the event of deprivation of essential nutrients or growth factors, or if a cell has reached a fully differentiated state such as a hepatocytes or neuron. In these conditions, the cell is sometimes referred to as in a quiescent state. Additionally, a cell can enter the G 0 phase and become senescent due to DNA damage or telomere attrition. This is often an alternative to self-destruction of the damaged cell by apoptosis.


In the M phase, mitosis commences when DNA condenses into visible chromosomes, followed by the separation of the chromosomes into two identical sets. Cytokinesis is the last phase of mitosis when the two daughter cells separate, each with a nucleus and cytoplasmic organelles. Mitosis begins with nuclear membrane breakdown followed by condensation of the chromosomes and separation of the centrosomes (prophase). This is accompanied by the formation of mitotic spindles, which are attached on one end, the centrosomes, and the other end, kinetochore, a protein structure located at or near the centromeres of mitotic chromosomes (prometaphase). At this point, kinetochores that are unattached to the mitotic spindles generate a “wait” signal that delays the onset of anaphase until all chromosomes are properly attached and aligned. This signal is also called mitotic checkpoint or spindle assembly checkpoint (SAC), which is satisfied once all chromosomes are congregated at the equatorial plate (metaphase). This is followed by separation of the chromosomes to the opposite poles (anaphase) and formation of new nuclear membranes around the daughter nuclei and uncoiling of the chromosomes (telophase), eventually forming a cleavage furrow that leads to the formation of two daughter cells (cytokinesis). Fig. 8.2 illustrates the various stages of mitosis.




Fig. 8.2


The stages of mitosis. The different phases of mitosis are illustrated. SAC is spindle assembly checkpoint.

(Adapted from reference Kops GJ, Weaver BA, Cleveland DW. On the road to cancer: aneuploidy and the mitotic checkpoint. Nat Rev Cancer 2005;5(10):773–85, with permission from Nature Publishing Group.)





Control of the Cell Cycle


The organization of the cell cycle and its control system are highly conserved among eukaryotic organisms. Abundant information on how the cell cycle is regulated in vertebrates was derived from earlier studies in yeast. These studies demonstrate that both cell cycle progression and cell division are driven by the sequential activation, and subsequent inactivation, of two key classes of regulatory molecules, cyclins, and cyclin-dependent kinases (CDKs) Cyclins and CDKs form heterodimers with the former acting as the regulatory subunits and the latter catalytic subunits; CDKs are inactive in the absence of their cyclin partners. Specific pairs of cyclin-CDK dimers function in specific phases of the cell cycle ( Fig. 8.3 ). For example, the cyclin E-CDK2 complex becomes active late in the G 1 phase and is responsible for the transition from G 1 into S phase. When activated by a bound cyclin, CDKs phosphorylate a series of downstream target proteins, either activating or inactivating them, which then orchestrate the coordinated entry into the next phase of the cell cycle. The identities of the target proteins depend on the particular combination of cyclin-CDK complexes. While CDKs are constitutively expressed throughout the cell cycle, cyclins are synthesized (and destroyed) in specific stages of the cell cycle (hence the name cyclin), which is often dependent upon various signaling molecules. This cyclic nature of cyclin expression ensures that CDKs are activated and inactivated in a precise manner and safeguards the orderly progression of the cell cycle ( Fig. 8.4 ).




Fig. 8.3


The cell cycle is driven by cyclin-cyclin-dependent (CDK) complexes. The various pairs of cyclin-CDK complexes and their respective positions in the cell cycle are shown in the figure.



Fig. 8.4


Relative expression levels of cyclins in the cell cycle. The relative protein levels of the four different cyclins in various phases of the cell cycle are demonstrated. The letters in parenthesis indicate the classes of cyclins. The levels of cyclin D often fluctuate, depending on whether growth factors or mitogens are present or not. “R” indicates the restriction point.



Cyclins


Cyclins were named because they undergo a cycle of synthesis and degradation in each cell cycle. Depending on the timing of their expression and functions in the cell cycle, cyclins are divided into four classes. Three of these classes, the G 1 /S cyclins, S cyclins, and M cyclins are directly involved in the control of cell cycle events. The fourth class, the G 1 cyclins, controls the entry into the cell cycle in response to extracellular growth factors or mitogens. In the G 1 phase, growth factors are necessary to initiate and maintain the proper transition to the S phase. Early in the G 1 phase, growth factors stimulate the synthesis of G 1 cyclins, represented by the cyclin D family of cyclins, which activates CDK4/6 to induce synthesis of downstream targets, one of which is cyclin E. The rise in cyclin E (a G 1 /S cyclin) levels and activity of its partner, CDK2, drive the cell past a restriction point (R in Fig. 8.4 ) in the cell cycle after which the cell is irreversibly committed to proceeding to DNA synthesis, even if the growth factors are withdrawn. Subsequently, the S cyclins (represented by cyclin A) and M cyclins (represented by cyclin B) are required for the initiation of DNA replication and entry into mitosis, respectively ( Fig. 8.4 ).



G 1 Cyclins


The G 1 cyclins are composed of the D-type cyclins that include cyclins D1, D2, and D3. Along with their partners, CDK4 and CDK6, G 1 cyclins act early in the G 1 phase of the cell cycle. The levels of G 1 cyclin are low in G 0 phase and increase progressively upon addition of growth factors or mitogens to the cells. The mechanisms by which mitogens or growth factors activate cyclin D1 are complex and occur at both transcriptional and posttranscriptional levels. At the transcriptional level, induction of cyclin D1 by growth factors is dependent on the RAS/RAF/mitogen- activated kinase (MEK)/extracellular signal-regulated kinase (ERK) signaling pathway. Once synthesized, cyclin D1 protein has a short half-life —its turnover being governed by ubiquitination and proteasomal degradation, which in turn are dependent on phosphorylation of cyclin D1 by glycogen synthase kinase-3β (GSK-3β). Growth factors prevent cyclin D1 degradation by inhibiting GSK-3β-dependent phosphorylation of cyclin D1 through the Ras/phosphatidylinositol-3-OH kinase (PI3K)/AKT pathway. It was subsequently demonstrated that growth factor induced cyclin D1 gene transcription by the MEK/ERK pathway and cyclin D1 protein stabilization by the PI3K/AKT pathway need to occur in a sequential manner with the former occurring early and the latter occurring late in the G 1 phase in order to drive the progression of the cell cycle.


One of the key targets of an activated cyclin D-CDK4/6 complex is the retinoblastoma (RB) protein. RB is one of three “pocket protein” family of cell cycle regulator proteins—the other two being p107 and p130—and has a major role in restraining the transition between G 1 and S phases of the cell cycle. In the absence of mitogenic stimuli, RB interacts with and inhibits the activity of the transcription factor E2F. As E2F-binding sites are present in the promoters of many genes required for cell cycle progression, the inhibition of E2F by RB prevents entry into the cell cycle. In addition to physically interacting with E2F, RB also recruits chromatin remodeling enzymes such as histone deacetylases (HDACs) that often serve as transcription corepressors. Thus, the binding of RB to E2F not only simply inhibits E2F activity, but the RB-E2F complex binds to promoters and actively represses transcription by blocking activity of the surrounding enhancers on the promoter.


The activity of RB is governed by phosphorylation catalyzed by CDKs. RB contains 16 potential phosphorylation sites by CDKs and oscillates between hypophosphorylated and hyperphosphorylated forms during the cell cycle. The form that inhibits E2F is the hypophosphorylated form. CDKs, in complex with their cyclin partners, phosphorylate the hypophosphorylated form of RB, leading to the hyperphosphorylated form. At least three different cyclin-CDK complexes are known to phosphorylate RB during the cell cycle-cyclin D-CDK4/6 acts early in G 1 ; cyclin E-CDK2 in late G 1 ; and cyclin A-CDK2 in the S phase. In this way, RB becomes sequentially phosphorylated in the cell cycle. It has been shown that successive phosphorylation of RB by cyclin D-CDK4/6 and cyclin E-CDK2 is necessary for the complete hyperphosphorylation of RB. The hyperphosphorylation of RB prevents its binding to E2F, thus releasing E2F from transcriptional repression. Once liberated from RB, E2F, along with its partner transcription factor DP, activate transcription of genes that are involved in nucleotide metabolism and DNA synthesis, thus allowing entry into the S phase . It is of interest to note that one of the target genes of an active E2F encodes cyclin E. It was also shown that phosphorylation of RB by cyclin D-CDK4/6 and cyclin E-CDK2 triggers sequential intramolecular interactions in RB that progressively block RB functions as cells move through G 1. The complete hyperphosphorylation of RB by cyclin D-CDK4/6 and cyclin E-CDK2 coincides with the R point ( Fig. 8.4 ), beyond which the cell is committed to completion of the cell cycle.



G 1 /S Cyclins


Cyclins E1 and E2 (collectively considered as cyclin E) are the G 1 /S cyclins. Transcription of the cyclin E gene is regulated by E2F, which, as described above, is activated due to cyclin D-CDK4/6-stimulated phosphorylation of RB. The amounts of cyclin E protein and its associated kinase (CDK2) activity are maximal in late G 1 and early S phases ( Fig. 8.4 ). Cyclin E-CDK2 completes RB phosphorylation in the G 1 phase and the transition from cyclin D-CDK4/6 to cyclin E-CDK2 accounts for the loss of dependency on growth factors. Cyclin E-CDK2 phosphorylates RB on different sites from cyclin D-CDK4/6, and these kinases may differentially impact the interaction between RB and E2Fs, HDACs, and other chromatin remodeling proteins. In contrast to cyclin D-CDK4/6, the functions of cyclin E-CDK2 are not limited to G 1 control. Thus, cyclin E-CDK2 phosphorylates other substrates that are more directly involved in the control of DNA replication, centrosome duplication, replication origin licensing and firing. The timing of cyclin E-CDK2 activation and its broader range of substrates suggest that cyclin E-CDK2 spans the interface between G 1 regulation and core cell cycle machinery during S phase.


In early S phase, cyclin E-CDK2 activity abruptly ceases as a consequence of cyclin E degradation ( Fig. 8.4 ). This is mediated by phosphorylation by both GSK-3β and CDK2, which target cyclin E for ubiquitination by the SCF Fbw7 E3 ligase, leading it to proteasomal degradation. Like cyclin D1, the involvement of GSK-3β, an enzyme that is inhibited by the PI3K/AKT signaling pathway, in regulating stability of cyclin E suggests that cyclin E can be influenced at least by one mitogen-dependent signaling cascade.



S Cyclins


The S cyclins include both cyclins A1 and A2. While cyclin A1 is restricted to the germ cell lineages, cyclin A2 is ubiquitously expressed in all cell types. Low levels of cyclin A2 are first detected at the G 1 /S boundary. The levels then rise steadily as cells begin to replicate their DNA and do not decline until cyclin A is degraded in late G 2 ( Fig. 8.4 ). In S phase, cyclin A and its partner, CDK2, phosphorylate substrates that commence DNA replication from preformed replication initiation complexes. Cyclin A-CDK2 are also required to coordinate the end of the S phase with activation of the mitotic cyclin-CDKs.



M Cyclins


During G 2 , A-type cyclins (the S cyclins) are degraded by ubiquitin-mediated proteolysis whereas B-type cyclins (the M cyclins) are actively synthesized ( Fig. 8.4 ). As a consequence, CDK1 (also known as Cdc2) binds to B-type cyclins—an association required for the commencement of mitosis. CDK1 preferentially binds to two main B-type cyclins, cyclins B1 and B2. In contrast, the third isoforms, cyclin B3, may have a function in the meiotic cell cycle. Cyclin B-CDK1 regulate events during both the G 2 /M transition and progression through mitosis. This is accomplished by the phosphorylation of over 70 proteins by the cyclin B-CDK1 complexes although the number of substrates could be much larger. Phosphorylation of target proteins leads to numerous events that include separation of centrosomes. condensation of chromosomes, breakdown of the nuclear lamina, and disassembly of the Golgi apparatus, among others. Finally, inactivation of the cyclin B-CDK1 complexes is required for proper exit from mitosis and this inactivation is achieved by the degradation of B-type cyclins by ubiquitin-mediated proteolysis that is regulated by the anaphase-promoting complex/cyclosome (APC/C).



Cyclin-Dependent Kinases (CDKs)


CDKs are the catalytic subunits of a relatively large family of serine/threonine protein kinases with a primary role in cell cycle progression. The mammalian genome contains 11 genes encoding CDKs and nine others encoding CDK-like proteins with conserved structure. The prototype CDK, CDK1, was first identified in yeasts and designated as Cdc2 in Saccharomyces pombe or Cdc28 in Saccharomyces cerevisiae . The mammalian homologue of yeast Cdc2, CDK1, was subsequently identified due to its ability to complement the yeast mutants. Other mammalian CDKs were then identified by a host of techniques including complementation, homology probing, differential display, and PCR amplification with degenerate primers. Again, as described above, progression through the G 1 phase of the cell cycle requires at least three CDKs—CDK4, CDK6, and CDK2—which, along with their regulatory cyclins, target the RB family of proteins that include RB, p107, and p130. CDK2 is also important for S phase progression and CDK1 for G 2 /M transition and mitosis.


The kinase activities of CDKs are regulated at multiple levels, including interaction with their regulatory subunits (cyclins), binding to negative regulators called CDK inhibitors or CKIs (see below), phosphorylation-dephosphorylation, folding and subcellular localization. Among these, phosphorylation-dephosphorylation can have activating or inhibitory effects on CDK activity. For example, active cyclin- CDK complexes need to be phosphorylated in the T-loop of the CDK subunit by CDK-activating kinase (CAK), which contains a complex of CDK7, cyclin H, and MAT1 ( Fig. 8.5 ). In contrast, cyclin-CDK complexes can be negatively regulated by phosphorylation in adjacent threonine and tyrosine residues of the CDK subunit by the dual-specificity protein kinases WEE1 and MYT1. Conversely, these inhibitory phosphorylations can be reversed by the ability of the dual-specificity CDC25 phosphatases (CDC25A, CDC25B, and CDC25C) to dephosphorylate the same threonine and tyrosine residues and thus act as positive regulators of cyclin-CDK activity ( Fig. 8.5 ). If both activating and inactivating phosphorylations exist in the same molecule, they result in an inactive kinase.




Fig. 8.5


The regulatory mechanisms of cell cycle CDKs. CDKs require binding to their cyclin partners for activation of kinase activity. The INK4 proteins inhibit CDK activity by directly binding to monomeric CDK4 or CDK6. In contrast, Cip and Kip inhibitors inactivate CDKs by binding to cyclin-CDK complexes. Cyclin-CDK complexes are activated by phosphorylation of the CDK subunit by cyclin-dependent activating kinase (CAK) that contains three subunits: CDK7, cyclin H, and MAT1. By contrast, cyclin-CDK complexes are inhibited by phosphorylation in adjacent threonine and tyrosine residues by the dual-specificity protein kinases, WEE1 and MYT1. These inhibitory phosphorylations can be reversed by the dual-specificity CDC25 phosphatases that dephosphorylate the CDKs at the same amino acid residues. Cyc is cyclin.

(Reproduced from reference Malumbres M, Barbacid M. Mammalian cyclin-dependent kinases. Trends Biochem Sci 2005;30(11):630–41, with permission from Elsevier).



In Vivo Functions of Cyclins and CDKs as Revealed by Gene Knockout in Mice


Numerous studies have investigated the in vivo functions of cyclins and CDKs in knockout mice with the genes deleted either singly or in combination. Since it is not the intent of this chapter to review all of the phenotypes of the knockout mice, the readers are referred to the many excellent recent review articles that summarize the findings of these studies. From these studies, it becomes clear that most of the cyclins and CDKs, although previously considered essential for cell proliferation, have turned out to be dispensable. Several compensatory mechanisms were uncovered among cyclins and CDKs. The particularly unexpected finding is that CDK2, thought to be a master regulator of the cell cycle, is dispensable for the regulation of the cell cycle with both CDK4 and CDK1 covering CDK2’s functions. In fact, CDK1 alone is able to drive the mammalian cell cycle, indicating that the regulation of the mammalian cell cycle is highly conserved.



CDK Inhibitors (CKIs)


CDK inhibitors (CKIs) are proteins that constrain the activities of CDKs. Two classes of CDK inhibitors have been described. The first class includes the INK4 proteins ( in hibitors of CD K4 ). Four such INK4 proteins have been identified: p16 INK4a (also known as CDK inhibitor 2A or CDKN2A), p15 INK4b (CDKN2B), p18 INK4c (CDKN2C), and p19 INK4d (CDKN2D). INK4 proteins specifically bind to and inhibit monomeric CDK4 and CDK6 proteins. The second class of CKIs includes the Cip/Kip ( C DK- i nteracting protein/CD K i nteracting p rotein) family of proteins which are more broadly acting than the INK4 family of proteins and do so by binding to cyclin-CDK complexes. There are three members of the Cip/Kip family of CKIs: p21 Cip1 (also called CDK inhibitor 1A or CDKN1A), p27 Kip1 (CDKN1B), and p57 Kip2 (CDKN1C). Cip and Kip inhibitors block CDK activity by forming inactive trimeric complexes (cyclin E-CDK2, cyclin A-CDK2, cyclin B-CDK1, and possibly cyclin D-CDK4 and cyclin D-CDK6), thus exerting a much broader effect on the progression of the cell cycle. Fig. 8.5 summarizes the multiple regulatory mechanisms by which CDK activities are regulated.





