You are on page 1of 17

Annu. Rev. Physiol. 1998.

60:60117

CELL CYCLE REGULATION AND APOPTOSIS1


of Molecular and Cellular Physiology, University of Cincinnati Medical Center, P. O. Box 670576, Cincinnati, Ohio 45267-0576; #National Institutes of Health, National Institute of Environmental Health Sciences, P.O. Box 12233 MD E2-02, Research Triangle Park, North Carolina 27709; e-mail: cidlowski@niehs.nih.gov
KEY WORDS: p53, pRb, E2F, cell death, cell cycle, lymphocytes
Department

K. L. King and J. A. Cidlowski#

ABSTRACT
Tissue homeostasis requires a balance between cell proliferation and death. Apoptosis and proliferation are linked by cell cycle regulators, and apoptotic stimuli affect both cell proliferation and death. Glucocorticoids induce G1 arrest and apoptosis in transformed lymphoid cells. Decreased expression of the cell cycle components c-myc and cyclin D3 is essential for glucocorticoid-induced growth arrest and death in dividing cells. Other G1 regulators, such as p53, pRb, and E2F, have also been implicated in apoptosis. Mice lacking either p53 or E2F display aberrant cell proliferation and tumor formation, suggesting that these proteins are involved in the elimination of abnormal cells through apoptosis. In contrast, pRb induces G1 arrest and suppresses apoptosis in cultured cells. Mice that lack pRb are nonviable and show ectopic mitosis and massive cell death, suggesting that pRb is an apoptotic suppressor. Further analysis of common components of apoptotic and cell cycle machinery may provide insight into the coordinated regulation of these antagonistic processes.

INTRODUCTION
Tissue homeostasis is dependent on the proper relationships among cell proliferation, differentiation, and cell death. As somatic cells proliferate, the cell
1 The US Government has the right to retain a nonexclusive, royalty-free license in and to any copyright covering this paper.

601

602

KING & CIDLOWSKI

mitotic cycle progression is tightly regulated by an intricate network of positive and negative signals, and much is known about the molecules involved in cell cycle control. Programmed cell death, or apoptosis, is also a highly regulated process by which an organism eliminates unwanted cells without eliciting an inammatory response. Apoptosis is involved in many physiological processes including tissue homeostasis, embryonic development, and the immune response (1). Mitosis and apoptosis display several similar morphological features. Both mitotic and apoptotic cells lose substrate attachment and become rounded. During both processes, cells shrink, condense their chromatin, and display rapid membrane blebbing. Although a number of similarities exist between mitotic and apoptotic cells, several distinct differences are also apparent. For example, only apoptotic cells display fragmentation of their DNA into approximately 200 base-pair fragments. At the end of apoptosis, the cell is broken into multiple apoptotic bodies that are phagocytosed by neighboring cells. Because cellular contents are not released, this occurs with little inammation. During mitosis, DNA is segregated and the nucleus is divided into two distinct but equal parts. The mitotic process ends with cytokinesis and the production of a new daughter cell. Furthermore, cell cycle components such as p53, pRb, and E2F, have been shown to participate in both cell cycle progression and apoptosis. Hence, comparison of apoptosis and cell proliferation may provide insight into the regulation and molecular mechanisms of these two antagonistic processes. Multiple inducers of apoptosis have been identied. These signals are often cell-type specic, and a partial list includes growth factor withdrawal, ionizing radiation, Ca2+ inux, tumor necrosis factor, viral infection, and glucocorticoids. In response to these signals, the cell induces a programmed cascade of events that results in the destruction of the cell. A number of catalytic pathways are induced during apoptosis, including protease activation, which leads to the destruction of cellular proteins, and nuclease activation, which results in DNA fragmentation and RNA degradation. Caspases, also called ICE proteases, are a family of cysteine proteases that act as effectors of the mammalian cell death pathway (2). Their activation leads to morphological changes characteristic of the apoptotic process.

CELL PROLIFERATION AND APOPTOSIS


The balance between proliferation and apoptosis must be strictly maintained to sustain tissue homeostasis. An imbalance between these two processes can result in either unwanted tissue atrophy or tissue growth. For example, when adrenalectomized rats are treated with the synthetic glucocorticoid dexamethasone, a steroid that induces thymocyte apoptosis, the wet weight of the rat

GLUCOCORTICOIDS AND APOPTOSIS

603

thymus decreases by 50% within 24 h owing to an increase in the rate of apoptosis of cortical thymocytes that is not offset by an increase in mitosis (3). Tumor formation can result from a decrease in cell death, as well as an increase in cell proliferation. For example, when bcl-2, an anti-apoptotic gene, is overexpressed, neoplasia results because of a reduction of apoptosis that is not offset by a decrease in cell proliferation (4, 4a,b).

Cell Cycle Checkpoints


Because it is essential to identify and eliminate cells proliferating inappropriately, apoptosis and proliferation are tightly coupled, and cell cycle regulators can inuence both cell division and cell death (5, 6). The timing and order of cell cycle events are monitored during cell cycle checkpoints that occur at the G1/S phase boundary, in S phase, and during the G2/M phases (7). These checkpoints ensure that critical events in a particular phase of the cell cycle are completed before a new phase is initiated, thereby preventing the formation of genetically abnormal cells. Cell cycle progression can be blocked at these checkpoints in response to the status of both the intracellular and extracellular environment. For example, the expression of certain genes required for DNA synthesis occurs only in the presence of growth factors, thus linking the extracellular environment to cell proliferation. Additionally, growth arrest can be induced when DNA damage is detected or when chromosomes are misaligned on the mitotic spindle (8). Upon repair of the damage, progression through the cell cycle resumes. An alternative to repairing damaged cells is simply to eliminate them through the process of apoptosis. Thus as a cell progresses through the cell cycle, it must determine whether to complete cell division, arrest growth to repair cellular damage, or undergo apoptosis if the damage is too severe to be repaired or if the cell is incapable of repairing the DNA. It is at the checkpoints that the cell determines which of these options is suitable. The cell cycle control system is based on two protein families: the cyclindependent protein kinases (Cdks) and the cyclins. Cdks allow progression through the different phases of the cell cycle by phosphorylating substrates. Their kinase activity is dependent on the presence of activating subunits known as cyclins. The abundance of specic cyclins increases during the phase of the cell cycle wherein they are required and decreases during phases in which they are not needed. For example, in most circumstances cyclin D associates with Cdk4 and Cdk6 during early G1, whereas cyclin E activates Cdk2 during G1 to S phase transition. Cyclin A binds to Cdk2 or Cdc2 during S phase and the G2 to M phase transition, and the cyclin B/Cdc2 complex functions during the G2 to M phase transition. Thus specic cyclin/Cdk complexes are activated, and their phosphorylation of particular proteins permits the cell cycle processes to continue. The transcription of genes necessary for S phase, for

