Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 The mitochondrial protein VDAC1 serves as cellular energy regulator and as of apoptotic regulation and thus VDAC1 appears to be promising targets for therapeutic intervention. + 2366 word(s) 2366 2020-11-10 10:56:21 |
2 format change -80 word(s) 2286 2020-11-26 05:09:28 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Shoshan-Barmatz, V.; Shteinfer-Kuzmine, A.; Verma, A. VDAC1. Encyclopedia. Available online: https://encyclopedia.pub/entry/3107 (accessed on 29 March 2024).
Shoshan-Barmatz V, Shteinfer-Kuzmine A, Verma A. VDAC1. Encyclopedia. Available at: https://encyclopedia.pub/entry/3107. Accessed March 29, 2024.
Shoshan-Barmatz, Varda, Anna Shteinfer-Kuzmine, Ankit Verma. "VDAC1" Encyclopedia, https://encyclopedia.pub/entry/3107 (accessed March 29, 2024).
Shoshan-Barmatz, V., Shteinfer-Kuzmine, A., & Verma, A. (2020, November 19). VDAC1. In Encyclopedia. https://encyclopedia.pub/entry/3107
Shoshan-Barmatz, Varda, et al. "VDAC1." Encyclopedia. Web. 19 November, 2020.
VDAC1
Edit

The voltage-dependent anion channel 1 (VDAC1) protein, is an important regulator of mitochondrial function, and serves as a mitochondrial gatekeeper, with responsibility for cellular fate.

Apoptosis VDAC metabolisim Mitochondria

1. Introduction

Mitochondria as cellular energy powerhouses provide a central location for the multiple metabolic reactions required to satisfy the energy and biomolecule demands of cells, and serve to integrate the diverse metabolic pathways and provide cells with metabolic flexibility. The appreciation of mitochondria as sites of biosynthesis and bioenergy production has dramatically expanded in recent years, and they are now known to play a crucial role in almost all aspects of cell biology and to regulate cellular homeostasis, metabolism, innate immunity, apoptosis, epigenetics, cellular fate, and more[1][2]. Mitochondrial dysfunction induces stress responses, which link the fitness of this organelle to the condition of the whole organism. Mitochondria are also integrators and amplifiers of the death program These functions typically center around the metabolic traffic in and out of the mitochondria, and they are likely to contribute to the exceptional variability of mitochondrial disease manifestations. The voltage-dependent anion channel 1 (VDAC1) protein, which is located in the outer mitochondrial membrane (OMM), serves as a gatekeeper that can regulate the metabolic and energetic cross-talk between mitochondria and the rest of the cell, and plays a role in mitochondria-mediated apoptosis[3][4][5].

Other different aspects of VDAC1 structures[6] and functions, such as in cell stress[5], Ca2+ regulation[7][8], metabolism[5], and apoptosis[7][9], and as a therapeutic target[10][11][12][13][14][15], were presented in ours and others’ recent reviews.

2. VDAC1 Structural Elements: The N-Terminal Domain and its Oligomeric State

Three mammalian isoforms of VDAC (VDAC1, VDAC2, and VDAC3) that share a number of functional and structural attributes have been identified[16][17][18]. Information about VDAC isoform function and structure was obtained from channel activity of purified and reconstituted protein and, using cell-based assays for survival, metabolism, reactive oxygen species (ROS), and cellular Ca2+ regulation, and by gene knockout mouse models[17]. VDAC1 is the most abundant isoform and the focus of this review. VDAC2 knock out is lethal and is considered to be an anti-apoptotic protein. VDAC3 is the least known, active as channel. While VDAC1 contains two cysteines, VDAC2 and VDAC3, with nine and six cysteines, respectively, are proposed to function as oxidative stress sensors[17]. The crystalline structure of the most prevalent and studied isoform, VDAC1, was solved at atomic resolution, revealing a β-barrel composed of 19 transmembrane β-strands connected by flexible loops. The β1 and β19 strands, together, are arranged in parallel, where the N-terminal region (26-residues) lies inside the pore[19][20][21], but can flick out of it[22][23] and interact with hexokinase (HK) [4][5][24][25][26][27][28][29][30], Aβ[31][32], and other proteins, such as the anti-apoptotic proteins, Bcl-2 and Bcl-xL[4][19][33][34][35][36][37]. Thus, this region of the protein is well positioned to regulate the traffic of materials through the VDAC1 channel[19][21].

The diameter of the channel pore has been estimated as 2.6–3.0 nm[26], but this can decrease to about 1.5 nm when the N-terminal flexible region is located inside the pore[26][27][28]. The sequence of rich with glycine residues (21GlyTyrGlyPheGly25)[26][27][28] connecting the N-terminal region to β1 strand is thought to confer the flexibility needed for this region to move in and out of the VDAC1 channel[30]. This mobility has been reported to be important for channel gating, dimerization of VDAC1[30], and interaction with hexokinase (HK) and members of the apoptosis regulating Bcl-2 family (i.e., Bax, Bcl-2, and Bcl-xL)[30][19][33][34][38][39].

Membranal and purified VDAC1, can form dimers, trimers, tetramers, hexamers, and higher-order oligomeric forms [4][23][40][41][42][43][44][45][46][47] through contact sites that have been identified[48]. We have demonstrated this oligomerization to be a dynamic process that occurs in response to a variety of apoptotic stimuli, acting through a range of signaling processes[28][40][44][45][47][48][49][50][51][52][53] , as presented below (Section 5.1).

3. VDAC1 Extra-Mitochondrial Localization, Function, and Association with Pathological Conditions

In addition to the mitochondrial membrane, VDAC1 has also been detected in other cell compartments[16] [54][55][56][57][58][59][60][61][62][63][64], including the plasma membrane[16][57], the sarcoplasmic reticulum (SR) of skeletal muscles[65], and the endoplasmic reticulum (ER) of the rat cerebellum[64][66].

Antibodies raised against the N-terminus of VDAC1 interacted with the plasma membrane of bovine astrocytes and blocked a high conductance anion channel[61]. Interestingly, when detected in the plasma membrane (pl-VDAC1), the amino acid residues that were exposed to the cytosol in the mitochondrial protein were found to face the extracellular space[55][60][67][68][69][70]. This was demonstrated in epithelial cells, astrocytes, and neurons[56][57]and in differentiated hippocampal neurons[55]. VDAC1 has also been identified in the brain post-synaptic membrane fraction[59] and in the caveolae or caveolae-related domains of established T-lymphoid-like cell lines[58]. The protein has also been found to participate in the multi-protein complexes found in lipid rafts, together with the estrogen receptor α (mERα) and insulin-growth factor-1 receptor (IGF-1R)[58][62][63][64]. In red blood cells that do not possess mitochondria, VDAC1 has been found in the endoplasmic reticulum, and Golgi apparatus[59].

The plasma membrane form of VDAC1 may have an extended N-terminal signal peptide which is responsible for its targeting to the cell membrane[70][71]. Alternatively, the human plasminogen kringle 5 (K5) may induce translocation of VDAC1 to the cell surface, where the protein was recently identified as the receptor for K5 on HUVEC membrane[72][73]. Additional mechanisms, such as the presence of alternative mRNA untranslated regions, have also been suggested[57][64].

The levels of pl-VDAC1 were reported to be increased under pathological conditions such as Alzheimer’s disease (AD)[73][74][75], where it has been suggested that pl-VDAC1 serves as an "amyloid-regulated" apoptosis related channel[61][62][63][64]. We recently described a direct interaction between Aβ and the N-terminal region of VDAC1, and demonstrated that VDAC1 is required for Aβ entry into the cells[76] and apoptosis induction[31][32]. We have also recently demonstrated that VDAC1 is overexpressed in type 2 diabetes (T2D)[77][78][79][80] and mistargeted to the β-cell plasma membrane[81]. This overexpression under pathological conditions is also seen in cancer[3][15][81][82], autoimmune diseases such as lupus[83], non-alcoholic steatohepatitis (NASH)[84], inflammatory bowel disease (IBD) (unpublished data), and cardiac diseases[85][86], as presented below (Section 8).

