Skip to main content

MINI REVIEW article

Front. Cell Death, 03 July 2023
Sec. Apoptosis
Volume 2 - 2023 | https://doi.org/10.3389/fceld.2023.1223926

Regulators mount up: the metabolic roles of apoptotic proteins

  • Department of Biological Sciences, University of Notre Dame, Notre Dame, IN, United States

The induction of apoptosis, a programmed cell death pathway governed by activation of caspases, can result in fundamental changes in metabolism that either facilitate or restrict the execution of cell death. In addition, metabolic adaptations can significantly impact whether cells in fact initiate the apoptotic cascade. In this mini-review, we will highlight and discuss the interconnectedness of apoptotic regulation and metabolic alterations, two biological outcomes whose regulators are intertwined.

Introduction

Apoptosis is a programmed cellular suicide mechanism with important roles in the maintenance of multicellular organisms. The classical intrinsic apoptotic pathway begins with an apoptotic stimulus that alters pro-apoptotic proteins at the mitochondria. Subsequently, this leads to outer membrane permeabilization (MOMP) as a consequence of Bax/Bak oligomerization (Oltval et al., 1993; Shi et al., 2003; Buytaert et al., 2006; McArthur et al., 2018). MOMP allows for the secretion of soluble cytochrome c into the cytosol where it binds to apoptotic protease activating factor 1 (Apaf-1) and causes the formation of the apoptosome (Schafer and Kornbluth, 2006; Riedl and Salvesen, 2007). Apoptosome formation triggers the concomitant cleavage and activation of caspase 9, which cleaves and activates caspases 3 and 7 ultimately culminating in the death of the cell (Xue and Robert Horvitz, 1995; Elmore, 2007; McIlwain et al., 2013). Each of the proteins listed are critically important for proper execution of the intrinsic apoptosis pathway. However, these proteins (and others known to regulate apoptosis) do not solely function as determinants of whether a cell lives or dies. Rather, many of the main players in apoptotic pathways have other, oftentimes vital, roles in cellular processes such as metabolism. Perhaps the best example of the dual role of certain apoptotic proteins within metabolic pathways is the aforementioned cytochrome c, whose apoptotic role was discovered in the laboratory of Xiaodong Wang (Liu et al., 1996; Li et al., 1997). While the discovery of a pro-apoptotic role for a protein best understood to function in the mitochondrial respiratory chain was (at the time) surprising, the capacity of cytosolic cytochrome c to activate caspases is now central to our understanding of programmed cell death. In this article, it is our objective to highlight the sometimes-underappreciated metabolic functions of apoptotic regulators and to discuss the circumstances in which metabolism directly impinges upon the execution of apoptosis.

Glucose flux and the capacity of p53 to induce apoptosis

As mentioned above, turnover of aged or defective cells in the human body can often occur through the induction of apoptosis (Kerr et al., 1972; Fawthrop et al., 1991). The execution of apoptosis is tightly regulated by a number highly conserved and distinct signal transduction networks. In addition, proteins which regulate apoptotic death can be broadly separated into two main categories: pro-apoptotic and anti-apoptotic. Anti-apoptotic proteins function to provide defenses against the activation of proteases that execute the apoptotic program. Meanwhile, pro-apoptotic effectors can sense deleterious events in the cell and sound the alarm to initiate the induction of apoptosis. Alternatively, pro-apoptotic effectors can function directly in the cleavage of target proteins that ultimately contribute to the orderly dismantling of the cell. Dysregulation of these coordinated pathways is linked to pathological conditions like fibrosis, autoimmune disease, neurodegeneration, and cancer (Elmore, 2007).

One of the most well-studied pro-apoptotic regulators is p53, a known tumor suppressor that integrates extracellular and intracellular signals to promote cell death (Surget et al., 2013; Aubrey et al., 2018). The best characterized activator of p53 is DNA damage, which ultimately causes phosphorylation and stabilization of the p53 protein and thus allows is to transcriptionally activate downstream targets. However, nutrient availability and energy demands can also impact p53 activity (Horn and Vousden, 2007). Previous work has demonstrated that high levels of ADP promote the binding of p53 to DNA whereas elevated ATP levels block this interaction (Okorokov and Milner, 1999). There are also significant links between the withdrawal of growth factors, changes in nutrient uptake and p53. For example, serum starvation can often promote p53-mediated increased apoptosis due to elevated expression of p53 upregulated modulator of apoptosis (PUMA) (Ekoff et al., 2007). In contrast, if glucose uptake is maintained (e.g., through localization of Glut1 to the plasma membrane) after growth factor withdrawal, apoptosis is attenuated as a consequence of glycolytic flux-mediated inhibition of p53 and PUMA (Zhao et al., 2008).

In addition to the links between glucose metabolism and PUMA, the withdrawal of extracellular glucose from cells acts as a trigger for the nucleocytoplasmic isoform of the metabolic enzyme malate dehydrogenase (Lee et al., 2009). This enzyme is normally involved in the malate-aspartate shuttle, and can function to stabilize and transactivate p53 to promote p53-dependent cell-cycle arrest. Similarly, withdrawal of glucose often leads to the activation of the AMP-activating protein kinase (AMPK), a cell cycle arrest at the G (1)/S stage, and the induction of p53 (Jones et al., 2005; Okoshi et al., 2008).

Oftentimes, the regulation of apoptosis by p53 is largely a consequence of post-translational modifications to key residues that affect protein stability and ultimately the genes targeted by this transcription factor. Indeed, the AMPK-mediated p53 activation during glucose deprivation is dependent upon the phosphorylation of serine 15 (Jones et al., 2005). Similarly, the capacity of glucose to prevent p53 induction of PUMA depends on a decrease in the phosphorylation of serine 46 (Garufi and D'Orazi, 2014). In addition to these roles, the decrease of glucose utilization can lower the apoptotic threshold of cells due to alterations in the cytoplasmic function of p53. A good example of this p53-mediated effect can be found in patient-derived models of glioblastoma which have been pharmacologically restricted in their glucose uptake. These tumors are primed for apoptosis in vivo but are kept alive owing to the sequestration of cytoplasmic p53 by Bcl-xL and the prevention of p53 mediated transcription of pro-apoptotic factors keeping the tumors just below the threshold of cell death (Mai et al., 2017). Lastly, p53 can function to ensure that elevated glucose flux cannot prevent the initiation of the apoptotic cascade by inhibiting the expression of the glucose transporters Glut1 and Glut4 (Schwartzenberg-Bar-Yoseph et al., 2004).