604

KING & CIDLOWSKI

example, is regulated by cyclinD/Cdk4-dependent phosphorylation of the cell cycle regulator pRb. For more extensive reviews of cyclins, Cdks, and cell cycle progression, see References 8a,b,c. Whereas cyclin binding is required for Cdk kinase activity, other proteins have been identied whose association leads to the inhibition of Cdk activity. Cyclin/Cdk complexes can be bound by Cdk inhibitor (CKI) proteins, which inhibit kinase activity and prevent cell cycle progression. Two separate families of CKI proteins have been identied (8b,d,e). The p21 family, composed of p21, p27, and p57, predominantly inhibits the Cdks of the G1 to S phase transition. The INK4 (inhibitors of Cdk4) family includes p15, p16, p18, and p19, several of which are mutated or deleted in certain types of human cancers. Numerous genes, including c-myc, are known to inhibit or activate cell proliferation by affecting the formation and activity of Cdk complexes. The protooncogene c-myc is an immediate early gene encoding a protein that functions as a transcription factor. Its activity is dependent on its association with another factor, Max, and this association is required for mitogenesis (9, 10). Expression of c-myc, which is exquisitely sensitive to growth factors, is required for quiescent cells to enter the cell cycle (11). The c-Myc protein has been shown to activate transcription of Cdc25, a phosphatase that activates Cdks, and expression of Cdc25 appears necessary for c-myc-induced cell cycle activation (12). The G1- to S-phase transition also requires c-Myc, and inhibition of c-myc expression leads to growth arrest (13). Conversely, if c-myc is expressed in G1-arrested cells in the absence of growth factors, the cells exit G1 and divide, a process that eventually leads to death (1417). Deregulated expression of this proto-oncogene has been implicated in a number of human malignancies (18, 19). A more detailed discussion of the roles of c-myc in apoptosis may be found in the chapter by EB Thompson (this volume).

GLUCOCORTICOIDS, THE CELL CYCLE, AND APOPTOSIS


Glucocorticoids induce death through at least two separate pathways. When proliferating thymocytes are treated with glucocorticoids, the signal to activate apoptosis involves changes in cell cycle components (Figure 1). Glucocorticoids also induce apoptosis in nonproliferating thymocytes, and while not identical, this signaling pathway shares some features in common with that in proliferating cells. In both systems, glucocorticoid binds to its receptor and induces changes in gene expression. However, in nonproliferating cells, the signaling pathway leading to apoptosis does not appear to involve changes in cell cycle proteins. Thus upstream signals induced by glucocorticoids in the circumstances differ depending on whether the cell is progressing through the

GLUCOCORTICOIDS AND APOPTOSIS

605

Figure 1 Possible pathways for glucocorticoid-induced apoptosis of lymphoid cells (2133). Although G1 arrest and cell death appear linked, the mechanism by which these two processes are coordinated remains unknown.

cell cycle. The downstream components of both signaling pathways ultimately appear to activate apoptotic effector molecules that include the caspases and nucleases. Many apoptotic stimuli induce cell cycle arrest before cell death, thereby affecting both cell cycle and apoptotic machinery. Growth suppressive and lytic effects of glucocorticoids on thymocytes and transformed lymphoid cells have been studied (20, 21). Glucocorticoid treatment of many lymphoid cell lines expressing markers of immature thymocytes induces G1 arrest and, in most cases, this growth suppression is followed by cell death. The precise mechanism by which glucocorticoids induce G1 arrest and apoptosis is currently being investigated but is largely unknown.

Glucocorticoids Regulate Cell Cycle Genes


Glucocorticoids elicit their effects by binding to their cognate receptors, which function as ligand-dependent transcription factors. Glucocorticoids can inhibit gene expression and alter mRNA stability of genes that are critical for the

606

KING & CIDLOWSKI

G1- to S-phase transition. One of the cell cycle genes regulated by this steroid is c-myc (2226). Although an inappropriate increase in c-myc expression usually leads to cell death, downregulation of c-myc by glucocorticoids causes apoptosis in a lymphoma cell line and in a leukemic cell line (22, 27, 28). Thus at least two pathways involving c-myc result in apoptosis. Although the precise molecular mechanism by which glucocorticoids decrease c-myc mRNA has not yet been elucidated, this decrease appears to be critical for induction of death. For example, sustained expression of c-myc blocks glucocorticoid-induced death in the human leukemic cell line CEM-C7, and antisense c-myc oligomers trigger apoptosis in these cells (27). Furthermore, the glucocorticoid antagonist RU 486 blocks steroid-induced cell death, and this inhibition is correlated with an increase in c-myc mRNA levels (28). Thus changes in this cell cycle regulator play a prominent role in glucocorticoid-induced apoptosis. Other cell cycle regulators are also sensitive to glucocorticoid treatment including cyclin D3, the primary isoform of cyclin D in thymocytes, and one of its catalytic subunits Cdk4 (23, 29, 30). In the murine lymphoma cell line P1798, glucocorticoids inhibit the transcription of Cdk4, a cyclin-dependent kinase thought to be involved in progression through G1. Additionally, cyclin D3 mRNA is rapidly downregulated in P1798 cells treated with the synthetic glucocorticoid dexamethasone (23, 29). Destabilization of this mRNA appears to be the main mechanism by which glucocorticoids decrease cyclin D3. In the absence of steroid, the half life of cyclin D3 mRNA is 8 h, whereas in the presence of glucocorticoid, half life is reduced to 1 h (29). The increase in turnover of mRNA is independent of cell cycle progression and dependent on protein synthesis. Thus glucocorticoids induce expression of proteins that accelerate the degradation of cyclin D3 mRNA. When P1798 lymphoma cells with normal cyclin D3 and c-myc are cultured in the absence of serum and treated with dexamethasone, they rapidly undergo apoptosis (23, 31, 32). In the presence of serum, these cells undergo G1 arrest. Cells that overexpress both cyclin D3 and c-myc genes override glucocorticoidinduced G1 arrest (23). Additionally, cells that overexpress both proteins in the absence of serum are resistant to glucocorticoid-induced death. Thus, cyclin D3 and c-Myc convey resistance to both G1 arrest and cell death, and glucocorticoid inhibition of both these genes appears to be critical for steroid-mediated growth suppression and apoptosis in P1798 cells. It is unclear whether G1 arrest and apoptosis are parallel, independent functions or sequential, dependent events induced by decreases in c-Myc and cyclin D. Glucocorticoid treatment of cultured lymphoid cells affects other G1 regulators including pRb and E2F (23, 33). For example, treatment of P1798 cells induced a 7590% reduction in pRb phosphorylation by a cyclin D3-associated kinase (23). Hypophosphorylation of pRb, G1 arrest, and apoptosis have been