The exact functions of extra-mitochondrial VDAC are unknown, although several possible roles have been proposed (reviewed in[63][86][87]), and include regulation of tissue volume in the brain[88], and other cell types [70,93,94], or release of ATP in β-cells and in human erythrocytes[89].

4. VDAC1, a Multi-Functional Channel Controlling Cell Energy, Metabolism, and Oxidative Stress

To reach the mitochondrion matrix or to be released to the cytosol, all metabolites and ions must traverse the OMM via VDAC1, the sole channel mediating the flux of ions, nucleotides, and other metabolites up to ~5000 Da. In this way, VDAC maintains control of the metabolic and ion cross-talk between the mitochondria and the rest of the cell (Figure 1). Nucleotides and metabolites transported include pyruvate, malate, succinate, and NADH/NAD+, as well as lipids, heme, cholesterol, and ions such as Ca2+[3][4][5][90]. In contrast, there are over 50 mitochondrial substrate-specific carrier proteins of the family solute carrier family 25 (SLC25) in the inner mitochondrial membrane (IMM), such as the (ADP/ATP) antiporter, the adenine nucleotide translocator (ANT) , the transporter of Pi (PiC), as well as transporters of aspartate/glutamate, pyruvate, acyl carnitine, and citrate, among others[91].

Figure 1. Voltage-dependent anion channel 1 (VDAC1) as a multi-functional channel mediates metabolites, nucleotides, and Ca2+ transport, controlling energy production, endoplasmic reticulum (ER)-mitochondria cross-talk, and apoptosis. VDAC1 is responsible for a number of functions in the cell and mitochondria including: (A) transfer of metabolites between the mitochondria and cytosol; (B) passage of Ca2+ to and from the intermembrane space (IMS) to facilitate Ca2+ signaling; (C) Mitochondrial antiviral-signaling protein (MAVS) associated with VDAC1 enable anti-viral signaling. (D) Transfer of  acetyl coenzyme-A (acyl-CoAs) across the outer mitochondrial membrane (OMM) to the IMS, for conversion into acylcarnitine by CPT1a for further processing by β-oxidation. Together with Star and translocator protein (TSPO), VDAC1 forms the multi-protein transduceosome, which transports cholesterol. (E) Recycling ATP/ADP, NAD+/NADH, and acyl-CoA between the cytosol and the IMS, and regulating glycolysis via association with HK; (F) contributing to ER-mitochondria contacts, where Ca2+ released by IP3 activation of inositol 1,4,5-trisphosphate receptors (IP3R) in the ER is directly transferred to IMS via VDAC1, and then is transported to the matrix by the Ca2+ uniporter (MCU complex). In the matrix Ca2+ regulates energy production via activation of the tricarboxylic acid cycle (TCA) cycle enzymes: pyruvate dehydrogenase (PDH), isocitrate dehydrogenase (ICDH), and α-ketoglutarate dehydrogenase (α-KGDH). The electron transport chain (ETC) and the ATP synthase (FoF1) are also presented. (G) VDAC1 oligomers forming a hydrophilic protein-conducting channel capable of mediating the release of apoptogenic proteins (e.g., Cyto c and apoptosis-inducing factor (AIF)) from the mitochondrial IMS to the cytosol, leading to apoptosis. (H). VDAC1 oligomers allow mtDNA release triggering inflammasome activation. Pathological conations lead to dysfunction of the mitochondria as reflected in the activities presented in the box on the right. These altered activities can be prevented by VDAC1-interacting molecules, such as DIDS, VBIT-4 and VBIT-12.

Closure [98] or down-regulation of VDAC1 channel expression reduced the exchange of metabolites between the mitochondria and the rest of the cell and inhibited cell growth[82][92], indicating the importance of the protein to the maintenance of physiological cellular function.

As already described, the mitochondria are the energy source of the cell. They are responsible for the ATP generated during glycolysis and oxidative phosphorylation (OXPHOS). This is then exported to the cytosol and exchanged for ADP, which is recycled again in the mitochondria to generate ATP. This shuttling process that also involves ANT and creatinine kinase (CrK), located between the IMM and OMM[93], and may be regulated by tubulin αβ heterodimers[94], but it is ultimately facilitated by VDAC1, which thereby controls the electron transport chain[4] (Figure 1) and the energy state of the cell [95].

In addition to ATP, VDAC1 is also involved in the transfer of a number of other essential molecules across the OMM. These include Ca2+, cholesterol, fatty acids, and reactive oxygen species (ROS). Controlling the flow of Ca2+ allows VDAC1 to regulate mitochondrial Ca2+ homeostasis, oxidative phosphorylation, and Ca2+ cross-talk between the VDAC1 in the OMM and the IP3 receptor in the ER. This takes place through the mitochondria associated membranes (MAM) and involves the chaperone GRP75[66][96][97].

VDAC1[98] is a necessary component of the cholesterol transport multi-protein complex, the transduceosome, composed of the translocator protein (TSPO), and the steroidogenic acute regulatory protein (STAR)[84][99] (Figure 1).

VDAC1 is also part of a complex mediating the transport of fatty acids through the OMM[84][99][100] composed of carnitine palmitoyltransferase 1a (CPT1a), which faces the intermembrane space (IMS), and the long chain acetyl coenzyme-A (acyl-CoA) synthetase (ACSL) protein. Once activated by ACSL, VDAC1 transfers acyl-CoAs across the OMM to the IMS, where they are converted into acylcarnitines by CPT1a (Figure 1). They are then transferred across the IMM by carnitine/acylcarnitine translocase, and converted back into acyl-CoA by CPT2 in the IMM, and, subsequently, undergo β-oxidation in the matrix[99][101]. In this respect, VDAC1 has recently been reported to serve as a lipid sensor[102]. Finally, VDAC is also involved in regulating oxidative stress[5]. ROS formed by reaction with O2- at complex III are released through VDAC1 where they activates c-Jun N-terminal kinase (JNK), the extracellular signal-regulated kinase (ERK 1/2), and p38, members of the mitogen-activated protein kinase (MAPK) family of serine/threonine kinases whose signaling may be detrimental to mitochondrial function[103][104]. Importantly, ROS release and consequent cytotoxicity are decreased when HK-I and HK-II bind to VDAC1[105][106][107][108].

VDAC1 is also affected by hypoxic conditions. The C-terminal end of VDAC1 is cleaved (VDAC1-ΔC), with silencing of hypoxia inducible factor 1A (HIF-1α) prevents such cleavage[109][110]. This formation of VDAC1-ΔC, is thought to prevent apoptosis and permit the maintenance of ATP and cell survival in hypoxia[111].

As described, the location of VDAC1 in the OMM provides the perfect opportunity to preside over the traffic of metabolites between the mitochondria and the cytosol, where it interacts with other proteins in order to orchestrate and integrate mitochondrial functions with other cellular activities[3][4][40][41][112][113](Figure 1).