TIGAR: a protein that functions downstream of p53 to impact metabolism

The understanding of how metabolic flux impacts p53 regulation of apoptosis also inspired efforts to study downstream effectors of p53 that impact metabolism. One particular example of p53-mediated metabolic regulators is TP53-inducible glycolysis and apoptosis regulator (TIGAR). TIGAR was discovered through microarray analysis of genes following p53 induction and functions to attenuate glycolytic flux through regulation of fructose-2,6-bisphosphatase (Bensaad et al., 2006). The capacity to regulate fructose-2,6-bisphosphatase is dependent on TIGAR’s phosphatase activity, which inhibits the flow of carbon units through glycolysis and prevents glycolysis from counteracting the action of p53. In addition, the activity of TIGAR not only blocks glycolysis, but also can promote cell survival by diverting glycolytic flux into biosynthetic or of antioxidant-generating pathways. One key determinant of TIGAR function is subcellular localization, which is highly dependent on various stress stimuli encountered by the cell. While in the cytoplasm, TIGAR antagonizes glycolysis and shifts carbon flux into the pentose phosphate pathway (PPP), which promotes production of NADPH for redox homeostasis or nucleotide synthesis. However, exposure to DNA damaging chemotherapeutic agents causes TIGAR to translocate to the nucleus where it halts the cell cycle and promotes DNA damage repair in a p53-dependent manner (Yu et al., 2015). Hypoxia can also impact the subcellular localization of TIGAR. Under hypoxic conditions, TIGAR can localize to the outer mitochondrial membrane where it binds to hexokinase 2 (HK2) and increases HK2 activity. The binding of TIGAR to HK2 increases glycolytic flux and helps to lower mitochondrial reactive oxygen species (ROS) with TIGAR functioning as a scaffold for HK2 activation (Cheung et al., 2012). Studies have also examined the loss of TIGAR, which can promote glucose oxidation and glycolysis in myocardial tissue demonstrating the importance this enzyme plays in throttling glycolytic flux following p53 activation (Okawa et al., 2019). Taking all of this together, it is clear that TIGAR is an important link between p53 and glycolytic metabolism. However, p53 induced genes can reciprocally counteract this effect to facilitate p53-mediated apoptosis. Thus, p53 exemplifies a “push-pull” dynamic between one of the best studied tumor suppressors (p53) and a common metabolic alteration observed in cancer cells (increased glycolysis) with important ramifications for disease progression.

MOMP and metabolism

Following the initiation of apoptotic signaling, pores are formed in the outer mitochondrial membrane during MOMP to allow proteins from the mitochondrial intermembrane space, including cytochrome c, to reach the cytosol and propagate the apoptotic cascade (Westphal et al., 2011). Although cytochrome c release was once considered the point of no return for cell death, research over the past decade indicates that sub-lethal amounts of MOMP may occur in response to certain apoptotic stimuli. This phenomenon is known as incomplete or “minority” MOMP (Tait et al., 2010; Ichim et al., 2015). For a cell to survive MOMP it must either limit the formation of the pore, neutralize the cytochrome c that is released, or functionally adapt to improve its fitness in this environment.

Blocking formation of the pore in the mitochondrial membrane is a logical place for the beginning of our discussion on links between metabolism and MOMP. Indeed, bioactive lipids, such as sphingolipids, have been implicated as a significant player in the regulation of MOMP in a variety of cell types. More specifically, ceramide, a class of sphingolipid, can induce apoptosis through a mechanism that is dependent on the molecular machinery required to promote MOMP (Taha et al., 2006; von Haefen et al., 2002). Furthermore, inhibition of sphingolipid metabolism through the use of pharmacological inhibitors is capable of preventing MOMP as it blocks the interaction between ceramide metabolites and inducers of MOMP (Chipuk et al., 2012). Relatedly, cancer cells can manipulate sphingolipid metabolism to decrease the intracellular levels of ceramide to maintain mitochondrial membrane integrity. As an example, sphingomyelin synthases are activated in leukemic cells to decrease the levels of ceramide and prevent stress-induced apoptosis (Dolgachev et al., 2004; Separovic et al., 2007; Lafont et al., 2010).

However, if a cell cannot alter metabolism to prevent MOMP, another strategy to survive involves the neutralization of the cytochrome c that is released from the inner membrane space. One such strategy to neutralize the execution of apoptosis employed by these cells is the rewiring of metabolism to defang cytosolic cytochrome c and conserve viability. Retroviral cDNA screens following MOMP have shown a role for glyceraldehyde-3-phosphate dehydrogenase (GAPDH) in promoting survival through increased glycolytic flux in the cell when caspase activation was inhibited (Colell et al., 2007). Interestingly, cytochrome c released from these cells following MOMP was re-localized back to the mitochondria only in cells with increased GAPDH expression. These interesting results indicate that GAPDH can promote recovery and function of those mitochondria post-MOMP. In addition, the pro-apoptotic activity of cytochrome c relies on its redox state and is known to be controlled by ROS signaling. Increased glucose flux into the PPP of cancer cells generates elevated levels of the antioxidant glutathione, which can also function to antagonize cell death by numerous mechanisms [e.g., ferroptosis (Dixon et al., 2012)]. However, with regards to apoptotic regulation, glutathione can inactivate cytochrome c following MOMP to prevents apoptosis (Vaughn and Deshmukh, 2008). Instead of attempting to inactivate cytochrome c, a cell can also survive during MOMP by simply eliminating the cytochrome c that is released. RNAi screens in both neurons and cancer cells uncovered a conserved strategy for dealing with cytoplasmic cytochrome c. p53-associated Parkin-like cytoplasmic protein (PARC) functions as an E3 ligase that targets cytochrome c for destruction following mitochondrial stress and minority MOMP to promote viability (Gama et al., 2014). In some cases, adaptation to a sub-lethal MOMP benefits a cell and may eventually promote tumorigenesis. Due to the fact that minority MOMP induces only limited caspase activation to cause DNA damage, this can contribute to genomic instability and result in oncogenic transformation (Ichim et al., 2015). Malignant cells that have been treated with agents to induce MOMP allows for cytosolic cytochrome c to active the heme-regulated inhibitor kinase (HRI) engaging the integrated stress response (ISR) (Kalkavan et al., 2022). The engagement of the ISR by cancerous cells generates a drug-resistant cell population that is not only protected against apoptosis but also abrogates the efficacy of the therapy. Therefore, minority MOMP can not only promote tumorigenesis but also contribute to therapeutic resistance in malignant cells.