GLUCOCORTICOIDS AND APOPTOSIS

607

observed in other cell types, including Burkitt lymphoma cells treated with an anti-immunoglobulin antibody (34) and in the myeloid leukemic cell lines HL 60 and U937 treated with the DNA-damaging agent cytosine arabinoside (35). Glucocorticoid treatment of P1798 cells induces G1 arrest and causes changes in the proteins associated with the transcription factor E2F (33). Although the role that these changes play in induction of G1 arrest and apoptosis is unclear, it is apparent that G1 checkpoint regulators play a role in both cell cycle regulation and glucocorticoid-induced apoptosis in those cells that are progressing through the cell cycle.

THE G1 CHECKPOINT AND APOPTOSIS p53


Insight into the roles of cell cycle checkpoints in apoptosis and proliferation has come from in vitro studies of cells and in vivo studies of mice lacking various cell cycle components. The G1 checkpoint regulators p53, pRb, and E2F have been extensively analyzed by both methods (36, 37). The well-characterized tumor suppressor p53 is either inactivated by mutation or sequestered by viral proteins in numerous tumors (3740). In response to cell damage, p53 has been implicated in controlling the G1- to S-phase transition, blocking cell cycle progression at G1 in response to DNA damage (41). Protein p53 mediates these effects through its transcriptional activation functions. A number of genes controlling cell cycle progression, including the cyclin-dependent kinase inhibitor p21, are transcribed in a p53-dependent manner (42, 43). It has been postulated that p53 induces the expression of p21 in response to ionizing radiation, resulting in G1 arrest. p53 also plays an important role in the induction of apoptosis. When wild-type p53 is transfected into some cell lines lacking p53, apoptosis is induced (44, 45). Within 15 min of exposure to the apoptotic signal delivered to oligodendrocytes by IL-2, p53 translocates from the cytoplasm, and a mutant p53 protects these cells from death (46). Consequently, in cell culture systems, p53 not only regulates progression through the cell cycle but under some circumstances can also induce apoptosis in damaged cells.
p53 KNOCKOUT MICE To further study the role of p53 in mammalian physiology, p53 knockout mice have been generated (47). These mice are viable but show a high incidence of cancer and typically die from T-cell lymphomas. Furthermore, thymocytes derived from animals lacking p53 are resistant to apoptosis induced by DNA damaging agents such as ionizing radiation and topoisomerase II inhibitors, but not to cell death induced by glucocorticoids or calcium ionophore/phorbol ester treatment (48, 49). Therefore, p53 participates in the induction of some, but not all, forms of apoptosis. Obviously p53

608

KING & CIDLOWSKI

is not an essential component of the machinery that carries out apoptosis but rather is an activator of its function. Both p53-dependent and -independent pathways may be necessary to transduce the large diversity of death signals, or these multiple signaling pathways may provide a fail-safe mechanism if one pathway becomes inactivated. The mechanism by which p53 induces apoptosis remains unclear. Because p53-dependent cell death may occur in the presence of transcriptional and translational inhibitors, p53 transcriptional activation apparently is not essential (5053). Additionally, p53 mutants that are unable to transactivate have been shown to induce apoptosis (51). However, others have found that p53 doublepoint mutants that cannot activate or repress transcription are compromised in their ability to induce apoptosis (52). Using loss- and gain-of-function p53 mutants, Attardi et al (53) have shown that transcriptional activation is necessary for both G1 arrest and apoptosis. However, p53 targets different genes for each process. Rowan et al (54) isolated a mutant that retains transcriptional activation abilities and can induce G1 arrest but does not induce apoptosis, suggesting that these two functions are separable. Discrepancies among results may be due to the existence of multiple transactivation-dependent and transactivationindependent pathways that lead to apoptosis. Additionally, interpretation of mutant studies may be complicated by the possibility that some mutants have lost the ability to activate genes required for cell death but retain the ability to induce genes required for other p53 functions. Thereby, mutants that are transactivation positive would be unable to induce apoptosis even though gene transcription would still be necessary for apoptotic induction. Finally, differences in model systems make comparisons of results difcult. For example, some cell lines employed contain endogenous wild-type p53 in addition to mutant p53 protein, and others have viral proteins that target p53. Further studies will be needed to resolve inconsistencies in results before the role of p53-mediated transcriptional activation and repression in apoptosis is understood.
p53 REGULATES APOPTOSIS GENES p53 regulates the expression of several proteins known to inuence the apoptotic process. In the murine leukemia cell line M1, a temperature-sensitive p53 mutant decreases bcl-2 expression and increases bax expression (55). Bcl-2 and Bax have opposing effects on cell death: Bcl-2 inhibits or delays cell death, and Bax accelerates apoptosis (56). Furthermore, the ratio of Bcl-2 to Bax has been proposed to inuence the propensity of a cell to undergo apoptosis. Additionally, Fas, a cell surface protein that triggers apoptosis upon ligand binding, is encoded by a target gene for transcriptional activation by p53 (57). Both wild-type p53 and a temperature-sensitive p53 mutant (at a permissive temperature) induced a three- to sixfold increase in fas expression in several human cell lines. Accordingly, if p53 transcriptional