VDAC1 activities are modulated by Ca2+, ATP, glutamate, and NADH, as well as by a variety of proteins (see Section 7)[114][115][116][117]. Using a photo-reactive ATP analog, we identified three potential nucleotide binding sites [115]. Subsequent NMR spectroscopy and site-directed mutagenesis revealed that hVDAC1 possesses one major binding region for ATP, UTP, and GTP that is formed by the N-terminal α-helix, the linker connecting the helix to the first β-strand, and adjacent barrel residues[118]. The crystal structure of mouse VDAC1 in the presence of ATP, revealed an additional low-affinity binding site[119]. With respect to a high and low affinity ATP binding site, it should be noted that the cellular concentration of ATP is 1–2 mM. 

In addition, Ca2+ binds to VDAC1 although the physiological function of this connection is not clear. Binding of Ca2+ to purified and bilayer reconstituted VDAC1 maintained the channel in an open configuration, which could be useful in upregulating the exchange of metabolites[120]. The divalent cation-binding sites bind the lanthanides, La3+ and Tb3+, as well as ruthenium red (RuR), and its analogue Ru360[121][122][123], the photo-reactive analogue azido ruthenium (AzRu)[124]. All reduce the conductance of native, but not mutant, VDAC1.

VDAC1 undergoes all known types of post-translational modifications (PTMs), including nitrosylation, acetylation, carbonylation, and phosphorylation[118]. TVDAC1 contains two cysteines; Cys232 is found in the carboxyamidomethylated form, while Cys127 is in the oxidized form of sulfonic acid[125]. VDAC1 possesses several potentially phosphorylatable serine and threonine residues, many of which have indeed been shown to undergo phosphorylation[126][127][128] by protein kinase A (PKA)[127], protein kinase C (PKC)ε[126], and GSK3b[129] Both VDAC1 and VDAC2 are phosphorylated at a specific Tyr residue under hypoxic conditions[128].

Under pathological conditions, such as oxidation, aging, or after ischemic reperfusion injury, VDAC was shown to undergo nitration[130][131][132], while the protein undergoes carbonylation in the Alzheimer's disease-affected brain or after exposure to acrolein, produced by lipid peroxidation[133].

Thus,  VDAC1, a protein that plays a pivotal role in regulating cellular energy and metabolism and when over-expressed, leads to cell death can be considered as a therapeutic target for initiating  or inhibiting cell death.