Checkpoints


The cell cycle contains several checkpoints to monitor and regulate its progression. Checkpoints are positioned at specific locations in the cell cycle to allow verification of phase processes and repair of DNA damage. A cell cannot proceed from one phase to the next without satisfying all of the checkpoint requirements. An important function of many of the cell cycle checkpoints is to assess DNA damage. Upon detection of DNA damage, the checkpoint initiates a signal cascade to either arrest the cell cycle until repairs are properly made, or if repairs are not possible, to target the cell for destruction via apoptosis as a means to maintain genomic integrity. In the cell cycle, there are three specific checkpoints for damaged or incompletely replicated DNA: G 1 /S, G 2 /M, and intra-S checkpoints. These checkpoints are patrolled by some of the CKIs described above. A fourth important and specific checkpoint occurs in mitosis, the so-called mitotic checkpoint or SAC. This checkpoint is designed to monitor proper alignment of the chromosomes during mitosis. Anaphase cannot proceed unless this checkpoint is satisfied. The locations of the various checkpoints in the cell cycle are illustrated in Fig. 8.6 . The various cell cycle checkpoint pathways in response to DNA damage are summarized in Fig. 8.7 .




Fig. 8.6


The cell cycle checkpoints. The locations of the various checkpoints (STOP signs) and the primary defects the checkpoints are designed to monitor are shown.



Fig. 8.7


A simplified scheme of the cell cycle checkpoint pathways in response to DNA damage. The black arrows indicate activating and red lines indicate inhibitory actions.



G 1 /S Checkpoint


The G 1 /S checkpoint (also called the G 1 checkpoint) is located near the end of the G 1 phase; just before the entry into S phase ( Fig. 8.6 ). In mammalian cells, the G 1 checkpoint is the restriction or R point ( Fig. 8.4 ). This is a point where cells typically arrest the cell cycle if environmental conditions are unfavorable for cell division, such as the presence of DNA damage or lack of growth factors. The G 1 checkpoint is controlled by both the INK4 and Cip/Kip families of CKIs. INK4 proteins specifically bind to CDK4 and CDK6 and inhibit their activity. Enforced expression of INK4 proteins arrest cells in the G 1 phase in an RB-dependent manner. Here CDK4 is redistributed from cyclin D-CDK4 complexes to INK4-CDK4 complexes, and unbound D-type cyclin is rapidly degraded by ubiquitination-mediated proteasomal pathway. Also, in early G 1 phase, the cyclin E-CDK2 and cyclin A-CDK2 complexes are inhibited by bound p21 Cip1 and p27 Kip1 . In addition, cyclin D-CDK4/6 complexes bind p21 Cip1 and p27 Kip1 . Loss of D-type cyclins therefore prevents the titration of p21 Cip1 and p27 Kip1 by cyclin D-CDK4/6 complexes away from the cyclin E-CDK2 and cyclin A-CDK2 complexes. As a result, there is complete inhibition of cyclin E-CDK2 and cyclin A-CDK2 activities as well as RB phosphorylation, leading to exit from the G 1 phase. Conversely, growth-induced or oncogenic-induced expression of cyclin D allows its interaction with CDK4/6 by competing with INK4 for binding. The binding of cyclin D-CDK4/6 complexes to p21 Cip1 and p27 Kip1 releases the inhibitors from the cyclin E-CDK2 and cyclin A-CDK2 complexes. This unleashes cyclin E-CDK2 and cyclin A-CDK2 activity, allowing further RB phosphorylation, exit from G 1 and entry into S phase.


The G 1 /S checkpoint is activated upon detection of DNA damage. The mammalian DNA-damage response is a complex network, involving a multitude of proteins that include “sensor” proteins that sense the damage and transmit signals to “transducer” proteins, which, in turn, convey the signals to numerous “effector” proteins implicated in specific cellular pathways, including DNA repair mechanisms, cell cycle checkpoints, cellular senescence, and apoptosis. In response to DNA damage, signals initiated by the sensors rapidly transduce to the ATM (ataxia telangiectasia, mutated) and ATR (ataxia telangiectasia and Rad3-related) kinases, which phosphorylate a great number of substrates. Among the substrates phosphorylated by activated ATM and ATR are the checkpoint serine/threonine kinases, CHK1 (checkpoint kinase 1) and CHK2 (checkpoint kinase 2). To prevent entry into S phase, CHK1 and CHK2 phosphorylate the cell cycle regulatory phosphatase CDC25A, leading to its ubiquitin-mediated proteolysis. Inactivation of CDC25A leads to sustained inhibitory phosphorylation of cyclin E-CDK2 complexes, thus preventing G 1 /S transition ( Fig. 8.5 ). Both ATM and ATR belong to the phosphatidylinositol-3-kinase-like kinase family (PIKKs), which also includes a third member of the DNA damage response transducer, DNA-PKcs (DNA-dependent protein kinase). DNA-PKcs appears to regulate a smaller number of targets and plays a role primarily in nonhomologous end joining (NHEJ) of double-strand DNA breaks. ATM/ATR and CHK1/CHK2 kinases also target the tumor suppressor, p53. The phosphorylation of p53 by ATM/ATR inhibits its association with MDM2, an E3 ubiquitin ligase normally functions to keep the p53 level low, and this results in stabilization of the p53 protein. Subsequently, p53 transcriptionally activates expression of the p21 Cip1 gene, which in turn inhibits cyclin E-CDK2 and prevents G 1 /S transition. Lastly, rapid degradation of cyclin D1 in response to DNA damage occurs that is independent of p53. The reduced cyclin D1 protein decreases the amount of cyclin D-CDK4/6 complexes, resulting in the redistribution of p21 Cip1 to cyclin E-CDK2 complexes, resulting in the latter’s inactivation.



G 2 /M Checkpoint


The G 2 /M checkpoint (also known as G 2 checkpoint) prevents cells from initiating mitosis when they experience DNA damage while in G 2 , when they progress into G 2 with either unrepaired DNA sustained during the previous S or G 1 phase, or when they possess incompletely replicated DNA from S phase. The critical target of the G 2 checkpoint is the mitosis-promoting activity of the cyclin B-CDK1 complexes, whose activation after genotoxic stresses is inhibited by ATM/ATR, CHK1/CHK2-mediated degradation of CDC25 family of phosphatases, which normally activate CDK1 at the G 2 /M boundary. In addition, other regulators of CDC25 and cyclin B-CDK1, such as the Polo-like kinases (PLKs) are targeted by DNA damage-induced mechanisms. Finally, the maintenance of the G 2 checkpoint is dependent on the transcriptional programs regulated by p53, leading to an induction of cell cycle inhibitors such as p21 Cip1 , growth arrest and DNA damage-inducible 45 (GADD45), and 14-3-3σ proteins. These proteins cooperatively inhibit cyclin B-CDK1 activity by directly binding to cyclin B-CDK1 (p21 Cip1 ), dissociating CDK1 from cyclin B (GADD45), and sequestering CDK1 in the cytoplasm (14-3-3σ), resulting in G 2 arrest.



Intra-S Checkpoint


While proliferating cells respond to genotoxic stresses by activating checkpoint responses that impose durable cell cycle arrest in G 1 or G 2 , before entry into S phase or mitosis, respectively, cells that experience genotoxic stresses during DNA replication can only delay their progression through the S phase in a transient fashion. If the damage is not repaired during the delay, cells exit S phase and arrest upon reaching the G 2 checkpoint. Nonetheless, the intra-S checkpoint is important for the maintenance of genomic stability as DNA replication is a vulnerable period of the cell cycle in which errors occur both endogenously or are introduced exogenously. When the intra-S checkpoint is activated, both replication initiation and fork progression are inhibited, thus reducing the rate of DNA replication. Studies have revealed numerous proteins that are involved in the control of intra-S checkpoint. Among these, activation of ATM and ATR is necessary to initiate intra-S checkpoint upon sensing DNA double-strand breaks (DSBs). Activated ATM and ATR then activates CHK1 and CHK2 kinases, which phosphorylate CDC25A, leading to its ubiquitin-proteosome-dependent degradation. This results in persistent inhibitory phosphorylation of CDK2 in the cyclin A-CDK2 complexes, which, in turn, prevents firing of the replication origins. A second independent pathway is involved in slowing down the replication rate in cells that have suffered from DNA damage. This effect is mediated by ATM-dependent phosphorylation of a cohesion protein, structural maintenance of chromosomes-1 (SMC-1). However, the mechanism by which phosphorylated SMC-1 interferes with DNA replication remains unknown.



Mitotic Checkpoint or SAC


Mitosis is the process in which a cell divides itself into two halves, each with an identical set of chromosomes ( Fig. 8.2 ). The central regulator of this process is the mitotic checkpoint, also known as the spindle assembly checkpoint or SAC, a signaling mechanism that arrests the progression of metaphase to anaphase until all chromosomes are attached to the mitotic spindles. This signal is akin to an “anaphase wait” signal that is generated at the kinetochores of unattached chromosomes and is extinguished once all kinetochores are properly attached to the spindles ( Fig. 8.2 ). Thus, sister chromatids are separated only when they are in a position to be equally distributed to the two daughter cells. Accordingly, the mitotic checkpoint serves to prevent chromosome mis-segregation.


The proteins that control mitotic checkpoint were originally identified by screens for mutations that bypassed the ability of wild type S. cerevisiae to arrest in mitosis in the presence of spindle poisons. The genes identified include MAD (mitotic-arrest deficient), MAD1, MAD2, MAD3 (BUBR1 in humans), and BUB1 (budding uninhibited by benzimidazole 1). It was later found that these genes are conserved in all eukaryotes. When activated, these SAC proteins target CDC20, which is a co-factor of the ubiquitin ligase anaphase promoting complex/cyclosome (APC/C). Specifically, SAC inhibits the ability of CDC20 to activate APC/C-mediated ubiquitination of two key substrates, cyclin B and securin, thus preventing their degradation by the 26S proteosome. Proteolysis of cyclin B inactivates CDK1, allowing the cells to exit from mitosis. On the other hand, destruction of securin leads to activation of separase, which is required to cleave the cohesion complex that holds sister chromatids together in order to execute anaphase. Therefore, by keeping CDC20 in check, the SAC allows the chromosomes to properly align on the metaphase plate and attached to the two spindle poles through mitotic spindles. Once the chromosomes are properly oriented, the checkpoint is extinguished, relieving the mitotic arrest and allowing anaphase to proceed.





Noncanonical Functions of Cyclins, CDKs, and CKIs


Close cooperation among cyclins, CDKs, and CKIs is necessary for ensuring orderly progression through the cell cycle. However, evidence has emerged that the roles of this trio are beyond cell cycle regulation. The “noncanonical” functions of cyclins, CDKs and CKIs involve myriads of other cellular processes such as transcription, DNA damage repair, apoptosis, cell differentiation, epigenetic regulation, stem cell self-renewal, metabolism, and the immune response. Some of these functions are performed by cyclins or CDKs independent of their respective cell cycle partners, suggesting that there is substantial divergence in their functions during evolution. For example, D-type cyclins, independent of any associated kinase activity, are known to have direct roles in regulating transcription by interacting with many transcription factors to activate or repress transcription of specific genes. Similarly CDK6, but not CDK4, can regulate angiogenesis and myloid differentiation, respectively, by modulating the transcriptional activity of JUN and RUNX1.


In addition to regulating the cell cycle, the trio of cyclins, CDKs, and CKIs exerts important functions in the repair of DNA damage sustained from DSBs. DNA DSBs are repaired by two different mechanisms: homologous recombination and NHEJ. Cyclin D1 has been shown to localize to DNA DSBs and to recruit RAD51, which activates HR-mediated DNA repair. CDK2 was also shown to support HR by promoting the interaction between breast cancer type 1 susceptibility protein (BRCA1) and the MRE11 exonuclease, leading to the resection of DSBs. These observations suggest that there is a profound, two-way connection between DNA damage repair and cell cycle control.


The Cip/Kip family of CKIs (p21 Cip1 , p27 Kip1 , and p57 Kip2 ) also exhibit functions independent of their ability to inhibit cyclin-CDK complexes. They have been found to be key regulators of transcription, apoptosis, and actin cytoskeletal dynamics. For example, p21 Cip1 binds to the transcription factors E2F-1, STAT3 and c-Myc to inhibit their transcriptional activities, and binds to p300/CBP to block its transcriptional repressor activity. p21 Cip1 also binds and inhibits JNK1/SAPK kinase and MAPK-kinase-kinase ASK1/MEKK5 to block stress-induced apoptosis. The antiapoptotic activities are attributed to the cytoplasmic pool of p21 Cip1 . Finally all three Cip/Kip CKIs regulate actin cytoskeletal dynamics by modulating the RhoA-ROCK-LIMK-cofilin pathway, leading to redistribution of monomeric actin and contributing to increase cell motility and morphological changes. These divergent functions are performed in distinct cellular compartments and contribute to the seemingly contradictory observation that the CKIs can both suppress or promote tumor formation.





Pathological Consequences of Cell Cycle Deregulation or Dysregulation


Because regulation of the cell cycle is central to the control of cell proliferation, it is not surprising that cancers are often the results of deregulation or dysregulation of the cell cycle. Take colorectal cancer, for example, recent genomic-scale sequencing studies have identified numerous somatic mutations in genes that possibly are involved in the formation of cancer. Among these, some of the most highly ranked “cancer genes” are either directly or indirectly involved in the regulation of the cell cycle. Examples include p53, adenomatous polyposis coli (APC), KRAS, F-box and WD40 domain protein 7 (FBXW7), and phosphatidylinositol 3-kinase, catalytic, alpha subunit (PI3KCA). Here we briefly review some of the common cancer genes with cell cycle regulatory functions.



Retinoblastoma (RB) Tumor Suppressor Gene


The RB tumor suppressor protein limits cell proliferation by preventing entry into the S phase of the cell cycle. RB achieves its inhibitory effect by blocking the activity of E2F. Progression into S phase occurs when the ability of RB to suppress E2F is disrupted by the hyperphosphorylation of RB by cyclin D- and cyclin E-dependent CDKs in the G 1 phase of the cell cycle. The INK4 family of CKIs, particularly p16 INKa , directly inhibits activities of the cyclin D-dependent kinases, CDK4 and CDK6, thus maintaining RB in its active, antiproliferative state. Functional disruption of the tumor suppressors, p16 INKa and RB, or overexpression of the proto-oncogene products, cyclin D1 and CDK4, occur in many human cancers, prompting the speculation that disabling the “RB pathway” is an essential part of cancer formation. In addition to being causative in RB, loss of RB has been found in many other cancers. Similarly, inactivating mutations of 16 INK4a have been identified in numerous tumors. The often exclusive nature of mutations that result in RB or p16 INK4a loss, and/or cyclin D1 or CDK4 overexpression, suggest that each of these cell cycle regulatory genes exerts a critical function in the RB pathway of cancer formation. As such RB and its effectors prove to be potentially powerful predictive, prognostic and therapeutic target in cancer.



p53 and INK4a/ARF Tumor Suppressor Genes


The tumor suppressor p53 is mutated in more than 50% of human cancers. It has been estimated that cancers derived from over 50 human cell types or tissues contain mutations in the p53 gene. p53 is a labile protein but accumulates in response to cellular stresses from DNA damage, hypoxia, or oncogenic activation. Upon stabilization and activation, p53 initiates a transcriptional program that triggers either cell cycle arrest or apoptosis. Among the p53-responsive genes are p21 Cip1 , BCL2-associated X protein (BAX), and mouse double-minute 2, homolog (MDM2). While p21 Cip1 regulates progression of the cell cycle by inhibiting cyclins (E, A, and B)-CDK2 complexes, BAX causes apoptosis. The transcriptional induction of MDM2 by p53 is a negative feedback mechanism as binding of MDM2, an E3 ubiquitin ligase, to p53 induces ubiquitination of p53 and subsequent degradation. MDM2, in turn, is negatively regulated by the ARF (alternative reading frame) tumor suppressor (p14 ARF in humans and p19 ARF in mice). ARF directly associates with MDM2 to block its interaction with p53, therefore stabilizing p53. Thus, disruption of the ARF-MDM2-p53 signaling pathway is a common feature in cancers. This is supported by the finding that MDM2 is overexpressed in 5%–10% of human cancers, whereas ARF is silenced or deleted in many others. It is of interest to note that ARF is derived from an alternative reading frame in exon 2 of the INK4a gene. Loss of the INK4a/ARF locus therefore predisposes to many tumor types due to the dual disruption of the RB-E2F and MDM2-p53 pathways.



hCDC4/FBXW7 Tumor Suppressor Gene


Abundant evidence indicates that ubiquitin-dependent proteolysis regulates many aspects of the cell cycle and that dysregulation of the ubiquitin-proteosome pathway contributes to the formation of cancers. The proteolytic regulation of cell division is primarily controlled by two ubiquitin ligase systems, APC/C and SCF. The APC/C ubiquitin ligase is a multimeric protein complex that regulates chromosomes segregation (see above, The Mitotic Checkpoint or Spindle Assembly Checkpoint). SCF ligase complexes are structurally similar to APC/C and composed of an invariable core complex of SKP1, CUL1, and RBX1, and associated with a variable member of the F-box protein family that serves as the substrate recognition component. Among the approximately 70 different F-box proteins that were identified, the human homologue of yeast CDC4 (hCDC), also known as F-box and WD40-domain protein 7 (FBXW7), has been extensively characterized and implicated in human tumorigenesis. For example, among the approximately 140 mutated “cancer genes” identified in a recent genomic screen in colorectal cancers, FBXW7 ranks fifth as the most frequently mutated gene (after APC, KRAS, p53, and PI3K). Recurrent alterations in FBXW7 are also found in colorectal adenomas.


Human FBXW7 exists in three different isoforms, α, β, and γ, each with a unique amino terminal end fused to a common carboxyl terminal. The interaction between FBXW7 and its substrates depends on phosphorylation of the substrate within a motif called the CDC4-phophodegredron or CPD. This feature enables FBXW7 to simultaneously regulate a host of substrates by ubiquitination. Among the many substrates for FBXW7, some are critically involved in the regulation of the cell cycle such as cyclin E, c-Myc, c-Jun, and Notch. Mutations in the FBXW7 gene therefore lead to stabilization and elevated levels of these substrates. It is no wonder that FBXW7 is such a commonly mutated gene in human cancers.