GLUCOCORTICOIDS AND APOPTOSIS

609

activation and repression functions are required to induce apoptosis, these effects may be mediated through changes in the expression of known apoptotic regulators.
p53, APOPTOSIS, AND CANCER THERAPY Because p53 is altered in a large number of human cancers, defects in the p53-dependent apoptotic pathway may be a signicant obstacle to overcome for successful cancer therapy. Indeed, p53 status in cancer cells does affect the efcacy of cancer treatment (58). Both ionizing radiation and the chemotherapeutic agent adriamycin were more effective in inducing apoptosis in tumor cells that were homozygous for wild-type p53. This information is currently being used to develop new cancer therapy strategies. Reintroduction of normal p53 into tumor cells lacking p53 function has been employed for human cancer treatment. Roth et al (59) injected a retroviral vector containing the wild-type p53 driven by the B actin promoter into lung tumors of human cancer victims. They found p53 DNA in tumor cells, and posttreatment biopsies showed that apoptosis occurred more frequently. Tumor regression or growth stabilization were observed in approximately 60% of the patients treated. This suggests that introducing components necessary for a functional apoptotic pathway into tumor cells can aid in cancer treatment.

pRb
The retinoblastoma protein (pRb), similar to p53, functions as a negative regulator of cell growth and is a tumor suppressor (60). pRb inactivation or deletion is found in many cancers, including retinoblastomas and carcinomas of the lung, breast, bladder, and prostate. Similar to the case with p53, viral-transforming proteins have been shown to inactivate pRb, thus leading to inappropriate cell proliferation and tumor formation (6163). By binding to and inhibiting transcription factors such as E2F, which are necessary for S-phase entry, pRb is believed to inhibit cell cycle progression. In mid-to-late G1, cyclin-dependent kinases along with their respective cyclins, phosphorylate pRb and the pRb-like proteins p107 and p130. Once phosphorylated, pRb and p130 release the bound transcription factors, including the E2F family. These transcription factors activate expression of genes necessary for the cell to traverse from G1 to S phase. In addition to playing a role in growth arrest, pRb has been shown to suppress apoptosis. In a pRb-defective bladder carcinoma cell line, IFN induces apoptosis, but in pRb-positive cells, IFN does not induce cell death (64). In the human osteosarcoma cell line SOAS, cells transiently or stably transfected with pRb are protected from radiation-induced apoptosis (65). These cells undergo G1 arrest after irradiation, and this block in cell cycle progression may be responsible for the decrease in susceptibility to apoptosis. Also, pRb inhibits TGF- 1 induced apoptosis in hepatoma cells (66). TGF- 1 treatment of

610

KING & CIDLOWSKI

hepatoma cells is associated with inhibition of both pRb expression and phosphorylation and with induction of cell death. Overexpression of pRb in these cells overrides the TGF- 1 death-stimulating pathway and blocks apoptosis. Recently, pRb was shown to be proteolytically cleaved during apoptosis (67, 68). A member of the caspase family of apoptotic proteases specically cleaves pRb in tumor necrosis factor, staurosporine, and cytosine arabinosideinduced apoptosis (67, 68). This proteolysis is blocked by tetrapeptide inhibitors of ICE-like enzymes. The destruction of pRb, which acts as an apoptosis suppressor, may be required to induce death.
pRB KNOCKOUT MICE

Deletion of pRb in transgenic mice is an embryoniclethal mutation that causes death in utero at 14 to 15 days (6971). The pRb null embryos are defective in erythropoiesis and also show massive cell death in the central and peripheral nervous systems, and in liver, lens, and skeletal muscle precursors. This cell death is associated with inappropriate S-phase entry and increased expression of cyclin E (72). Transgenic mice that express low levels of pRb have also been generated (73). These mice die at birth due to skeletal muscle defects. There is a large increase in myoblast apoptosis, and surviving myoblasts do not differentiate properly. These results suggest a role for pRb in both cell survival and terminal differentiation (74).

p53 and pRB Interaction


There may be complementary roles for p53 and pRb as tumor suppressors, and each may be able to compensate for the loss of the others tumor suppressor activity. The basis for pRb and p53 cooperativity is being analyzed in several model systems. For example, the interaction between p53 and pRb has been examined in HeLa cells that lack both p53 and pRb (51). When p53 is transiently overexpressed in these cells, apoptosis is induced. However, cell death can be inhibited when pRb is coexpressed with p53. In this case, cell cycle arrest occurs instead of apoptosis. Thus pRb protects the cells from apoptosis, possibly by inducing cell cycle arrest. This lack of aberrant cell proliferation may prevent the induction of p53-dependent apoptosis. In addition to inducing G1 arrest, p53 can induce cell death in response to DNA damage and inappropriate proliferation signals. There appears to be a direct link between p53 and pRb in cell proliferation and apoptosis. Interactions between p53 and pRb have also been examined in transgenic animals. Mice that are pRb+/ are viable but develop thyroid and pituitary tumors (6971). Because pRb null mice are nonviable, pRb heterozygotes have been used to analyze pRb and p53 regulation of apoptosis. pRb+//p53/ mice show tumors common to each knockout alone, plus additional tumors such as pinealoblastomas and islet tumors (74). Loss of the wild-type Rb

GLUCOCORTICOIDS AND APOPTOSIS

611

allele was associated with tumorigenesis in some tissues. For example, Rb inactivation was frequently seen in pineal blastomas but not in lymphomas. Thus it appears that in some tissues each tumor suppressor can compensate for the loss of the other. Apoptosis is dependent on p53 because embryos that are null for p53 and heterozygous for pRb do not show apoptotic effects seen in mice decient in pRb alone. Thus pRb deciency leads to induction of apoptosis in a p53-dependent manner. In lens tissue, p53 and pRb functions have been extremely well characterized. In pRb null mice, lens ber cells fail to differentiate and undergo apoptosis (75). In a p53 null background, lens ber cells still do not differentiate properly, but they are not removed by apoptosis; consequently, tumors form in the lens. These results have been conrmed using DNA tumor viruses expressing proteins that inactivate p53 or pRb. When the human papilloma virus oncoprotein E7 is expressed in the developing lens, pRb is inactivated and ectopic mitosis and apoptosis is evident (76). Conversely, if E7 and E6, oncoproteins that inactivate p53, are coexpressed, apoptosis is inhibited and tumor formation occurs. E7 has also been expressed in lens cells in a p53 null background with similar results (77). Similar interactions between p53 and pRb function are also seen in the choroid plexus. A simian virus 40 (SV40) T antigen fragment that inactivates pRb was targeted to epithelial cells in the choroid plexus (78). Inactivation of pRb led to slow growing tumors and an increase in p53-dependent apoptosis. In contrast, when pRb was inactivated in a p53 null background, rapidly growing tumors formed, and there was a decrease in apoptosis. Symonds et al concluded that p53-dependent apoptosis occurring in response to tumorigenic events is a key regulator of tumor formation (78). Therefore, deregulation of the cell cycle by pRb mutation or inactivation and inhibition of apoptosis by p53 disruption result in malignant transformation in numerous tissues. These properties have been exploited by DNA tumor viruses that can simultaneously inactivate both tumor suppressors and cause deregulation of cell cycle progression and apoptosis.