References

  1. eth, K.; Zachariae, U. Ten Years of High Resolution Structural Research on the Voltage Dependent Anion Channel (VDAC)-Recent Developments and Future Directions. Front. Physiol. 2018, 9, 108, doi:10.3389/fphys.2018.00108.
  2. Shoshan-Barmatz, V.; Maldonado, E.N.; Krelin, Y. VDAC1 at the crossroads of cell metabolism, apoptosis and cell stress. Cell Stress 2017, 1, 11–36, doi:10.15698/cst2017.10.104.
  3. Shoshan-Barmatz, V.; De, S.; Meir, A. The Mitochondrial Voltage-Dependent Anion Channel 1, Ca2+ Transport, Apoptosis, and Their Regulation. Front. Oncol. 2017, 7, 60, doi:10.3389/fonc.2017.00060.
  4. Shoshan-Barmatz, V.; Krelin, Y.; Shteinfer-Kuzmine, A. VDAC1 functions in Ca(2+) homeostasis and cell life and death in health and disease. Cell Calcium. 2018, 69, 81–100, doi:10.1016/j.ceca.2017.06.007.
  5. Shoshan-Barmatz, V.; Krelin, Y.; Chen, Q. VDAC1 as a Player in Mitochondria-Mediated Apoptosis and Target for Modulating Apoptosis. Curr. Med. Chem. 2017, 24, 4435–4446, doi:10.2174/0929867324666170616105200.
  6. Camara, A.K.S.; Zhou, Y.; Wen, P.C.; Tajkhorshid, E.; Kwok, W.M. Mitochondrial VDAC1: A Key Gatekeeper as Potential Therapeutic Target. Front. Physiol. 2017, 8, 460, doi:10.3389/fphys.2017.00460.
  7. Fang, D.; Maldonado, E.N. VDAC Regulation: A Mitochondrial Target to Stop Cell Proliferation. Adv. Cancer Res. 2018, 138, 41–69, doi:10.1016/bs.acr.2018.02.002.
  8. Karachitos, A.; Jordan, J.; Kmita, H. VDAC-Targeted Drugs Affecting Cytoprotection and Mitochondrial Physiology in Cerebrovascular and Cardiovascular Diseases. Curr. Med. Chem. 2017, 24, 4419–4434, doi:10.2174/0929867324666170530073238.
  9. Mazure, N.M. VDAC in cancer. Biochim. Biophys. Acta Bioenerg. 2017, 1858, 665–673, doi:10.1016/j.bbabio.2017.03.002.
  10. Reina, S.; De Pinto, V. Anti-Cancer Compounds Targeted to VDAC: Potential and Perspectives. Curr. Med. Chem. 2017, 24, 4447–4469, doi:10.2174/0929867324666170530074039.
  11. Shoshan-Barmatz, V.; Krelin, Y.; Shteinfer-Kuzmine, A.; Arif, T. Voltage-Dependent Anion Channel 1 As an Emerging Drug Target for Novel Anti-Cancer Therapeutics. Front. Oncol. 2017, 7, 154, doi:10.3389/fonc.2017.00154.
  12. Wallace, D.C. A mitochondrial paradigm of metabolic and degenerative diseases, aging, and cancer: A dawn for evolutionary medicine. Annu. Rev. Genet. 2005, 39, 359–407, doi:10.1146/annurev.genet.39.110304.095751.
  13. McBride, H.M.; Neuspiel, M. ; Wasiak, S. Mitochondria: More than just a powerhouse. Curr. Biol. 2006, 16, R551–R560, doi:10.1016/j.cub.2006.06.054.
  14. Shoshan-Barmatz, V.; Ben-Hail, D.; Admoni, L.; Krelin, Y.; Tripathi, S.S. The mitochondrial voltage-dependent anion channel 1 in tumor cells. Biochim. Biophys. Acta 2015, 1848 (10 Pt B), 2547–2575, doi:10.1016/j.bbamem.2014.10.040.
  15. Shoshan-Barmatz, V.; De Pinto, V.; Zweckstetter, M.; Raviv, Z.; Keinan, N.; Arbel, N. VDAC, a multi-functional mitochondrial protein regulating cell life and death. Mol. Aspects Med. 2010, 31, 227–285, doi:10.1016/j.mam.2010.03.002.
  16. De Pinto, V.; Guarino, F.; Guarnera, A.A.; Messina, S.; Reina, F.M.; Palermo, T.V.; Mazzoni, C. Characterization of human VDAC isoforms: A peculiar function for VDAC3? Biochim. Biophys. Acta 2010, 1797, 1268–1275, doi:10.1016/j.bbabio.2010.01.031.
  17. Messina, A.; Reina, S.; Guarino, F.; De Pinto, V. VDAC isoforms in mammals. Biochim. Biophys. Acta 2012, 1818, 1466–1476, doi:10.1016/j.bbamem.2011.10.005.
  18. Raghavan, A.; Sheiko, T.; Graham, B.H.; Craigen, W.J. Voltage-dependant anion channels: Novel insights into isoform function through genetic models. Biochim. Biophys. Acta 2012, 1818, 1477–1485, doi:10.1016/j.bbamem.2011.10.019.
  19. Bayrhuber, M.; Meins, T.; Habeck, M.; Becker, S.; Giller, K.; Villinger, S.; Vonrhein, C.; Griesinger, C.; Zweckstetter, M.; Zeth, K. Structure of the human voltage-dependent anion channel. Proc. Natl. Acad. Sci. USA 2008, 105, 15370–15375, doi:10.1073/pnas.0808115105.
  20. Hiller, S.; Garces, R.G.; Malia, T.J.; Orekhov, V.Y.; Colombini, M.; Wagner, G. Solution structure of the integral human membrane protein VDAC-1 in detergent micelles. Science 2008, 321, 1206–1210, doi:10.1126/science.1161302.
  21. Ujwal, R.; Cascio, D.; Colletier, J.P.; Faham, S.; Zhang, J.; Toro, L.; Ping, P.; Abramson, J. The crystal structure of mouse VDAC1 at 2.3 A resolution reveals mechanistic insights into metabolite gating. Proc. Natl. Acad. Sci. USA 2008, 105, 17742–17747, doi:10.1073/pnas.0809634105.
  22. Hiller, S.; Wagner, G. The role of solution NMR in the structure determinations of VDAC-1 and other membrane proteins. Curr. Opin. Struct. Biol. 2009, 19, 396–401, doi: 10.1016/j.sbi.2009.07.013.
  23. Geula, S.; Ben-Hail, D.; Shoshan-Barmatz, V. Structure-based analysis of VDAC1: N-terminus location, translocation, channel gating and association with anti-apoptotic proteins. Biochem. J. 2012, 444, 475–485, doi:10.1042/BJ20112079.
  24. Abu-Hamad, S.; Zaid, H.; Israelson, A.; Nahon, E.; Shoshan-Barmatz, V. Hexokinase-I protection against apoptotic cell death is mediated via interaction with the voltage-dependent anion channel-1: Mapping the site of binding. J. Biol. Chem. 2008, 283, 13482–13490, doi:10.1074/jbc.M708216200.
  25. Neumann, D.; Buckers, J.; Kastrup, L.; Hell, S.W.; Jakobs, S. Two-color STED microscopy reveals different degrees of colocalization between hexokinase-I and the three human VDAC isoforms. PMC Biophys. 2010, 3, 4, doi: 10.1186/1757–5036–3-4.
  26. Azoulay-Zohar, H.; Israelson, A.; Abu-Hamad, S.; Shoshan-Barmatz, V. In self-defence: Hexokinase promotes voltage-dependent anion channel closure and prevents mitochondria-mediated apoptotic cell death. Biochem. J. 2004, 377 Pt 2, 347–355, doi:10.1042/BJ20031465 BJ20031465.
  27. Arbel, N.; Ben-Hail, D.; Shoshan-Barmatz, V. Mediation of the antiapoptotic activity of Bcl-xL protein upon interaction with VDAC1 protein. J. Biol. Chem. 2012, 287, 23152–23161, doi:10.1074/jbc.M112.345918.
  28. Shoshan-Barmatz, V.; Arbel, N.; Arzoine, L. VDAC, the voltage-dependent anion channel: Function, regulation & mitochondrial signaling in cell life and death. Cell Sci. 2008, 4, 74–118.
  29. Zaid, H.; Abu-Hamad, S.; Israelson, A.; Nathan, I.; Shoshan-Barmatz, V. The voltage-dependent anion channel-1 modulates apoptotic cell death. Cell Death Differ. 2005, 12, 751–760, doi:10.1038/sj.cdd.4401599.
  30. Shoshan-Barmatz, V.; Mizrachi, D. VDAC1: From structure to cancer therapy. Front. Oncol. 2012, 2, 164, doi:10.3389/fonc.2012.00164.
