Next Article in Journal
Patient-Derived Glioma Models: From Patients to Dish to Animals
Next Article in Special Issue
The Deubiquitinating Enzyme Inhibitor PR-619 Enhances the Cytotoxicity of Cisplatin via the Suppression of Anti-Apoptotic Bcl-2 Protein: In Vitro and In Vivo Study
Previous Article in Journal
Photosensitizer Activation Drives Apoptosis by Interorganellar Ca2+ Transfer and Superoxide Production in Bystander Cancer Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Perspective

The Incomplete Puzzle of the BCL2 Proteins

by
Hector Flores-Romero
and
Ana J. García-Sáez
*
Interfaculty Institute of Biochemistry, Eberhard-Karls-Universität Tübingen, 72076 Tübingen, Germany
*
Author to whom correspondence should be addressed.
Cells 2019, 8(10), 1176; https://doi.org/10.3390/cells8101176
Submission received: 3 September 2019 / Revised: 25 September 2019 / Accepted: 26 September 2019 / Published: 29 September 2019
(This article belongs to the Special Issue Regulation of Apoptosis by the Bcl-2 Family of Proteins)

Abstract

:
The proteins of the BCL2 family are key players in multiple cellular processes, chief amongst them being the regulation of mitochondrial integrity and apoptotic cell death. These proteins establish an intricate interaction network that expands both the cytosol and the surface of organelles to dictate the cell fate. The complexity and unpredictability of the BCL2 interactome resides in the large number of family members and of interaction surfaces, as well as on their different behaviours in solution and in the membrane. Although our current structural knowledge of the BCL2 proteins has been proven therapeutically relevant, the precise structure of membrane-bound complexes and the regulatory effect that membrane lipids exert over these proteins remain key questions in the field. Here, we discuss the complexity of BCL2 interactome, the new insights, and the black matter in the field.