E2F
The pRb protein is a transcriptional repressor that regulates gene expression by physically associating with transcription factors such as those of the E2F family (79). The function of E2Fs in cell proliferation and apoptosis has been characterized in cell culture systems and in knockout mice. The E2F family consists of ve closely related transcription factors (36). Some differences in biochemical properties of the E2F members have been noted. E2F-1, -2, and -3 prefer to bind to pRB, whereas E2F-4 and -5 predominantly associate with pRB-related proteins p107 and p130. It is likely that the different isoforms of E2F have different functions or are used to differing extents in specic tissues. E2F-1, the most well characterized member of this family, activates genes whose

612

KING & CIDLOWSKI

products are important in the G1- to S-phase transition (60, 79, 80). E2F binds to the DP family of transcription factors, and heterodimers of E2F and DP regulate gene expression necessary for G1/S-phase transition. E2F-1 plays a prominent role in G1/S-phase transition: The ectopic expression of E2F-1 can drive serum-starved, growth-arrested cells through the G1 to S phase, and expression of dominant-negative mutants of E2F-1 inhibits cell cycle progression during this phase of the cell cycle (8184). Moreover, viral oncoproteins such as SV40 large T antigen and human papilloma virus E7 can transform cells when they bind to pRb and release activated E2F-1 (79). Furthermore, E2F expression alone, and in conjunction with the oncogene ras, can transform cells that are then tumorigenic in nude mice (81, 85, 86). E2F-1 has also been shown to play a role in apoptosis in cell culture. Deregulated E2F-1 expression in rat 2 broblasts causes early S-phase entry and subsequent apoptosis (8284), which is dependent on p53 (82, 87). When E2F-1 and p53 are coexpressed in a mouse broblast cell line, apoptotic cell death is induced and p53-mediated G1 arrest is overridden. Overexpression of E2F-1 may induce apoptosis by altering transcription of genes necessary for cell survival or by inducing inappropriate progression through the cell cycle.
E2F-1 KNOCKOUT MICE Mice that are null for E2F-1 have been generated (88, 89). Because E2F is required for S-phase progression, it is perhaps surprising that E2F knockout mice do not show hypoproliferation. Instead, they show enlarged thymus glands and lymph nodes owing to the excess of single positive (CD4+/CD8 or CD4/CD8+) thymocytes. Additionally, lymphomas, lung tumors, and reproductive tract tumors are detected. Thus, in addition to its role during S phase, E2F-1 appears to suppress inappropriate cell proliferation, perhaps because E2F null mice are defective in normal apoptotic pathways. Depending on the cellular context, E2F-1 may function as an oncogene to promote cell growth or may function as a tumor suppressor to induce apoptosis. For example, when E2F-1 is bound to pRb, cell cycle progression is repressed. However, when pRb is absent, E2F-1 stimulates proliferation. The cellular concentration of E2F or the presence or absence of other cell cycle regulators may inuence whether E2F-1 functions to promote cell growth or death.

CONCLUSIONS
Apoptotic stimuli often arrest growth before inducing cell death. Glucocorticoid treatment of transformed lymphoid cells, for example, causes cells to arrest in G1 before entering apoptosis. Glucocorticoids have been shown to inhibit the expression of G1-phase components such as cdk4, cycD3, and c-myc. Downregulation of c-myc is necessary for G1 arrest and the induction of

GLUCOCORTICOIDS AND APOPTOSIS

613

Figure 2 Effects of p53, pRb, and E2F mutations in cell culture systems and transgenic mice. All these G1 regulators are capable of altering cell cycle progression and affecting the propensity of a cell to undergo apoptosis (45, 4749, 51, 6466, 6977, 8185, 88, 89).

apoptosis in CEM cells, and P1798 cells require decreased expression of both cyclin D3 and c-myc, which suggests that these cell cycle regulators inuence cell death as well as proliferation. Glucocorticoids appear to act through two separate pathways to induce apoptosisone in proliferating cells and another in nonproliferating cells. These pathways share common upstream and downstream components that include glucocorticoid binding to its receptor, changes in gene expression, and activation of apoptotic effector molecules. During mouse development, p53, pRb, and E2F function as cell cycle regulators. This is readily apparent in the phenotypes of knockout mice (Figure 2). In all three mouse models, abnormal cell proliferation is visible in some tissues; however, the response to this uncontrolled cell division differs among the three different knockout mice. In p53 and E2F knockout models, proliferation is allowed to continue unchecked, resulting in the formation of tumors. In contrast, pRb knockout mice are nonviable and display massive cell death in response to inappropriate mitosis. Thus it appears that p53 and E2F provide positive signals for apoptosis in response to inappropriate cell proliferation, and removal of these regulators results in lack of cell death and tumor formation. It appears that pRb inhibits inappropriate apoptosis from occurring during embryogenesis by inducing alternative, non-apoptotic pathways to prevent undue growth. When this protein is deleted from cells, growth arrest cannot occur and apoptosis is induced. Animals that are heterozygous for pRb display a different phenotype. These mice show unregulated proliferation and tumor formation, suggesting that pRb has other effects that may be masked by the early death of pRb null embryos. Thus although these three proteins serve as both cell cycle

614

KING & CIDLOWSKI

and apoptotic regulators, their functions differ greatly. Whereas pRb induces growth arrest to suppress apoptosis, p53 and E2F are necessary for induction of apoptosis in response to unregulated cell growth. The mechanisms by which these cell cycle regulators inuence the apoptotic process remain unknown. In summary, the cell has integrated control of two antagonistic processes cell proliferation and cell deathat cell cycle checkpoints. This coordinated regulation provides an effective means to control inappropriate cell cycle progression.
Visit the Annual Reviews home page at http://www.AnnualReviews.org.