Metabolic functions of Bcl-2 family members

The formation of pores in the outer mitochondrial membrane as a consequence of the regulation of Bcl-2 family members is often critical to the induction of apoptosis. As discussed above, the Bcl-2 proteins have multiple structural and functional similarities with the ability to be either pro- or anti-apoptotic. For example, Bcl-2 family members are well known to contain BH3 motifs (Bcl-2 homology 3) (Blaineau and Aouacheria, 2009), which have diverse biological functions in the regulation of apoptosis. However, many of the Bcl-2 family members also have a significant role in the regulation of metabolism. Notably, while Bcl-2 family members have been shown to regulate mitochondrial dynamics and thereby indirectly alter metabolic pathway utilization (Autret and Martin, 2010)), we have chosen to focus the section below on the involvement of Bcl-2 family members on metabolic flux directly.

One such Bcl-2 family member with a significant metabolic function is Myeloid cell leukemia-1 (MCL1 or Mcl-1) which localizes to the mitochondria to prevent MOMP and blocks apoptosis (Kozopas et al., 1993; Reynolds et al., 1994; Zhou et al., 1997; Opferman et al., 2003; Opferman et al., 2005). Mcl-1 is among the most overexpressed survival proteins across all human cancers and is linked to poor survival and therapeutic resistance (Wuillème-Toumi et al., 2005; Wei et al., 2006; Beroukhim et al., 2010). In addition to its anti-apoptotic function, Mcl-1 has been shown to facilitate nutrient recycling through the induction of mitophagy. Although Mcl-1 inhibits nonselective autophagy caused by nutrient starvation, Mcl-1 promotes the targeted destruction of depolarized mitochondria ensuring a functional mitochondrial pool to meet energetic needs (Moyzis et al., 2022). However, increased mitophagy is not always beneficial to the cell. Recent investigations have suggested that overactive mitophagy may prove detrimental to the cell and compromise viability depending on the context of the cellular stressor (Hawk et al., 2018). One can therefore postulate that perhaps a sustained increase in Mcl-1 activity would deplete the mitochondrial pool below required levels.

Mcl-1 also helps cancerous cells meet energetic needs through direct regulation of metabolic pathways (see Box 2 in Figure 1). These functions can be tied to alternative splicing as distinct Mcl-1 spliceoforms can impact metabolism in unique manners. The full-length isoform associates with the outer mitochondrial membrane where it regulates cell death. However, the short isoform of Mcl-1 is imported into the inner mitochondrial matrix where it regulates fusion, ATP production, membrane potential, and maintenance of ATP synthase to support the adenylate energy charge of the cell (Perciavalle et al., 2012). The short isoform of Mcl-1 also has a direct impact on fatty acid oxidation (FAO). The α helix of Mcl-1 directly interacts with very long-chain acyl-CoA dehydrogenase (VLCAD) in nutrient deprived conditions to dynamically tune FAO (Escudero et al., 2018). This metabolic action of Mcl-1 during periods of stress may provide additional support to cancer cells that extend beyond survival. In fact, in B Cell acute lymphoblastic leukemias (B-ALL), where Mcl-1 overexpression is a defining characteristic, FAO is a critical fuel source. Expression of Mcl-1 correlates with elevated FAO gene signatures in these malignancies and loss of Mcl-1 rewires fuel utilization from catabolism of fatty acids to a reliance upon glycolysis (Prew et al., 2022). As such, in the absence of apoptotic stimuli, Mcl-1 can functions as a significant regulator of the FAO program.

FIGURE 1
www.frontiersin.org

FIGURE 1. Metabolic function of apoptotic regulators.

Pro-apoptotic Bcl-2 family members also have well-defined functions in metabolic pathways. For example, Bcl-2 associated agonist of cell death (BAD) promotes programmed cell death by binding to Bcl-2 and preventing its ability to block MOMP and cytochrome c release (Yang et al., 1995). BAD is often regulated by phosphorylation which determines not only its role in apoptosis, but also the specific function it plays in metabolism (See Box 1 in Figure 1). For example, BAD affects glucose utilization by forming a complex containing glucokinase at the mitochondria due to its phosphorylation at S112. As a consequence, the activity of glucokinase is increased in a fashion that drives mitochondrial oxidative phosphorylation (Danial et al., 2003). Meanwhile, phosphorylation of BAD at S118 and S99 has been shown to affect metabolism due to altered Akt signaling, stimulation of complex I and elevated mitochondrial oxygen consumption (Mann et al., 2019). BAD can also bind to other metabolic enzymes as immunoprecipitation experiments have demonstrated that it can associate with phosphofructokinase 1 (PFK1). Functional assays demonstrate that JNK1-mediated phosphorylation of BAD at T201 increases PFK1 function to raise fructose-1,6-bisphosphate generation, a key rate-limiting step of glycolysis (Deng et al., 2008). Some of the post-translational modifications to BAD are cell type-specific for distinct tissue metabolic functions. For example, modification to the S155 site in the BH3 domain of BAD allows the protein to control glucose stimulated insulin secretion in the beta cells of the pancreas (Danial et al., 2008). Like the pancreas, liver-specific activity of BAD modulates the metabolism of hepatic cells. The liver must precisely balance gluconeogenesis, the generation of glucose from other metabolic substrates, and glycolysis to maintain homeostasis. BAD maintains this physiological balance by employing phosphorylation of its BH3 domain as a sensor to activate glucokinase and suppress gluconeogenesis to properly coordinate hepatic glucose output (Giménez-Cassina et al., 2014).

Links between dietary metabolism, apoptosis, and disease

The interplay between metabolism and apoptosis extends beyond the individual cell as whole-organism metabolism impacts death across tissue systems. For example, it has long been known that heightened availability of the sugar glucose is beneficial to tumorigenesis and cancer cell survival (Warburg, 1956; Liberti and Locasale, 2016). In fact, the metabolic rewiring of cells to promote increased glucose uptake and glycolytic flux is a hallmark of cancer (Hanahan and Weinberg, 2011). Monosaccharides in general are not universal in their impact on cell survival versus death, and thus, the specific form of sugar taken up by cellscould alter the balance between tumor promotion and tumor suppression.