  31. Smilansky, A.; Dangoor, L.; Nakdimon, I.; Ben-Hail, D.; Mizrachi, D.; Shoshan-Barmatz, V. The Voltage-dependent Anion Channel 1 Mediates Amyloid beta Toxicity and Represents a Potential Target for Alzheimer Disease Therapy. J. Biol. Chem. 2015, 290, 30670–30683, doi:10.1074/jbc.M115.691493.
  32. Thinnes, F.P. Apoptogenic interactions of plasmalemmal type-1 VDAC and Abeta peptides via GxxxG motifs induce Alzheimer's disease—a basic model of apoptosis? Wien. Med. Wochenschr. 2011, 161, 274–276, doi:10.1007/s10354–011–0887–5.
  33. Abu-Hamad, S.; Arbel, N.; Calo, D.; Arzoine, L.; Israelson, A.; Keinan, N.; Ben-Romano, R.; Friedman, O.; Shoshan-Barmatz, V. The VDAC1 N-terminus is essential both for apoptosis and the protective effect of anti-apoptotic proteins. J. Cell Sci. 2009, 122 Pt 11, 1906–1916, doi:10.1242/jcs.040188.
  34. Arbel, N.; Shoshan-Barmatz, V. Voltage-dependent anion channel 1-based peptides interact with Bcl-2 to prevent antiapoptotic activity. J. Biol. Chem. 2010, 285, 6053–6062, doi:10.1074/jbc.M109.082990.
  35. Malia, T.J.; Wagner, G. NMR structural investigation of the mitochondrial outer membrane protein VDAC and its interaction with antiapoptotic Bcl-xL. Biochemistry 2007, 46, 514–525, doi:10.1021/bi061577h.
  36. Shimizu, S.; Ide, T.; Yanagida, T.; Tsujimoto, Y. Electrophysiological study of a novel large pore formed by Bax and the voltage-dependent anion channel that is permeable to cytochrome c. J. Biol. Chem. 2000, 275, 12321–12325, doi:10.1074/jbc.275.16.12321.
  37. Shimizu, S.; Narita, M.; Tsujimoto, Y. Bcl-2 family proteins regulate the release of apoptogenic cytochrome c by the mitochondrial channel VDAC. Nature 1999, 399, 483–487, doi:10.1038/20959.
  38. Shi, Y.; Chen, J.; Weng, C.; Chen, R.; Zheng, Y.; Chen, Q.; Tang, H. Identification of the protein-protein contact site and interaction mode of human VDAC1 with Bcl-2 family proteins. Biochem. Biophys. Res. Commun. 2003, 305, 989–996, doi:10.1016/S0006–291X(03)00871–4.
  39. Arzoine, L.; Zilberberg, N.; Ben-Romano, R.; Shoshan-Barmatz, V. Voltage-dependent anion channel 1-based peptides interact with hexokinase to prevent its anti-apoptotic activity. J. Biol. Chem. 2009, 284, 3946–3955, doi:10.1074/jbc.M803614200M803614200.
  40. Keinan, N.; Tyomkin, D.; Shoshan-Barmatz, V. Oligomerization of the mitochondrial protein voltage-dependent anion channel is coupled to the induction of apoptosis. Mol. Cell Biol. 2010, 30, 5698–5709, doi:10.1128/MCB.00165–10.
  41. Shoshan-Barmatz, V.; Israelson, A.; Brdiczka, D.; Sheu, S.S. The voltage-dependent anion channel (VDAC): Function in intracellular signalling, cell life and cell death. Curr. Pharm. Des. 2006, 12, 2249–2270, doi:10.2174/138161206777585111.
  42. Shoshan-Barmatz, V.; Mizrachi, D.; Keinan, N. Oligomerization of the mitochondrial protein VDAC1: From structure to function and cancer therapy. Prog. Mol. Biol. Transl. Sci. 2013, 117, 303–334, doi: 10.1016/B978-0-12-386931-9.00011-8.
  43. Shoshan-Barmatz, V.; Golan, M. Mitochondrial VDAC1: Function in cell life and death and a target for cancer therapy. Curr. Med. Chem. 2012, 19, 714–735.
  44. Shoshan-Barmatz, V.; Keinan, N.; Zaid, H. Uncovering the role of VDAC in the regulation of cell life and death. J. Bioenerg. Biomembr. 2008, 40, 183–191, doi:10.1007/s10863–008–9147–9.
  45. Zeth, K.; Meins, T.; Vonrhein, C. Approaching the structure of human VDAC1, a key molecule in mitochondrial cross-talk. J. Bioenerg. Biomembr. 2008, 40, 127–132, doi:10.1007/s10863–008–9144-z.
  46. Schlattner, U.; Tokarska-Schlattner, M.; Wallimann, T. Mitochondrial creatine kinase in human health and disease. Biochim. Biophys. Acta 2006, 1762, 164–180, doi:10.1016/j.bbadis.2005.09.004.
  47. Zalk, R.; Israelson, A.; Garty, E.S.; Azoulay-Zohar, H.; Shoshan-Barmatz, V. Oligomeric states of the voltage-dependent anion channel and cytochrome c release from mitochondria. Biochem. J. 2005, 386 Pt 1, 73–83, doi:10.1042/BJ20041356BJ20041356.
  48. Ujwal, R.; Cascio, D.; Chaptal, V.; Ping, P.; Abramson, J. Crystal packing analysis of murine VDAC1 crystals in a lipidic environment reveals novel insights on oligomerization and orientation. Channels (Austin) 2009, 3, 167–170.
  49. Keinan, N.; Pahima, H.; Ben-Hail, D.; Shoshan-Barmatz, V. The role of calcium in VDAC1 oligomerization and mitochondria-mediated apoptosis. Biochim. Biophys. Acta 2013, 1833, 1745–1754, doi: 10.1016/j.bbamcr.2013.03.017.
  50. Weisthal, S.; Keinan, N.; Ben-Hail, D.; Arif, T.; Shoshan-Barmatz, V. Ca(2+)-mediated regulation of VDAC1 expression levels is associated with cell death induction. Biochim. Biophys. Acta 2014, 1843, 2270–2281, doi:10.1016/j.bbamcr.2014.03.021.
  51. Huang, L.; Han, J.; Ben-Hail, D.; He, L.; Li, B.; Chen, Z.; Wang, Y.; Yang, Y.; Liu, L.; Zhu, Y.; et al. A new fungal diterpene induces VDAC1-dependent apoptosis in Bax/Bak-deficient cells. J. Biol. Chem. 2015, 290, 23563–23578, doi:10.1074/jbc.M115.648774.
  52. Ben-Hail, D.; Begas-Shvartz, R.; Shalev, M.; Shteinfer-Kuzmine, A.; Gruzman, A.; Reina, S.; De Pinto, V.; Shoshan-Barmatz, V. Novel Compounds Targeting the Mitochondrial Protein VDAC1 Inhibit Apoptosis and Protect against Mitochondrial Dysfunction. J. Biol. Chem. 2016, 291, 24986–25003, doi:10.1074/jbc.M116.744284.
  53. Ben-Hail, D.; Shoshan-Barmatz, V. VDAC1-interacting anion transport inhibitors inhibit VDAC1 oligomerization and apoptosis. Biochim. Biophys. Acta 2016. 1863 (7 Pt A), 1612–1623, doi:10.1016/j.bbamcr.2016.04.002.
  54. Akanda, N.; Tofighi, R.; Brask, J.; Tamm, C.; Elinder, F.; Ceccatelli, S. Voltage-dependent anion channels (VDAC) in the plasma membrane play a critical role in apoptosis in differentiated hippocampal neurons but not in neural stem cells. Cell Cycle 2008, 7, 3225–3234, doi:10.4161/cc.7.20.6831.
  55. Bahamonde, M.I.; Valverde, M.A. Voltage-dependent anion channel localises to the plasma membrane and peripheral but not perinuclear mitochondria. Pflugers Arch. 2003, 446, 309–313, doi:10.1007/s00424-003-1054-1057.
  56. Bathori, G.; Parolini, I.; Szabo, I.; Tombola, F.; Messina, A.; Oliva, M.; Sargiacomo, M.; De Pinto, V.; Zoratti, M. Extramitochondrial porin: Facts and hypotheses. J. Bioenerg. Biomembr. 2000, 32, 79–89, doi:10.1023/a:1005516513313.
  57. Bathori, G.; Parolini, I.; Tombola, F.; Szabo, I.; Messina, A.; Oliva, M.; De Pinto, V.; Lisanti, M.; Sargiacomo, M.; Zoratti, M. Porin is present in the plasma membrane where it is concentrated in caveolae and caveolae-related domains. J. Biol. Chem. 1999, 274, 29607–29612, doi:10.1074/jbc.274.42.29607.