Perspective in BCL2 Universe

The proteins of the BCL2 family are the main regulators of the intrinsic apoptotic pathway and constitute a fundamental part of tumorigenic cell dismissal and cancer treatment effectiveness [1,2]. Apoptosis effectors, or BAX-type proteins, are the most effective removers of damaged cells, while their antiapoptotic counterparts, or BCL2-type proteins, inhibit apoptotic cell death and play a role in chemotherapeutic resistance [3,4]. The opposing forces between BAX- and BCL2-type proteins are tuned by the so-called BH3-only proteins, a third subgroup of this family of proteins that promotes apoptosis by activating BAX-type proteins and/or blocking antiapoptotic proteins [5].
The proteins of the BCL2 family interact with each other by a BH3-into-groove mechanism, where the BH3 domain of one protomer binds to the hydrophobic groove of another protomer, thereby forming homo and heterodimers to control their apoptotic function. Under this premise several models have emerged, with differences in binding affinities amongst subgroups and the relevancy of the membrane environment. These models propose that antiapoptotic proteins repress apoptosis neutralizing either BH3-only activators (direct model, MODE 1) or BAX-type proteins (indirect model, MODE 2) [6,7,8,9,10,11,12,13,14]. In addition, retrotranslocation or inhibition MODE 0 postulates that BCL2-type proteins inhibit apoptosis by keeping BAX-type proteins inactive through continuous retrotranslocation from the mitochondrial surface into the cytosol [15,16,17,18,19]. These models, however, do not consider an enigmatic property shared by all BCL2-type proteins, which is their ability to promote, rather than inhibit, apoptosis under specific conditions (PRODEATH MODE) [15,20,21,22]. The complex interaction network that orchestrates these proteins’ actions is commonly termed the BCL2 interactome, which constitutes an intricate puzzle yet unresolved (Figure 1).
There are multiple reasons why, in spite of the pieces of this puzzle being defined long ago, it remains impossible to unequivocally model BCL2-mediated cell fate. First, BCL2 proteins perform their function, that is, regulating mitochondrial outer membrane (MOM) permeabilization (MOMP) to release apoptogenic factors into the cytosol and therefore induce apoptosis, when targeted to the membrane. The membrane and its constituting lipids affect the pieces of the BCL2 puzzle by modulating the affinities between the different family members [23,24] or by altering their canonical phenotype or function, for example switching their antiapoptotic nature to proapoptotic activity [15]. In addition to the membrane environment, posttranslational modifications have also been shown to modify the affinity and function of the BCL2 family members [13,25,26]. These modifications include phosphorylation, proteolytic cleavage, ubiquitination, and proteosomal degradation [14,25]. Second, despite common consensus on the importance of the BH3:groove in mediating the interaction between proteins, additional non-canonical surfaces exist that regulate the BCL2 interactome (e.g., rear binding site, N-terminal alpha helix 1 and tail anchoring domain) [27,28,29,30,31]. Third, these proteins can assemble into defined supramolecular structures that expand their role in cell death beyond cytochrome c release [32,33]. Indeed, the oligomeric apoptosis effectors BAX and BAK are able to mediate mitochondrial DNA (mtDNA) release and an immunological response [34,35,36]. Finally, some of the BCL2 puzzle’s pieces participate in other cellular puzzles, as BCL2 proteins are reported to have many other functions which are not directly related to apoptotic cell death. For example, these proteins elicit critical roles in normal cell physiology related to metabolism, mitophagy, mitochondrial dynamics and energetics, and calcium homeostasis, amongst others [37,38,39,40].
Years of extensive research efforts have shed light on important mechanistic and structural details of the BCL2 family proteins [41,42,43]. Understanding the atomic structure and interaction network of these proteins has provided fundamental opportunities for the rational design of drugs that specifically target them. These compounds, commonly known as BH3 mimetics, are molecules based on the BH3 domain of BH3-only proteins designed to interact with specific BCL2 family members [44]. BH3 mimetics exhibit enhanced lethal activity in primed cells, which contain high levels of antiapoptotic and proapoptotic effectors [44,45,46]. Chief amongst them is Venetoclax (or ABT-199); based on the BH3 only protein BAD, this compound efficiently neutralizes BCL2, thereby leading to BAX/BAK activation and to MOMP to induce apoptosis [47,48]. Although this drug has been recently approved to treat chronic lymphoid leukaemia (CLL), acute myeloid leukaemia (AML) and small lymphocytic lymphoma (SLL) [47,48,49,50,51,52,53,54], its applicability for cancer treatment is limited and chemotherapy still remains the most frequent alternative [55]. There are many possible explanations for the partial efficiency of BH3 mimetics, including cancer heterogeneity, the lack on specific BH3-mimetics optimized for the different BCL2 members governing cell death resistance in the tumor, and mutations or posttranslational modification in the BCL2 family members that alter their canonical function and structure [45,49,55]. On the other hand, these drugs were designed based on solution studies where the regulatory role of the membrane is neglected. Moreover, some BH3-only proteins like BIMs, a shorter isoform of the BH3-only protein BIM, are reported to kill mainly due to their membrane targeting, rather than due to interaction with the antiapoptotic family members [56].
The exact pattern of interactions comprising the BCL2 interactome and the precise structure of the membrane-bound complexes, particularly considering the MOM environment, remains controversial [14,57,58]. Regarding the proapoptotic effectors BAX and BAK, there is solid evidence suggesting that their active conformations arrange into toroidal pores of proteo-lipidic nature and tunable size [59,60,61,62]. The structural reorganization driving BAX-type proteins from the inactive to the fully activated conformation at the MOM is considered the “holy grail” of apoptosis research [63]. These events are usually divided into: (i) early activation steps; involving TM dislodgement and N terminal exposure [64,65,66], (ii) BH3 domain exposure, which occurs due to BAX/BAK reorganization in two different parts (dimerization and piercing domains) [5,67,68,69,70], (iii) oligomerization and redistribution into apoptotic foci [33,70,71,72,73] and (iv) pore formation [59,74] (Figure 2). Importantly, these events are regulated, at least partially, by mitochondrial membranes. Although there is literature describing the topology of active BAX/BAK in the membrane [70,71,73], we still fail to understand the contribution of mitochondrial lipids in modulating their activation, oligomerization and formation of supramolecular structures at apoptotic foci during and after MOMP. BAX was recently reported to induce mtDNA release [34,36]. This renders mitochondrial apoptosis an unexpected immunological relevance, which changes the current paradigm and expands the horizons of BCL2-based therapeutics.
Concerning the antiapoptotic members of the BCL2 family, membrane lipid composition can enhance their binding affinity for the proapoptotic members [24,75] or ablate their inhibition capacity and release a hidden pore forming activity [15,20,21,76]. Although the transition of antiapoptotic BCL2 members to pro-death molecules remains poorly understood, the therapeutic potential of this phenotypic reversion should not be neglected, given that their overexpression is key in promoting resistance to chemotherapy. The membrane permeabilizing activity of BCL2-type proteins has similarities and differences to that of BAX-type proteins. Structurally, the pores formed by BCL2-type proteins are smaller, and do not require canonical BH3:groove interactions for oligomerization and pore opening [15,20,77]. Similarly to BAX-type proteins, the amphipathic alpha helix 5 of some antiapoptotic member has been reported to mediate their membrane-permeabilizing function [15,20]. Mechanistically, stimulation of the phenotypic reversion of BCL2- type proteins is diverse, including changes in pH, caspase and µcalpain cleavage, and membrane lipid composition amongst others [15,76,78,79].
Particularly, the mitochondrion-specific lipid cardiolipin (CL) has been postulated as a key regulatory element in BCL2 protein activity [15,24]. CL is implicated in many mitochondrial functions such as normal organelle ultrastructure, mitochondrial dynamics, energy metabolism and apoptosis [80,81]. Indeed, different lines of evidence indicate that the net content of CL at the MOM increases during apoptosis [82,83]. Because of its unique structural properties (e.g., two negative charges, a relatively small head group and four acyl chains), CL can form highly-curved inverted hexagonal structures [84,85,86] and laterally segregate into defined nanodomains [87,88]. These elements support the concept that CL potentially creates a unique environment for BCL2 family proteins and promotes mitochondrial membrane alterations that facilitate bilayer structure remodeling, deformation, and ultimately permeabilization. Moreover, the peroxidized isoform of CL (CLox) weakens the interaction of cytochrome c with the MIM, a process that may also contribute to ease MOMP [82,89,90].
Beyond their role in cell death, BCL2 family proteins participate in several cellular processes, including the regulation of mitochondrial dynamics [91]. New insights into the link between shape and function of mitochondria in health and disease (mitopathology) is beginning to unravel on several fronts [92]. A new connection between mitochondrial dynamics and not only cellular metabolism but also cell fate pathways may emerge from the intersection of BCL2 family proteins and mitochondrial reshaping machinery [91]. In vertebrates, the fundamental protein for mitochondrial fission is a large GTPase termed dynamin-related protein 1 (DRP1) [93]. The localization of DRP1 at constriction points to induce membrane fission is not random, but it seems to be mainly associated with MERCS (mitochondria ER contact sites) and to colocalize to apoptotic foci with the proapoptotic effectors BAX/BAK [94,95]. BCL2 proteins have been also related to Mitofusins 1 and 2 (MFN1/2), dynamin-like proteins involved in mitochondrial fusion [96]. Finally, mitochondrial cristae remodeling appears to be a fundamental step for the BAX-induced differential release of apoptotic factors at the apoptotic foci [34,36,97]. OPA1 is a key regulator of mitochondrial cristae remodeling [98], and its function appears to be regulated by the BH3 only protein tBID [99,100,101]. Thus, it is conceivable that BCL2 family proteins can elicit a direct regulatory effect over mitochondrial dynamics with diverse effects on cell death and survival.
All in all, it is striking that, after more than 30 years of BCL2 research, it is still unclear how these proteins behave specially at the membrane and how we could efficiently guide them in death and disease. Although many important mechanistic details have been uncovered during these years, the puzzle remains challenging to complete. However, we should not forget that therapeutic-regulation of apoptosis particularly by modulating the BCL2 interactome has strong therapeutic potential to combat human disease, including cancer and neurodegenerative disorders. Indeed, the design of Venetoclax based on BCL2 knowledge is the best evidence that a treatment targeting apoptotic proteins can get us closer to curing cancer. In spite of this, BCL2 regulation and drug targeting at the mitochondrial membrane remain intangible. Mitochondrial lipids regulate BCL2 proteins, both indirectly by changing the mechanical properties of the membrane, or directly by specifically modulating protein targeting, structure and function. Therefore, the understanding of BCL2 action in the membrane context appears to be compulsory, particularly in the light of two recent activities unveiled by these proteins that occur in the membrane, supramolecular organization into defined structures and mtDNA release [32,33,34,36]. In this context, it remains to be understood if the phenotypically-reverted antiapoptotic proteins share these activities with the proapoptotic effector BAX. On the other hand, function and abundance of BCL2 proteins are influenced by posttranslational modifications. As many of these modifications are governed by enzymes, their modulation could be efficiently achieved using small molecules, a suitable scenario for drug design and therapy. Moreover, there is a growing body of evidence suggesting that BCL2 proteins regulate metabolism and mitochondrial function, which are dysregulated in many disease pathologies [39,92,102]. The role of BCL2 in mitopathology, or mitochondria-related diseases, also provides new therapeutic opportunities [92,102]. Finally, as BCL2 proteins are all highly overexpressed in cancers, they represent prime candidates as antigens for anti-cancer therapy. Importantly, cellular immune responses against the BCL2 family proteins have been reported as common features in cancer patients, highlighting that these proteins are natural targets for the immune system and tumor microenvironment [103,104]. Taken together, comprehensive knowledge of the BCL2 family proteins is a highly reliable option to rationally design specific treatments that can cure on demand, alone or combined, adjusted depending on the specific BCL2 profile of patients. As BCL2 family proteins are reported to mediate many cellular processes in healthy and pathological situations, their targeting holds the potential to be unmatched.