In addition to glucose, fructose has also been shown to promote cancer through inhibition of apoptotic pathways. Nutrients taken in through diet enter the tissue of mammals at the epithelium of the small intestine and colon. In the small intestine, sugars from the diet (e.g., glucose and fructose) enter the cell through the Glut family of transporters where they are phosphorylated by hexokinase (HK) and ketohexokinase (KHK) respectively (Miller et al., 1956; Roberts and Miyamoto, 2015; Jang et al., 2018). The uptake of these metabolites is governed by the villus structure within the small intestine. Intestinal villus length is balanced by proliferation and death of intestinal epithelial cells (IECs) (Hall et al., 1994). The migration of IECs up the villus separates them from their blood supply leading to hypoxia which ultimately can cause their apoptotic death. However, fructose increases the survival of hypoxic cells as well as IECs of the small intestine in mice (Taylor et al., 2021). Fructose in the IECs results in the accumulation of fructose-1-phosphate which competes with fructose-1,6-bisphosphate for the binding pocket of the glycolytic enzyme pyruvate kinase (PKM2). In this setting, inhibition of glycolytic flux, through decreased PKM2 activity, allows for increased upstream glycolytic metabolites to counteract the hypoxic insult and thereby dampen the apoptotic stimulus (Luo et al., 2011; Taylor et al., 2021). The cancer-causing potential of fructose is solidified by the finding that excessive levels of fructose in the colon, owing to the consumption of high fructose corn syrup, increases tumorigenesis in murine models (Goncalves et al., 2019). Importantly in this study, fructose was found to saturate the Glut receptors on the IECs causing fructose levels to still be high even into the colon where it could exert its neoplastic potential. Both glucose and fructose benefit malignant cells by promoting growth while simultaneously blunting apoptotic signaling.

Alternatively, other sugar molecules—including those deemed “rare” sugars—are deleterious to cancer cells. Rare monosaccharides with structural similarity to glucose or fructose induce apoptosis of cancer cells and may function as tumor suppressors. One such rare sugar, D-allose, has been shown in human head and neck cancers to induce apoptosis due to competitive inhibition of glucose uptake and lowering of the apoptotic threshold to improve efficacy of the taxane docetaxel (Mitani et al., 2009; Indo et al., 2014). Another rare sugar, L-sorbose, the C-3 epimer of fructose, is taken up and initially phosphorylated similarly to fructose by KHK to produce L-sorbose-1-phosphate (S-1-P). However, unlike fructose, S-1-P inhibits the action of HK at the top of glycolytic flux thereby preventing the synthesis of essential glycolytic metabolites and promoting apoptosis (Xu et al., 2023). In this same vein, mannose, the C-2 epimer of glucose, is also tumor suppressive. Mannose enters the cell through Glut transporters and is phosphorylated by HK to mannose-6-phosphate (M6P). M6P decreases the levels of the anti-apoptotic proteins Mcl-1 and Bcl-XL by inhibiting key glycolytic enzymes including HK and sensitizing cells to the chemotherapeutic cisplatin (Gonzalez et al., 2018). These findings importantly show that it is not necessarily the abundance of sugar but rather the specific sugar source which tips the balance between tumorigenesis and cell death.

Conclusion and perspectives

As discussed here, apoptosis and metabolism are tightly interwoven cellular processes. The choice of whether or not to activate the cellular suicide program can be dictated by nutrient availability and metabolic pathway engagement. Apoptosis is kept in check through the sequestration of important apoptotic players (e.g., cytochrome c in the mitochondria) during normal metabolic flux with robust nutrient sources. Members of the Bcl-2 family of proteins play integral roles in properly trafficking nutrients through their requisite pathways to ensure proper growth and division of the cell. Apoptosis-inducing stimuli alter metabolic flux modifying the function, localization, or both of apoptotic proteins which then either must be blocked or the cell will succumb to this insult. There is inherent difficulty in decoupling the metabolic and death role of apoptotic proteins. The examination of the metabolic role of these proteins through loss of function studies can prove lethal to the cell. Therefore, thorough consideration must be taken to safeguard cell viability when attempting to disentangle the metabolic role from the cell death aspect. Mutations in these proteins which specifically inactivate the apoptotic domain (e.g., altering the BH3 domain of Bcl-2 family proteins) are a useful tool for elucidating the metabolic pathways regulated by the protein of interest. Likewise, in basal conditions or those that do not meet the apoptotic threshold, such as minority MOMP, the metabolic impact of apoptotic regulators could better be considered. This is particularly important given the prevalence of chemotherapeutics which specifically target Bcl-2 family members to kill cancer cells while simultaneously having a metabolic impact on non-malignant tissues (such as immune cells, cancer-associated fibroblasts, etc.). On the other hand, the advent of recently approved cancer therapeutics targeting metabolism may inadvertently trigger pathways that regulate apoptosis (Stine et al., 2022). These alterations to apoptotic signaling may be synergistic or may function to counteract the efficacy of these drugs, but possible knock-on impacts to apoptosis must be considered. Ideally, future ventures to treat cancer will exploit the relationship between apoptosis and metabolism to simultaneously activate the cellular death programs while starving transformed cells of nutrients required for survival.

Author contributions

JS and ZS jointly conceptualized, drafted, and edited this contribution. All authors contributed to the article and approved the submitted version.

Acknowledgments

We thank Veronica Schafer and all current/past ZS lab members for helpful comments and valuable discussion. We are grateful for support from the National Institutes of Health/National Cancer Institute (R01CA262439), the Coleman Foundation, the Malanga Family Excellence Fund for Cancer Research at Notre Dame, the College of Science at Notre Dame, the Department of Biological Sciences at Notre Dame, and funds from Mr. Nick L. Petroni.

Conflict of interest

The author ZS declared that they were an editorial board member of Frontiers, at the time of submission. This had no impact on the peer review process and the final decision.

The remaining author declares that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

Aubrey, B. J., Kelly, G. L., Janic, A., Herold, M. J., and Strasser, A. (2018). How does p53 induce apoptosis and how does this relate to p53-mediated tumour suppression? Cell. Death Differ. 25, 104–113. doi:10.1038/cdd.2017.169

PubMed Abstract | CrossRef Full Text | Google Scholar

Autret, A., and Martin, S. J. (2010). Bcl-2 family proteins and mitochondrial fission/fusion dynamics. Cell. Mol. Life Sci. 67, 1599–1606. doi:10.1007/s00018-010-0286-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Bensaad, K., Tsuruta, A., Selak, M. A., Vidal, M. N. C., Nakano, K., Bartrons, R., et al. (2006). TIGAR, a p53-inducible regulator of glycolysis and apoptosis. Cell. 126, 107–120. doi:10.1016/j.cell.2006.05.036