Literature Cited 1. Schwartzman RA, Cidlowski JA. 1993. Apoptosis: the biochemistry and molecular biology of programmed cell death. Endocr. Rev. 14:13350 2. Golstein P. 1997. Controlling cell death. Science 275:108182 3. Compton MM, Cidlowski JA. 1986. Rapid in vivo effects of glucocorticoids on the integrity of rat lymphocyte genomic deoxyribonucleic acid. Endocrinology 118:3845 4. McDonnell TJ, Deane N, Platt M, Nunez G, Jaeger U, et al. 1989. bcl-2 immunoglobulin transgenic mice demonstrate extended B cell survival and follicular lympho-proliferation. Cell 57:7988 4a. McDonnell TJ, Korsmeyer SJ. 1991. Progression from lymphoid hyperplasia to high grade malignant lymphoma in mice transgenic for the t(14;18). Nature 349:25456 4b. Stasser A, Harris AW, Cory S. 1991. bcl-2 transgene inhibits T cell death and perturbs thymic self-censorship. Cell 67:88999 5. Meikrantz W, Schlegel R. 1995. Apoptosis and the cell cycle. J. Cell. Biochem. 58:16074 6. King KL, Schwartzman RA, Cidlowski JA. 1996. Apoptosis in life, death, and the cell cycle. Endocrinol. Metab. 3 Suppl. A:9397 7. Murray A, Hunt T. 1993. The Cell Cycle. New York: Oxford Univ. Press. 251 pp. 8. Murray A. 1994. Cell cycle checkpoints. Curr. Opin. Cell Biol. 6:87276 8a. Morgan DO. 1995. Principles of Cdk regulation. Nature 374:13134 8b. Hunter T, Pines J. 1994. Cyclins and cancer II: cyclin D and Cdk inhibitors come of age. Cell 79:57382 8c. Fisher RP. 1997. Cdks and cyclins in transition. Curr. Opin. Gene Dev. 7:3238 8d. Harper JW, Elledge SJ. 1994. Cdk inhibitors in development and cancer. Curr. Opin. Gene Dev. 6:5664 8e. Xiong Y. 1996. Why are there so many Cdk inhibitors? Biochim. Biophys. Acta 1288:15 9. Amati B, Brooks MW, Levy N, Littlewood TD, Evan GI, et al. 1993. Oncogenic activity of the c-Myc protein requires dimerization with Max. Cell 72:23345 10. Amati B, Littlewood TD, Evan GI, Land H. 1993. The c-Myc protein induces cell cycle progression and apoptosis through dimerization with Max. EMBO J. 12:508387 11. Eilers M, Schirm S, Bishop JM. 1991. The myc protein activates transcription of the -prothymosin gene. EMBO J. 10:133 41 12. Galaktionov K, Chen X, Beach D. 1996. Cdc25 cell-cycle phosphatase as a target of c-myc. Nature 382:51117 13. Heikkila R, Schwab G, Wickstrom E, Loke SL, Pluznik DH, et al. 1987. A cmyc antisense oligodeoxynucleotide inhibits entry into S phase but not progression from G0 to G1. Nature 328:44549 14. Askew D, Ashmun R, Simmons B, Cleveland J. 1991. Constitutive c-myc expression in IL-3 dependent myeloid cell line suppresses cycle cell arrest and accelerates apoptosis. Oncogene 6:191522 15. Evan G, Wyllie A, Gilbert C, Littlewood T, Land H, et al. 1992. Induction of apoptosis in broblasts by c-myc protein. Cell 63:11925 16. Wyllie AH, Rose KA, Morris RG, Steel

GLUCOCORTICOIDS AND APOPTOSIS


CM, Foster E, et al. 1987. Rodent broblast tumours expressing human myc and ras genes: growth, metastasis, and endogenous oncogene expression. Br. J. Cancer 56:25159 Vogt M, Lesley J, Bogenberger JM, Haggblom C, Swift S, et al. 1987. The induction of growth factor independence in murine myelocytes by oncogenes results in monoclonal cell lines and is correlated with cell crisis and karyotypic instability. Oncogene Res. 2:4963 Ryan KM, Birnie GD. 1996. myc oncogenes: the enigmatic family. Biochem. J. 314:71321 Evan GI, Littlewood TD. 1993. The role of c-myc in cell growth. Curr. Opin. Genet. Dev. 3:4449 Munck A, Crabtree GR. 1981. Glucocorticoid-induced lymphocyte death. In Cell Death in Biology and Pathology, ed. ID Bowen, RA Lockshin, pp. 32959. New York: Chapman & Hall Cidlowski JA, King KL, Evans-Storms RB, Montague JW, Bortner CD, et al. 1996. The biochemistry and molecular biology of glucocorticoid-induced apoptosis in the immune system. Rec. Prog. Horm. Res. 51:45791 Eastman-Reks SB, Vedeckis WV. 1986. Glucocorticoid inhibition of c-myc, cmyb, and c-Ki-ras expression in a mouse lymphoma cell line. Cancer Res. 46:245762 Rhee K, Bresnahan W, Hirai A, Hirai M, Thompson EA. 1995. c-myc and cyclin D (CcnD3) genes are independent targets for glucocorticoid inhibition of lymphoid cell proliferation. Cancer Res. 55:418895 Ma T, Mahajan PB, Thompson EA. 1992. Glucocorticoid regulation of c-myc promoter utilization in P1798 T-lymphoma cells. Mol. Endocrinol. 6:96068 Forsthoefel AM, Thompson EA. 1987. Glucocorticoid regulation of transcription of the c-myc cellular protooncogene in P1798 cells. Mol. Endocrinol. 1:899907 Yuh Y-S, Thompson EB. 1987. Glucocorticoid effect on oncogene/growth related gene expression in human lymphoblastic leukemic cell line CCRF-CEM. J. Biol. Chem. 264:1090410 Thulasi R, Harbour DV, Thompson EB. 1993. Supression of c-myc is a critical step in glucocorticoid-induced human leukemic cell lysis. J. Biol. Chem. 268:1830612 Thompson EB, Thulasi R, Saeed MF, Johnson BH. 1995. Glucocorticoid antagonist RU 486 reverses agonist-induced apoptosis and c-myc repression in human

615

29.