Acknowledgments

We thank John S. Danial for carefully reading our manuscript and for his comments and Uris Ros for her help in figure design.

Funding

This work in our lab is supported by the European Research Council (APOSITE ERC-Co 817758) and by the Deutsche Forschungsgemeinschaft (GA1641/2-2).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kalkavan, H.; Green, D.R. MOMP, cell suicide as a BCL-2 family business. Cell Death Differ. 2017. [Google Scholar] [CrossRef] [PubMed]
  2. Delbridge, A.R.; Grabow, S.; Strasser, A.; Vaux, D.L. Thirty years of BCL-2: Translating cell death discoveries into novel cancer therapies. Nat. Rev. Cancer 2016, 16, 99–109. [Google Scholar] [CrossRef] [PubMed]
  3. Kontos, C.K.; Avgeris, M.; Vassilacopoulou, D.; Ardavanis, A.; Scorilas, A. Molecular Effects of Treatment of Human Colorectal Cancer Cells with Natural and Classical Chemotherapeutic Drugs: Alterations in the Expression of Apoptosis-related BCL2 Family Members, Including BCL2L12. Curr. Pharm. Biotechnol. 2018, 19, 1064–1075. [Google Scholar] [CrossRef] [PubMed]
  4. Kontos, C.K.; Christodoulou, M.I.; Scorilas, A. Apoptosis-related BCL2-family members: Key players in chemotherapy. Anticancer Agents Med. Chem. 2014, 14, 353–374. [Google Scholar] [CrossRef] [PubMed]
  5. Moldoveanu, T.; Grace, C.R.; Llambi, F.; Nourse, A.; Fitzgerald, P.; Gehring, K.; Kriwacki, R.W.; Green, D.R. BID-induced structural changes in BAK promote apoptosis. Nat. Struct. Mol. Biol. 2013, 20, 589–597. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Fletcher, J.I.; Meusburger, S.; Hawkins, C.J.; Riglar, D.T.; Lee, E.F.; Fairlie, W.D.; Huang, D.C.; Adams, J.M. Apoptosis is triggered when prosurvival Bcl-2 proteins cannot restrain Bax. Proc. Natl. Acad. Sci. USA 2008, 105, 18081–18087. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Willis, S.N.; Fletcher, J.I.; Kaufmann, T.; van Delft, M.F.; Chen, L.; Czabotar, P.E.; Ierino, H.; Lee, E.F.; Fairlie, W.D.; Bouillet, P.; et al. Apoptosis initiated when BH3 ligands engage multiple Bcl-2 homologs, not Bax or Bak. Science 2007, 315, 856–859. [Google Scholar] [CrossRef]
  8. Chen, L.; Willis, S.N.; Wei, A.; Smith, B.J.; Fletcher, J.I.; Hinds, M.G.; Colman, P.M.; Day, C.L.; Adams, J.M.; Huang, D.C. Differential targeting of prosurvival Bcl-2 proteins by their BH3-only ligands allows complementary apoptotic function. Mol. Cell 2005, 17, 393–403. [Google Scholar] [CrossRef]
  9. Chen, H.C.; Kanai, M.; Inoue-Yamauchi, A.; Tu, H.C.; Huang, Y.; Ren, D.; Kim, H.; Takeda, S.; Reyna, D.E.; Chan, P.M.; et al. An interconnected hierarchical model of cell death regulation by the BCL-2 family. Nat. Cell Biol. 2015, 17, 1270–1281. [Google Scholar] [CrossRef] [Green Version]
  10. Leber, B.; Lin, J.; Andrews, D.W. Embedded together: The life and death consequences of interaction of the Bcl-2 family with membranes. Apoptosis Int. J. Program. Cell Death 2007, 12, 897–911. [Google Scholar] [CrossRef] [Green Version]
  11. Llambi, F.; Moldoveanu, T.; Tait, S.W.; Bouchier-Hayes, L.; Temirov, J.; McCormick, L.L.; Dillon, C.P.; Green, D.R. A unified model of mammalian BCL-2 protein family interactions at the mitochondria. Mol. Cell 2011, 44, 517–531. [Google Scholar] [CrossRef] [PubMed]
  12. Basu, A.; DuBois, G.; Haldar, S. Posttranslational modifications of Bcl2 family members—A potential therapeutic target for human malignancy. Front. Biosci. J. Virtual Libr. 2006, 11, 1508–1521. [Google Scholar] [CrossRef] [PubMed]
  13. Kale, J.; Kutuk, O.; Brito, G.C.; Andrews, T.S.; Leber, B.; Letai, A.; Andrews, D.W. Phosphorylation switches Bax from promoting to inhibiting apoptosis thereby increasing drug resistance. EMBO Rep. 2018, 19, e45235. [Google Scholar] [CrossRef] [PubMed]
  14. Kale, J.; Osterlund, E.J.; Andrews, D.W. BCL-2 family proteins: Changing partners in the dance towards death. Cell Death Differ. 2018, 25, 65–80. [Google Scholar] [CrossRef]
  15. Flores-Romero, H.; Landeta, O.; Ugarte-Uribe, B.; Cosentino, K.; Garcia-Porras, M.; Garcia-Saez, A.J.; Basanez, G. BFL1 modulates apoptosis at the membrane level through a bifunctional and multimodal mechanism showing key differences with BCLXL. Cell Death Differ. 2018, 1880–1894. [Google Scholar] [CrossRef]
  16. Edlich, F.; Banerjee, S.; Suzuki, M.; Cleland, M.M.; Arnoult, D.; Wang, C.X.; Neutzner, A.; Tjandra, N.; Youle, R.J. Bcl-x(L) Retrotranslocates Bax from the Mitochondria into the Cytosol. Cell 2011, 145, 104–116. [Google Scholar] [CrossRef]
  17. Todt, F.; Cakir, Z.; Reichenbach, F.; Emschermann, F.; Lauterwasser, J.; Kaiser, A.; Ichim, G.; Tait, S.W.; Frank, S.; Langer, H.F.; et al. Differential retrotranslocation of mitochondrial Bax and Bak. EMBO J. 2015, 34, 67–80. [Google Scholar] [CrossRef]
  18. Edlich, F. The great migration of Bax and Bak. Mol. Cell. Oncol. 2015, 2, e995029. [Google Scholar] [CrossRef] [Green Version]