  58. Bouyer, G.; Cueff, A.; Egee, S.; Kmiecik, J.; Maksimova, Y.; Glogowska, E.; Gallagher, P.G.; Thomas, S.L. Erythrocyte peripheral type benzodiazepine receptor/voltage-dependent anion channels are upregulated by Plasmodium falciparum. Blood 2011, 118, 2305–2312, doi:10.1182/blood-2011–01–329300.
  59. De Pinto, V.; Messina, A.; Lane, D.J.; Lawen, A. Voltage-dependent anion-selective channel (VDAC) in the plasma membrane. FEBS Lett. 2010, 584, 1793–1799, doi:10.1016/j.febslet.2010.02.049.
  60. Dermietzel, R.; Hwang, T.K.; Buettner, R.; Hofer, A.; Dotzler, E.; Kremer, M.; Deutzmann, R.; Thinnes, F.P.; Fishman, G.I.; Spray, D.C.; et al. Cloning and in situ localization of a brain-derived porin that constitutes a large-conductance anion channel in astrocytic plasma membranes. Proc. Natl. Acad. Sci. USA 1994, 91, 499–503, doi:10.1073/pnas.91.2.499.
  61. Marin, R.; Ramirez, C.M.; Gonzalez, M.; Gonzalez-Munoz, E.; Zorzano, A.; Camps, M.; Alonso, R.; Diaz, M. Voltage-dependent anion channel (VDAC) participates in amyloid beta-induced toxicity and interacts with plasma membrane estrogen receptor alpha in septal and hippocampal neurons. Mol. Membr. Biol. 2007, 24, 148–160, doi:10.1080/09687860601055559.
  62. Ramirez, C.M.; Gonzalez, M.; D'iaz, M.; Alonso, R.; Marin, R. VDAC and ERalpha interaction in caveolae from human cortex is altered in Alzheimer's disease. Mol. Cell Neurosci. 2009, 42, 172–183, doi:10.1016/j.mcn.2009.07.001.
  63. Sabirov, R.Z.; Merzlyak, P. Plasmalemmal VDAC controversies and maxi-anion channel puzzle. Biochim. Biophys. Acta 2012, 1818, 1570–1580, doi:10.1016/j.bbamem.2011.09.024.
  64. Thinnes, F.P. Neuroendocrine differentiation of LNCaP cells suggests: VDAC in the cell membrane is involved in the extrinsic apoptotic pathway. Mole. Genet. Metabol. 2009, 97, 241–243, doi:10.1016/j.ymgme.2009.04.010.
  65. Shoshan-Barmatz, V.N.; Hadad, F.I.; Shafir, I.; Orr, M.V.; Heilmeyer, L.M. VDAC/porin is present in sarcoplasmic reticulum from skeletal muscle. FEBS Lett. 1996, 386, 205–210, doi:0014–579300442–5.
  66. Shoshan-Barmatz, V.N.; Zalk, R.; Gincel, D.; Vardi, N. Subcellular localization of VDAC in mitochondria and ER in the cerebellum. Biochim. Biophys. Acta 2004, 1657, 105–114, doi:10.1016/j.bbabio.2004.02.009S000527280400043X.
  67. Stadtmuller, U.; Eben-Brunnen, J.; Schmid, A.; Hesse, D.; Klebert, S.; Kratzin, H.D.; Hesse, J.; Zimmermann, B.; Reymann, S.; Thinnes, F.P.; et al. Mitochondria-derived and extra-mitochondrial human type-1 porin are identical as revealed by amino acid sequencing and electrophysiological characterisation. Biol. Chem. 1999, 380, 1461–1466, doi:10.1515/BC.1999.189.
  68. Okada, S.F.; O'Neal, W.K.; Huang, P.; Nicholas, R.A.; Ostrowski, L.E.; Craigen, W.J.; Lazarowski, E.R.; Boucher, R.C. Voltage-dependent anion channel-1 (VDAC-1) contributes to ATP release and cell volume regulation in murine cells. J. Gen. Physiol. 2004, 124, 513–526, doi:10.1085/jgp.200409154.
  69. Buettner, R.; Papoutsoglou, G.; Scemes, E.; Spray, D.C.; Dermietzel, R. Evidence for secretory pathway localization of a voltage-dependent anion channel isoform. Proc. Natl. Acad. Sci. USA 2000, 97, 3201–3206, doi:10.1073/pnas.060242297.
  70. Thinnes, F.P.; Gotz, H.; Kayser, H.; Benz, R.; Schmidt, W.E.; Kratzin, H.D.; Hilschmann, N. Identification of human porins. I. Purification of a porin from human B-lymphocytes (Porin 31HL) and the topochemical proof of its expression on the plasmalemma of the progenitor cell. Biol. Chem. Hoppe. Seyler. 1989, 370, 1253–1264, doi:10.1515/bchm3.1989.370.2.1253.
  71. Gonzalez-Gronow, M.; Kalfa, T.; Johnson, C.E.; Gawdi, G.; Pizzo, S.V. The voltage-dependent anion channel is a receptor for plasminogen kringle 5 on human endothelial cells. J. Biol. Chem. 2003, 278, 27312–27318, doi:10.1074/jbc.M303172200.
  72. Li, L.; Yao, Y.C.; Gu, X.Q.; Che, D.; Ma, C.Q.; Dai, Z.Y.; Li, C.; Zhou, T.; Cai, W.B.; Yang, Z.H.; et al. Plasminogen kringle 5 induces endothelial cell apoptosis by triggering a voltage-dependent anion channel 1 (VDAC1) positive feedback loop. J. Biol. Chem. 2014, 289, 32628–32638, doi:10.1074/jbc.M114.567792.
  73. Manczak, M.; Reddy, P.H. Abnormal interaction of VDAC1 with amyloid beta and phosphorylated tau causes mitochondrial dysfunction in Alzheimer's disease. Hum. Mol. Genet. 2012, 21, 5131–5146, doi:10.1093/hmg/dds360.
  74. Cuadrado-Tejedor, M.; Vilarino, M.; Cabodevilla, F.; Del Rio, J.; Frechilla, D.; Perez-Mediavilla, A. Enhanced expression of the voltage-dependent anion channel 1 (VDAC1) in Alzheimer's disease transgenic mice: An insight into the pathogenic effects of amyloid-beta. J. Alzheimers Dis. 2011, 23, 195–206, doi:10.3233/JAD-2010–100966.
  75. Perez-Gracia, E.; Torrejon-Escribano, B.; Ferrer, I. Dystrophic neurites of senile plaques in Alzheimer's disease are deficient in cytochrome c oxidase. Acta Neuropathol. 2008, 116, 261–268, doi:10.1007/s00401–008–0370–6.
  76. Inoue, M.; Hur, J.Y.; Kihara, T.; Teranishi, Y.; Yamamoto, N.G.; Ishikawa, T.; Wiehager, B.; Winblad, B.; Tjernberg, L.O.; Schedin-Weiss, S. Human brain proteins showing neuron-specific interactions with gamma-secretase. FEBS J. 2015, 282, 2587–2599, doi:10.1111/febs.13303.
  77. Ahmed, M.; Muhammed, S.; Kessler, B.; Salehi, A. Mitochondrial proteome analysis reveals altered expression of voltage dependent anion channels in pancreatic beta-cells exposed to high glucose. Islets 2010, 2, 283–292, doi:10.4161/isl.2.5.12639.
  78. Gong, D.; Chen, X.; Middleditch, M.; Huang, L.; Vazhoor Amarsingh, G.; Reddy, S.; Lu, J.; Zhang, S.; Phillips, R.; et al. Quantitative proteomic profiling identifies new renal targets of copper(II)-selective chelation in the reversal of diabetic nephropathy in rats. Proteomics 2009, 9, 4309–4320, doi:10.1002/pmic.200900285.
  79. Pittala, S.; Levy, I.; De, S.; Kumar Pandey, S.; Melnikov, N.; Shoshan-Barmatz, V. The VDAC1-based R-Tf-D-LP4 Peptide as a Potential Treatment for Diabetes Mellitus. Cells 2020, 9, 481, doi:10.3390/cells9020481.
  80. Zhong, Z.; Lemasters, J.J. A Unifying Hypothesis Linking Hepatic Adaptations for Ethanol Metabolism to the Proinflammatory and Profibrotic Events of Alcoholic Liver Disease. Alcohol. Clin. Exp. Res. 2018, 42, 2072–2089, doi:10.1111/acer.13877.
  81. Arif, T.; Krelin, Y.; Nakdimon, I.; Benharroch, D.; Paul, A.; Dadon-Klein, D.; Shoshan-Barmatz, V. VDAC1 is a molecular target in glioblastoma, with its depletion leading to reprogrammed metabolism and reversed oncogenic properties. Neuro. Oncol. 2017, 19, 951–964, doi:10.1093/neuonc/now297.
  82. Arif, T.; Vasilkovsky, L.; Refaely, K.A.; Shoshan-Barmatz, V. Silencing VDAC1 Expression by siRNA Inhibits Cancer Cell Proliferation and Tumor Growth In Vivo. Mol. Ther. Nucleic Acids 2014, 3, e159, doi:10.1038/mtna.2014.9.