One of the best characterized substrates of FBXW7 is cyclin E, which is essential for entry into S phase from G 1 phase in the cell cycle. Cyclin E level is elevated or dysregulated in many human cancers, resulting in dysfunction of the cell cycle. The consequences of cyclin E deregulation are multitude and include genetic instability, centrosome amplification, and fork collapse during DNA replication. Several studies subsequently identified cyclin E as a substrate for FBXW7, which mediates the phosphorylation- and ubiquitination-dependent degradation of cyclin E. Thus, it appears that tumorigenesis secondary to cyclin E deregulation is linked to altered function of FBXW7/hCDC4.



Chromosomal Instability (CIN) and Aneuploidy as Consequences of an Aberrant Mitotic Checkpoint


Genetic instability has long been recognized as an integral part of human cancers. In colorectal cancer (CRC), there are two major forms of genetic instability: microsatellite instability (MIN) and chromosomal instability (CIN). MIN tumors have mutations in the DNA mismatch repair (MMR) genes and accounts for approximately 15%–20% of sporadic CRC. The remainders of the sporadic CRC have CIN, which are frequently aneuploid, that is, they exhibit alterations in the number of chromosomes. It has been suggested that CIN is the driving force for the formation of aneuploidy and tumorigenesis. Although there has not been a unified mechanism responsible for CIN, defects in several cellular processes have been causally linked to its formation. These include chromosome dynamics (e.g., chromosome condensation, segregation, cohesion, and kinetochore-spindle interaction), centrosome duplication, cell cycle checkpoints (include G 1 , S, G 2, and the SACs), DNA damage repair pathway, and telomere functions. Among these potential factors contributing to CIN, the mitotic checkpoint is probably the most important one since it is an essential part of the cell cycle that ensures equal distribution of chromosomes upon the conclusion of cell division. Studies in mice lacking specific components of the mitotic checkpoint support this view. Mice with genetically reduced levels of mitotic checkpoint proteins including MAD1, MAD2, BUB1, BUB3, BUBR1, and centromeres protein E (CENP-E) all have increased level of aneuploidy and CIN, with the eventual formation of tumors in some animals. Importantly, somatic mutations of many of the same genes have been identified in human cancers, indicating the importance of the mitotic checkpoint in maintaining genomic integrity. These findings render the concept of exploiting the genetic instability of cancer cells through therapeutic intervention of the mitotic checkpoint a distinct possibility.



The Adenomatous Polyposis Coli Tumor Suppressor Gene


Germline mutations of the APC (note: different from APC/C) tumor suppressor gene cause familial adenomatous polyposis (FAP), an autosomal dominant disorder characterized by the presence of numerous colonic adenomas and colon cancer early in life. Somatic mutations in the APC gene were subsequently found to be present in the majority of sporadic colorectal cancer. Its involvement in tumor initiation leads to the hypothesis that APC functions as a “gatekeeper” of colonic epithelial cell proliferation and is responsible for the maintenance of colonic epithelial cell renewal. A telling insight into the function of APC came from the observation that it interacts with β-catenin. β-Catenin is implicated in regulating the Wnt signaling pathway of growth control. Particularly, Wnt signaling is now a well-established central mechanism for regulating proliferation of intestinal epithelial stem cells that are the precursors to all subsequent intestinal epithelial cell lineages.


Wnt signaling is normally absent in a quiescent, noncycling cell. This is accompanied by the sequestration of β-catenin in a “destruction complex” in the cytoplasm that includes APC, Axin, casein kinase 1 (CK1), and glycogen synthase kinase 3 (GSK3) ( Fig. 8.8 A). This complex leads to the phosphorylation of β-catenin, which is then degraded by ubiquitin-mediated proteasomal degradation. When Wnt, a secreted glycoprotein, is present, it binds to the cell surface receptors Frizzled (Fz) and lipoprotein receptor-related protein (LRP). This leads to activation of the protein disheveled (Dsh) and subsequent release of β-catenin from the destruction complex, resulting in accumulation of free β-catenin in the cytoplasm ( Fig. 8.8 B). Some of this free β-catenin is shuttled into the nucleus where it becomes associated with the transcription factor, T-cell factor (TCF) to activate target gene expression ( Fig. 8.8 B). Among the genes stimulated by the β-catenin/TCF complexes are those encoding cyclin D1 and c-Myc, both of which are critical for the progression of the cell cycle. It is important to note that mutational inactivation of APC or constitutional mutational activation of β-catenin, two common events in the pathogenesis of colorectal cancer, results in a similar net effect of nuclear β-catenin accumulation to that seen during Wnt signaling. These mutations then lead to unregulated cell proliferation.




Fig. 8.8


Schematic presentation of the Wnt signaling pathway. (A) Wnt is absent and (B) Wnt is present. β-cat = β-catenin; Fz = Frizzled; LRP = lipoprotein receptor-related protein; Dsh = disheveled; CK1 = casein kinase 1; GSK3 = glycogen synthase kinase 3; TCF = T-cell factor.

(Adapted with from reference Yang VW. APC as a checkpoint gene: the beginning or the end? Gastroenterology 2002;123(3):935–9, with permission from Elsevier.)


APC has also been implicated in the regulation of mitosis in a β-catenin-independent manner. Here APC is localized to the ends of microtubules embedded in kinetochores and forms a complex with several SAC proteins. This finding suggests that APC is involved in chromosome segregation. The role of APC in regulation of mitosis is consistent with the finding that mutations in the APC gene cause chromosomal instability. Thus, APC may fulfill two requirements for a colonic epithelial cell to develop into colon cancer—the cell must acquire a selective advantage to allow for an initial clonal expansion, and generate genetic instability to allow for multiple hits in other genes that are responsible for tumor progression and malignant transformation.



Cell Cycle Regulators as Targets for Cancer Treatment


Because deregulation or dysregulation of the cell cycle is frequently found in cancer, the cell cycle regulatory proteins are logical targets for development of novel theories for cancer. Recent studies have demonstrated that orally available small-molecule inhibitors of the cyclin D-dependent CDK4 and CDK6, when combined with established therapies, have potential in the treatment of certain cancers. A recently completed phase 3 clinical trial is an example of a success story that the combination of palbociclib, a CDK4-CDK6 inhibitor, and letrozole, an aromatase inhibitor, is highly effective in the treatment of advanced breast cancer. Given that many additional novel similar agents are in various stages of development, cell cycle-based targeted therapy promises to become part of a new armamentarium in the combat against cancer.





Conclusion


The cell cycle is a fundamental cellular process in which a cell duplicates itself in a highly faithful manner to maintain genomic integrity. The cell cycle is divided into several distinct phases that are controlled by specific pairs of cyclin-CDK. The cell cycle also contains specific checkpoints with which to monitor the presence of damaged or incompletely replicated DNA in interphase, or unaligned chromosomes in mitosis. Dysregulation or deregulation of the cell cycle often have pathological consequences and a good example is cancer. Hence, many of the cell cycle regulators act as tumor suppressors or oncoproteins such as Rb, p53, FBXW7, and cyclins. The relevance of cell cycle regulation also has an impact on the physiology of the gastrointestinal tract as the epithelium is composed of rapidly proliferating cells. A prominent example is the crucial role of the Wnt signaling pathway in regulating proliferation of intestinal stem cells and that a critical component of the Wnt pathway, the APC protein, is mutated in the majority of colorectal cancer. Recent clinical studies have also shown promising results in cancer treatment by targeting cell cycle regulatory proteins. Thus, a clear understanding of the cell cycle machinery is essential not only for understanding the mechanisms regulating proliferation of intestinal epithelial cells, but developing novel therapies for cancer.


References



  1. 1. Sanso M., and Fisher R.P.: Modelling the CDK-dependent transcription cycle in fission yeast. Biochem Soc Trans 2013; 41s: pp. 1660-1665

  2. 2. Chang F., and Nurse P.: Finishing the cell cycle: control of mitosis and cytokinesis in fission yeast. Trends Genet 1993; 9: pp. 333-335

  3. 3. Bartlett R., and Nurse P.: Yeast as a model system for understanding the control of DNA replication in Eukaryotes. BioEssays 1990; 12: pp. 457-463

  4. 4. Nurse P.: The Josef Steiner lecture: CDKs and cell-cycle control in fission yeast: relevance to other eukaryotes and cancer. Int J Cancer 1997; 71: pp. 707-708

  5. 5. Woo R.A., and Poon R.Y.: Cyclin-dependent kinases and S phase control in mammalian cells. Cell Cycle 2003; 2: pp. 316-324

  6. 6. Malumbres M., and Barbacid M.: Mammalian cyclin-dependent kinases. Trends Biochem Sci 2005; 30: pp. 630-641

  7. 7. Sherr C.J., and Roberts J.M.: Living with or without cyclins and cyclin-dependent kinases. Genes Dev 2004; 18: pp. 2699-2711

  8. 8. Sanchez I., and Dynlacht B.D.: New insights into cyclins, CDKs, and cell cycle control. Semin Cell Dev Biol 2005; 16: pp. 311-321

  9. 9. Harper J.V., and Brooks G.: The mammalian cell cycle: an overview. Methods Mol Biol 2005; 296: pp. 113-153

  10. 10. Nurse P.M.: Nobel lecture. Cyclin dependent kinases and cell cycle control. Biosci Rep 2002; 22: pp. 487-499

  11. 11. Doree M., and Hunt T.: From Cdc2 to Cdk1: when did the cell cycle kinase join its cyclin partner? J Cell Sci 2002; 115: pp. 2461-2464

  12. 12. Blagosklonny M.V., and Pardee A.B.: The restriction point of the cell cycle. Cell Cycle 2002; 1: pp. 103-110

  13. 13. Musgrove E.A.: Cyclins: roles in mitogenic signaling and oncogenic transformation. Growth Factors 2006; 24: pp. 13-19

  14. 14. Choi Y.J., and Anders L.: Signaling through cyclin D-dependent kinases. Oncogene 2014; 33: pp. 1890-1903

  15. 15. Sherr C.J., and Roberts J.M.: CDK inhibitors: positive and negative regulators of G1-phase progression. Genes Dev 1999; 13: pp. 1501-1512

  16. 16. Filmus J., Robles A.I., Shi W., Wong M.J., Colombo L.L., and Conti C.J.: Induction of cyclin D1 overexpression by activated ras. Oncogene 1994; 9: pp. 3627-3633

  17. 17. Winston J.T., Coats S.R., Wang Y.Z., and Pledger W.J.: Regulation of the cell cycle machinery by oncogenic ras. Oncogene 1996; 12: pp. 127-134

  18. 18. Lavoie J.N., L’Allemain G., Brunet A., Muller R., and Pouyssegur J.: Cyclin D1 expression is regulated positively by the p42/p44MAPK and negatively by the p38/HOGMAPK pathway. J Biol Chem 1996; 271: pp. 20608-20616

  19. 19. Aktas H., Cai H., and Cooper G.M.: Ras links growth factor signaling to the cell cycle machinery via regulation of cyclin D1 and the Cdk inhibitor p27KIP1. Mol Cell Biol 1997; 17: pp. 3850-3857

  20. 20. Koepp D.M., Harper J.W., and Elledge S.J.: How the cyclin became a cyclin: regulated proteolysis in the cell cycle. Cell 1999; 97: pp. 431-434

  21. 21. Diehl J.A., Zindy F., and Sherr C.J.: Inhibition of cyclin D1 phosphorylation on threonine-286 prevents its rapid degradation via the ubiquitin- proteasome pathway. Genes Dev 1997; 11: pp. 957-972

  22. 22. Diehl J.A., Cheng M., Roussel M.F., and Sherr C.J.: Glycogen synthase kinase-3beta regulates cyclin D1 proteolysis and subcellular localization. Genes Dev 1998; 12: pp. 3499-3511

  23. 23. Jones S.M., and Kazlauskas A.: Growth-factor-dependent mitogenesis requires two distinct phases of signalling. Nat Cell Biol 2001; 3: pp. 165-172

  24. 24. Bartek J., Bartkova J., and Lukas J.: The retinoblastoma protein pathway and the restriction point. Curr Opin Cell Biol 1996; 8: pp. 805-814

  25. 25. Weinberg R.A.: The retinoblastoma protein and cell cycle control. Cell 1995; 81: pp. 323-330

  26. 26. Harbour J.W., and Dean D.C.: The Rb/E2F pathway: expanding roles and emerging paradigms. Genes Dev 2000; 14: pp. 2393-2409

  27. 27. Cobrinik D.: Pocket proteins and cell cycle control. Oncogene 2005; 24: pp. 2796-2809

  28. 28. Chellappan S.P., Hiebert S., Mudryj M., Horowitz J.M., and Nevins J.R.: The E2F transcription factor is a cellular target for the RB protein. Cell 1991; 65: pp. 1053-1061

  29. 29. Nevins J.R.: Toward an understanding of the functional complexity of the E2F and retinoblastoma families. Cell Growth Differ 1998; 9: pp. 585-593

  30. 30. Dyson N.: The regulation of E2F by pRB-family proteins. Genes Dev 1998; 12: pp. 2245-2262

  31. 31. Blake M.C., and Azizkhan J.C.: Transcription factor E2F is required for efficient expression of the hamster dihydrofolate reductase gene in vitro and in vivo. Mol Cell Biol 1989; 9: pp. 4994-5002

  32. 32. Thalmeier K., Synovzik H., Mertz R., Winnacker E.L., and Lipp M.: Nuclear factor E2F mediates basic transcription and trans-activation by E1a of the human MYC promoter. Genes Dev 1989; 3: pp. 527-536

  33. 33. Dalton S.: Cell cycle regulation of the human cdc2 gene. EMBO J 1992; 11: pp. 1797-1804

  34. 34. DeGregori J., Leone G., Ohtani K., Miron A., and Nevins J.R.: E2F-1 accumulation bypasses a G1 arrest resulting from the inhibition of G1 cyclin-dependent kinase activity. Genes Dev 1995; 9: pp. 2873-2887

  35. 35. Magnaghi-Jaulin L., Groisman R., Naguibneva I., Robin P., Lorain S., Le Villain J.P., et al: Retinoblastoma protein represses transcription by recruiting a histone deacetylase. Nature 1998; 391: pp. 601-605

  36. 36. Kingston R.E., and Narlikar G.J.: ATP-dependent remodeling and acetylation as regulators of chromatin fluidity. Genes Dev 1999; 13: pp. 2339-2352

  37. 37. Kornberg R.D., and Lorch Y.: Chromatin-modifying and -remodeling complexes. Curr Opin Genet Dev 1999; 9: pp. 148-151

  38. 38. Collingwood T.N., Urnov F.D., and Wolffe A.P.: Nuclear receptors: coactivators, corepressors and chromatin remodeling in the control of transcription. J Mol Endocrinol 1999; 23: pp. 255-275

  39. 39. Wolffe A.P., and Hayes J.J.: Chromatin disruption and modification. Nucleic Acids Res 1999; 27: pp. 711-720

  40. 40. Weintraub S.J., Chow K.N., Luo R.X., Zhang S.H., He S., and Dean D.C.: Mechanism of active transcriptional repression by the retinoblastoma protein. Nature 1995; 375: pp. 812-815

  41. 41. Meloni A.R., Smith E.J., and Nevins J.R.: A mechanism for Rb/p130-mediated transcription repression involving recruitment of the CtBP corepressor. Proc Natl Acad Sci U S A 1999; 96: pp. 9574-9579

  42. 42. Ferreira R., Magnaghi-Jaulin L., Robin P., Harel-Bellan A., and Trouche D.: The three members of the pocket proteins family share the ability to repress E2F activity through recruitment of a histone deacetylase. Proc Natl Acad Sci U S A 1998; 95: pp. 10493-10498

  43. 43. Chow K.N., Starostik P., and Dean D.C.: The Rb family contains a conserved cyclin-dependent-kinase-regulated transcriptional repressor motif. Mol Cell Biol 1996; 16: pp. 7173-7181

  44. 44. Hinds P.W., Mittnacht S., Dulic V., Arnold A., Reed S.I., and Weinberg R.A.: Regulation of retinoblastoma protein functions by ectopic expression of human cyclins. Cell 1992; 70: pp. 993-1006

  45. 45. Lundberg A.S., and Weinberg R.A.: Functional inactivation of the retinoblastoma protein requires sequential modification by at least two distinct cyclin-cdk complexes. Mol Cell Biol 1998; 18: pp. 753-761

  46. 46. Chen P.L., Scully P., Shew J.Y., Wang J.Y., and Lee W.H.: Phosphorylation of the retinoblastoma gene product is modulated during the cell cycle and cellular differentiation. Cell 1989; 58: pp. 1193-1198

  47. 47. Sherr C.J.: Cancer cell cycles. Science 1996; 274: pp. 1672-1677

  48. 48. Bertoli C., and de Bruin R.A.: Turning cell cycle entry on its head. elife 2014; 3:

  49. 49. Narasimha A.M., Kaulich M., Shapiro G.S., Choi Y.J., Sicinski P., and Dowdy S.F.: Cyclin D activates the Rb tumor suppressor by mono-phosphorylation. elife 2014; 3:

  50. 50. Lam E.W., and La Thangue N.B.: DP and E2F proteins: coordinating transcription with cell cycle progression. Curr Opin Cell Biol 1994; 6: pp. 859-866

  51. 51. La Thangue N.B.: DP and E2F proteins: components of a heterodimeric transcription factor implicated in cell cycle control. Curr Opin Cell Biol 1994; 6: pp. 443-450

  52. 52. Adams P.D., and Kaelin W.G.: Transcriptional control by E2F. Semin Cancer Biol 1995; 6: pp. 99-108

  53. 53. La Thangue N.B.: E2F and the molecular mechanisms of early cell-cycle control. Biochem Soc Trans 1996; 24: pp. 54-59

  54. 54. Lavia P., and Jansen-Durr P.: E2F target genes and cell-cycle checkpoint control. BioEssays 1999; 21: pp. 221-230