  19. Schellenberg, B.; Wang, P.; Keeble, J.A.; Rodriguez-Enriquez, R.; Walker, S.; Owens, T.W.; Foster, F.; Tanianis-Hughes, J.; Brennan, K.; Streuli, C.H.; et al. Bax exists in a dynamic equilibrium between the cytosol and mitochondria to control apoptotic priming. Mol. Cell 2013, 49, 959–971. [Google Scholar] [CrossRef]
  20. Landeta, O.; Garcia Valero, J.; Flores-Romero, H.; Bustillo-Zabalbeitia, I.; Landajuela, A.; Garcia-Porras, M.; Terrones, O.; Basanez, G. Lipid-dependent bimodal MCL1 membrane activity. ACS Chem. Biol. 2014, 9, 2852–2863. [Google Scholar] [CrossRef]
  21. Basanez, G.; Zhang, J.; Chau, B.N.; Maksaev, G.I.; Frolov, V.A.; Brandt, T.A.; Burch, J.; Hardwick, J.M.; Zimmerberg, J. Pro-apoptotic cleavage products of Bcl-xL form cytochrome c-conducting pores in pure lipid membranes. J. Biol. Chem. 2001, 276, 31083–31091. [Google Scholar] [CrossRef] [PubMed]
  22. Lin, B.; Kolluri, S.K.; Lin, F.; Liu, W.; Han, Y.H.; Cao, X.; Dawson, M.I.; Reed, J.C.; Zhang, X.K. Conversion of Bcl-2 from protector to killer by interaction with nuclear orphan receptor Nur77/TR3. Cell 2004, 116, 527–540. [Google Scholar] [CrossRef]
  23. Garcia-Saez, A.J.; Ries, J.; Orzaez, M.; Perez-Paya, E.; Schwille, P. Membrane promotes tBID interaction with BCL(XL). Nat. Struct. Mol. Biol. 2009, 16, 1178–1185. [Google Scholar] [CrossRef] [PubMed]
  24. Bleicken, S.; Hantusch, A.; Das, K.K.; Frickey, T.; Garcia-Saez, A.J. Quantitative interactome of a membrane Bcl-2 network identifies a hierarchy of complexes for apoptosis regulation. Nat. Commun. 2017, 8, 73. [Google Scholar] [CrossRef]
  25. Kutuk, O.; Letai, A. Regulation of Bcl-2 family proteins by posttranslational modifications. Curr. Mol. Med. 2008, 8, 102–118. [Google Scholar]
  26. Cui, J.; Placzek, W.J. Post-Transcriptional Regulation of Anti-Apoptotic BCL2 Family Members. Int. J. Mol. Sci. 2018, 19, 308. [Google Scholar] [CrossRef]
  27. Barclay, L.A.; Wales, T.E.; Garner, T.P.; Wachter, F.; Lee, S.; Guerra, R.M.; Stewart, M.L.; Braun, C.R.; Bird, G.H.; Gavathiotis, E.; et al. Inhibition of Pro-apoptotic BAX by a noncanonical interaction mechanism. Mol. Cell 2015, 57, 873–886. [Google Scholar] [CrossRef]
  28. Gavathiotis, E.; Suzuki, M.; Davis, M.L.; Pitter, K.; Bird, G.H.; Katz, S.G.; Tu, H.C.; Kim, H.; Cheng, E.H.; Tjandra, N.; et al. BAX activation is initiated at a novel interaction site. Nature 2008, 455, 1076–1081. [Google Scholar] [CrossRef] [Green Version]
  29. Vasquez-Montes, V.; Vargas-Uribe, M.; Pandey, N.K.; Rodnin, M.V.; Langen, R.; Ladokhin, A.S. Lipid-modulation of membrane insertion and refolding of the apoptotic inhibitor Bcl-xL. Biochim. Biophys. Acta Proteins Proteom. 2019, 1867, 691–700. [Google Scholar] [CrossRef]
  30. Todt, F.; Cakir, Z.; Reichenbach, F.; Youle, R.J.; Edlich, F. The C-terminal helix of Bcl-x(L) mediates Bax retrotranslocation from the mitochondria. Cell Death Diff. 2013, 20, 333–342. [Google Scholar] [CrossRef]
  31. Andreu-Fernandez, V.; Sancho, M.; Genoves, A.; Lucendo, E.; Todt, F.; Lauterwasser, J.; Funk, K.; Jahreis, G.; Perez-Paya, E.; Mingarro, I.; et al. Bax transmembrane domain interacts with prosurvival Bcl-2 proteins in biological membranes. Proc. Natl. Acad. Sci. USA 2017, 114, 310–315. [Google Scholar] [CrossRef] [PubMed]
  32. Grosse, L.; Wurm, C.A.; Bruser, C.; Neumann, D.; Jans, D.C.; Jakobs, S. Bax assembles into large ring-like structures remodeling the mitochondrial outer membrane in apoptosis. EMBO J. 2016, 35, 402–413. [Google Scholar] [CrossRef] [PubMed]
  33. Salvador-Gallego, R.; Mund, M.; Cosentino, K.; Schneider, J.; Unsay, J.; Schraermeyer, U.; Engelhardt, J.; Ries, J.; Garcia-Saez, A.J. Bax assembly into rings and arcs in apoptotic mitochondria is linked to membrane pores. EMBO J. 2016, 35, 389–401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. McArthur, K.; Whitehead, L.W.; Heddleston, J.M.; Li, L.; Padman, B.S.; Oorschot, V.; Geoghegan, N.D.; Chappaz, S.; Davidson, S.; San Chin, H.; et al. BAK/BAX macropores facilitate mitochondrial herniation and mtDNA efflux during apoptosis. Science 2018, 359. [Google Scholar] [CrossRef] [PubMed]
  35. Flores-Romero, H.; Garcia-Saez, A.J. MAVS-induced mitochondrial membrane remodeling. FEBS J. 2019, 286, 1540–1542. [Google Scholar] [CrossRef]
  36. Riley, J.S.; Quarato, G.; Cloix, C.; Lopez, J.; O’Prey, J.; Pearson, M.; Chapman, J.; Sesaki, H.; Carlin, L.M.; Passos, J.F.; et al. Mitochondrial inner membrane permeabilisation enables mtDNA release during apoptosis. EMBO J. 2018, 37, e99238. [Google Scholar] [CrossRef]
  37. Pattingre, S.; Tassa, A.; Qu, X.; Garuti, R.; Liang, X.H.; Mizushima, N.; Packer, M.; Schneider, M.D.; Levine, B. Bcl-2 antiapoptotic proteins inhibit Beclin 1-dependent autophagy. Cell 2005, 122, 927–939. [Google Scholar] [CrossRef]
  38. Hardwick, J.M.; Soane, L. Multiple functions of BCL-2 family proteins. Cold Spring Harb. Perspect. Biol. 2013, 5. [Google Scholar] [CrossRef]
  39. Srivastava, R.; Cao, Z.; Nedeva, C.; Naim, S.; Bachmann, D.; Rabachini, T.; Gangoda, L.; Shahi, S.; Glab, J.; Menassa, J.; et al. BCL-2 family protein BOK is a positive regulator of uridine metabolism in mammals. Proc. Natl. Acad. Sci. USA 2019, 116, 15469–15474. [Google Scholar] [CrossRef] [Green Version]