17.

30.

31.

18. 19. 20.

32.

33.

21.

34.

22.

35. 36. 37. 38. 39. 40. 41.

23.

24.

25.

26.

42.

27.

28.

43.

leukemic CEM-C7 cells. Ann. NY Acad. Sci. 761:26175 Reisman D, Thompson EA. 1995. Glucocorticoid regulation of cyclin D3 gene transcription and mRNA stability in lymphoid cells. Mol. Endocrinol. 9:15009 Rhee K, Reisman D, Bresnahan W, Thompson EA. 1995. Glucocorticoid regulation of G1 cyclin-dependent kinase genes in lymphoid cells. Cell Growth Diff. 6:69198 Thompson EA. 1991. Insensitivity to the cytolytic effects of glucocorticoids in vivo is associated with a novel slow death phenotype. Cancer Res. 51:554450 Thompson EA. 1991. Glucocorticoid insensitivity of P1798 lymphoma cells is associated with a factor that attenuates the lytic response. Cancer Res. 51:555156 Rhee K, Ma T, Thompson EA. 1994. The macromolecular state of the transcription factor E2F and glucocorticoid regulation of c-myc transcription. J. Biol. Chem. 269:1703542 Wang H, Grand RJ, Milner AE, Armitage RJ, Gordon J, et al. 1996. Repression of apoptosis in human B-lymphoma cells by CD40-ligand and Bcl-2: relationship to the cell cycle and role of the retinoblastoma protein. Oncogene 13:37379 Dou QP, Lui VW. 1995. Failure to dephosphorylate retinoblastoma protein in drug resistant cells. Cancer Res. 55:522225 Weinberg RA. 1996. E2F and cell proliferation: a world turned upside down. Cell 85:45759 Picksley SW, Lane DP. 1994. p53 and Rb: their cellular roles. Curr. Opin. Cell Biol. 6:85358 Oren M. 1992. p53: the ultimate tumor suppressor gene? FASEB J. 6:316976 Friend S. 1994. p53: a glimpse at the puppet behind the shadow play. Science 265:33435 Berns A. 1994. Is p53 the only real tumor suppressor gene? Curr. Biol. 4:137139 Kuerbitz SJ, Plunkett BS, Walsh WV, Kastan MB. 1992. Wild-type p53 is a cell cycle checkpoint determinant following irradiation. Proc. Natl. Acad. Sci. USA 89:749195 Dulic V, Kaufmann WK, Wilson SJ, Tlsty TD, Lees E, et al. 1994. p53dependent inhibition of cyclin-dependent kinase activities in human broblasts during radiation-induced G1 arrest. Cell 76:101323 El-Deiry WS, Tokino T, Velculescu VE, Levy DB, Parsons R, et al. 1993. WAF1, a potential mediator of p53 tumor suppression. Cell 75:81725

616

KING & CIDLOWSKI


57. Owen-Schaub LB, Zhang W, Cusack JC, Angelo LS, Santee SM, et al. 1995. Wildtype human p53 and a temperature sensitive mutant induce Fas/APO-1 expression. Mol. Cell. Biol. 15:303240 58. Lowe SW, Bodis S, McClatchey A, Remington L, Ruley HE, et al. 1994. p53 status and the efcacy of cancer therapy in vivo. Science 266:80710 59. Roth JA, Nguyen D, Lawrence DD, Kemp BL, Carrasco CH, et al. 1996. Retrovirusmediated wild-type p53 gene transfer to tumors of patients with lung cancer. Nat. Med. 2:98591 60. Weinberg R. 1995. The retinoblastoma protein and cell cycle control. Cell 81:32330 61. DeCaprio JA, Ludlow JW, Figge J, Shew JY, Huang CM, et al. 1988. SV40 large tumor antigen forms a specic complex with the product of the retinoblastoma susceptibility gene. Cell 54:27583 62. Dyson N, Howley PM, Munger K, Harlow E. 1989. The human papilloma virus16 E7 oncoprotein is able to bind to the retinoblastoma protein. J. Virol. 64:1353 56 63. Whyte P, Buchkovich KJ, Horowitz JM, Friend SH, Raybuck M, et al. 1988. Association between an oncogene and an antioncogene: The adenovirus E1A proteins bind to the retinoblastoma gene product. Nature 334:12429 64. Berry DE, Lu Y, Schmidt B, Fallon PG, OConnell C, et al. 1996. Retinoblastoma protein inhibits IFN- induced apoptosis. Oncogene 12:180919 65. Haas-Kogan DA, Kogan SC, Levi D, Dazin p, TAng A, et al. 1995. Inhibition of apoptosis by the retinoblastoma gene product. EMBO J. 14:46172 66. Fan G, Ma X, Kren BT, Steer CJ. 1996. The retinoblastoma gene product inhibits TGF- 1 induced apoptosis in primary rat hepatocytes and human HuH-7 hepatoma cells. Oncogene 12:190919 67. An B, Dou QP. 1996. Cleavage of retinoblastoma protein during apoptosis: an interleukin 1 -converting enzymelike protease as candidate. Cancer Res. 56:43842 68. Janicke RU, Walker PA, Lin XY, Porter AG. 1996. Specic cleavage of the retinoblastoma protein by an ICE-like protease in apoptosis. EMBO J. 15:6969 78 69. Clarke AR, Maandag ER, van Roon M, van der Lugt NMT, van der Valk M, et al. 1992. Requirement for a functional Rb1 gene in murine development. Nature 359:32830