  55. 55. Berckmans B., and De Veylder L.: Transcriptional control of the cell cycle. Curr Opin Plant Biol 2009; 12: pp. 599-605

  56. 56. Brehm A., Miska E., Reid J., Bannister A., and Kouzarides T.: The cell cycle-regulating transcription factors E2F-RB. Br J Cancer 1999; 80: pp. 38-41

  57. 57. Hateboer G., Wobst A., Petersen B.O., Le Cam L., Vigo E., Sardet C., et al: Cell cycle-regulated expression of mammalian CDC6 is dependent on E2F. Mol Cell Biol 1998; 18: pp. 6679-6697

  58. 58. Black A.R., and Azizkhan-Clifford J.: Regulation of E2F: a family of transcription factors involved in proliferation control. Gene 1999; 237: pp. 281-302

  59. 59. Dimova D.K., and Dyson N.J.: The E2F transcriptional network: old acquaintances with new faces. Oncogene 2005; 24: pp. 2810-2826

  60. 60. Bracken A.P., Ciro M., Cocito A., and Helin K.: E2F target genes: unraveling the biology. Trends Biochem Sci 2004; 29: pp. 409-417

  61. 61. Ohtani K.: Implication of transcription factor E2F in regulation of DNA replication. Front Biosci 1999; 4: pp. D793-D804

  62. 62. Ohtani K., Iwanaga R., Nakamura M., Ikeda M., Yabuta N., Tsuruga H., et al: Cell growth-regulated expression of mammalian MCM5 and MCM6 genes mediated by the transcription factor E2F. Oncogene 1999; 18: pp. 2299-2309

  63. 63. Helin K., CL W., Fattaey A.R., Lees J.A., Dynlacht B.D., Ngwu C., et al: Heterodimerization of the transcription factors E2F-1 and DP-1 leads to cooperative trans-activation. Genes Dev 1993; 7: pp. 1850-1861

  64. 64. Harbour J.W., Luo R.X., Dei Santi A., Postigo A.A., and Dean D.C.: Cdk phosphorylation triggers sequential intramolecular interactions that progressively block Rb functions as cells move through G1. Cell 1999; 98: pp. 859-869

  65. 65. Zhang H.S., Postigo A.A., and Dean D.C.: Active transcriptional repression by the Rb-E2F complex mediates G1 arrest triggered by p16INK4a, TGFbeta, and contact inhibition. Cell 1999; 97: pp. 53-61

  66. 66. Geng Y., Whoriskey W., Park M.Y., Bronson R.T., Medema R.H., Li T., et al: Rescue of cyclin D1 deficiency by knockin cyclin E. Cell 1999; 97: pp. 767-777

  67. 67. Bertoli C., Skotheim J.M., and de Bruin R.A.: Control of cell cycle transcription during G1 and S phases. Nat Rev Mol Cell Biol 2013; 14: pp. 518-528

  68. 68. Dulic V., Lees E., and Reed S.I.: Association of human cyclin E with a periodic G1-S phase protein kinase. Science 1992; 257: pp. 1958-1961

  69. 69. Koff A., Giordano A., Desai D., Yamashita K., Harper J.W., Elledge S., et al: Formation and activation of a cyclin E-cdk2 complex during the G1 phase of the human cell cycle. Science 1992; 257: pp. 1689-1694

  70. 70. Yu Q., and Sicinski P.: Mammalian cell cycles without cyclin E-CDK2. Cell Cycle 2004; 3: pp. 292-295

  71. 71. Teixeira L.K., Wang X., Li Y., Ekholm-Reed S., Wu X., Wang P., et al: Cyclin E deregulation promotes loss of specific genomic regions. Curr Biol 2015; 25: pp. 1327-1333

  72. 72. Hu W., Nevzorova Y.A., Haas U., Moro N., Sicinski P., Geng Y., et al: Concurrent deletion of cyclin E1 and cyclin-dependent kinase 2 in hepatocytes inhibits DNA replication and liver regeneration in mice. Hepatology 2014; 59: pp. 651-660

  73. 73. Clurman B.E., Sheaff R.J., Thress K., Groudine M., and Roberts J.M.: Turnover of cyclin E by the ubiquitin-proteasome pathway is regulated by cdk2 binding and cyclin phosphorylation. Genes Dev 1996; 10: pp. 1979-1990

  74. 74. Won K.A., and Reed S.I.: Activation of cyclin E/CDK2 is coupled to site-specific autophosphorylation and ubiquitin-dependent degradation of cyclin E. EMBO J 1996; 15: pp. 4182-4193

  75. 75. Welcker M., Singer J., Loeb K.R., Grim J., Bloecher A., Gurien-West M., et al: Multisite phosphorylation by Cdk2 and GSK3 controls cyclin E degradation. Mol Cell 2003; 12: pp. 381-392

  76. 76. Sweeney C., Murphy M., Kubelka M., Ravnik S.E., Hawkins C.F., Wolgemuth D.J., et al: A distinct cyclin A is expressed in germ cells in the mouse. Development 1996; 122: pp. 53-64

  77. 77. Krude T., Jackman M., Pines J., and Laskey R.A.: Cyclin/Cdk-dependent initiation of DNA replication in a human cell-free system. Cell 1997; 88: pp. 109-119

  78. 78. Hua X.H., and Newport J.: Identification of a preinitiation step in DNA replication that is independent of origin recognition complex and cdc6, but dependent on cdk2. J Cell Biol 1998; 140: pp. 271-281

  79. 79. Coverley D., Laman H., and Laskey R.A.: Distinct roles for cyclins E and A during DNA replication complex assembly and activation. Nat Cell Biol 2002; 4: pp. 523-528

  80. 80. Mitra J., and Enders G.H.: Cyclin A/Cdk2 complexes regulate activation of Cdk1 and Cdc25 phosphatases in human cells. Oncogene 2004; 23: pp. 3361-3367

  81. 81. Nguyen T.B., Manova K., Capodieci P., Lindon C., Bottega S., Wang X.Y., et al: Characterization and expression of mammalian cyclin b3, a prepachytene meiotic cyclin. J Biol Chem 2002; 277: pp. 41960-41969

  82. 82. Nigg E.A.: Mitotic kinases as regulators of cell division and its checkpoints. Nat Rev Mol Cell Biol 2001; 2: pp. 21-32

  83. 83. Ubersax J.A., Woodbury E.L., Quang P.N., Paraz M., Blethrow J.D., Shah K., et al: Targets of the cyclin-dependent kinase Cdk1. Nature 2003; 425: pp. 859-864

  84. 84. Fisher D., Krasinska L., Coudreuse D., and Novak B.: Phosphorylation network dynamics in the control of cell cycle transitions. J Cell Sci 2012; 125: pp. 4703-4711

  85. 85. Crasta K., Lim H.H., Giddings T.H., Winey M., and Surana U.: Inactivation of Cdh1 by synergistic action of Cdk1 and polo kinase is necessary for proper assembly of the mitotic spindle. Nat Cell Biol 2008; 10: pp. 665-675

  86. 86. McHugh B., and Heck M.M.: Regulation of chromosome condensation and segregation. Curr Opin Genet Dev 2003; 13: pp. 185-190

  87. 87. Meyer H., Drozdowska A., and Dobrynin G.: A role for Cdc48/p97 and Aurora B in controlling chromatin condensation during exit from mitosis. Biochem Cell Biol 2010; 88: pp. 23-28

  88. 88. Gavet O., and Pines J.: Activation of cyclin B1-Cdk1 synchronizes events in the nucleus and the cytoplasm at mitosis. J Cell Biol 2010; 189: pp. 247-259

  89. 89. Draviam V.M., Orrechia S., Lowe M., Pardi R., and Pines J.: The localization of human cyclins B1 and B2 determines CDK1 substrate specificity and neither enzyme requires MEK to disassemble the Golgi apparatus. J Cell Biol 2001; 152: pp. 945-958

  90. 90. Harper J.W., Burton J.L., and Solomon M.J.: The anaphase-promoting complex: iťs not just for mitosis any more. Genes Dev 2002; 16: pp. 2179-2206

  91. 91. Noton E., and Diffley J.F.: CDK inactivation is the only essential function of the APC/C and the mitotic exit network proteins for origin resetting during mitosis. Mol Cell 2000; 5: pp. 85-95

  92. 92. Peters J.M.: The anaphase-promoting complex: proteolysis in mitosis and beyond. Mol Cell 2002; 9: pp. 931-943

  93. 93. Castro A., Bernis C., Vigneron S., Labbe J.C., and Lorca T.: The anaphase-promoting complex: a key factor in the regulation of cell cycle. Oncogene 2005; 24: pp. 314-325

  94. 94. Teixeira L.K., and Reed S.I.: Ubiquitin ligases and cell cycle control. Annu Rev Biochem 2013; 82: pp. 387-414

  95. 95. Mocciaro A., and Rape M.: Emerging regulatory mechanisms in ubiquitin- dependent cell cycle control. J Cell Sci 2012; 125: pp. 255-263

  96. 96. Russell P., and Nurse P.: [object Object]. Cell 1986; 45: pp. 781-782

  97. 97. Lee M.G., and Nurse P.: Cell cycle genes of the fission yeast. Sci Prog 1987; 71: pp. 1-14

  98. 98. Draetta G., Brizuela L., Potashkin J., and Beach D.: Identification of p34 and p13, human homologs of the cell cycle regulators of fission yeast encoded by cdc2 . Cell 1987; 50: pp. 319-325

  99. 99. Lee M.G., and Nurse P.: Complementation used to clone a human homologue of the fission yeast cell cycle control gene cdc2. Nature 1987; 327: pp. 31-35

  100. 100. Hanks S.K.: Homology probing: identification of cDNA clones encoding members of the protein-serine kinase family. Proc Natl Acad Sci U S A 1987; 84: pp. 388-392

  101. 101. Elledge S.J., and Spottswood M.R.: A new human p34 protein kinase, CDK2, identified by complementation of a cdc28 mutation in . EMBO J 1991; 10: pp. 2653-2659

  102. 102. Paris J., Le Guellec R., Couturier A., Le Guellec K., Omilli F., Camonis J., et al: Cloning by differential screening of a Xenopus cDNA coding for a protein highly homologous to cdc2. Proc Natl Acad Sci U S A 1991; 88: pp. 1039-1043

  103. 103. Tsai L.H., Harlow E., and Meyerson M.: Isolation of the human cdk2 gene that encodes the cyclin A- and adenovirus E1A-associated p33 kinase. Nature 1991; 353: pp. 174-177

  104. 104. Ninomiya-Tsuji J., Nomoto S., Yasuda H., Reed S.I., and Matsumoto K.: Cloning of a human cDNA encoding a CDC2-related kinase by complementation of a budding yeast cdc28 mutation. Proc Natl Acad Sci U S A 1991; 88: pp. 9006-9010

  105. 105. Matsushime H., Ewen M.E., Strom D.K., Kato J.Y., Hanks S.K., Roussel M.F., et al: Identification and properties of an atypical catalytic subunit (p34PSK-J3/cdk4) for mammalian D type G1 cyclins. Cell 1992; 71: pp. 323-334

  106. 106. Xiong Y., Zhang H., and Beach D.: D type cyclins associate with multiple protein kinases and the DNA replication and repair factor PCNA. Cell 1992; 71: pp. 505-514

  107. 107. Malumbres M., and Barbacid M.: To cycle or not to cycle: a critical decision in cancer. Nat Rev Cancer 2001; 1: pp. 222-231

  108. 108. Bendris N., Lemmers B., and Blanchard J.M.: Cell cycle, cytoskeleton dynamics and beyond: the many functions of cyclins and CDK inhibitors. Cell Cycle 2015; 14: pp. 1786-1798

  109. 109. Lolli G., and Johnson L.N.: CAK-Cyclin-dependent activating kinase: a key kinase in cell cycle control and a target for drugs? Cell Cycle 2005; 4: pp. 572-577

  110. 110. Kaldis P.: The cdk-activating kinase (CAK): from yeast to mammals. Cell Mol Life Sci 1999; 55: pp. 284-296

  111. 111. Fisher R.P.: Secrets of a double agent: CDK7 in cell-cycle control and transcription. J Cell Sci 2005; 118: pp. 5171-5180

  112. 112. Yang L., MacLellan W.R., Han Z., Weiss J.N., and Qu Z.: Multisite phosphorylation and network dynamics of cyclin-dependent kinase signaling in the eukaryotic cell cycle. Biophys J 2004; 86: pp. 3432-3443

  113. 113. Watanabe N., Broome M., and Hunter T.: Regulation of the human WEE1Hu CDK tyrosine 15-kinase during the cell cycle. EMBO J 1995; 14: pp. 1878-1891

  114. 114. Berry L.D., and Gould K.L.: Regulation of Cdc2 activity by phosphorylation at T14/Y15. Prog Cell Cycle Res 1996; 2: pp. 99-105

  115. 115. Parker L.L., and Piwnica-Worms H.: Inactivation of the p34cdc2-cyclin B complex by the human WEE1 tyrosine kinase. Science 1992; 257: pp. 1955-1957

  116. 116. McGowan C.H., and Russell P.: Human Wee1 kinase inhibits cell division by phosphorylating p34cdc2 exclusively on Tyr15. EMBO J 1993; 12: pp. 75-85

  117. 117. Fattaey A., and Booher R.N.: Myt1: a Wee1-type kinase that phosphorylates Cdc2 on residue Thr14. Prog Cell Cycle Res 1997; 3: pp. 233-240

  118. 118. Leise W., and Mueller P.R.: Multiple Cdk1 inhibitory kinases regulate the cell cycle during development. Dev Biol 2002; 249: pp. 156-173

  119. 119. Sorensen C.S., and Syljuasen R.G.: Safeguarding genome integrity: the checkpoint kinases ATR, CHK1 and WEE1 restrain CDK activity during normal DNA replication. Nucleic Acids Res 2012; 40: pp. 477-486

  120. 120. Aressy B., and Ducommun B.: Cell cycle control by the CDC25 phosphatases. Anti Cancer Agents Med Chem 2008; 8: pp. 818-824

  121. 121. Rudolph J.: Cdc25 phosphatases: structure, specificity, and mechanism. Biochemistry 2007; 46: pp. 3595-3604

  122. 122. Coleman T.R., and Dunphy W.G.: Cdc2 regulatory factors. Curr Opin Cell Biol 1994; 6: pp. 877-882

  123. 123. Lavecchia A., Di Giovanni C., and Novellino E.: CDC25 phosphatase inhibitors: an update. Mini-Rev Med Chem 2012; 12: pp. 62-73

  124. 124. Santamaria D., and Ortega S.: Cyclins and CDKS in development and cancer: lessons from genetically modified mice. Front Biosci 2006; 11: pp. 1164-1188

  125. 125. Satyanarayana A., and Kaldis P.: Mammalian cell-cycle regulation: several Cdks, numerous cyclins and diverse compensatory mechanisms. Oncogene 2009; 28: pp. 2925-2939

  126. 126. Aleem E., and Kaldis P.: Mouse models of cell cycle regulators: new paradigms. Results Probl Cell Differ 2006; 42: pp. 271-328

  127. 127. Risal S., Adhikari D., and Liu K.: Animal models for studying the in vivo functions of cell cycle CDKs. Methods Mol Biol 2016; 1336: pp. 155-166

  128. 128. Santamaria D., Barriere C., Cerqueira A., Hunt S., Tardy C., Newton K., et al: Cdk1 is sufficient to drive the mammalian cell cycle. Nature 2007; 448: pp. 811-815

  129. 129. Serrano M., Hannon G.J., and Beach D.: A new regulatory motif in cell-cycle control causing specific inhibition of cyclin D/CDK4. Nature 1993; 366: pp. 704-707

  130. 130. Hannon G.J., and Beach D.: p15INK4B is a potential effector of TGF-beta-induced cell cycle arrest. Nature 1994; 371: pp. 257-261

  131. 131. Guan K.L., Jenkins C.W., Li Y., Nichols M.A., Wu X., O’Keefe C.L., et al: Growth suppression by p18, a p16INK4/MTS1- and p14INK4B/MTS2-related CDK6 inhibitor, correlates with wild-type pRb function. Genes Dev 1994; 8: pp. 2939-2952

  132. 132. Hirai H., Roussel M.F., Kato J.Y., Ashmun R.A., and Sherr C.J.: Novel INK4 proteins, p19 and p18, are specific inhibitors of the cyclin D-dependent kinases CDK4 and CDK6. Mol Cell Biol 1995; 15: pp. 2672-2681

  133. 133. Chan F.K., Zhang J., Cheng L., Shapiro D.N., and Winoto A.: Identification of human and mouse p19, a novel CDK4 and CDK6 inhibitor with homology to p16ink4. Mol Cell Biol 1995; 15: pp. 2682-2688

  134. 134. Sherr C.J., and Roberts J.M.: Inhibitors of mammalian G1 cyclin-dependent kinases. Genes Dev 1995; 9: pp. 1149-1163

  135. 135. Pavletich N.P.: Mechanisms of cyclin-dependent kinase regulation: structures of Cdks, their cyclin activators, and Cip and INK4 inhibitors. J Mol Biol 1999; 287: pp. 821-828

  136. 136. Gu Y., Turck C.W., and Morgan D.O.: Inhibition of CDK2 activity in vivo by an associated 20K regulatory subunit. Nature 1993; 366: pp. 707-710

  137. 137. Harper J.W., Adami G.R., Wei N., Keyomarsi K., and Elledge S.J.: The p21 Cdk-interacting protein Cip1 is a potent inhibitor of G1 cyclin-dependent kinases. Cell 1993; 75: pp. 805-816

  138. 138. el-Deiry W.S., Tokino T., Velculescu V.E., Levy D.B., Parsons R., Trent J.M., et al: WAF1, a potential mediator of p53 tumor suppression. Cell 1993; 75: pp. 817-825