  40. Schulman, J.J.; Szczesniak, L.M.; Bunker, E.N.; Nelson, H.A.; Roe, M.W.; Wagner, L.E., 2nd; Yule, D.I.; Wojcikiewicz, R.J.H. Bok regulates mitochondrial fusion and morphology. Cell Death Diff. 2019, 826–836. [Google Scholar] [CrossRef]
  41. Suzuki, M.; Youle, R.J.; Tjandra, N. Structure of Bax: Coregulation of dimer formation and intracellular localization. Cell 2000, 103, 645–654. [Google Scholar] [CrossRef]
  42. Muchmore, S.W.; Sattler, M.; Liang, H.; Meadows, R.P.; Harlan, J.E.; Yoon, H.S.; Nettesheim, D.; Chang, B.S.; Thompson, C.B.; Wong, S.L.; et al. X-ray and NMR structure of human Bcl-xL, an inhibitor of programmed cell death. Nature 1996, 381, 335–341. [Google Scholar] [CrossRef] [PubMed]
  43. Chou, J.J.; Li, H.; Salvesen, G.S.; Yuan, J.; Wagner, G. Solution structure of BID, an intracellular amplifier of apoptotic signaling. Cell 1999, 96, 615–624. [Google Scholar] [CrossRef]
  44. Opydo-Chanek, M.; Gonzalo, O.; Marzo, I. Multifaceted anticancer activity of BH3 mimetics: Current evidence and future prospects. Biochem. Pharmacol. 2017, 136, 12–23. [Google Scholar] [CrossRef] [PubMed]
  45. Kotschy, A.; Szlavik, Z.; Murray, J.; Davidson, J.; Maragno, A.L.; Le Toumelin-Braizat, G.; Chanrion, M.; Kelly, G.L.; Gong, J.N.; Moujalled, D.M.; et al. The MCL1 inhibitor S63845 is tolerable and effective in diverse cancer models. Nature 2016, 538, 477–482. [Google Scholar] [CrossRef] [PubMed]
  46. Villalobos-Ortiz, M.; Ryan, J.; Mashaka, T.N.; Opferman, J.T.; Letai, A. BH3 profiling discriminates on-target small molecule BH3 mimetics from putative mimetics. Cell Death Differ. 2019, 391–402. [Google Scholar] [CrossRef]
  47. Souers, A.J.; Leverson, J.D.; Boghaert, E.R.; Ackler, S.L.; Catron, N.D.; Chen, J.; Dayton, B.D.; Ding, H.; Enschede, S.H.; Fairbrother, W.J.; et al. ABT-199, a potent and selective BCL-2 inhibitor, achieves antitumor activity while sparing platelets. Nat. Med. 2013, 19, 202–208. [Google Scholar] [CrossRef]
  48. Zigart, N.; Casar, Z. A literature review of the patent publications on venetoclax - a selective Bcl-2 inhibitor: Discovering the therapeutic potential of a novel chemotherapeutic agent. Expert Opin. Ther. Pat. 2019, 29, 487–496. [Google Scholar] [CrossRef]
  49. DiNardo, C.D.; Pratz, K.; Pullarkat, V.; Jonas, B.A.; Arellano, M.; Becker, P.S.; Frankfurt, O.; Konopleva, M.; Wei, A.H.; Kantarjian, H.M.; et al. Venetoclax combined with decitabine or azacitidine in treatment-naive, elderly patients with acute myeloid leukemia. Blood 2019, 133, 7–17. [Google Scholar] [CrossRef] [Green Version]
  50. Roberts, A.W.; Davids, M.S.; Pagel, J.M.; Kahl, B.S.; Puvvada, S.D.; Gerecitano, J.F.; Kipps, T.J.; Anderson, M.A.; Brown, J.R.; Gressick, L.; et al. Targeting BCL2 with Venetoclax in Relapsed Chronic Lymphocytic Leukemia. N. Engl. J. Med. 2016, 374, 311–322. [Google Scholar] [CrossRef]
  51. Korycka-Wolowiec, A.; Wolowiec, D.; Kubiak-Mlonka, A.; Robak, T. Venetoclax in the treatment of chronic lymphocytic leukemia. Expert Opin. Drug Metab. Toxicol. Expert. 2019, 15, 353–366. [Google Scholar] [CrossRef] [PubMed]
  52. Ashkenazi, A.; Fairbrother, W.J.; Leverson, J.D.; Souers, A.J. From basic apoptosis discoveries to advanced selective BCL-2 family inhibitors. Nat. Rev. Drug Discov. 2017, 16, 273–284. [Google Scholar] [CrossRef] [PubMed]
  53. Borg, M.A.; Clemmons, A. Venetoclax: A Novel Treatment for Patients With del(17p) Chronic Lymphocytic Leukemia. J. Adv. Pract. Oncol. 2017, 8, 647–652. [Google Scholar] [PubMed]
  54. Gentile, M.; Petrungaro, A.; Uccello, G.; Vigna, E.; Recchia, A.G.; Caruso, N.; Bossio, S.; De Stefano, L.; Palummo, A.; Storino, F.; et al. Venetoclax for the treatment of chronic lymphocytic leukemia. Expert Opin. Investig. Drugs 2017, 26, 1307–1316. [Google Scholar] [CrossRef] [PubMed]
  55. Yecies, D.; Carlson, N.E.; Deng, J.; Letai, A. Acquired resistance to ABT-737 in lymphoma cells that up-regulate MCL-1 and BFL-1. Blood 2010, 115, 3304–3313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Weber, A.; Paschen, S.A.; Heger, K.; Wilfling, F.; Frankenberg, T.; Bauerschmitt, H.; Seiffert, B.M.; Kirschnek, S.; Wagner, H.; Hacker, G. BimS-induced apoptosis requires mitochondrial localization but not interaction with anti-apoptotic Bcl-2 proteins. J. Cell Biol. 2007, 177, 625–636. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Strasser, A.; Vaux, D.L. Viewing BCL2 and cell death control from an evolutionary perspective. Cell Death Differ. 2018, 25, 13–20. [Google Scholar] [CrossRef]
  58. Iyer, S.; Bell, F.; Westphal, D.; Anwari, K.; Gulbis, J.; Smith, B.J.; Dewson, G.; Kluck, R.M. Bak apoptotic pores involve a flexible C-terminal region and juxtaposition of the C-terminal transmembrane domains. Cell Death Differ. 2015, 22, 1665–1675. [Google Scholar] [CrossRef] [Green Version]