  83. Kim, J.R.; Gupta, L.P.; Blanco, S.; Yang, A.; Shteinfer-Kuzmine, W.; Kang, Z.; Zhu, X.; Park, S.-J.; et al. VDAC oligomers form mitochondrial pores to release mtDNA fragments and promote lupus-like disease. Science 2019, 366, 1531–1536, doi:10.1126/science.aav4011.
  84. Pittala, S.; Krelin, Y.; Kuperman, Y.; Shoshan-Barmatz, V. A Mitochondrial VDAC1-Based Peptide Greatly Suppresses Steatosis and NASH-Associated Pathologies in a Mouse Model. Mol. Ther. 2019, 27, 1848–1862, doi:10.1016/j.ymthe.2019.06.017.
  85. Branco, A.F.; Pereira, S.L.; Moreira, A.C.; Holy, J.; Sardao, V.A.; Oliveira, P.J. Isoproterenol cytotoxicity is dependent on the differentiation state of the cardiomyoblast H9c2 cell line. Cardiovasc Toxicol. 2011, 11, 191–203, doi:10.1007/s12012–011–9111–5.
  86. Klapper-Goldstein, H.; Verma, A.; Elyagon, S.; Gillis, R.; Murninkas, M.; Pittala, S.; Paul, A.; Shoshan-Barmatz, V.; a.E. Y. VDAC1 overexpression in the diseased myocardium of humans and rats and the effect of VDAC1-interacting compound on atrial fibrosis induced by hyperaldosteronism. 2020, under Revision.
  87. Shoshan-Barmatz, V.; Israelson, A. The voltage-dependent anion channel in endoplasmic/sarcoplasmic reticulum: Characterization, modulation and possible function. J. Membr. Biol. 2005, 204, 57–66, doi:10.1007/s00232–005–0749–4.
  88. Dermietzel, R.; Hwang, T.K.; Buettner, R.; Hofer, A.; Dotzler, E.; Kremer, M.; Deutzmann, R.; Thinnes, F.P.; Fishman, G.I.; Spray, D.C.; et al. Cloning and in situ localization of a brain-derived porin that constitutes a large-conductance anion channel in astrocytic plasma membranes. Proc. Natl. Acad. Sci. USA 1994, 91, 499–503, doi:10.1073/pnas.91.2.499
  89. Marginedas-Freixa, I.; Alvarez, C.L.; Moras, M.; Leal Denis, M.F.; Hattab, C.; Halle, F.; Bihel, F.; Mouro-Chanteloup, I.; Lefevre, S.D.; Le Van Kim, C.; et al. Human erythrocytes release ATP by a novel pathway involving VDAC oligomerization independent of pannexin-1. Sci. Rep. 2018, 8, 11384, doi:10.1038/s41598–018–29885–7.
  90. Shoshan-Barmatz, V.; Ben-Hail, D. VDAC, a multi-functional mitochondrial protein as a pharmacological target. Mitochondrion 2012, 12, 24–34, doi:10.1016/j.mito.2011.04.001.
  91. Palmieri, F.; Pierri, C.L. Mitochondrial metabolite transport. Essays Biochem. 2010, 47, 37–52, doi:10.1042/bse0470037bse0470037.
  92. Abu-Hamad, S.; Sivan, S.; Shoshan-Barmatz, V. The expression level of the voltage-dependent anion channel controls life and death of the cell. Proc. Natl. Acad. Sci. USA 2006, 103, 5787–5792, doi:10.1073/pnas.0600103103.
  93. Dolder, M.; Wendt, S.; Wallimann, T. Mitochondrial creatine kinase in contact sites: Interaction with porin and adenine nucleotide translocase, role in permeability transition and sensitivity to oxidative damage. Biol. Signals Recept. 2001, 10, 93–111.
  94. Gurnev, P.A.; Rostovtseva, T.K.; Bezrukov, S.M. Tubulin-blocked state of VDAC studied by polymer and ATP partitioning. FEBS Lett. 2011, 585, 2363–2366, doi:10.1016/j.febslet.2011.06.008.
  95. Vander Heiden, M.G.; Chandel, N.S.; Li, X.X.; Schumacker, P.T.; Colombini, M.; Thompson, C.B. Outer mitochondrial membrane permeability can regulate coupled respiration and cell survival. Proc. Natl. Acad. Sci. USA 2000, 97, 4666–4671, doi:10.1073/pnas.090082297090082297.
  96. Csordás, G.; Renken, C.; Várnai, P.; Walter, L.; Weaver, D.; Buttle, K.F.; Balla, T.; Mannella, C.A.; Hajnóczky, G. Structural and functional features and significance of the physical linkage between ER and mitochondria. J. Cell Biol. 2006, 174, 915–921, doi:10.1083/jcb.200604016.
  97. Marchi, S.; Patergnani, S.; Pinton, P. The endoplasmic reticulum-mitochondria connection: One touch, multiple functions. Biochim. Biophys. Acta 2014, 1837, 461–469, doi:10.1016/j.bbabio.2013.10.015.
  98. Rone, M.B.; Fan, J.; Papadopoulos, V. Cholesterol transport in steroid biosynthesis: Role of protein-protein interactions and implications in disease states. Biochim. Biophys. Acta 2009, 1791, 646–658, doi: 10.1016/j.bbalip.2009.03.001.
  99. Lee, K.; Kerner, J.; Hoppel, C.L. Mitochondrial carnitine palmitoyltransferase 1a (CPT1a) is part of an outer membrane fatty acid transfer complex. J. Biol. Chem. 2011, 286, 25655–62562, doi:10.1074/jbc.M111.228692.
  100. Paillard, M.; Tubbs, E.; Thiebaut, P.A.; Gomez, L.; Fauconnier, J.; Da Silva, C.C.; Teixeira, G.; Mewton, N.; Belaidi, E.; Durand, A.; et al. Depressing mitochondria-reticulum interactions protects cardiomyocytes from lethal hypoxia-reoxygenation injury. Circulation 2013, 128, 1555–1565, doi:10.1161/circulationaha.113.001225.
  101. Tonazzi, A.; Giangregorio, N.; Console, L.; Indiveri, C. Mitochondrial carnitine/acylcarnitine translocase: Insights in structure/ function relationships. Basis for drug therapy and side effects prediction. Mini Rev. Med. Chem. 2015, 15, 396–405, doi:10.2174/138955751505150408142032.
  102. Qiu, J.; Tan, Y.-W.; Hagenston, A.M.; Martel, M.-A.; Kneisel, N.; Skehel, P.A.; Wyllie, D.J.A.; Bading, H.; Hardingham, G.E. Mitochondrial calcium uniporter Mcu controls excitotoxicity and is transcriptionally repressed by neuroprotective nuclear calcium signals. Nat. Commun. 2013, 4, 2034, doi:10.1038/ncomms3034.
  103. Kamata, H.; Honda, S.; Maeda, S.; Chang, L.; Hirata, H.; Karin, M. Reactive oxygen species promote TNFalpha-induced death and sustained JNK activation by inhibiting MAP kinase phosphatases. Cell 2005, 120, 649–661, doi:10.1016/j.cell.2004.12.041.
  104. Son, Y.; Cheong, Y.K.; Kim, N.H.; Chung, H.T.; Kang, D.G.; Pae, H.O. Mitogen-Activated Protein Kinases and Reactive Oxygen Species: How Can ROS Activate MAPK Pathways? J. Signal. Transduct. 2011, 2011, 792639, doi:10.1155/2011/792639.
  105. Ahmad, A.; Ahmad, S.; Schneider, B.K.; Allen, C.B.; Chang, L.Y.; White, C.W. Elevated expression of hexokinase II protects human lung epithelial-like A549 cells against oxidative injury. Am. J. Physiol. Lung. Cell Mol. Physiol. 2002, 283, L573–L584, doi:10.1152/ajplung.00410.2001.
  106. Da-Silva, W.S.; Gomez-Puyou, A.; de Gomez-Puyou, M.T.; Moreno-Sanchez, R.; De Felice, F.G.; de Meis, L.; Oliveira, M.F.; Galina, A. Mitochondrial bound hexokinase activity as a preventive antioxidant defense: Steady-state ADP formation as a regulatory mechanism of membrane potential and reactive oxygen species generation in mitochondria. J. Biol. Chem. 2004, 279, 39846–39855, doi:10.1074/jbc.M403835200M403835200.