  139. 139. Xiong Y., Hannon G.J., Zhang H., Casso D., Kobayashi R., and Beach D.: p21 is a universal inhibitor of cyclin kinases. Nature 1993; 366: pp. 701-704

  140. 140. Dulic V., Kaufmann W.K., Wilson S.J., Tlsty T.D., Lees E., Harper J.W., et al: p53-dependent inhibition of cyclin-dependent kinase activities in human fibroblasts during radiation-induced G1 arrest. Cell 1994; 76: pp. 1013-1023

  141. 141. Noda A., Ning Y., Venable S.F., Pereira-Smith O.M., and Smith J.R.: Cloning of senescent cell-derived inhibitors of DNA synthesis using an expression screen. Exp Cell Res 1994; 211: pp. 90-98

  142. 142. Polyak K., Lee M.H., Erdjument-Bromage H., Koff A., Roberts J.M., Tempst P., et al: Cloning of p27Kip1, a cyclin-dependent kinase inhibitor and a potential mediator of extracellular antimitogenic signals. Cell 1994; 78: pp. 59-66

  143. 143. Polyak K., Kato J.Y., Solomon M.J., Sherr C.J., Massague J., Roberts J.M., et al: p27Kip1, a cyclin-Cdk inhibitor, links transforming growth factor-beta and contact inhibition to cell cycle arrest. Genes Dev 1994; 8: pp. 9-22

  144. 144. Toyoshima H., and Hunter T.: p27, a novel inhibitor of G1 cyclin-Cdk protein kinase activity, is related to p21. Cell 1994; 78: pp. 67-74

  145. 145. Lee M.H., Reynisdottir I., and Massague J.: Cloning of p57KIP2, a cyclin-dependent kinase inhibitor with unique domain structure and tissue distribution. Genes Dev 1995; 9: pp. 639-649

  146. 146. Matsuoka S., Edwards M.C., Bai C., Parker S., Zhang P., Baldini A., et al: p57KIP2, a structurally distinct member of the p21CIP1 Cdk inhibitor family, is a candidate tumor suppressor gene. Genes Dev 1995; 9: pp. 650-662

  147. 147. Chen J., Jackson P.K., Kirschner M.W., and Dutta A.: Separate domains of p21 involved in the inhibition of Cdk kinase and PCNA. Nature 1995; 374: pp. 386-388

  148. 148. Chen J., Saha P., Kornbluth S., Dynlacht B.D., and Dutta A.: Cyclin-binding motifs are essential for the function of p21CIP1. Mol Cell Biol 1996; 16: pp. 4673-4682

  149. 149. Nakanishi M., Robetorye R.S., Adami G.R., Pereira-Smith O.M., and Smith J.R.: Identification of the active region of the DNA synthesis inhibitory gene p21Sdi1/CIP1/WAF1. EMBO J 1995; 14: pp. 555-563

  150. 150. Warbrick E., Lane D.P., Glover D.M., and Cox L.S.: A small peptide inhibitor of DNA replication defines the site of interaction between the cyclin-dependent kinase inhibitor p21WAF1 and proliferating cell nuclear antigen. Curr Biol 1995; 5: pp. 275-282

  151. 151. Lin J., Reichner C., Wu X., and Levine A.J.: Analysis of wild-type and mutant p21WAF-1 gene activities. Mol Cell Biol 1996; 16: pp. 1786-1793

  152. 152. Russo A.A., Jeffrey P.D., Patten A.K., Massague J., and Pavletich N.P.: Crystal structure of the p27Kip1 cyclin-dependent-kinase inhibitor bound to the cyclin A-Cdk2 complex. Nature 1996; 382: pp. 325-331

  153. 153. Elledge S.J.: Cell cycle checkpoints: preventing an identity crisis. Science 1996; 274: pp. 1664-1672

  154. 154. Murray A.: Cell cycle checkpoints. Curr Opin Cell Biol 1994; 6: pp. 872-876

  155. 155. Pietenpol J.A., and Stewart Z.A.: Cell cycle checkpoint signaling: cell cycle arrest versus apoptosis. Toxicology 2002; 181–182: pp. 475-481

  156. 156. Xie S., Xie B., Lee M.Y., and Dai W.: Regulation of cell cycle checkpoints by polo-like kinases. Oncogene 2005; 24: pp. 277-286

  157. 157. O’Connor P.M.: Mammalian G1 and G2 phase checkpoints. Cancer Surv 1997; 29: pp. 151-182

  158. 158. Donjerkovic D., and Scott D.W.: Regulation of the G1 phase of the mammalian cell cycle. Cell Res 2000; 10: pp. 1-16

  159. 159. Malumbres M.: Revisiting the “Cdk-centric” view of the mammalian cell cycle. Cell Cycle 2005; 4: pp. 206-210

  160. 160. Holland A.J., and Cleveland D.W.: Boveri revisited: chromosomal instability, aneuploidy and tumorigenesis. Nat Rev Mol Cell Biol 2009; 10: pp. 478-487

  161. 161. Yen T.J., and Kao G.D.: Mitotic checkpoint, aneuploidy and cancer. Adv Exp Med Biol 2005; 570: pp. 477-499

  162. 162. Qi W., and Yu H.: The spindle checkpoint and chromosomal stability. Genome Dyn 2006; 1: pp. 116-130

  163. 163. Kops G.J.: The kinetochore and spindle checkpoint in mammals. Front Biosci 2008; 13: pp. 3606-3620

  164. 164. Suijkerbuijk S.J., and Kops G.J.: Preventing aneuploidy: the contribution of mitotic checkpoint proteins. Biochim Biophys Acta 2008; 1786: pp. 24-31

  165. 165. Sczaniecka M.M., and Hardwick K.G.: The spindle checkpoint: how do cells delay anaphase onset? SEB Exp Biol Ser 2008; 59: pp. 243-256

  166. 166. Decordier I., Cundari E., and Kirsch-Volders M.: Mitotic checkpoints and the maintenance of the chromosome karyotype. Mutat Res 2008; 651: pp. 3-13

  167. 167. Musacchio A., and Salmon E.D.: The spindle-assembly checkpoint in space and time. Nat Rev Mol Cell Biol 2007; 8: pp. 379-393

  168. 168. May K.M., and Hardwick K.G.: The spindle checkpoint. J Cell Sci 2006; 119: pp. 4139-4142

  169. 169. Hartwell L.H., and Weinert T.A.: Checkpoints: controls that ensure the order of cell cycle events. Science 1989; 246: pp. 629-634

  170. 170. Lukas J., Parry D., Aagaard L., Mann D.J., Bartkova J., Strauss M., et al: Retinoblastoma-protein-dependent cell-cycle inhibition by the tumour suppressor p16. Nature 1995; 375: pp. 503-506

  171. 171. Medema R.H., Herrera R.E., Lam F., and Weinberg R.A.: Growth suppression by p16ink4 requires functional retinoblastoma protein. Proc Natl Acad Sci U S A 1995; 92: pp. 6289-6293

  172. 172. Blain S.W., Montalvo E., and Massague J.: Differential interaction of the cyclin-dependent kinase (Cdk) inhibitor p27Kip1 with cyclin A-Cdk2 and cyclin D2-Cdk4. J Biol Chem 1997; 272: pp. 25863-25872

  173. 173. LaBaer J., Garrett M.D., Stevenson L.F., Slingerland J.M., Sandhu C., Chou H.S., et al: New functional activities for the p21 family of CDK inhibitors. Genes Dev 1997; 11: pp. 847-862

  174. 174. Cheng M., Olivier P., Diehl J.A., Fero M., Roussel M.F., Roberts J.M., et al: The p21(Cip1) and p27(Kip1) CDK ‘inhibitors’ are essential activators of cyclin D-dependent kinases in murine fibroblasts. EMBO J 1999; 18: pp. 1571-1583

  175. 175. Roussel M.F.: The INK4 family of cell cycle inhibitors in cancer. Oncogene 1999; 18: pp. 5311-5317

  176. 176. Schmitt E., Paquet C., Beauchemin M., and Bertrand R.: DNA-damage response network at the crossroads of cell-cycle checkpoints, cellular senescence and apoptosis. J Zhejiang Univ Sci B 2007; 8: pp. 377-397

  177. 177. Sancar A., Lindsey-Boltz L.A., Unsal-Kacmaz K., and Linn S.: Molecular mechanisms of mammalian DNA repair and the DNA damage checkpoints. Annu Rev Biochem 2004; 73: pp. 39-85

  178. 178. Motoyama N., and Naka K.: DNA damage tumor suppressor genes and genomic instability. Curr Opin Genet Dev 2004; 14: pp. 11-16

  179. 179. Bartek J., and Lukas J.: DNA damage checkpoints: from initiation to recovery or adaptation. Curr Opin Cell Biol 2007; 19: pp. 238-245

  180. 180. Ishikawa K., Ishii H., and Saito T.: DNA damage-dependent cell cycle checkpoints and genomic stability. DNA Cell Biol 2006; 25: pp. 406-411

  181. 181. Niida H., and Nakanishi M.: DNA damage checkpoints in mammals. Mutagenesis 2006; 21: pp. 3-9

  182. 182. Nojima H.: Protein kinases that regulate chromosome stability and their downstream targets. Genome Dyn 2006; 1: pp. 131-148

  183. 183. Cimprich K.A., and Cortez D.: ATR: an essential regulator of genome integrity. Nat Rev Mol Cell Biol 2008; 9: pp. 616-627

  184. 184. Lavin M.F., and Kozlov S.: ATM activation and DNA damage response. Cell Cycle 2007; 6: pp. 931-942

  185. 185. Paulsen R.D., and Cimprich K.A.: The ATR pathway: fine-tuning the fork. DNA Repair (Amst) 2007; 6: pp. 953-966

  186. 186. Hurley P.J., and Bunz F.: ATM and ATR: components of an integrated circuit. Cell Cycle 2007; 6: pp. 414-417

  187. 187. Awasthi P., Foiani M., and Kumar A.: ATM and ATR signaling at a glance. J Cell Sci 2015; 128: pp. 4255-4262

  188. 188. Marechal A., and Zou L.: DNA damage sensing by the ATM and ATR kinases. Cold Spring Harb Perspect Biol 2013; 5:

  189. 189. Falck J., Mailand N., Syljuasen R.G., Bartek J., and Lukas J.: The ATM-Chk2-Cdc25A checkpoint pathway guards against radioresistant DNA synthesis. Nature 2001; 410: pp. 842-847

  190. 190. Bartek J., and Lukas J.: Chk1 and Chk2 kinases in checkpoint control and cancer. Cancer Cell 2003; 3: pp. 421-429

  191. 191. Donzelli M., and Draetta G.F.: Regulating mammalian checkpoints through Cdc25 inactivation. EMBO Rep 2003; 4: pp. 671-677

  192. 192. Lempiainen H., and Halazonetis T.D.: Emerging common themes in regulation of PIKKs and PI3Ks. EMBO J 2009; 28: pp. 3067-3073

  193. 193. Lovejoy C.A., and Cortez D.: Common mechanisms of PIKK regulation. DNA Repair (Amst) 2009; 8: pp. 1004-1008

  194. 194. Matsuoka S., Ballif B.A., Smogorzewska A., ER M.D., Hurov K.E., Luo J., et al: ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science 2007; 316: pp. 1160-1166

  195. 195. Smolka M.B., Albuquerque C.P., Chen S.H., and Zhou H.: Proteome-wide identification of in vivo targets of DNA damage checkpoint kinases. Proc Natl Acad Sci U S A 2007; 104: pp. 10364-10369

  196. 196. Bensimon A., Schmidt A., Ziv Y., Elkon R., Wang S.Y., Chen D.J., et al: ATM-dependent and -independent dynamics of the nuclear phosphoproteome after DNA damage. Sci Signal 2010; 3: pp. rs3

  197. 197. Beli P., Lukashchuk N., Wagner S.A., Weinert B.T., Olsen J.V., Baskcomb L., et al: Proteomic investigations reveal a role for RNA processing factor THRAP3 in the DNA damage response. Mol Cell 2012; 46: pp. 212-225

  198. 198. Farnebo M., Bykov V.J., and Wiman K.G.: The p53 tumor suppressor: a master regulator of diverse cellular processes and therapeutic target in cancer. Biochem Biophys Res Commun 2010; 396: pp. 85-89

  199. 199. Boehme K.A., and Blattner C.: Regulation of p53–insights into a complex process. Crit Rev Biochem Mol Biol 2009; 44: pp. 367-392

  200. 200. Meek D.W.: Tumour suppression by p53: a role for the DNA damage response? Nat Rev Cancer 2009; 9: pp. 714-723

  201. 201. Vazquez A., Bond E.E., Levine A.J., and Bond G.L.: The genetics of the p53 pathway, apoptosis and cancer therapy. Nat Rev Drug Discov 2008; 7: pp. 979-987

  202. 202. Viadiu H.: Molecular architecture of tumor suppressor p53. Curr Top Med Chem 2008; 8: pp. 1327-1334

  203. 203. Varley J.: TP53, hChk2, and the Li-Fraumeni syndrome. Methods Mol Biol 2003; 222: pp. 117-129

  204. 204. Schwartz D., and Rotter V.: p53-dependent cell cycle control: response to genotoxic stress. Semin Cancer Biol 1998; 8: pp. 325-336

  205. 205. Shieh S.Y., Ikeda M., Taya Y., and Prives C.: DNA damage-induced phosphorylation of p53 alleviates inhibition by MDM2. Cell 1997; 91: pp. 325-334

  206. 206. Agami R., and Bernards R.: Distinct initiation and maintenance mechanisms cooperate to induce G1 cell cycle arrest in response to DNA damage. Cell 2000; 102: pp. 55-66

  207. 207. Xu B., Kim S.T., Lim D.S., and Kastan M.B.: Two molecularly distinct G(2)/M checkpoints are induced by ionizing irradiation. Mol Cell Biol 2002; 22: pp. 1049-1059

  208. 208. Nyberg K.A., Michelson R.J., Putnam C.W., and Weinert T.A.: Toward maintaining the genome: DNA damage and replication checkpoints. Annu Rev Genet 2002; 36: pp. 617-656

  209. 209. Mailand N., Podtelejnikov A.V., Groth A., Mann M., Bartek J., and Lukas J.: Regulation of G(2)/M events by Cdc25A through phosphorylation-dependent modulation of its stability. EMBO J 2002; 21: pp. 5911-5920

  210. 210. Taylor W.R., and Stark G.R.: Regulation of the G2/M transition by p53. Oncogene 2001; 20: pp. 1803-1815

  211. 211. Grallert B., and Boye E.: The multiple facets of the intra-S checkpoint. Cell Cycle 2008; 7: pp. 2315-2320

  212. 212. Bartek J., Lukas C., and Lukas J.: Checking on DNA damage in S phase. Nat Rev Mol Cell Biol 2004; 5: pp. 792-804

  213. 213. Canman C.E.: Replication checkpoint: preventing mitotic catastrophe. Curr Biol 2001; 11: pp. R121-R124

  214. 214. Friedel A.M., Pike B.L., and Gasser S.M.: ATR/Mec1: coordinating fork stability and repair. Curr Opin Cell Biol 2009; 21: pp. 237-244

  215. 215. Segurado M., and Tercero J.A.: The S-phase checkpoint: targeting the replication fork. Biol Cell 2009; 101: pp. 617-627

  216. 216. Bartek J., and Lukas J.: Mammalian G1- and S-phase checkpoints in response to DNA damage. Curr Opin Cell Biol 2001; 13: pp. 738-747

  217. 217. Shiloh Y.: ATM and related protein kinases: safeguarding genome integrity. Nat Rev Cancer 2003; 3: pp. 155-168

  218. 218. Bakkenist C.J., and Kastan M.B.: DNA damage activates ATM through intermolecular autophosphorylation and dimer dissociation. Nature 2003; 421: pp. 499-506

  219. 219. Zou L., and Elledge S.J.: Sensing DNA damage through ATRIP recognition of RPA-ssDNA complexes. Science 2003; 300: pp. 1542-1548

  220. 220. Helt C.E., Cliby W.A., Keng P.C., Bambara R.A., and O’Reilly M.A.: Ataxia telangiectasia mutated (ATM) and ATM and Rad3-related protein exhibit selective target specificities in response to different forms of DNA damage. J Biol Chem 2005; 280: pp. 1186-1192

  221. 221. Heffernan T.P., Simpson D.A., Frank A.R., Heinloth A.N., Paules R.S., Cordeiro-Stone M., et al: An ATR- and Chk1-dependent S checkpoint inhibits replicon initiation following UVC-induced DNA damage. Mol Cell Biol 2002; 22: pp. 8552-8561

  222. 222. Gatei M., Sloper K., Sorensen C., Syljuasen R., Falck J., Hobson K., et al: Ataxia-telangiectasia-mutated (ATM) and NBS1-dependent phosphorylation of Chk1 on Ser-317 in response to ionizing radiation. J Biol Chem 2003; 278: pp. 14806-14811

  223. 223. Wang X.Q., Redpath J.L., Fan S.T., and Stanbridge E.J.: ATR dependent activation of Chk2. J Cell Physiol 2006; 208: pp. 613-619

  224. 224. Broderick R., and Nasheuer H.P.: Regulation of Cdc45 in the cell cycle and after DNA damage. Biochem Soc Trans 2009; 37: pp. 926-930

  225. 225. Kitagawa R., Bakkenist C.J., McKinnon P.J., and Kastan M.B.: Phosphorylation of SMC1 is a critical downstream event in the ATM-NBS1-BRCA1 pathway. Genes Dev 2004; 18: pp. 1423-1438

  226. 226. Kim S.T., Xu B., and Kastan M.B.: Involvement of the cohesin protein, Smc1, in Atm-dependent and independent responses to DNA damage. Genes Dev 2002; 16: pp. 560-570

  227. 227. Yazdi P.T., Wang Y., Zhao S., Patel N., Lee E.Y., and Qin J.: SMC1 is a downstream effector in the ATM/NBS1 branch of the human S-phase checkpoint. Genes Dev 2002; 16: pp. 571-582