  59. Bleicken, S.; Landeta, O.; Landajuela, A.; Basanez, G.; Garcia-Saez, A.J. Proapoptotic Bax and Bak proteins form stable protein-permeable pores of tunable size. J. Biol. Chem. 2013, 288, 33241–33252. [Google Scholar] [CrossRef]
  60. Basanez, G.; Sharpe, J.C.; Galanis, J.; Brandt, T.B.; Hardwick, J.M.; Zimmerberg, J. Bax-type apoptotic proteins porate pure lipid bilayers through a mechanism sensitive to intrinsic monolayer curvature. J. Biol. Chem. 2002, 277, 49360–49365. [Google Scholar] [CrossRef]
  61. Basanez, G.; Nechushtan, A.; Drozhinin, O.; Chanturiya, A.; Choe, E.; Tutt, S.; Wood, K.A.; Hsu, Y.; Zimmerberg, J.; Youle, R.J. Bax, but not Bcl-xL, decreases the lifetime of planar phospholipid bilayer membranes at subnanomolar concentrations. Proc. Natl. Acad. Sci. USA. 1999, 96, 5492–5497. [Google Scholar] [CrossRef] [PubMed]
  62. Garcia-Saez, A.J.; Mingarro, I.; Perez-Paya, E.; Salgado, J. Membrane-insertion fragments of Bcl-xL, Bax, and Bid. Biochemistry 2004, 43, 10930–10943. [Google Scholar] [CrossRef] [PubMed]
  63. Youle, R.J.; Strasser, A. The BCL-2 protein family: Opposing activities that mediate cell death. Nat. Rev. Mol. Cell Biol. 2008, 9, 47–59. [Google Scholar] [CrossRef] [PubMed]
  64. Kim, H.; Tu, H.C.; Ren, D.; Takeuchi, O.; Jeffers, J.R.; Zambetti, G.P.; Hsieh, J.J.; Cheng, E.H. Stepwise activation of BAX and BAK by tBID, BIM, and PUMA initiates mitochondrial apoptosis. Mol. Cell 2009, 36, 487–499. [Google Scholar] [CrossRef] [PubMed]
  65. Nechushtan, A.; Smith, C.L.; Hsu, Y.T.; Youle, R.J. Conformation of the Bax C-terminus regulates subcellular location and cell death. EMBO J. 1999, 18, 2330–2341. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Yethon, J.A.; Epand, R.F.; Leber, B.; Epand, R.M.; Andrews, D.W. Interaction with a membrane surface triggers a reversible conformational change in Bax normally associated with induction of apoptosis. J. Biol. Chem. 2003, 278, 48935–48941. [Google Scholar] [CrossRef]
  67. Czabotar, P.E.; Westphal, D.; Dewson, G.; Ma, S.; Hockings, C.; Fairlie, W.D.; Lee, E.F.; Yao, S.; Robin, A.Y.; Smith, B.J.; et al. Bax crystal structures reveal how BH3 domains activate Bax and nucleate its oligomerization to induce apoptosis. Cell 2013, 152, 519–531. [Google Scholar] [CrossRef]
  68. Cartron, P.F.; Arokium, H.; Oliver, L.; Meflah, K.; Manon, S.; Vallette, F.M. Distinct domains control the addressing and the insertion of Bax into mitochondria. J. Biol. Chem. 2005, 280, 10587–10598. [Google Scholar] [CrossRef]
  69. Dewson, G.; Kratina, T.; Sim, H.W.; Puthalakath, H.; Adams, J.M.; Colman, P.M.; Kluck, R.M. To trigger apoptosis, Bak exposes its BH3 domain and homodimerizes via BH3:groove interactions. Mol. Cell 2008, 30, 369–380. [Google Scholar] [CrossRef]
  70. Bleicken, S.; Jeschke, G.; Stegmueller, C.; Salvador-Gallego, R.; Garcia-Saez, A.J.; Bordignon, E. Structural model of active Bax at the membrane. Mol. Cell 2014, 56, 496–505. [Google Scholar] [CrossRef]
  71. Flores-Romero, H.; Garcia-Porras, M.; Basanez, G. Membrane insertion of the BAX core, but not latch domain, drives apoptotic pore formation. Sci. Rep. 2017, 7, 16259. [Google Scholar] [CrossRef] [PubMed]
  72. Subburaj, Y.; Cosentino, K.; Axmann, M.; Pedrueza-Villalmanzo, E.; Hermann, E.; Bleicken, S.; Spatz, J.; Garcia-Saez, A.J. Bax monomers form dimer units in the membrane that further self-assemble into multiple oligomeric species. Nat. Commun. 2015, 6, 8042. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Westphal, D.; Dewson, G.; Menard, M.; Frederick, P.; Iyer, S.; Bartolo, R.; Gibson, L.; Czabotar, P.E.; Smith, B.J.; Adams, J.M.; et al. Apoptotic pore formation is associated with in-plane insertion of Bak or Bax central helices into the mitochondrial outer membrane. Proc. Natl. Acad. Sci. USA 2014, 111, E4076–E4085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Terrones, O.; Antonsson, B.; Yamaguchi, H.; Wang, H.G.; Liu, J.; Lee, R.M.; Herrmann, A.; Basanez, G. Lipidic pore formation by the concerted action of proapoptotic BAX and tBID. J. Biol. Chem. 2004, 279, 30081–30091. [Google Scholar] [CrossRef]
  75. Landeta, O.; Landajuela, A.; Gil, D.; Taneva, S.; Di Primo, C.; Sot, B.; Valle, M.; Frolov, V.A.; Basanez, G. Reconstitution of proapoptotic BAK function in liposomes reveals a dual role for mitochondrial lipids in the BAK-driven membrane permeabilization process. J. Biol. Chem. 2011, 286, 8213–8230. [Google Scholar] [CrossRef]
  76. Cheng, E.H.; Kirsch, D.G.; Clem, R.J.; Ravi, R.; Kastan, M.B.; Bedi, A.; Ueno, K.; Hardwick, J.M. Conversion of Bcl-2 to a Bax-like death effector by caspases. Science 1997, 278, 1966–1968. [Google Scholar] [CrossRef]
  77. Bleicken, S.; Hofhaus, G.; Ugarte-Uribe, B.; Schroder, R.; Garcia-Saez, A.J. cBid, Bax and Bcl-xL exhibit opposite membrane remodeling activities. Cell Death Dis. 2016, 7, e2121. [Google Scholar] [CrossRef]
  78. Valero, J.G.; Cornut-Thibaut, A.; Juge, R.; Debaud, A.L.; Gimenez, D.; Gillet, G.; Bonnefoy-Berard, N.; Salgado, J.; Salles, G.; Aouacheria, A.; et al. micro-Calpain conversion of antiapoptotic Bfl-1 (BCL2A1) into a prodeath factor reveals two distinct alpha-helices inducing mitochondria-mediated apoptosis. PLoS ONE 2012, 7, e38620. [Google Scholar] [CrossRef]