PubMed Abstract | CrossRef Full Text | Google Scholar

Beroukhim, R., Mermel, C. H., Porter, D., Wei, G., Raychaudhuri, S., Donovan, J., et al. (2010). The landscape of somatic copy-number alteration across human cancers. Nature 463, 899–905. doi:10.1038/nature08822

PubMed Abstract | CrossRef Full Text | Google Scholar

Blaineau, S. V., and Aouacheria, A. (2009). BCL2DB: Moving 'helix-bundled' BCL-2 family members to their database. Apoptosis 14, 923–925. doi:10.1007/s10495-009-0376-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Buytaert, E., Callewaert, G., Vandenheede, J. R., and Agostinis, P. (2006). Deficiency in apoptotic effectors BAX and BAK reveals an autophagic cell death pathway initiated by photodamage to the endoplasmic reticulum. Autophagy 2, 238–240. doi:10.4161/auto.2730

PubMed Abstract | CrossRef Full Text | Google Scholar

Cheung, E. C., Ludwig, R. L., and Vousden, K. H. (2012). Mitochondrial localization of TIGAR under hypoxia stimulates HK2 and lowers ROS and cell death. Proc. Natl. Acad. Sci. U. S. A. 109, 20491–20496. doi:10.1073/pnas.1206530109

PubMed Abstract | CrossRef Full Text | Google Scholar

Chipuk, J. E., McStay, G. P., Bharti, A., Kuwana, T., Clarke, C. J., Siskind, L. J., et al. (2012). Sphingolipid metabolism cooperates with BAK and BAX to promote the mitochondrial pathway of apoptosis. Cell. 148, 988–1000. doi:10.1016/j.cell.2012.01.038

PubMed Abstract | CrossRef Full Text | Google Scholar

Colell, A., Ricci, J. E., Tait, S., Milasta, S., Maurer, U., Bouchier-Hayes, L., et al. (2007). GAPDH and autophagy preserve survival after apoptotic cytochrome c release in the absence of caspase activation. Cell. 129, 983–997. doi:10.1016/j.cell.2007.03.045

PubMed Abstract | CrossRef Full Text | Google Scholar

Danial, N. N., Gramm, C. F., Scorrano, L., Zhang, C. Y., Krauss, S., Ranger, A. M., et al. (2003). BAD and glucokinase reside in a mitochondrial complex that integrates glycolysis and apoptosis. Nature 424, 952–956. doi:10.1038/nature01825

PubMed Abstract | CrossRef Full Text | Google Scholar

Danial, N. N., Walensky, L. D., Zhang, C. Y., Choi, C. S., Fisher, J. K., Molina, A. J. A., et al. (2008). Dual role of proapoptotic BAD in insulin secretion and beta cell survival. Nat. Med. 14, 144–153. doi:10.1038/nm1717

PubMed Abstract | CrossRef Full Text | Google Scholar

Deng, H., Yu, F., Chen, J., Zhao, Y., Xiang, J., and Lin, A. (2008). Phosphorylation of Bad at Thr-201 by JNK1 promotes glycolysis through activation of phosphofructokinase-1. J. Biol. Chem. 283, 20754–20760. doi:10.1074/jbc.M800024200

PubMed Abstract | CrossRef Full Text | Google Scholar

Dixon, S. J., Lemberg, K. M., Lamprecht, M. R., Skouta, R., Zaitsev, E. M., Gleason, C. E., et al. (2012). Ferroptosis: An iron-dependent form of nonapoptotic cell death. Cell. 149, 1060–1072. doi:10.1016/j.cell.2012.03.042

PubMed Abstract | CrossRef Full Text | Google Scholar

Dolgachev, V., Farooqui, M. S., Kulaeva, O. I., Tainsky, M. A., Nagy, B., Hanada, K., et al. (2004). De novo ceramide accumulation due to inhibition of its conversion to complex sphingolipids in apoptotic photosensitized cells. J. Biol. Chem. 279, 23238–23249. doi:10.1074/jbc.M311974200

PubMed Abstract | CrossRef Full Text | Google Scholar

Ekoff, M., Kaufmann, T., Engström, M., Motoyama, N., Villunger, A., Jönsson, J. I., et al. (2007). The BH3-only protein Puma plays an essential role in cytokine deprivation induced apoptosis of mast cells. Blood 110, 3209–3217. doi:10.1182/blood-2007-02-073957

PubMed Abstract | CrossRef Full Text | Google Scholar

Elmore, S. (2007). Apoptosis: A review of programmed cell death. Toxicol. Pathol. 35, 495–516. doi:10.1080/01926230701320337

PubMed Abstract | CrossRef Full Text | Google Scholar

Escudero, S., Zaganjor, E., Lee, S., Mill, C. P., Morgan, A. M., Crawford, E. B., et al. (2018). Dynamic regulation of long-chain fatty acid oxidation by a noncanonical interaction between the MCL-1 BH3 helix and VLCAD. Mol. Cell. 69, 729–743.e7. doi:10.1016/j.molcel.2018.02.005

PubMed Abstract | CrossRef Full Text | Google Scholar

Fawthrop, D. J., Boobis, A. R., and Davies, D. S. (1991). Mechanisms of cell death. Arch. Toxicol. 65, 437–444. doi:10.1007/bf01977355

PubMed Abstract | CrossRef Full Text | Google Scholar

Gama, V., Swahari, V., Schafer, J., Kole, A. J., Evans, A., Huang, Y., et al. (2014). The E3 ligase PARC mediates the degradation of cytosolic cytochrome c to promote survival in neurons and cancer cells. Sci. Signal 7, ra67. doi:10.1126/scisignal.2005309

PubMed Abstract | CrossRef Full Text | Google Scholar

Garufi, A., and D'Orazi, G. (2014). High glucose dephosphorylates serine 46 and inhibits p53 apoptotic activity. J. Exp. Clin. Cancer Res. 33, 79. doi:10.1186/s13046-014-0079-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Giménez-Cassina, A., Garcia-Haro, L., Choi, C. S., Osundiji, M. A., Lane, E. A., Huang, H., et al. (2014). Regulation of hepatic energy metabolism and gluconeogenesis by BAD. Cell. Metab. 19, 272–284. doi:10.1016/j.cmet.2013.12.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Goncalves, M. D., Lu, C., Tutnauer, J., Hartman, T. E., Hwang, S. K., Murphy, C. J., et al. (2019). High-fructose corn syrup enhances intestinal tumor growth in mice. Science 363, 1345–1349. doi:10.1126/science.aat8515