  228. 228. Kops G.J., Weaver B.A., and Cleveland D.W.: On the road to cancer: aneuploidy and the mitotic checkpoint. Nat Rev Cancer 2005; 5: pp. 773-785

  229. 229. Rieder C.L., and Maiato H.: Stuck in division or passing through: what happens when cells cannot satisfy the spindle assembly checkpoint. Dev Cell 2004; 7: pp. 637-651

  230. 230. Malmanche N., Maia A., and Sunkel C.E.: The spindle assembly checkpoint: preventing chromosome mis-segregation during mitosis and meiosis. FEBS Lett 2006; 580: pp. 2888-2895

  231. 231. Shah J.V., and Cleveland D.W.: Waiting for anaphase: Mad2 and the spindle assembly checkpoint. Cell 2000; 103: pp. 997-1000

  232. 232. Sacristan C., and Kops G.J.: Joined at the hip: kinetochores, microtubules, and spindle assembly checkpoint signaling. Trends Cell Biol 2015; 25: pp. 21-28

  233. 233. Wang Y., Jin F., Higgins R., and McKnight K.: The current view for the silencing of the spindle assembly checkpoint. Cell Cycle 2014; 13: pp. 1694-1701

  234. 234. Hoyt M.A., Totis L., and Roberts B.T.: S. cerevisiae genes required for cell cycle arrest in response to loss of microtubule function. Cell 1991; 66: pp. 507-517

  235. 235. Li R., and Murray A.W.: Feedback control of mitosis in budding yeast. Cell 1991; 66: pp. 519-531

  236. 236. Musacchio A., and Hardwick K.G.: The spindle checkpoint: structural insights into dynamic signalling. Nat Rev Mol Cell Biol 2002; 3: pp. 731-741

  237. 237. Taylor S.S., Scott M.I., and Holland A.J.: The spindle checkpoint: a quality control mechanism which ensures accurate chromosome segregation. Chromosom Res 2004; 12: pp. 599-616

  238. 238. London N., and Biggins S.: Signalling dynamics in the spindle checkpoint response. Nat Rev Mol Cell Biol 2014; 15: pp. 736-747

  239. 239. Joglekar A.P.: A cell biological perspective on past, present and future investigations of the spindle assembly checkpoint. Biology 2016; 5:

  240. 240. Lischetti T., and Nilsson J.: Regulation of mitotic progression by the spindle assembly checkpoint. Mol Cell Oncol 2015; 2:

  241. 241. Kim S.H., Lin D.P., Matsumoto S., Kitazono A., and Matsumoto T.: Fission yeast Slp1: an effector of the Mad2-dependent spindle checkpoint. Science 1998; 279: pp. 1045-1047

  242. 242. Hwang L.H., Lau L.F., Smith D.L., Mistrot C.A., Hardwick K.G., Hwang E.S., et al: Budding yeast Cdc20: a target of the spindle checkpoint. Science 1998; 279: pp. 1041-1044

  243. 243. Simpson-Lavy K.J., Oren Y.S., Feine O., Sajman J., Listovsky T., and Brandeis M.: Fifteen years of APC/cyclosome: a short and impressive biography. Biochem Soc Trans 2010; 38: pp. 78-82

  244. 244. Wasch R., Robbins J.A., and Cross F.R.: The emerging role of APC/CCdh1 in controlling differentiation, genomic stability and tumor suppression. Oncogene 2010; 29: pp. 1-10

  245. 245. Pesin J.A., and Orr-Weaver T.L.: Regulation of APC/C activators in mitosis and meiosis. Annu Rev Cell Dev Biol 2008; 24: pp. 475-499

  246. 246. Lindon C.: Control of mitotic exit and cytokinesis by the APC/C. Biochem Soc Trans 2008; 36: pp. 405-410

  247. 247. Baker D.J., Dawlaty M.M., Galardy P., and van Deursen J.M.: Mitotic regulation of the anaphase-promoting complex. Cell Mol Life Sci 2007; 64: pp. 589-600

  248. 248. Peters J.M.: The anaphase promoting complex/cyclosome: a machine designed to destroy. Nat Rev Mol Cell Biol 2006; 7: pp. 644-656

  249. 249. Acquaviva C., and Pines J.: The anaphase-promoting complex/cyclosome: APC/C. J Cell Sci 2006; 119: pp. 2401-2404

  250. 250. Vanoosthuyse V., and Hardwick K.G.: Bub1 and the multilayered inhibition of Cdc20-APC/C in mitosis. Trends Cell Biol 2005; 15: pp. 231-233

  251. 251. Vodermaier H.C.: APC/C and SCF: controlling each other and the cell cycle. Curr Biol 2004; 14: pp. R787-R796

  252. 252. Meadows J.C., and Millar J.B.: Sharpening the anaphase switch. Biochem Soc Trans 2015; 43: pp. 19-22

  253. 253. Sivakumar S., and Gorbsky G.J.: Spatiotemporal regulation of the anaphase-promoting complex in mitosis. Nat Rev Mol Cell Biol 2015; 16: pp. 82-94

  254. 254. Hershko A.: Mechanisms and regulation of the degradation of cyclin B. Philos Trans R Soc Lond Ser B Biol Sci 1999; 354: pp. 1571-1575

  255. 255. Wolf F., Sigl R., and Geley S.: ‘…The end of the beginning’: cdk1 thresholds and exit from mitosis. Cell Cycle 2007; 6: pp. 1408-1411

  256. 256. Irniger S.: Cyclin destruction in mitosis: a crucial task of Cdc20. FEBS Lett 2002; 532: pp. 7-11

  257. 257. John P.C., Mews M., and Moore R.: Cyclin/Cdk complexes: their involvement in cell cycle progression and mitotic division. Protoplasma 2001; 216: pp. 119-142

  258. 258. Stemmann O., Gorr I.H., and Boos D.: Anaphase topsy-turvy: Cdk1 a securin, separase a CKI. Cell Cycle 2006; 5: pp. 11-13

  259. 259. Yanagida M.: Cell cycle mechanisms of sister chromatid separation; roles of Cut1/separin and Cut2/securin. Genes Cells 2000; 5: pp. 1-8

  260. 260. Nagao K., and Yanagida M.: Regulating sister chromatid separation by separase phosphorylation. Dev Cell 2002; 2: pp. 2-4

  261. 261. Uhlmann F.: Chromosome cohesion and separation: from men and molecules. Curr Biol 2003; 13: pp. R104-R114

  262. 262. Pellman D., and Christman M.F.: Separase anxiety: dissolving the sister bond and more. Nat Cell Biol 2001; 3: pp. E207-E209

  263. 263. Agarwal R., and Cohen-Fix O.: Mitotic regulation: the fine tuning of separase activity. Cell Cycle 2002; 1: pp. 255-257

  264. 264. Uhlmann F.: Separase regulation during mitosis. Biochem Soc Symp 2003; 70: pp. 243-251

  265. 265. Hydbring P., Malumbres M., and Sicinski P.: Non-canonical functions of cell cycle cyclins and cyclin-dependent kinases. Nat Rev Mol Cell Biol 2016; 17: pp. 280-292

  266. 266. Lim S., and Kaldis P.: Cdks, cyclins and CKIs: roles beyond cell cycle regulation. Development 2013; 140: pp. 3079-3093

  267. 267. Fujimoto T., Anderson K., Jacobsen S.E., Nishikawa S.I., and Nerlov C.: Cdk6 blocks myeloid differentiation by interfering with Runx1 DNA binding and Runx1-C/EBPalpha interaction. EMBO J 2007; 26: pp. 2361-2370

  268. 268. Kollmann K., Heller G., Schneckenleithner C., Warsch W., Scheicher R., Ott R.G., et al: A kinase-independent function of CDK6 links the cell cycle to tumor angiogenesis. Cancer Cell 2013; 24: pp. 167-181

  269. 269. Le Guen T., Ragu S., Guirouilh-Barbat J., and Lopez B.S.: Role of the double-strand break repair pathway in the maintenance of genomic stability. Mol Cell Oncol 2015; 2:

  270. 270. Jirawatnotai S., Hu Y., Michowski W., Elias J.E., Becks L., Bienvenu F., et al: A function for cyclin D1 in DNA repair uncovered by protein interactome analyses in human cancers. Nature 2011; 474: pp. 230-234

  271. 271. Chen L., Nievera C.J., Lee A.Y., and Wu X.: Cell cycle-dependent complex formation of BRCA1.CtIP.MRN is important for DNA double-strand break repair. J Biol Chem 2008; 283: pp. 7713-7720

  272. 272. Hustedt N., and Durocher D.: The control of DNA repair by the cell cycle. Nat Cell Biol 2016; 19: pp. 1-9

  273. 273. Starostina N.G., and Kipreos E.T.: Multiple degradation pathways regulate versatile CIP/KIP CDK inhibitors. Trends Cell Biol 2012; 22: pp. 33-41

  274. 274. Abbas T., and Dutta A.: p21 in cancer: intricate networks and multiple activities. Nat Rev Cancer 2009; 9: pp. 400-414

  275. 275. Huang S., Shu L., Dilling M.B., Easton J., Harwood F.C., Ichijo H., et al: Sustained activation of the JNK cascade and rapamycin-induced apoptosis are suppressed by p53/p21(Cip1). Mol Cell 2003; 11: pp. 1491-1501

  276. 276. Dotto G.P.: p21(WAF1/Cip1): more than a break to the cell cycle? Biochim Biophys Acta 2000; 1471: pp. M43-M56

  277. 277. Narumiya S., Tanji M., and Ishizaki T.: Rho signaling, ROCK and mDia1, in transformation, metastasis and invasion. Cancer Metastasis Rev 2009; 28: pp. 65-76

  278. 278. Bernstein B.W., and Bamburg J.R.: ADF/cofilin: a functional node in cell biology. Trends Cell Biol 2010; 20: pp. 187-195

  279. 279. Wood L.D., Parsons D.W., Jones S., Lin J., Sjoblom T., Leary R.J., et al: The genomic landscapes of human breast and colorectal cancers. Science 2007; 318: pp. 1108-1113

  280. 280. Sjoblom T., Jones S., Wood L.D., Parsons D.W., Lin J., Barber T.D., et al: The consensus coding sequences of human breast and colorectal cancers. Science 2006; 314: pp. 268-274

  281. 281. Vogelstein B., Papadopoulos N., Velculescu V.E., Zhou S., Diaz L.A., and Kinzler K.W.: Cancer genome landscapes. Science 2013; 339: pp. 1546-1558

  282. 282. Copeland N.G., and Jenkins N.A.: Deciphering the genetic landscape of cancer—from genes to pathways. Trends Genet 2009; 25: pp. 455-462

  283. 283. Schweiger M.R., Hussong M., Rohr C., and Lehrach H.: Genomics and epigenomics of colorectal cancer. Wiley Interdiscip Rev Syst Biol Med 2013; 5: pp. 205-219

  284. 284. Beijersbergen R.L., and Bernards R.: Cell cycle regulation by the retinoblastoma family of growth inhibitory proteins. Biochim Biophys Acta 1996; 1287: pp. 103-120

  285. 285. Kaelin W.G.: Recent insights into the functions of the retinoblastoma susceptibility gene product. Cancer Investig 1997; 15: pp. 243-254

  286. 286. Sellers W.R., and Kaelin W.G.: Role of the retinoblastoma protein in the pathogenesis of human cancer. J Clin Oncol 1997; 15: pp. 3301-3312

  287. 287. Nevins J.R.: The Rb/E2F pathway and cancer. Hum Mol Genet 2001; 10: pp. 699-703

  288. 288. Hahn W.C., and Weinberg R.A.: Modelling the molecular circuitry of cancer. Nat Rev Cancer 2002; 2: pp. 331-341

  289. 289. Ortega S., Malumbres M., and Barbacid M.: Cyclin D-dependent kinases, INK4 inhibitors and cancer. Biochim Biophys Acta 2002; 1602: pp. 73-87

  290. 290. Sherr C.J., and McCormick F.: The RB and p53 pathways in cancer. Cancer Cell 2002; 2: pp. 103-112

  291. 291. Sherr C.J.: The INK4a/ARF network in tumour suppression. Nat Rev Mol Cell Biol 2001; 2: pp. 731-737

  292. 292. Wikenheiser-Brokamp K.A.: Retinoblastoma family proteins: insights gained through genetic manipulation of mice. Cell Mol Life Sci 2006; 63: pp. 767-780

  293. 293. Dyson N.J.: RB1: a prototype tumor suppressor and an enigma. Genes Dev 2016; 30: pp. 1492-1502

  294. 294. Indovina P., Pentimalli F., Casini N., Vocca I., and Giordano A.: RB1 dual role in proliferation and apoptosis: cell fate control and implications for cancer therapy. Oncotarget 2015; 6: pp. 17873-17890

  295. 295. Dick F.A., and Rubin S.M.: Molecular mechanisms underlying RB protein function. Nat Rev Mol Cell Biol 2013; 14: pp. 297-306

  296. 296. Kamb A., Shattuck-Eidens D., Eeles R., Liu Q., Gruis N.A., Ding W., et al: Analysis of the p16 gene (CDKN2) as a candidate for the chromosome 9p melanoma susceptibility locus. Nat Genet 1994; 8: pp. 23-26

  297. 297. Ruas M., and Peters G.: The p16INK4a/CDKN2A tumor suppressor and its relatives. Biochim Biophys Acta 1998; 1378: pp. F115-F177

  298. 298. Li J., Poi M.J., and Tsai M.D.: Regulatory mechanisms of tumor suppressor P16(INK4A) and their relevance to cancer. Biochemistry 2011; 50: pp. 5566-5582

  299. 299. LaPak K.M., and Burd C.E.: The molecular balancing act of p16(INK4a) in cancer and aging. Mol Cancer Res 2014; 12: pp. 167-183

  300. 300. Johnson J., Thijssen B., McDermott U., Garnett M., Wessels L.F., and Bernards R.: Targeting the RB-E2F pathway in breast cancer. Oncogene 2016; 35: pp. 4829-4835

  301. 301. Hutcheson J., Witkiewicz A.K., and Knudsen E.S.: The RB tumor suppressor at the intersection of proliferation and immunity: relevance to disease immune evasion and immunotherapy. Cell Cycle 2015; 14: pp. 3812-3819

  302. 302. Radulescu R.T.: From the RB tumor suppressor to MCR peptides. Protein Pept Lett 2014; 21: pp. 589-592

  303. 303. Levine A.J.: p53, the cellular gatekeeper for growth and division. Cell 1997; 88: pp. 323-331

  304. 304. Hollstein M., Rice K., Greenblatt M.S., Soussi T., Fuchs R., Sorlie T., et al: Database of p53 gene somatic mutations in human tumors and cell lines. Nucleic Acids Res 1994; 22: pp. 3551-3555

  305. 305. Giaccia A.J., and Kastan M.B.: The complexity of p53 modulation: emerging patterns from divergent signals. Genes Dev 1998; 12: pp. 2973-2983

  306. 306. Gottifredi V., Shieh S.Y., and Prives C.: Regulation of p53 after different forms of stress and at different cell cycle stages. Cold Spring Harb Symp Quant Biol 2000; 65: pp. 483-488

  307. 307. Helton E.S., and Chen X.: p53 modulation of the DNA damage response. J Cell Biochem 2007; 100: pp. 883-896

  308. 308. Williams A.B., and Schumacher B.: p53 in the DNA-damage-repair process. Cold Spring Harb Perspect Med 2016; 6:

  309. 309. Speidel D.: The role of DNA damage responses in p53 biology. Arch Toxicol 2015; 89: pp. 501-517

  310. 310. Jackson S.P., and Bartek J.: The DNA-damage response in human biology and disease. Nature 2009; 461: pp. 1071-1078

  311. 311. Bieging K.T., Mello S.S., and Attardi L.D.: Unravelling mechanisms of p53-mediated tumour suppression. Nat Rev Cancer 2014; 14: pp. 359-370

  312. 312. Sionov R.V., and Haupt Y.: The cellular response to p53: the decision between life and death. Oncogene 1999; 18: pp. 6145-6157

  313. 313. Bates S., and Vousden K.H.: Mechanisms of p53-mediated apoptosis. Cell Mol Life Sci 1999; 55: pp. 28-37

  314. 314. Corn P.G., and El-Deiry W.S.: Microarray analysis of p53-dependent gene expression in response to hypoxia and DNA damage. Cancer Biol Ther 2007; 6: pp. 1858-1866

  315. 315. Riley T., Sontag E., Chen P., and Levine A.: Transcriptional control of human p53-regulated genes. Nat Rev Mol Cell Biol 2008; 9: pp. 402-412

  316. 316. Horvath M.M., Wang X., Resnick M.A., and Bell D.A.: Divergent evolution of human p53 binding sites: cell cycle versus apoptosis. PLoS Genet 2007; 3:

  317. 317. Jen K.Y., and Cheung V.G.: Identification of novel p53 target genes in ionizing radiation response. Cancer Res 2005; 65: pp. 7666-7673

  318. 318. Mirza A., Wu Q., Wang L., McClanahan T., Bishop W.R., Gheyas F., et al: Global transcriptional program of p53 target genes during the process of apoptosis and cell cycle progression. Oncogene 2003; 22: pp. 3645-3654

  319. 319. Rashi-Elkeles S., Elkon R., Shavit S., Lerenthal Y., Linhart C., Kupershtein A., et al: Transcriptional modulation induced by ionizing radiation: p53 remains a central player. Mol Oncol 2011; 5: pp. 336-348

  320. 320. Beckerman R., and Prives C.: Transcriptional regulation by p53. Cold Spring Harb Perspect Biol 2010; 2:

  321. 321. McCubrey J.A., Lertpiriyapong K., Fitzgerald T.L., Martelli A.M., Cocco L., Rakus D., et al: Roles of TP53 in determining therapeutic sensitivity, growth, cellular senescence, invasion and metastasis. Adv Biol Regul 2017; 63: pp. 32-48