  79. Xie, Z.; Schendel, S.; Matsuyama, S.; Reed, J.C. Acidic pH promotes dimerization of Bcl-2 family proteins. Biochemistry 1998, 37, 6410–6418. [Google Scholar] [CrossRef]
  80. Schlame, M.; Ren, M. The role of cardiolipin in the structural organization of mitochondrial membranes. Biochim. Biophys. Acta 2009, 1788, 2080–2083. [Google Scholar] [CrossRef] [Green Version]
  81. Tamura, Y.; Endo, T.; Iijima, M.; Sesaki, H. Ups1p and Ups2p antagonistically regulate cardiolipin metabolism in mitochondria. J. Cell Biol. 2009, 185, 1029–1045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Kagan, V.E.; Tyurin, V.A.; Jiang, J.; Tyurina, Y.Y.; Ritov, V.B.; Amoscato, A.A.; Osipov, A.N.; Belikova, N.A.; Kapralov, A.A.; Kini, V.; et al. Cytochrome c acts as a cardiolipin oxygenase required for release of proapoptotic factors. Nat. Chem. Biol. 2005, 1, 223–232. [Google Scholar] [CrossRef] [PubMed]
  83. Kagan, V.E.; Borisenko, G.G.; Tyurina, Y.Y.; Tyurin, V.A.; Jiang, J.; Potapovich, A.I.; Kini, V.; Amoscato, A.A.; Fujii, Y. Oxidative lipidomics of apoptosis: Redox catalytic interactions of cytochrome c with cardiolipin and phosphatidylserine. Free Radic. Biol. Med. 2004, 37, 1963–1985. [Google Scholar] [CrossRef] [PubMed]
  84. Unsay, J.D.; Cosentino, K.; Subburaj, Y.; Garcia-Saez, A.J. Cardiolipin effects on membrane structure and dynamics. Langmuir ACS J. Surf. Colloids 2013, 29, 15878–15887. [Google Scholar] [CrossRef]
  85. Grijalba, M.T.; Vercesi, A.E.; Schreier, S. Ca2+-induced increased lipid packing and domain formation in submitochondrial particles. A possible early step in the mechanism of Ca2+-stimulated generation of reactive oxygen species by the respiratory chain. Biochemistry 1999, 38, 13279–13287. [Google Scholar] [CrossRef]
  86. Ortiz, A.; Killian, J.A.; Verkleij, A.J.; Wilschut, J. Membrane fusion and the lamellar-to-inverted-hexagonal phase transition in cardiolipin vesicle systems induced by divalent cations. Biophys. J. 1999, 77, 2003–2014. [Google Scholar] [CrossRef]
  87. Kawai, F.; Shoda, M.; Harashima, R.; Sadaie, Y.; Hara, H.; Matsumoto, K. Cardiolipin domains in Bacillus subtilis marburg membranes. J. Bacteriol. 2004, 186, 1475–1483. [Google Scholar] [CrossRef]
  88. Sorice, M.; Manganelli, V.; Matarrese, P.; Tinari, A.; Misasi, R.; Malorni, W.; Garofalo, T. Cardiolipin-enriched raft-like microdomains are essential activating platforms for apoptotic signals on mitochondria. FEBS Lett. 2009, 583, 2447–2450. [Google Scholar] [CrossRef] [Green Version]
  89. Ott, M.; Robertson, J.D.; Gogvadze, V.; Zhivotovsky, B.; Orrenius, S. Cytochrome c release from mitochondria proceeds by a two-step process. Proc. Natl. Acad. Sci. USA 2002, 99, 1259–1263. [Google Scholar] [CrossRef]
  90. Li, M.; Mandal, A.; Tyurin, V.A.; DeLucia, M.; Ahn, J.; Kagan, V.E.; van der Wel, P.C.A. Surface-Binding to Cardiolipin Nanodomains Triggers Cytochrome c Pro-apoptotic Peroxidase Activity via Localized Dynamics. Structure 2019, 27, 806–815.e4. [Google Scholar] [CrossRef] [Green Version]
  91. Aouacheria, A.; Baghdiguian, S.; Lamb, H.M.; Huska, J.D.; Pineda, F.J.; Hardwick, J.M. Connecting mitochondrial dynamics and life-or-death events via Bcl-2 family proteins. Neurochem. Int. 2017, 109, 141–161. [Google Scholar] [CrossRef] [PubMed]
  92. Picard, M.; Wallace, D.C.; Burelle, Y. The rise of mitochondria in medicine. Mitochondrion 2016, 30, 105–116. [Google Scholar] [CrossRef] [PubMed]
  93. Fonseca, T.B.; Sanchez-Guerrero, A.; Milosevic, I.; Raimundo, N. Mitochondrial fission requires DRP1 but not dynamins. Nature 2019, 570, E34–E42. [Google Scholar] [CrossRef] [PubMed]
  94. Ugarte-Uribe, B.; Garcia-Saez, A.J. Apoptotic foci at mitochondria: In and around Bax pores. Philos. Trans. R. Soc. Lond. Ser. B Biol. Sci. 2017, 372. [Google Scholar] [CrossRef] [PubMed]
  95. Friedman, J.R.; Lackner, L.L.; West, M.; DiBenedetto, J.R.; Nunnari, J.; Voeltz, G.K. ER tubules mark sites of mitochondrial division. Science 2011, 334, 358–362. [Google Scholar] [CrossRef] [PubMed]
  96. Cleland, M.M.; Norris, K.L.; Karbowski, M.; Wang, C.; Suen, D.F.; Jiao, S.; George, N.M.; Luo, X.; Li, Z.; Youle, R.J. Bcl-2 family interaction with the mitochondrial morphogenesis machinery. Cell Death Differ. 2011, 18, 235–247. [Google Scholar] [CrossRef] [PubMed]
  97. Scorrano, L.; Ashiya, M.; Buttle, K.; Weiler, S.; Oakes, S.A.; Mannella, C.A.; Korsmeyer, S.J. A distinct pathway remodels mitochondrial cristae and mobilizes cytochrome c during apoptosis. Dev. Cell 2002, 2, 55–67. [Google Scholar] [CrossRef]
  98. Cipolat, S.; Rudka, T.; Hartmann, D.; Costa, V.; Serneels, L.; Craessaerts, K.; Metzger, K.; Frezza, C.; Annaert, W.; D’Adamio, L.; et al. Mitochondrial rhomboid PARL regulates cytochrome c release during apoptosis via OPA1-dependent cristae remodeling. Cell 2006, 126, 163–175. [Google Scholar] [CrossRef]