PubMed Abstract | CrossRef Full Text | Google Scholar

Gonzalez, P. S., O'Prey, J., Cardaci, S., Barthet, V. J. A., Sakamaki, J. I., Beaumatin, F., et al. (2018). Mannose impairs tumour growth and enhances chemotherapy. Nature 563, 719–723. doi:10.1038/s41586-018-0729-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Hall, P. A., Coates, P. J., Ansari, B., and Hopwood, D. (1994). Regulation of cell number in the mammalian gastrointestinal tract: The importance of apoptosis. J. Cell. Sci. 107 (12), 3569–3577. doi:10.1242/jcs.107.12.3569

PubMed Abstract | CrossRef Full Text | Google Scholar

Hanahan, D., and Weinberg, R. A. (2011). Hallmarks of cancer: The next generation. Cell. 144, 646–674. doi:10.1016/j.cell.2011.02.013

PubMed Abstract | CrossRef Full Text | Google Scholar

Hawk, M. A., Gorsuch, C. L., Fagan, P., Lee, C., Kim, S. E., Hamann, J. C., et al. (2018). RIPK1-mediated induction of mitophagy compromises the viability of extracellular-matrix-detached cells. Nat. Cell. Biol. 20, 272–284. doi:10.1038/s41556-018-0034-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Horn, H. F., and Vousden, K. H. (2007). Coping with stress: Multiple ways to activate p53. Oncogene 26, 1306–1316. doi:10.1038/sj.onc.1210263

PubMed Abstract | CrossRef Full Text | Google Scholar

Ichim, G., Lopez, J., Ahmed, S. U., Muthalagu, N., Giampazolias, E., Delgado, M. E., et al. (2015). Limited mitochondrial permeabilization causes DNA damage and genomic instability in the absence of cell death. Mol. Cell. 57, 860–872. doi:10.1016/j.molcel.2015.01.018

PubMed Abstract | CrossRef Full Text | Google Scholar

Indo, K., Hoshikawa, H., Kamitori, K., Yamaguchi, F., Mori, T., Tokuda, M., et al. (2014). Effects of D-allose in combination with docetaxel in human head and neck cancer cells. Int. J. Oncol. 45, 2044–2050. doi:10.3892/ijo.2014.2590

PubMed Abstract | CrossRef Full Text | Google Scholar

Jang, C., Hui, S., Lu, W., Cowan, A. J., Morscher, R. J., Lee, G., et al. (2018). The small intestine converts dietary fructose into glucose and organic acids. Cell. Metab. 27, 351–361.e3. doi:10.1016/j.cmet.2017.12.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Jones, R. G., Plas, D. R., Kubek, S., Buzzai, M., Mu, J., Xu, Y., et al. (2005). AMP-activated protein kinase induces a p53-dependent metabolic checkpoint. Mol. Cell. 18, 283–293. doi:10.1016/j.molcel.2005.03.027

PubMed Abstract | CrossRef Full Text | Google Scholar

Kalkavan, H., Chen, M. J., Crawford, J. C., Quarato, G., Fitzgerald, P., Tait, S. W. G., et al. (2022). Sublethal cytochrome c release generates drug-tolerant persister cells. Cell. 185, 3356–3374.e22. doi:10.1016/j.cell.2022.07.025

PubMed Abstract | CrossRef Full Text | Google Scholar

Kerr, J. F., Wyllie, A. H., and Currie, A. R. (1972). Apoptosis: A basic biological phenomenon with wide-ranging implications in tissue kinetics. Br. J. Cancer 26, 239–257. doi:10.1038/bjc.1972.33

PubMed Abstract | CrossRef Full Text | Google Scholar

Kozopas, K. M., Yang, T., Buchan, H. L., Zhou, P., and Craig, R. W. (1993). MCL1, a gene expressed in programmed myeloid cell differentiation, has sequence similarity to BCL2. Proc. Natl. Acad. Sci. U. S. A. 90, 3516–3520. doi:10.1073/pnas.90.8.3516

PubMed Abstract | CrossRef Full Text | Google Scholar

Lafont, E., Milhas, D., Carpentier, S., Garcia, V., Jin, Z. X., Umehara, H., et al. (2010). Caspase-mediated inhibition of sphingomyelin synthesis is involved in FasL-triggered cell death. Cell. Death Differ. 17, 642–654. doi:10.1038/cdd.2009.130

PubMed Abstract | CrossRef Full Text | Google Scholar

Lee, S. M., Kim, J. H., Cho, E. J., and Youn, H. D. (2009). A nucleocytoplasmic malate dehydrogenase regulates p53 transcriptional activity in response to metabolic stress. Cell. Death Differ. 16, 738–748. doi:10.1038/cdd.2009.5

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, P., Nijhawan, D., Budihardjo, I., Srinivasula, S. M., Ahmad, M., Alnemri, E. S., et al. (1997). Cytochrome c and dATP-dependent formation of apaf-1/caspase-9 complex initiates an apoptotic protease cascade. Cell. 91, 479–489. doi:10.1016/S0092-8674(00)80434-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Liberti, M. V., and Locasale, J. W. (2016). The warburg effect: How does it benefit cancer cells? Trends Biochem. Sci. 41, 211–218. doi:10.1016/j.tibs.2015.12.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, X., Kim, C. N., Yang, J., Jemmerson, R., and Wang, X. (1996). Induction of apoptotic program in cell-free extracts: Requirement for dATP and cytochrome c. Cell. 86, 147–157. doi:10.1016/s0092-8674(00)80085-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Luo, W., Hu, H., Chang, R., Zhong, J., Knabel, M., O'Meally, R., et al. (2011). Pyruvate kinase M2 is a PHD3-stimulated coactivator for hypoxia-inducible factor 1. Cell. 145, 732–744. doi:10.1016/j.cell.2011.03.054

PubMed Abstract | CrossRef Full Text | Google Scholar

Mai, W. X., Gosa, L., Daniels, V. W., Ta, L., Tsang, J. E., Higgins, B., et al. (2017). Cytoplasmic p53 couples oncogene-driven glucose metabolism to apoptosis and is a therapeutic target in glioblastoma. Nat. Med. 23, 1342–1351. doi:10.1038/nm.4418

PubMed Abstract | CrossRef Full Text | Google Scholar

Mann, J., Githaka, J. M., Buckland, T. W., Yang, N., Montpetit, R., Patel, N., et al. (2019). Non-canonical BAD activity regulates breast cancer cell and tumor growth via 14-3-3 binding and mitochondrial metabolism. Oncogene 38, 3325–3339. doi:10.1038/s41388-018-0673-6

PubMed Abstract | CrossRef Full Text | Google Scholar

McArthur, K., Whitehead, L. W., Heddleston, J. M., Li, L., Padman, B. S., Oorschot, V., et al. (2018). BAK/BAX macropores facilitate mitochondrial herniation and mtDNA efflux during apoptosis. Science 359, eaao6047. doi:10.1126/science.aao6047

PubMed Abstract | CrossRef Full Text | Google Scholar

McIlwain, D. R., Berger, T., and Mak, T. W. (2013). Caspase functions in cell death and disease. Cold Spring Harb. Perspect. Biol. 5, a008656. doi:10.1101/cshperspect.a008656

PubMed Abstract | CrossRef Full Text | Google Scholar

Miller, M., Craig, J. W., Drucker, W. R., and Woodward, H. (1956). The metabolism of fructose in man. Yale J. Biol. Med. 29, 335–360.