  322. 322. Chipuk J.E., Kuwana T., Bouchier-Hayes L., Droin N.M., Newmeyer D.D., Schuler M., et al: Direct activation of Bax by p53 mediates mitochondrial membrane permeabilization and apoptosis. Science 2004; 303: pp. 1010-1014

  323. 323. Oltvai Z.N., Milliman C.L., and Korsmeyer S.J.: Bcl-2 heterodimerizes in vivo with a conserved homolog, Bax, that accelerates programmed cell death. Cell 1993; 74: pp. 609-619

  324. 324. Luna-Vargas M.P., and Chipuk J.E.: Physiological and Pharmacological Control of BAK, BAX, and Beyond. Trends Cell Biol 2016; 26: pp. 906-917

  325. 325. Renault T.T., Teijido O., Antonsson B., Dejean L.M., and Manon S.: Regulation of Bax mitochondrial localization by Bcl-2 and Bcl-x(L): keep your friends close but your enemies closer. Int J Biochem Cell Biol 2013; 45: pp. 64-67

  326. 326. Juven-Gershon T., and Oren M.: Mdm2: the ups and downs. Mol Med 1999; 5: pp. 71-83

  327. 327. Zhang Y., and Xiong Y.: Control of p53 ubiquitination and nuclear export by MDM2 and ARF. Cell Growth Differ 2001; 12: pp. 175-186

  328. 328. Michael D., and Oren M.: The p53-Mdm2 module and the ubiquitin system. Semin Cancer Biol 2003; 13: pp. 49-58

  329. 329. Prives C.: Signaling to p53: breaking the MDM2-p53 circuit. Cell 1998; 95: pp. 5-8

  330. 330. Piette J., Neel H., and Marechal V.: Mdm2: keeping p53 under control. Oncogene 1997; 15: pp. 1001-1010

  331. 331. Lowe S.W., and Sherr C.J.: Tumor suppression by Ink4a-Arf: progress and puzzles. Curr Opin Genet Dev 2003; 13: pp. 77-83

  332. 332. James M.C., and Peters G.: Alternative product of the p16/CKDN2A locus connects the Rb and p53 tumor suppressors. Prog Cell Cycle Res 2000; 4: pp. 71-81

  333. 333. Serrano M.: The INK4a/ARF locus in murine tumorigenesis. Carcinogenesis 2000; 21: pp. 865-869

  334. 334. Sharpless N.E.: INK4a/ARF: a multifunctional tumor suppressor locus. Mutat Res 2005; 576: pp. 22-38

  335. 335. Ko A., Han S.Y., and Song J.: Dynamics of ARF regulation that control senescence and cancer. BMB Rep 2016; 49: pp. 598-606

  336. 336. Zhao Y., Yu H., and Hu W.: The regulation of MDM2 oncogene and its impact on human cancers. Acta Biochim Biophys Sin Shanghai 2014; 46: pp. 180-189

  337. 337. Bueso-Ramos C.E., Yang Y., deLeon E., McCown P., Stass S.A., and Albitar M.: The human MDM-2 oncogene is overexpressed in leukemias. Blood 1993; 82: pp. 2617-2623

  338. 338. Oliner J.D., Kinzler K.W., Meltzer P.S., George D.L., and Vogelstein B.: Amplification of a gene encoding a p53-associated protein in human sarcomas. Nature 1992; 358: pp. 80-83

  339. 339. Borg A., Johannsson U., Johannsson O., Hakansson S., Westerdahl J., Masback A., et al: Novel germline p16 mutation in familial malignant melanoma in southern Sweden. Cancer Res 1996; 56: pp. 2497-2500

  340. 340. Cairns P., Mao L., Merlo A., Lee D.J., Schwab D., Eby Y., et al: Rates of p16 (MTS1) mutations in primary tumors with 9p loss. Science 1994; 265: pp. 415-417

  341. 341. Caldas C., Hahn S.A., da Costa L.T., Redston M.S., Schutte M., Seymour A.B., et al: Frequent somatic mutations and homozygous deletions of the p16 (MTS1) gene in pancreatic adenocarcinoma. Nat Genet 1994; 8: pp. 27-32

  342. 342. Meng X., Franklin D.A., Dong J., and Zhang Y.: MDM2-p53 pathway in hepatocellular carcinoma. Cancer Res 2014; 74: pp. 7161-7167

  343. 343. Quelle D.E., Zindy F., Ashmun R.A., and Sherr C.J.: Alternative reading frames of the INK4a tumor suppressor gene encode two unrelated proteins capable of inducing cell cycle arrest. Cell 1995; 83: pp. 993-1000

  344. 344. Bashir T., and Pagano M.: Aberrant ubiquitin-mediated proteolysis of cell cycle regulatory proteins and oncogenesis. Adv Cancer Res 2003; 88: pp. 101-144

  345. 345. Devoy A., Soane T., Welchman R., and Mayer R.J.: The ubiquitin-proteasome system and cancer. Essays Biochem 2005; 41: pp. 187-203

  346. 346. Mani A., and Gelmann E.P.: The ubiquitin-proteasome pathway and its role in cancer. J Clin Oncol 2005; 23: pp. 4776-4789

  347. 347. Reed S.I.: The ubiquitin-proteasome pathway in cell cycle control. Results Probl Cell Differ 2006; 42: pp. 147-181

  348. 348. Yamasaki L., and Pagano M.: Cell cycle, proteolysis and cancer. Curr Opin Cell Biol 2004; 16: pp. 623-628

  349. 349. Cenciarelli C., Chiaur D.S., Guardavaccaro D., Parks W., Vidal M., and Pagano M.: Identification of a family of human F-box proteins. Curr Biol 1999; 9: pp. 1177-1179

  350. 350. Kipreos E.T., and Pagano M.: The F-box protein family. Genome Biol 2000; 1:

  351. 351. Skaar J.R., and Pagano M.: Control of cell growth by the SCF and APC/C ubiquitin ligases. Curr Opin Cell Biol 2009; 21: pp. 816-824

  352. 352. Nakayama K.I., and Nakayama K.: Regulation of the cell cycle by SCF-type ubiquitin ligases. Semin Cell Dev Biol 2005; 16: pp. 323-333

  353. 353. Peters J.M.: SCF and APC: the Yin and Yang of cell cycle regulated proteolysis. Curr Opin Cell Biol 1998; 10: pp. 759-768

  354. 354. Cardozo T., and Pagano M.: The SCF ubiquitin ligase: insights into a molecular machine. Nat Rev Mol Cell Biol 2004; 5: pp. 739-751

  355. 355. Nishitani H., Sugimoto N., Roukos V., Nakanishi Y., Saijo M., Obuse C., et al: Two E3 ubiquitin ligases, SCF-Skp2 and DDB1-Cul4, target human Cdt1 for proteolysis. EMBO J 2006; 25: pp. 1126-1136

  356. 356. Cao J., Ge M.H., and Ling Z.Q.: Fbxw7 tumor suppressor: a vital regulator contributes to human tumorigenesis. Medicine (Baltimore) 2016; 95:

  357. 357. Xu W., Taranets L., and Popov N.: Regulating Fbw7 on the road to cancer. Semin Cancer Biol 2016; 36: pp. 62-70

  358. 358. Wang H., and Maitra A.: The emerging roles of F-box proteins in pancreatic tumorigenesis. Semin Cancer Biol 2016; 36: pp. 88-94

  359. 359. Heestand G.M., and Kurzrock R.: Molecular landscape of pancreatic cancer: implications for current clinical trials. Oncotarget 2015; 6: pp. 4553-4561

  360. 360. Borras E., San Lucas F.A., Chang K., Zhou R., Masand G., Fowler J., et al: Genomic landscape of colorectal mucosa and adenomas. Cancer Prev Res (Phila) 2016; 9: pp. 417-427

  361. 361. Zhang W., and Koepp D.M.: Fbw7 isoform interaction contributes to cyclin E proteolysis. Mol Cancer Res 2006; 4: pp. 935-943

  362. 362. Nash P., Tang X., Orlicky S., Chen Q., Gertler F.B., Mendenhall M.D., et al: Multisite phosphorylation of a CDK inhibitor sets a threshold for the onset of DNA replication. Nature 2001; 414: pp. 514-521

  363. 363. Tan Y., Sangfelt O., and Spruck C.: The Fbxw7/hCdc4 tumor suppressor in human cancer. Cancer Lett 2008; 271: pp. 1-12

  364. 364. Fujii Y., Yada M., Nishiyama M., Kamura T., Takahashi H., Tsunematsu R., et al: Fbxw7 contributes to tumor suppression by targeting multiple proteins for ubiquitin-dependent degradation. Cancer Sci 2006; 97: pp. 729-736

  365. 365. Fuchs S.Y.: Tumor suppressor activities of the Fbw7 E3 ubiquitin ligase receptor. Cancer Biol Ther 2005; 4: pp. 506-508

  366. 366. Minella A.C., and Clurman B.E.: Mechanisms of tumor suppression by the SCF(Fbw7). Cell Cycle 2005; 4: pp. 1356-1359

  367. 367. Wang Z., Inuzuka H., Zhong J., Wan L., Fukushima H., Sarkar F.H., et al: Tumor suppressor functions of FBW7 in cancer development and progression. FEBS Lett 2012; 586: pp. 1409-1418

  368. 368. Akhoondi S., Sun D., von der Lehr N., Apostolidou S., Klotz K., Maljukova A., et al: FBXW7/hCDC4 is a general tumor suppressor in human cancer. Cancer Res 2007; 67: pp. 9006-9012

  369. 369. Wang L., Ye X., Liu Y., Wei W., and Wang Z.: Aberrant regulation of FBW7 in cancer. Oncotarget 2014; 5: pp. 2000-2015

  370. 370. Lau A.W., Fukushima H., and Wei W.: The Fbw7 and betaTRCP E3 ubiquitin ligases and their roles in tumorigenesis. Front Biosci 2012; 17: pp. 2197-2212

  371. 371. Furstenthal L., Kaiser B.K., Swanson C., and Jackson P.K.: Cyclin E uses Cdc6 as a chromatin-associated receptor required for DNA replication. J Cell Biol 2001; 152: pp. 1267-1278

  372. 372. Ekholm-Reed S., Mendez J., Tedesco D., Zetterberg A., Stillman B., and Reed S.I.: Deregulation of cyclin E in human cells interferes with prereplication complex assembly. J Cell Biol 2004; 165: pp. 789-800

  373. 373. Rajagopalan H., and Lengauer C.: hCDC4 and genetic instability in cancer. Cell Cycle 2004; 3: pp. 693-694

  374. 374. Rajagopalan H., Jallepalli P.V., Rago C., Velculescu V.E., Kinzler K.W., Vogelstein B., et al: Inactivation of hCDC4 can cause chromosomal instability. Nature 2004; 428: pp. 77-81

  375. 375. Spruck C.H., Won K.A., and Reed S.I.: Deregulated cyclin E induces chromosome instability. Nature 1999; 401: pp. 297-300

  376. 376. Hanashiro K., Kanai M., Geng Y., Sicinski P., and Fukasawa K.: Roles of cyclins A and E in induction of centrosome amplification in p53-compromised cells. Oncogene 2008; 27: pp. 5288-5302

  377. 377. Mussman J.G., Horn H.F., Carroll P.E., Okuda M., Tarapore P., Donehower L.A., et al: Synergistic induction of centrosome hyperamplification by loss of p53 and cyclin E overexpression. Oncogene 2000; 19: pp. 1635-1646

  378. 378. Kawamura K., Izumi H., Ma Z., Ikeda R., Moriyama M., Tanaka T., et al: Induction of centrosome amplification and chromosome instability in human bladder cancer cells by p53 mutation and cyclin E overexpression. Cancer Res 2004; 64: pp. 4800-4809

  379. 379. Simone C., Resta N., Bagella L., Giordano A., and Guanti G.: Cyclin E and chromosome instability in colorectal cancer cell lines. Mol Pathol 2002; 55: pp. 200-203

  380. 380. Koepp D.M., Schaefer L.K., Ye X., Keyomarsi K., Chu C., Harper J.W., et al: Phosphorylation-dependent ubiquitination of cyclin E by the SCFFbw7 ubiquitin ligase. Science 2001; 294: pp. 173-177

  381. 381. Hao B., Oehlmann S., Sowa M.E., Harper J.W., and Pavletich N.P.: Structure of a Fbw7-Skp1-cyclin E complex: multisite-phosphorylated substrate recognition by SCF ubiquitin ligases. Mol Cell 2007; 26: pp. 131-143

  382. 382. Strohmaier H., Spruck C.H., Kaiser P., Won K.A., Sangfelt O., and Reed S.I.: Human F-box protein hCdc4 targets cyclin E for proteolysis and is mutated in a breast cancer cell line. Nature 2001; 413: pp. 316-322

  383. 383. Lau A.W., Inuzuka H., Fukushima H., Wan L., Liu P., Gao D., et al: Regulation of APC(Cdh1) E3 ligase activity by the Fbw7/cyclin E signaling axis contributes to the tumor suppressor function of Fbw7. Cell Res 2013; 23: pp. 947-961

  384. 384. Bhaskaran N., van Drogen F., Ng H.F., Kumar R., Ekholm-Reed S., Peter M., et al: Fbw7alpha and Fbw7gamma collaborate to shuttle cyclin E1 into the nucleolus for multiubiquitylation. Mol Cell Biol 2013; 33: pp. 85-97

  385. 385. Grady W.M., and Carethers J.M.: Genomic and epigenetic instability in colorectal cancer pathogenesis. Gastroenterology 2008; 135: pp. 1079-1099

  386. 386. Lengauer C., Kinzler K.W., and Vogelstein B.: Genetic instability in colorectal cancers. Nature 1997; 386: pp. 623-627

  387. 387. Lengauer C., Kinzler K.W., and Vogelstein B.: Genetic instabilities in human cancers. Nature 1998; 396: pp. 643-649

  388. 388. Gelsomino F., Barbolini M., Spallanzani A., Pugliese G., and Cascinu S.: The evolving role of microsatellite instability in colorectal cancer: a review. Cancer Treat Rev 2016; 51: pp. 19-26

  389. 389. Yamamoto H., and Imai K.: Microsatellite instability: an update. Arch Toxicol 2015; 89: pp. 899-921

  390. 390. Rajagopalan H., and Lengauer C.: CIN-ful cancers. Cancer Chemother Pharmacol 2004; 54: pp. S65-S68

  391. 391. Michor F., Iwasa Y., Vogelstein B., Lengauer C., and Nowak M.A.: Can chromosomal instability initiate tumorigenesis? Semin Cancer Biol 2005; 15: pp. 43-49

  392. 392. Rajagopalan H., and Lengauer C.: Aneuploidy and cancer. Nature 2004; 432: pp. 338-341

  393. 393. Rajagopalan H., Nowak M.A., Vogelstein B., and Lengauer C.: The significance of unstable chromosomes in colorectal cancer. Nat Rev Cancer 2003; 3: pp. 695-701

  394. 394. Shih I.M., Zhou W., Goodman S.N., Lengauer C., Kinzler K.W., and Vogelstein B.: Evidence that genetic instability occurs at an early stage of colorectal tumorigenesis. Cancer Res 2001; 61: pp. 818-822

  395. 395. Sansregret L., and Swanton C.: The role of aneuploidy in cancer evolution. Cold Spring Harb Perspect Med 2017; 7:

  396. 396. Nicholson J.M., and Cimini D.: Link between aneuploidy and chromosome instability. Int Rev Cell Mol Biol 2015; 315: pp. 299-317

  397. 397. Potapova T.A., Zhu J., and Li R.: Aneuploidy and chromosomal instability: a vicious cycle driving cellular evolution and cancer genome chaos. Cancer Metastasis Rev 2013; 32: pp. 377-389

  398. 398. Thompson S.L., Bakhoum S.F., and Compton D.A.: Mechanisms of chromosomal instability. Curr Biol 2010; 20: pp. R285-R295

  399. 399. Lingle W.L., Lukasiewicz K., and Salisbury J.L.: Deregulation of the centrosome cycle and the origin of chromosomal instability in cancer. Adv Exp Med Biol 2005; 570: pp. 393-421

  400. 400. Pihan G.A., and Doxsey S.J.: The mitotic machinery as a source of genetic instability in cancer. Semin Cancer Biol 1999; 9: pp. 289-302

  401. 401. Tomonaga T., and Nomura F.: Chromosome instability and kinetochore dysfunction. Histol Histopathol 2007; 22: pp. 191-197

  402. 402. Bharadwaj R., and Yu H.: The spindle checkpoint, aneuploidy, and cancer. Oncogene 2004; 23: pp. 2016-2027

  403. 403. Michel L.S., Liberal V., Chatterjee A., Kirchwegger R., Pasche B., Gerald W., et al: MAD2 haplo-insufficiency causes premature anaphase and chromosome instability in mammalian cells. Nature 2001; 409: pp. 355-359

  404. 404. Kalitsis P., Earle E., Fowler K.J., and Choo K.H.: Bub3 gene disruption in mice reveals essential mitotic spindle checkpoint function during early embryogenesis. Genes Dev 2000; 14: pp. 2277-2282

  405. 405. Iwanaga Y., Chi Y.H., Miyazato A., Sheleg S., Haller K., Peloponese J.M., et al: Heterozygous deletion of mitotic arrest-deficient protein 1 (MAD1) increases the incidence of tumors in mice. Cancer Res 2007; 67: pp. 160-166

  406. 406. Babu J.R., Jeganathan K.B., Baker D.J., Wu X., Kang-Decker N., and van Deursen J.M.: Rae1 is an essential mitotic checkpoint regulator that cooperates with Bub3 to prevent chromosome missegregation. J Cell Biol 2003; 160: pp. 341-353