  107. Sun, L.; Shukair, S.; Naik, T.J.; Moazed, F.; Ardehali, H. Glucose phosphorylation and mitochondrial binding are required for the protective effects of hexokinases I and II. Mol. Cell Biol. 2008, 28, 1007–1017, doi:10.1128/MCB.00224-07.
  108. Bryson, J.M.; Coy, P.E.; Gottlob, K.; Hay, N.; Robey, R.B. Increased hexokinase activity, of either ectopic or endogenous origin, protects renal epithelial cells against acute oxidant-induced cell death. J. Biol. Chem. 2002, 277, 11392–11400, doi:10.1074/jbc.M110927200.
  109. Brahimi-Horn, M.C.; Lacas-Gervais, S.; Adaixo, R.; Ilc, K.; Rouleau, M.; Notte, A.; Dieu, M.; Michiels, C.; Voeltzel, T.; Maguer-Satta, V.; et al. Local mitochondrial-endolysosomal microfusion cleaves voltage-dependent anion channel 1 to promote survival in hypoxia. Mol. Cell Biol. 2015, 35, 1491–1505, doi:10.1128/MCB.01402-14MCB.01402-14.
  110. Mazure, N.M. News about VDAC1 in Hypoxia. Front. Oncol. 2016, 6, 193, doi:10.3389/Fonc.2016.00193.
  111. Brahimi-Horn, M.C.; Ben-Hail, D.; Ilie, M.; Gounon, P.; Rouleau, M.; Hofman, V.; Doyen, J.; Mari, B.; Shoshan-Barmatz, V.; Hofman, P.; et al. Expression of a truncated active form of VDAC1 in lung cancer associates with hypoxic cell survival and correlates with progression to chemotherapy resistance. Cancer Res. 2012, 72, 2140–2150, doi:10.1158/0008–5472.CAN-11–3940.
  112. Aram, L.; Geula, S.; Arbel, N.; Shoshan-Barmatz, V. VDAC1 cysteine residues: Topology and function in channel activity and apoptosis. Biochem. J. 2010, 427, 445–454, doi:10.1042/BJ20091690.
  113. Colombini, M. VDAC structure, selectivity, and dynamics. Biochim. Biophys. Acta 2012, 1818, 1457–1465, doi:10.1016/j.bbamem.2011.12.026.
  114. Gincel, D.; Shoshan-Barmatz, V. Glutamate interacts with VDAC and modulates opening of the mitochondrial permeability transition pore. J. Bioenerg. Biomembr. 2004, 36, 179–186, doi:10.1023/B:JOBB.0000023621.72873.9e.
  115. Rostovtseva, T.K.; Komarov, A.; Bezrukov, S.M.; Colombini, M. VDAC channels differentiate between natural metabolites and synthetic molecules. J. Membr. Biol. 2002, 187, 147–156, doi:0.1007/s00232–001–0159–1.
  116. Yehezkel, G.; Hadad, N.; Zaid, H.; Sivan, S.; Shoshan-Barmatz, V. Nucleotide-binding sites in the voltage-dependent anion channel: Characterization and localization. J. Biol. Chem. 2006, 281, 5938–5946, doi:10.1074/jbc.M510104200.
  117. Shoshan-Barmatz, V.; Gincel, D. The voltage-dependent anion channel: Characterization, modulation, and role in mitochondrial function in cell life and death. Cell Biochem. Biophys. 2003, 39, 279–292, doi: 10.1385/CBB:39:3:279.
  118. Villinger, S.; Giller, K.; Bayrhuber, M.; Lange, A.; Griesinger, C.; Becker, S.; Zweckstetter, M. Nucleotide interactions of the human voltage-dependent anion channel. J. Biol. Chem. 2014, 289, 13397–13406, doi:10.1074/jbc.M113.524173.
  119. Choudhary, O.P.; Paz, A.; Adelman, J.L.; Colletier, J.P.; Abramson, J.; Grabe, M. Structure-guided simulations illuminate the mechanism of ATP transport through VDAC1. Nat. Struct. Mol. Biol. 2014, 21, 626–632, doi:10.1038/nsmb.2841.
  120. Bathori, G.; Csordas, G.; Garcia-Perez, C.; Davies, E.; Hajnoczky, G. Ca2+-dependent control of the permeability properties of the mitochondrial outer membrane and voltage-dependent anion-selective channel (VDAC). J. Biol. Chem. 2006, 281, 17347–17358, doi:10.1074/jbc.M600906200.
  121. Gincel, D.; Zaid, H.; Shoshan-Barmatz, V. Calcium binding and translocation by the voltage-dependent anion channel: A possible regulatory mechanism in mitochondrial function. Biochem. J. 2001, 358 Pt 1, 147–155, doi:10.1042/0264–6021:3580147.
  122. Israelson, A.; Abu-Hamad, S.; Zaid, H.; Nahon, E.; Shoshan-Barmatz, V. Localization of the voltage-dependent anion channel-1 Ca2+-binding sites. Cell Calcium. 2007, 41, 235–544, doi: 10.1016/j.ceca.2006.06.005.
  123. Gincel, D.; Vardi, N.; Shoshan-Barmatz, V. Retinal voltage-dependent anion channel: Characterization and cellular localization. Invest. Ophthalmol. Vis. Sci. 2002, 43, 2097–2104.
  124. Israelson, A.; Arzoine, L.; Abu-hamad, S.; Khodorkovsky, V.; Shoshan-Barmatz, V. A photoactivable probe for calcium binding proteins. Chem. Biol. 2005, 12, 1169–1178, doi:10.1016/j.chembiol.2005.08.006.
  125. Pittala, M.G.G.; Saletti, R.; Reina, S.; Cunsolo, V.; De Pinto, V.; Foti, S. A High Resolution Mass Spectrometry Study Reveals the Potential of Disulfide Formation in Human Mitochondrial Voltage-Dependent Anion Selective Channel Isoforms (hVDACs). Int. J. Mol. Sci. 2020, 21, 1468, doi:10.3390/ijms21041468.
  126. Baines, C.P.; Song, C.X.; Zheng, Y.T.; Wang, G.W.; Wang, Z.O.; Guo, Y.; Bolli, R.; Cardwell, E.M. Protein kinase Cepsilon interacts with and inhibits the permeability transition pore in cardiac mitochondria. Circ. Res. 2003, 92, 873–880.
  127. Bera, A.K.; Ghosh, S.; Das, S. Mitochondrial VDAC can be phosphorylated by cyclic AMP-dependent protein kinase. Biochem. Biophys. Res. Commun. 1995, 209, 213–217, doi: 10.1006/bbrc.1995.1491.
  128. Liberatori, S.; Canas, B.; Tani, C.; Bini, L.; Buonocore, G.; Godovac-Zimmermann, J.; Mishra, O.P.; Delivoria-Papadopoulos, M.; Bracci, R.; Pallini, V. Proteomic approach to the identification of voltage-dependent anion channel protein isoforms in guinea pig brain synaptosomes. Proteomics 2004, 4, 1335–1340, doi:10.1002/pmic.200300734.
  129. Kerner, J.; Lee, K.; Tandler, B.; Hoppel, C.L. VDAC proteomics: Post-translation modifications. Biochim. Biophys. Acta 2012, 1818, 1520–1525, doi:10.1016/j.bbamem.2011.11.013.
  130. Aulak, K.S.; Koeck, T.; Crabb, J.W.; Stuehr, D.J. Dynamics of protein nitration in cells and mitochondria. Am. J. Physiol. Heart Circul. Physiol. 2004, 286, H30–H38, doi:10.1152/ajpheart.00743.2003.
  131. Kanski, J.; Behring, A.; Pelling, J.; Schoneich, C. Proteomic identification of 3-nitrotyrosine-containing rat cardiac proteins: Effects Biol. Aging 2005, 288, H371–H381, doi:10.1152/ajpheart.01030.2003.
  132. Turko, I.V.; Li, L.; Aulak, K.S.; Stuehr, D.J.; Chang, J.Y.; Murad, F. Protein tyrosine nitration in the mitochondria from diabetic mouse heart. Implications to dysfunctional mitochondria in diabetes. J. Biol. Chem. 2003, 278, 33972–33977, doi:10.1074/jbc.M303734200.
  133. Mello, C.F.; Sultana, R.; Piroddi, M.; Cai, J.; Pierce, W.M.; Klein, J.B.; Butterfield, D.A. Acrolein induces selective protein carbonylation in synaptosomes. Neuroscience 2007, 147, 674–967, doi:10.1016/j.neuroscience.2007.04.003.
More
Information
Subjects: Cell Biology
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , ,
View Times: 766
Revisions: 2 times (View History)
Update Date: 26 Nov 2020
1000/1000