PubMed Abstract | Google Scholar

Mitani, T., Hoshikawa, H., Mori, T., Hosokawa, T., Tsukamoto, I., Yamaguchi, F., et al. (2009). Growth inhibition of head and neck carcinomas by D-allose. Head. Neck 31, 1049–1055. doi:10.1002/hed.21070

PubMed Abstract | CrossRef Full Text | Google Scholar

Moyzis, A. G., Lally, N. S., Liang, W., Najor, R. H., and Gustafsson Å, B. (2022). Mcl-1 differentially regulates autophagy in response to changes in energy status and mitochondrial damage. Cells 11, 1469. doi:10.3390/cells11091469

PubMed Abstract | CrossRef Full Text | Google Scholar

Okawa, Y., Hoshino, A., Ariyoshi, M., Kaimoto, S., Tateishi, S., Ono, K., et al. (2019). Ablation of cardiac TIGAR preserves myocardial energetics and cardiac function in the pressure overload heart failure model. Am. J. Physiol. Heart Circ. Physiol. 316, H1366–h1377. doi:10.1152/ajpheart.00395.2018

PubMed Abstract | CrossRef Full Text | Google Scholar

Okorokov, A. L., and Milner, J. (1999). An ATP/ADP-dependent molecular switch regulates the stability of p53-DNA complexes. Mol. Cell. Biol. 19, 7501–7510. doi:10.1128/mcb.19.11.7501

PubMed Abstract | CrossRef Full Text | Google Scholar

Okoshi, R., Ozaki, T., Yamamoto, H., Ando, K., Koida, N., Ono, S., et al. (2008). Activation of AMP-activated protein kinase induces p53-dependent apoptotic cell death in response to energetic stress. J. Biol. Chem. 283, 3979–3987. doi:10.1074/jbc.M705232200

PubMed Abstract | CrossRef Full Text | Google Scholar

Oltval, Z. N., Milliman, C. L., and Korsmeyer, S. J. (1993). Bcl-2 heterodimerizes in vivo with a conserved homolog, Bax, that accelerates programmed cell death. Cell. 74, 609–619. doi:10.1016/0092-8674(93)90509-O

PubMed Abstract | CrossRef Full Text | Google Scholar

Opferman, J. T., Iwasaki, H., Ong, C. C., Suh, H., Mizuno, S. i., Akashi, K., et al. (2005). Obligate role of anti-apoptotic MCL-1 in the survival of hematopoietic stem cells. Science 307, 1101–1104. doi:10.1126/science.1106114

PubMed Abstract | CrossRef Full Text | Google Scholar

Opferman, J. T., Letai, A., Beard, C., Sorcinelli, M. D., Ong, C. C., and Korsmeyer, S. J. (2003). Development and maintenance of B and T lymphocytes requires antiapoptotic MCL-1. Nature 426, 671–676. doi:10.1038/nature02067

PubMed Abstract | CrossRef Full Text | Google Scholar

Perciavalle, R. M., Stewart, D. P., Koss, B., Lynch, J., Milasta, S., Bathina, M., et al. (2012). Anti-apoptotic MCL-1 localizes to the mitochondrial matrix and couples mitochondrial fusion to respiration. Nat. Cell. Biol. 14, 575–583. doi:10.1038/ncb2488

PubMed Abstract | CrossRef Full Text | Google Scholar

Prew, M. S., Adhikary, U., Choi, D. W., Portero, E. P., Paulo, J. A., Gowda, P., et al. (2022). MCL-1 is a master regulator of cancer dependency on fatty acid oxidation. Cell. Rep. 41, 111445. doi:10.1016/j.celrep.2022.111445

PubMed Abstract | CrossRef Full Text | Google Scholar

Reynolds, J. E., Yang, T., Qian, L., Jenkinson, J. D., Zhou, P., Eastman, A., et al. (1994). Mcl-1, a member of the Bcl-2 family, delays apoptosis induced by c-Myc overexpression in Chinese hamster ovary cells. Cancer Res. 54, 6348–6352.

PubMed Abstract | Google Scholar

Riedl, S. J., and Salvesen, G. S. (2007). The apoptosome: Signalling platform of cell death. Nat. Rev. Mol. Cell. Biol. 8, 405–413. doi:10.1038/nrm2153

PubMed Abstract | CrossRef Full Text | Google Scholar

Roberts, D. J., and Miyamoto, S. (2015). Hexokinase II integrates energy metabolism and cellular protection: Akting on mitochondria and TORCing to autophagy. Cell. Death Differ. 22, 248–257. doi:10.1038/cdd.2014.173

PubMed Abstract | CrossRef Full Text | Google Scholar

Schafer, Z. T., and Kornbluth, S. (2006). The apoptosome: Physiological, developmental, and pathological modes of regulation. Dev. Cell. 10, 549–561. doi:10.1016/j.devcel.2006.04.008

PubMed Abstract | CrossRef Full Text | Google Scholar

Schwartzenberg-Bar-Yoseph, F., Armoni, M., and Karnieli, E. (2004). The tumor suppressor p53 down-regulates glucose transporters GLUT1 and GLUT4 gene expression. Cancer Res. 64, 2627–2633. doi:10.1158/0008-5472.can-03-0846

PubMed Abstract | CrossRef Full Text | Google Scholar

Separovic, D., Hanada, K., Maitah, M. Y. A., Nagy, B., Hang, I., Tainsky, M. A., et al. (2007). Sphingomyelin synthase 1 suppresses ceramide production and apoptosis post-photodamage. Biochem. Biophys. Res. Commun. 358, 196–202. doi:10.1016/j.bbrc.2007.04.095