  407. 407. Perera D., Tilston V., Hopwood J.A., Barchi M., Boot-Handford R.P., and Taylor S.S.: Bub1 maintains centromeric cohesion by activation of the spindle checkpoint. Dev Cell 2007; 13: pp. 566-579

  408. 408. Putkey F.R., Cramer T., Morphew M.K., Silk A.D., Johnson R.S., McIntosh J.R., et al: Unstable kinetochore-microtubule capture and chromosomal instability following deletion of CENP-E. Dev Cell 2002; 3: pp. 351-365

  409. 409. Dai W., Wang Q., Liu T., Swamy M., Fang Y., Xie S., et al: Slippage of mitotic arrest and enhanced tumor development in mice with BubR1 haploinsufficiency. Cancer Res 2004; 64: pp. 440-445

  410. 410. Wang Q., Liu T., Fang Y., Xie S., Huang X., Mahmood R., et al: BUBR1 deficiency results in abnormal megakaryopoiesis. Blood 2004; 103: pp. 1278-1285

  411. 411. Baker D.J., Jeganathan K.B., Cameron J.D., Thompson M., Juneja S., Kopecka A., et al: BubR1 insufficiency causes early onset of aging-associated phenotypes and infertility in mice. Nat Genet 2004; 36: pp. 744-749

  412. 412. Jeganathan K., Malureanu L., Baker D.J., Abraham S.C., and van Deursen J.M.: Bub1 mediates cell death in response to chromosome missegregation and acts to suppress spontaneous tumorigenesis. J Cell Biol 2007; 179: pp. 255-267

  413. 413. Funk L.C., Zasadil L.M., and Weaver B.A.: Living in CIN: mitotic infidelity and its consequences for tumor promotion and suppression. Dev Cell 2016; 39: pp. 638-652

  414. 414. Schuyler S.C., YF W., and Kuan V.J.: The Mad1-Mad2 balancing act—a damaged spindle checkpoint in chromosome instability and cancer. J Cell Sci 2012; 125: pp. 4197-4206

  415. 415. Cahill D.P., Lengauer C., Yu J., Riggins G.J., Willson J.K., Markowitz S.D., et al: Mutations of mitotic checkpoint genes in human cancers. Nature 1998; 392: pp. 300-303

  416. 416. Shichiri M., Yoshinaga K., Hisatomi H., Sugihara K., and Hirata Y.: Genetic and epigenetic inactivation of mitotic checkpoint genes hBUB1 and hBUBR1 and their relationship to survival. Cancer Res 2002; 62: pp. 13-17

  417. 417. Gemma A., Seike M., Seike Y., Uematsu K., Hibino S., Kurimoto F., et al: Somatic mutation of the hBUB1 mitotic checkpoint gene in primary lung cancer. Genes Chromosom Cancer 2000; 29: pp. 213-218

  418. 418. Hempen P.M., Kurpad H., Calhoun E.S., Abraham S., and Kern S.E.: A double missense variation of the BUB1 gene and a defective mitotic spindle checkpoint in the pancreatic cancer cell line Hs766T. Hum Mutat 2003; 21: pp. 445

  419. 419. Imai Y., Shiratori Y., Kato N., Inoue T., and Omata M.: Mutational inactivation of mitotic checkpoint genes, hsMAD2 and hBUB1, is rare in sporadic digestive tract cancers. Jpn J Cancer Res 1999; 90: pp. 837-840

  420. 420. Ohshima K., Haraoka S., Yoshioka S., Hamasaki M., Fujiki T., Suzumiya J., et al: Mutation analysis of mitotic checkpoint genes (hBUB1 and hBUBR1) and microsatellite instability in adult T-cell leukemia/lymphoma. Cancer Lett 2000; 158: pp. 141-150

  421. 421. Nomoto S., Haruki N., Takahashi T., Masuda A., Koshikawa T., Fujii Y., et al: Search for in vivo somatic mutations in the mitotic checkpoint gene, hMAD1, in human lung cancers. Oncogene 1999; 18: pp. 7180-7183

  422. 422. Tsukasaki K., Miller C.W., Greenspun E., Eshaghian S., Kawabata H., Fujimoto T., et al: Mutations in the mitotic check point gene, MAD1L1, in human cancers. Oncogene 2001; 20: pp. 3301-3305

  423. 423. Wang Z., Cummins J.M., Shen D., Cahill D.P., Jallepalli P.V., Wang T.L., et al: Three classes of genes mutated in colorectal cancers with chromosomal instability. Cancer Res 2004; 64: pp. 2998-3001

  424. 424. Dominguez-Brauer C., Thu K.L., Mason J.M., Blaser H., Bray M.R., and Mak T.W.: Targeting mitosis in cancer: emerging strategies. Mol Cell 2015; 60: pp. 524-536

  425. 425. Gavriilidis P., Giakoustidis A., and Giakoustidis D.: Aurora kinases and potential medical applications of aurora kinase inhibitors: a review. J Clin Med Res 2015; 7: pp. 742-751

  426. 426. Penna L.S., Henriques J.A., and Bonatto D.: Anti-mitotic agents: are they emerging molecules for cancer treatment? Pharmacol Ther 2017; undefined: pp. 67-82

  427. 427. Katsha A., Belkhiri A., Goff L., and El-Rifai W.: Aurora kinase A in gastrointestinal cancers: time to target. Mol Cancer 2015; 14: pp. 106

  428. 428. Malumbres M., and Perez de Castro I.: Aurora kinase A inhibitors: promising agents in antitumoral therapy. Expert Opin Ther Targets 2014; 18: pp. 1377-1393

  429. 429. Chan K.S., Koh C.G., and Li H.Y.: Mitosis-targeted anti-cancer therapies: where they stand. Cell Death Dis 2012; 3:

  430. 430. Kinzler K.W., Nilbert M.C., LK S., Vogelstein B., Bryan T.M., Levy D.B., et al: Identification of FAP locus genes from chromosome 5q21. Science 1991; 253: pp. 661-665

  431. 431. Nishisho I., Nakamura Y., Miyoshi Y., Miki Y., Ando H., Horii A., et al: Mutations of chromosome 5q21 genes in FAP and colorectal cancer patients. Science 1991; 253: pp. 665-669

  432. 432. Nakamura Y., Nishisho I., Kinzler K.W., Vogelstein B., Miyoshi Y., Miki Y., et al: Mutations of the APC (adenomatous polyposis coli) gene in FAP (familial polyposis coli) patients and in sporadic colorectal tumors. Tohoku J Exp Med 1992; 168: pp. 141-147

  433. 433. Kinzler K.W., and Vogelstein B.: Lessons from hereditary colorectal cancer. Cell 1996; 87: pp. 159-170

  434. 434. LK S., Vogelstein B., and Kinzler K.W.: Association of the APC tumor suppressor protein with catenins. Science 1993; 262: pp. 1734-1737

  435. 435. Rubinfeld B., Souza B., Albert I., Muller O., Chamberlain S.H., Masiarz F.R., et al: Association of the APC gene product with beta-catenin. Science 1993; 262: pp. 1731-1734

  436. 436. Gumbiner B.M.: Signal transduction of beta-catenin. Curr Opin Cell Biol 1995; 7: pp. 634-640

  437. 437. Bienz M.: Beta-catenin: a pivot between cell adhesion and Wnt signalling. Curr Biol 2005; 15: pp. R64-R67

  438. 438. Willert K., and Nusse R.: Beta-catenin: a key mediator of Wnt signaling. Curr Opin Genet Dev 1998; 8: pp. 95-102

  439. 439. Novak A., and Dedhar S.: Signaling through beta-catenin and Lef/Tcf. Cell Mol Life Sci 1999; 56: pp. 523-537

  440. 440. Moon R.T., Bowerman B., Boutros M., and Perrimon N.: The promise and perils of Wnt signaling through beta-catenin. Science 2002; 296: pp. 1644-1646

  441. 441. Kikuchi A.: Regulation of beta-catenin signaling in the Wnt pathway. Biochem Biophys Res Commun 2000; 268: pp. 243-248

  442. 442. Haegebarth A., and Clevers H.: Wnt signaling, lgr5, and stem cells in the intestine and skin. Am J Pathol 2009; 174: pp. 715-721

  443. 443. van der Flier L.G., and Clevers H.: Stem cells, self-renewal, and differentiation in the intestinal epithelium. Annu Rev Physiol 2009; 71: pp. 241-260

  444. 444. Barker N., van de Wetering M., and Clevers H.: The intestinal stem cell. Genes Dev 2008; 22: pp. 1856-1864

  445. 445. de Lau W., Barker N., and Clevers H.: WNT signaling in the normal intestine and colorectal cancer. Front Biosci 2007; 12: pp. 471-491

  446. 446. Pinto D., and Clevers H.: Wnt control of stem cells and differentiation in the intestinal epithelium. Exp Cell Res 2005; 306: pp. 357-363

  447. 447. Pinto D., and Clevers H.: Wnt, stem cells and cancer in the intestine. Biol Cell 2005; 97: pp. 185-196

  448. 448. Reya T., and Clevers H.: Wnt signalling in stem cells and cancer. Nature 2005; 434: pp. 843-850

  449. 449. Clevers H., Loh K.M., and Nusse R.: Stem cell signaling. An integral program for tissue renewal and regeneration: wnt signaling and stem cell control. Science 2014; 346:

  450. 450. Krausova M., and Korinek V.: Wnt signaling in adult intestinal stem cells and cancer. Cell Signal 2014; 26: pp. 570-579

  451. 451. Yang V.W.: APC as a checkpoint gene: the beginning or the end? Gastroenterology 2002; 123: pp. 935-939

  452. 452. Bustos V.H., Ferrarese A., Venerando A., Marin O., Allende J.E., and Pinna L.A.: The first armadillo repeat is involved in the recognition and regulation of beta-catenin phosphorylation by protein kinase CK1. Proc Natl Acad Sci U S A 2006; 103: pp. 19725-19730

  453. 453. Marikawa Y., and Elinson R.P.: beta-TrCP is a negative regulator of Wnt/beta-catenin signaling pathway and dorsal axis formation in Xenopus embryos. Mech Dev 1998; 77: pp. 75-80

  454. 454. Winston J.T., Strack P., Beer-Romero P., Chu C.Y., Elledge S.J., and Harper J.W.: The SCFbeta-TRCP-ubiquitin ligase complex associates specifically with phosphorylated destruction motifs in IkappaBalpha and beta-catenin and stimulates IkappaBalpha ubiquitination in vitro. Genes Dev 1999; 13: pp. 270-283

  455. 455. Latres E., Chiaur D.S., and Pagano M.: The human F box protein beta-Trcp associates with the Cul1/Skp1 complex and regulates the stability of beta-catenin. Oncogene 1999; 18: pp. 849-854

  456. 456. Sakanaka C.: Phosphorylation and regulation of beta-catenin by casein kinase I epsilon. J Biochem 2002; 132: pp. 697-703

  457. 457. Gao Z.H., Seeling J.M., Hill V., Yochum A., and Virshup D.M.: Casein kinase I phosphorylates and destabilizes the beta-catenin degradation complex. Proc Natl Acad Sci U S A 2002; 99: pp. 1182-1187

  458. 458. Rubinfeld B., Albert I., Porfiri E., Fiol C., Munemitsu S., and Polakis P.: Binding of GSK3beta to the APC-beta-catenin complex and regulation of complex assembly. Science 1996; 272: pp. 1023-1026

  459. 459. Hart M.J., de los Santos R., Albert I.N., Rubinfeld B., and Polakis P.: Downregulation of beta-catenin by human Axin and its association with the APC tumor suppressor, beta-catenin and GSK3 beta. Curr Biol 1998; 8: pp. 573-581

  460. 460. Shtutman M., Zhurinsky J., Simcha I., Albanese C., D’Amico M., Pestell R., et al: The cyclin D1 gene is a target of the beta-catenin/LEF-1 pathway. Proc Natl Acad Sci U S A 1999; 96: pp. 5522-5527

  461. 461. He T.C., Sparks A.B., Rago C., Hermeking H., Zawel L., da Costa L.T., et al: Identification of c-MYC as a target of the APC pathway. Science 1998; 281: pp. 1509-1512

  462. 462. Tetsu O., and McCormick F.: Beta-catenin regulates expression of cyclin D1 in colon carcinoma cells. Nature 1999; 398: pp. 422-426

  463. 463. Green R.A., and Kaplan K.B.: Chromosome instability in colorectal tumor cells is associated with defects in microtubule plus-end attachments caused by a dominant mutation in APC. J Cell Biol 2003; 163: pp. 949-961

  464. 464. Green R.A., Wollman R., and Kaplan K.B.: APC and EB1 function together in mitosis to regulate spindle dynamics and chromosome alignment. Mol Biol Cell 2005; 16: pp. 4609-4622

  465. 465. Kaplan K.B., Burds A.A., Swedlow J.R., Bekir S.S., Sorger P.K., and Nathke I.S.: A role for the Adenomatous Polyposis Coli protein in chromosome segregation. Nat Cell Biol 2001; 3: pp. 429-432

  466. 466. Kita K., Wittmann T., Nathke I.S., and Waterman-Storer C.M.: Adenomatous polyposis coli on microtubule plus ends in cell extensions can promote microtubule net growth with or without EB1. Mol Biol Cell 2006; 17: pp. 2331-2345

  467. 467. Mogensen M.M., Tucker J.B., Mackie J.B., Prescott A.R., and Nathke I.S.: The adenomatous polyposis coli protein unambiguously localizes to microtubule plus ends and is involved in establishing parallel arrays of microtubule bundles in highly polarized epithelial cells. J Cell Biol 2002; 157: pp. 1041-1048

  468. 468. Bahmanyar S., Nelson W.J., and Barth A.I.: Role of APC and its binding partners in regulating microtubules in mitosis. Adv Exp Med Biol 2009; 656: pp. 65-74

  469. 469. Fodde R., Smits R., and Clevers H.: APC, signal transduction and genetic instability in colorectal cancer. Nat Rev Cancer 2001; 1: pp. 55-67

  470. 470. Fodde R., Kuipers J., Rosenberg C., Smits R., Kielman M., Gaspar C., et al: Mutations in the APC tumour suppressor gene cause chromosomal instability. Nat Cell Biol 2001; 3: pp. 433-438

  471. 471. Caldwell C.M., and Kaplan K.B.: The role of APC in mitosis and in chromosome instability. Adv Exp Med Biol 2009; 656: pp. 51-64

  472. 472. Tighe A., Johnson V.L., and Taylor S.S.: Truncating APC mutations have dominant effects on proliferation, spindle checkpoint control, survival and chromosome stability. J Cell Sci 2004; 117: pp. 6339-6353

  473. 473. Abal M., Obrador-Hevia A., Janssen K.P., Casadome L., Menendez M., Carpentier S., et al: APC inactivation associates with abnormal mitosis completion and concomitant BUB1B/MAD2L1 up-regulation. Gastroenterology 2007; 132: pp. 2448-2458

  474. 474. Aoki K., Aoki M., Sugai M., Harada N., Miyoshi H., Tsukamoto T., et al: Chromosomal instability by beta-catenin/TCF transcription in APC or beta-catenin mutant cells. Oncogene 2007; 26: pp. 3511-3520

  475. 475. Stumpff J., Ghule P.N., Shimamura A., Stein J.L., and Greenblatt M.: Spindle microtubule dysfunction and cancer predisposition. J Cell Physiol 2014; 229: pp. 1881-1883

  476. 476. Sherr C.J.: A new cell-cycle target in cancer—inhibiting cyclin D-dependent kinases 4 and 6. N Engl J Med 2016; 375: pp. 1920-1923

  477. 477. Sherr C.J., Beach D., and Shapiro G.I.: Targeting CDK4 and CDK6: from discovery to therapy. Cancer Discov 2016; 6: pp. 353-367

  478. 478. Finn R.S., Crown J.P., Ettl J., Schmidt M., Bondarenko I.M., Lang I., et al: Efficacy and safety of palbociclib in combination with letrozole as first-line treatment of ER-positive, HER2-negative, advanced breast cancer: expanded analyses of subgroups from the randomized pivotal trial PALOMA-1/TRIO-18. Breast Cancer Res 2016; 18: pp. 67

  479. 479. Casimiro M.C., Velasco-Velazquez M., Aguirre-Alvarado C., and Pestell R.G.: Overview of cyclins D1 function in cancer and the CDK inhibitor landscape: past and present. Expert Opin Investig Drugs 2014; 23: pp. 295-304

  480. 480. Finn R.S., Martin M., Rugo H.S., Jones S., Im S.A., Gelmon K., et al: Palbociclib and letrozole in advanced breast cancer. N Engl J Med 2016; 375: pp. 1925-1936

  481. 481. VanArsdale T., Boshoff C., Arndt K.T., and Abraham R.T.: Molecular pathways: targeting the cyclin D-CDK4/6 axis for cancer treatment. Clin Cancer Res 2015; 21: pp. 2905-2910

  482. 482. Dickson M.A.: Molecular pathways: CDK4 inhibitors for cancer therapy. Clin Cancer Res 2014; 20: pp. 3379-3383

  483. 483. Hamilton E., and Infante J.R.: Targeting CDK4/6 in patients with cancer. Cancer Treat Rev 2016; 45: pp. 129-138

  484. 484. Roskoski R.: Cyclin-dependent protein kinase inhibitors including palbociclib as anticancer drugs. Pharmacol Res 2016; 107: pp. 249-275

  485. 485. DiPippo A.J., Patel N.K., and Barnett C.M.: Cyclin-dependent kinase inhibitors for the treatment of breast cancer: past, present, and future. Pharmacotherapy 2016; 36: pp. 652-667

Only gold members can continue reading. Log In or Register to continue

Stay updated, free articles. Join our Telegram channel

Apr 21, 2019 | Posted by in ABDOMINAL MEDICINE | Comments Off on The Cell Cycle

Full access? Get Clinical Tree

Get Clinical Tree app for offline access