44. Yonish-Rouach E, Grunwald D, Wilder S, Kimchi A, May E, et al. 1993. p53mediated cell death: relationship to cell cycle control. Mol. Cell. Biol. 13:1415 23 45. Yonish-Rouach E, Resnitzky D, Lotem J, Sachs L, Kimchi A, et al. 1993. Wildtype p53 induces apoptosis of myeloid leukaemic cells that is inhibited by interleukin-6. Nature 352:34547 46. Eizenberg O, Faber-Elman A, Gottlieb E, Oren M, Rotter V, et al. 1995. Direct involvement of p53 in programmed cell death of oligodendrocytes. EMBO J. 14:113644 47. Donehower LA, Harvey M, Slagle BL, McArthur MJ, Montgomery CAJ, et al. 1992. Mice decient for p53 are developmentally normal but susceptible to spontaneous tumors. Nature 356:21521 48. Clarke AR, Purdie CA, Harrison DJ, Morris RG, Bird CC, et al. 1993. Thymocyte apoptosis induced by p53-dependent and independent pathways. Nature 362:849 52 49. Lowe SW, Schmitt EM, Smith SW, Osborne BA, Jacks T. 1993. p53 is required for radiation-induced apoptosis in mouse thymocytes. Nature 362:84749 50. Wagner AJ, Kokontis JM, Hays N. 1994. Myc-mediated apoptosis requries wildtype p53 in a manner independent of cell cycle arrest and the ability of p53 to induce p21 WAF1/CIP1. Genes Dev. 8:281730 51. Haupt Y, Rowan S, Oren M. 1995. p53mediated apoptosis in HeLa cells can be overcome by excess pRb. Oncogene 10:156371 52. Sabbatini P, Lin J, Levine AJ, White E. 1995. Essential role for p53-mediated transcription in E1A-induced apoptosis. Genes Dev. 9:218492 53. Attardi LD, Lowe SW, Brugarolas J, Jacks T. 1996. Transcriptional activation by p53, but not induction of the p21 gene, is essential for oncogene-mediated apoptosis. EMBO J. 15:3693701 54. Rowan S, Ludwig RL, Haupt Y, Bates S, Lu X, et al. 1996. Specic loss of apoptotic but not cell-cycle arrest function in a human tumor derived p53 mutant. EMBO J. 15:82738 55. Miyashita T, Krajewski S, Krajewski M, Wang HG, Lin HK, et al. 1994. Tumor suppressor p53 is a regulator of bcl-2 and bax gene expression in vitro and in vivo. Oncogene 9:1799805 56. Reed JC. 1994. Bcl-2 and the regulation of programmed cell death. J. Cell Biol. 124:16

GLUCOCORTICOIDS AND APOPTOSIS


70. Jacks T, Fazeli A, Schmitt EM, Bronson RT, Goodell MA, et al. 1992. Effects of an Rb mutation in the mouse. Nature 359:295300 71. Lee EY-HP, Chang C-Y, Hu N, Wang YCJ, Lai C-C, et al. 1992. Mice decient for Rb are nonviable and show defects in neurogenesis and haematopoiesis. Nature 359:28894 72. Macleod KF, Hu Y, Jacks T. 1996. Loss of Rb activates both p53-dependent and independent cell death pathways in the developing mouse nervous system. EMBO J. 15:617888 73. Zacksenhaus E, Jiang Z, Chung D, Marth JD, Phillips RA, et al. 1996. pRb controls proliferation, differentiation, and death of skeletal muscle cells and other lineages during embryogenesis. Genes Dev. 10:305164 74. Williams BO, Remington L, Alberts DM, Mukai S, Bronson RT, et al. 1994. Cooperative tumorigenic effects of germline mutations in RB and p53. Nat. Genet. 7:480 84 75. Morgenbesser SD, Williams BO, Jacks T, Depino RA. 1994. p35-dependent apoptosis produced by Rb deciency in the eye. Nature 371:7274 76. Pan H, Griep AE. 1994. Altered cell cycle regulation in the lens of HPV-16 E6 or E7 transgenic mice: implications for tumor suppressor function in development. Genes Dev. 8:128599 77. Howes KA, Ransom N, Papermaster DS, Lasudry JGH, Alberts DM, et al. 1994. Apoptosis or retinoblastoma: alternative fates of photoreceptors expressing the HPV-16 E7 gene in the presence or absence of p53. Genes Dev. 8:130010 78. Symonds H, Krall L, Remington L, Saenz-Robles M, Lowe S, et al. 1994. p53-dependent apoptosis suppresses tumor growth and progression in vivo. Cell 78:70311

617

79. Lam EW-F, Thangue NBL. 1994. DP and E2F proteins: coordinating transcription with cell cycle progression. Curr. Opin. Cell Biol. 6:85966 80. Sherr CJ. 1993. Mammalian G1 cyclins. Cell 73:105965 81. Johnson DG, Schwarz JK, Cress WD, Nevins JR. 1993. Expression of transcription factor E2F-1 induces quiescent cells to enter S phase. Nature 365:34952 82. Qin X-Q, Livingston DM, WG, Kaelin J, Adams PD. 1994. Deregulated transcription factor E2F-1 expression leads to Sphase entry and p53-mediated apoptosis. Proc. Natl. Acad. Sci. USA 91:10918 22 83. Shan B, Lee W-H. 1994. Deregulated expression of E2F-1 induces S phase entry and leads to apoptosis. Mol. Cell. Biol. 14:816673 84. Shan B, Durfee T, Lee W-H. 1996. Disruption of RB/E2F-1 interaction by single point mutations in E2F-1 enhances S-phase entry and apoptosis. Proc. Natl. Acad. Sci. USA 93:67984 85. Singh P, Wong SH, Hong W. 1994. Overexpression of E2F-1 in rat embryo broblasts leads to neoplastic transformation. EMBO J. 13:332938 86. Xu G, Livingston DM, Krek W. 1995. Multiple members of the E2F transcription factor family are the products of oncogenes. Proc. Natl. Acad. Sci. USA 92:135761 87. Wu X, Levine AJ. 1994. p53 and E2F1 cooperate to mediate apoptosis. Proc. Natl. Acad. Sci. USA 91:36026 88. Field SJ, Tsai F-Y, Kuo F, Zubiaga AM, W.G. Kaelin J, et al. 1996. E2F-1 functions in mice to promote apoptosis and suppress proliferation. Cell 85:54961 89. Yamasaki L, Jacks T, Bronson R, Goillot E, Harlow E, et al. 1996. Tumor induction and tissue atrophy in mice lacking E2F-1. Cell 85:53748

You might also like