PubMed Abstract | CrossRef Full Text | Google Scholar

Shi, Y., Chen, J., Weng, C., Chen, R., Zheng, Y., Chen, Q., et al. (2003). Identification of the protein–protein contact site and interaction mode of human VDAC1 with Bcl-2 family proteins. Biochem. Biophysical Res. Commun. 305, 989–996. doi:10.1016/S0006-291X(03)00871-4

CrossRef Full Text | Google Scholar

Stine, Z. E., Schug, Z. T., Salvino, J. M., and Dang, C. V. (2022). Targeting cancer metabolism in the era of precision oncology. Nat. Rev. Drug Discov. 21, 141–162. doi:10.1038/s41573-021-00339-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Surget, S., Khoury, M. P., and Bourdon, J. C. (2013). Uncovering the role of p53 splice variants in human malignancy: A clinical perspective. Onco Targets Ther. 7, 57–68. doi:10.2147/ott.S53876

PubMed Abstract | CrossRef Full Text | Google Scholar

Taha, T. A., Mullen, T. D., and Obeid, L. M. (2006). A house divided: Ceramide, sphingosine, and sphingosine-1-phosphate in programmed cell death. Biochim. Biophys. Acta 1758, 2027–2036. doi:10.1016/j.bbamem.2006.10.018

PubMed Abstract | CrossRef Full Text | Google Scholar

Tait, S. W., Parsons, M. J., Llambi, F., Bouchier-Hayes, L., Connell, S., Muñoz-Pinedo, C., et al. (2010). Resistance to caspase-independent cell death requires persistence of intact mitochondria. Dev. Cell. 18, 802–813. doi:10.1016/j.devcel.2010.03.014

PubMed Abstract | CrossRef Full Text | Google Scholar

Taylor, S. R., Ramsamooj, S., Liang, R. J., Katti, A., Pozovskiy, R., Vasan, N., et al. (2021). Dietary fructose improves intestinal cell survival and nutrient absorption. Nature 597, 263–267. doi:10.1038/s41586-021-03827-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Vaughn, A. E., and Deshmukh, M. (2008). Glucose metabolism inhibits apoptosis in neurons and cancer cells by redox inactivation of cytochrome c. Nat. Cell. Biol. 10, 1477–1483. doi:10.1038/ncb1807

PubMed Abstract | CrossRef Full Text | Google Scholar

von Haefen, C., Wieder, T., Gillissen, B., Stärck, L., Graupner, V., Dörken, B., et al. (2002). Ceramide induces mitochondrial activation and apoptosis via a Bax-dependent pathway in human carcinoma cells. Oncogene 21, 4009–4019. doi:10.1038/sj.onc.1205497

PubMed Abstract | CrossRef Full Text | Google Scholar

Warburg, O. (1956). On the origin of cancer cells. Science 123, 309–314. doi:10.1126/science.123.3191.309

PubMed Abstract | CrossRef Full Text | Google Scholar

Wei, G., Twomey, D., Lamb, J., Schlis, K., Agarwal, J., Stam, R. W., et al. (2006). Gene expression-based chemical genomics identifies rapamycin as a modulator of MCL1 and glucocorticoid resistance. Cancer Cell. 10, 331–342. doi:10.1016/j.ccr.2006.09.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Westphal, D., Dewson, G., Czabotar, P. E., and Kluck, R. M. (2011). Molecular biology of Bax and Bak activation and action. Biochimica Biophysica Acta (BBA) - Mol. Cell. Res. 1813, 521–531. doi:10.1016/j.bbamcr.2010.12.019

CrossRef Full Text | Google Scholar

Wuillème-Toumi, S., Robillard, N., Gomez, P., Moreau, P., Le Gouill, S., Avet-Loiseau, H., et al. (2005). Mcl-1 is overexpressed in multiple myeloma and associated with relapse and shorter survival. Leukemia 19, 1248–1252. doi:10.1038/sj.leu.2403784

PubMed Abstract | CrossRef Full Text | Google Scholar

Xu, H. L., Zhou, X., Chen, S., Xu, S., Li, Z., Nakanishi, H., et al. (2023). Rare sugar L-sorbose exerts antitumor activity by impairing glucose metabolism. Commun. Biol. 6, 259. doi:10.1038/s42003-023-04638-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Xue, D., and Robert Horvitz, H. (1995). Inhibition of the Caenorhabditis elegans cell-death protease CED-3 by a CED-3 cleavage site in baculovirus p35 protein. Nature 377, 248–251. doi:10.1038/377248a0

PubMed Abstract | CrossRef Full Text | Google Scholar

Yang, E., Zha, J., Jockel, J., Boise, L. H., Thompson, C. B., and Korsmeyer, S. J. (1995). Bad, a heterodimeric partner for Bcl-XL and Bcl-2, displaces Bax and promotes cell death. Cell. 80, 285–291. doi:10.1016/0092-8674(95)90411-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Yu, H.-P., Xie, J. M., Li, B., Sun, Y. H., Gao, Q. G., Ding, Z. H., et al. (2015). TIGAR regulates DNA damage and repair through pentosephosphate pathway and Cdk5-ATM pathway. Sci. Rep. 5, 9853. doi:10.1038/srep09853

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhao, Y., Coloff, J. L., Ferguson, E. C., Jacobs, S. R., Cui, K., and Rathmell, J. C. (2008). Glucose metabolism attenuates p53 and Puma-dependent cell death upon growth factor deprivation. J. Biol. Chem. 283, 36344–36353. doi:10.1074/jbc.M803580200

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhou, P., Qian, L., Kozopas, K. M., and Craig, R. W. (1997). Mcl-1, a Bcl-2 family member, delays the death of hematopoietic cells under a variety of apoptosis-inducing conditions. Blood 89, 630–643. doi:10.1182/blood.v89.2.630

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: apoptosis, metabolism, caspases, mitochondria, MOMP

Citation: Schofield JH and Schafer ZT (2023) Regulators mount up: the metabolic roles of apoptotic proteins. Front. Cell. Death 2:1223926. doi: 10.3389/fceld.2023.1223926

Received: 16 May 2023; Accepted: 19 June 2023;
Published: 03 July 2023.

Edited by:

Olivier Micheau, Institut National de la Santé et de la Recherche Médicale (INSERM), France

Reviewed by:

Nieves Peltzer, University of Cologne, Germany
Atan Gross, Weizmann Institute of Science, Israel

Copyright © 2023 Schofield and Schafer. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Zachary T. Schafer, zschafe1@nd.edu

Download