  99. Frezza, C.; Cipolat, S.; Scorrano, L. Measuring mitochondrial shape changes and their consequences on mitochondrial involvement during apoptosis. Methods Mol. Biol. 2007, 372, 405–420. [Google Scholar] [CrossRef]
  100. Frezza, C.; Cipolat, S.; Martins de Brito, O.; Micaroni, M.; Beznoussenko, G.V.; Rudka, T.; Bartoli, D.; Polishuck, R.S.; Danial, N.N.; De Strooper, B.; et al. OPA1 controls apoptotic cristae remodeling independently from mitochondrial fusion. Cell 2006, 126, 177–189. [Google Scholar] [CrossRef]
  101. Ban, T.; Heymann, J.A.; Song, Z.; Hinshaw, J.E.; Chan, D.C. OPA1 disease alleles causing dominant optic atrophy have defects in cardiolipin-stimulated GTP hydrolysis and membrane tubulation. Hum. Mol. Genet. 2010, 19, 2113–2122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Gross, A.; Katz, S.G. Non-apoptotic functions of BCL-2 family proteins. Cell Death Differ. 2017, 24, 1348–1358. [Google Scholar] [CrossRef] [PubMed]
  103. Buggins, A.G.; Pepper, C.J. The role of Bcl-2 family proteins in chronic lymphocytic leukaemia. Leuk. Res. 2010, 34, 837–842. [Google Scholar] [CrossRef] [PubMed]
  104. Straten, P.; Andersen, M.H. The anti-apoptotic members of the Bcl-2 family are attractive tumor-associated antigens. Oncotarget 2010, 1, 239–245. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The BCL2 puzzle. Canonical BAX/BAK activation. Activation of BAX-type proteins at the mitochondrial outer membrane (MOM) by the BH3 only proteins induces their oligomerization, formation of supramolecular structures (lines, arcs and rings) and pore formation with the consequent release of apoptogenic factors. The apoptotic repressors, block this process by either interacting with BH3 only proteins (MODE1) or with BAX-type proteins in the membrane (MODE 2) or translocating them to the cytosol (MODE 0). Non canonical cell death or PRODEATH MODE of BCL2-type proteins. Under cellular stress, BCL2-type proteins can switch their antiapoptotic phenotype, directly eliciting rather than inhibiting membrane permeabilization. PL: phospholipids; grey balls: apoptogenic factors.
Figure 1. The BCL2 puzzle. Canonical BAX/BAK activation. Activation of BAX-type proteins at the mitochondrial outer membrane (MOM) by the BH3 only proteins induces their oligomerization, formation of supramolecular structures (lines, arcs and rings) and pore formation with the consequent release of apoptogenic factors. The apoptotic repressors, block this process by either interacting with BH3 only proteins (MODE1) or with BAX-type proteins in the membrane (MODE 2) or translocating them to the cytosol (MODE 0). Non canonical cell death or PRODEATH MODE of BCL2-type proteins. Under cellular stress, BCL2-type proteins can switch their antiapoptotic phenotype, directly eliciting rather than inhibiting membrane permeabilization. PL: phospholipids; grey balls: apoptogenic factors.
Cells 08 01176 g001
Figure 2. BAX/BAK structural organization during their activation process. (a) Protein disposition in solution. BAX is represented with nine cylinders corresponding to its nine α-helixes and based on [41]. (b) BAX/BAK early activation steps: including TM dislodgement and N terminal exposure (depicted in green and cyan respectively). (c) BAX/BAK reorganization in two different parts (dimerization and piercing domains) and BH3 domain exposure (depicted in orange). (d) Oligomerization and pore formation, structural representation of membrane embedded BAX/BAK in the context of toroidal pore (clamp model, based on [70]). One monomer is showed in grey (α1–9) and the other is depicted in dark grey (α1′–9′). The relative orientation of the helices 9 remains unresolved.
Figure 2. BAX/BAK structural organization during their activation process. (a) Protein disposition in solution. BAX is represented with nine cylinders corresponding to its nine α-helixes and based on [41]. (b) BAX/BAK early activation steps: including TM dislodgement and N terminal exposure (depicted in green and cyan respectively). (c) BAX/BAK reorganization in two different parts (dimerization and piercing domains) and BH3 domain exposure (depicted in orange). (d) Oligomerization and pore formation, structural representation of membrane embedded BAX/BAK in the context of toroidal pore (clamp model, based on [70]). One monomer is showed in grey (α1–9) and the other is depicted in dark grey (α1′–9′). The relative orientation of the helices 9 remains unresolved.
Cells 08 01176 g002

Share and Cite

MDPI and ACS Style

Flores-Romero, H.; García-Sáez, A.J. The Incomplete Puzzle of the BCL2 Proteins. Cells 2019, 8, 1176. https://doi.org/10.3390/cells8101176

AMA Style

Flores-Romero H, García-Sáez AJ. The Incomplete Puzzle of the BCL2 Proteins. Cells. 2019; 8(10):1176. https://doi.org/10.3390/cells8101176

Chicago/Turabian Style

Flores-Romero, Hector, and Ana J. García-Sáez. 2019. "The Incomplete Puzzle of the BCL2 Proteins" Cells 8, no. 10: 1176. https://doi.org/10.3390/cells8101176

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop