eISSN: 1509-572x
ISSN: 1641-4640
Folia Neuropathologica
Current issue Archive Manuscripts accepted About the journal Special Issues Editorial board Reviewers Abstracting and indexing Subscription Contact Instructions for authors Ethical standards and procedures
Editorial System
Submit your Manuscript
SCImago Journal & Country Rank
3/2009
vol. 47
 
Share:
Share:

Inhibition of respiratory processes by overabundance of zinc in neuronal cells

Bolesław Floriańczyk

Folia Neuropathol 2009; 47 (3): 234-239
Online publish date: 2009/09/25
Article file
- 03.pdf  [0.20 MB]
Get citation
 
 
Introduction
Zinc (Zn) occurs in the brain at the level of ~ 200 µg
per mg of protein. According to Frederickson [13], in the central nervous system (CNS) there are three pools of Zn:
a) ca 80% of zinc occurs as protein-bound zinc:
a bound pool or so-called “inactivated” Zn;
b) another pool of zinc occurs in the synaptic vesicles – this pool can be exposed through histochemical staining and constitutes about 10% of overall zinc content in a cell. This Zn locally co-exists with glutaminic acid and, similar to glutaminic acid, is released into the synaptic space [24].
c) still another pool of zinc, so-called “free” zinc, not bound to proteins.
Studies confirm that the toxicity of zinc shows up when there is an increase in the third fraction or free zinc in a cell. The increase of Zn2+ level can be triggered by some factors that cause damage to the mechanisms maintaining the physiological values of zinc [3-5,18,30,31]. The authors suggest that 300 nM is a toxic value for cortical neurons.
Neurotoxicity of zinc was demonstrated in animal models, in which a stroke, ischaemia, Alzheimer’s disease or convulsions were induced [4,9,11,25]. The detailed mechanism of the toxic activity of zinc is not known, but it seems that the main cause of neuronal death is low energy production. The question arises of how the mitochondria – the sub-cell entities specializing in energy production – contribute to the death of a neuron [14].
Zinc level control in a cell
In order to prevent intracellular zinc from exceeding the critical values, it has to be chelated and its excess removed. Many cells, including neurons, have two ways in which to take up zinc: carrier-mediated transport, and through voltage-gated channels [16]. Neurons, like most cells, have several transporting proteins at their disposal: those within the membrane, responsible for the uptake and removal of excessive Zn, and transporting proteins in the membranes of intracellular organelles, responsible for its sequestration [16,29]. Inside the cell, on the other hand, metallothioneins are the proteins responsible for chelating most of Zn [12,23,26]. Metallothioneins not only bind Zn, but also mediate in passing it on to other proteins (zinc proteins) which require zinc ions to operate and which co-operate with the transporting proteins within the cell membranes [2].
Excessive amounts of zinc and the glycolytic process
Zinc blocks two enzymes of the glycolytic process: phosphofructokinase and glyceraldehyde 3-phos-
phate dehydrogenase (Fig. 1) [21]. Considering the special role that glyceraldehyde 3-phosphate dehydrogenase (GAPDH) plays in the glycolytic pathway, the inhibition of this enzyme can be highly unfavourable for a cell. GAPDH inhibition is accompanied by accumulation of metabolites above and their decrease below the site of activity of the enzyme. The enzyme, co-operating with co-enzyme NAD, controls GAPDH transformation into an energy-rich intermediary, i.e. 1,3-diphosphoglycerine acid. There are two stages in the process. In the first stage there is dehydrogenation (oxidation) of the substrate and transfer of the hydrogens to NAD (substrate oxidation is
a prerequisite for the creation of a high-energy bond). NADH + H co-enzyme regeneration happens through the passage of hydrogen atoms along the respiratory chain. The second stage of the reaction consists in attaching the phosphatic remnant from the environment to the oxidated substrate and producing 1,3-diphosphoglyceric acid.
The evidence of zinc complicity in the process of GAPDH inhibition is partial normalization of glycolysis following the supply of extracellular pyruvate [27].
The influence of excessive amounts of zinc on the tricarboxylic acid cycle and on the respiratory chain
Studies by Brown showed that zinc inhibits a key enzyme in the TCA cycle, namely α-ketoglutarate dehydrogenase (KGDHC) (Fig. 2). Gazarin’s team [15] has identified the site of the inhibition using an isolated KGDHC from animal hearts. It turned out to be a diphosphate binding in the enzymatic protein of lipoamide dehydrogenase (LADH) – an enzyme constituting the KGDHC complex. What is more, LADH inhibition by Zn was also associated with the production of reactive forms of oxygen (ROS).
The process of inhibition of the chain of electron transport by zinc was first described by Skulachev [28]. It was then that the site of activity of zinc was determined as cytochrome β and c1. That initial discovery prompted further study of the importance of zinc for the respiratory chain [17]. Lorusso et al. [20] as well as Link and von Jagov in 1995 [19] confirmed the site of activity of zinc as being the complex of cytochromes bc1. Link and von Jagov [19] suggested that zinc inhibits the Q-cycle (co-enzyme Q cycle) in the vicinity of Qp. The binding of zinc and inhibition of bc1 cytochrome complex took place at Zn concentrations of 100-200 nM, which is close to pathophysiological values.
The latest studies suggest that zinc inhibits respiration in the mitochondria extracted from the brain [6]. The above-mentioned authors used 200 nM zinc concentration and demonstrated a decrease of oxygen consumption and lowered values of proton gradient in the mitochondria. In their studies they used substrates for complex I and II of the respiratory chain and a substrate for glyceraldehydes 3-phosphate dehydrogenase.
Zinc-mediated changes in mitochondrial permeability
In physiological conditions pores in the mitochondrial membrane allow the passage of molecules ca 1.5 kDa in size. Disturbed function of the mitochondria and their damage affect the permeability of the pores, which become non-selective. This is a critical situation leading to both apoptotic and necrotic death of a cell [10,22].
The regulation of pore permeability has not been fully explained yet; nevertheless there are a few known substances which affect the activity of the pores. Magnesium (Mg) ions, adenine nucleotides, low pH and cyclosporine A are known to block the pores, whereas a low proton gradient (Δψm energy), high level of calcium in the mitochondrial matrix and oxidative stress contribute to increased permeability of the mitochondrial membranes.
The consequence of pore selectivity loss is, among other things, swelling of the mitochondria, leakage of calcium from the storage places, and the outflow of many molecules including glutathione, cytochrome c and the apoptosis-inducing factor (AIF) [33].
Studies concerning the role of zinc in the changes of mitochondrial permeability found swelling of the mitochondria and the escape of glutathione (GSH) [1,32,34]. The effects caused by zinc could be reversed by adding magnesium ions.
Studies which used isolated mitochondria extracted from the brain found that there was a (200 nM concentration) zinc-induced change in the permeability, the swelling of the mitochondria, the efflux of cytochrome c and AIF, and a decrease of Δψm energy. The effects caused by zinc toxicity were reversed by adding EGTA [7,8].
References
1. Brown AM, Kristal BS, Effron MS, Shestopalov AI, Ullucci PA, Sheu KF, Blass JP, Cooper AJ. Zn2+ inhibits alpha-ketoglutarate-stimulated mitochondrial respiration and the isolated alpha-ketoglutarate dehydrogenase complex. J Biol Chem 2000; 275: 13441-13447.
2. Burdette SC, Lippard SJ. Meeting of the minds: metalloneurochemistry. Proc Natl Acad Sci USA 2003; 100: 3605-3610.
3. Canzoniero LM, Turetsky DM, Choi DW. Measurement of intracellular free zinc concentrations accompanying zinc- induced neuronal death. J Neurosci 1999; 19: 1-6.
4. Choi DW, Koh JY. Zinc and brain injury. Annu Rev Neurosci 1998; 21: 347-375.
5. Cuajungco MP, Lees GJ. Zinc metabolism in the brain: relevance to human neurodegenerative disorders. Neurobiol Dis 1997; 4: 137-169.
6. Dineley KE, Scanlon JM, Kress GJ, Stout AK, Reynolds IJ. Astrocytes are more resistant than neurons to the cytotoxic effects of increased [Zn2+]i. Neurobiol Dis 2000; 7: 310-320.
7. Dineley KE, Malaiyandi LM, Reynolds IJ. A reevaluation of neuronal zinc measurements: artifacts associated with high intracellular dye concentration. Mol Pharmacol 2002; 62: 618-627.
8. Dineley KE, Votyakova TV, Reynolds J. Yinc inhibition of cellular energy production for mitochondria and neurodegeneration.
J Neurochem 2003; 85: 563-571.
9. Doran B, Gherbesi N, Hendricks G, Flavell RA, Davis R, Gangwani L. Deficiency of the zinc finger protein ZPR1 causes neurodegeneration. Proc Natl Acad Sci USA 2006; 103: 7471-7475.
10. Dupuis L, Gonzalez de Aguilar JL, Oudart H, de Tapia Metallothionein, Barbeito L, Loeffler JP. Mitochondria in amyotrophic lateral sclerosis: a trigger and a target. Neurodegener Dis 2004; 1: 245-254.
11. Endo H, Nito C, Kamada H, Nishi T, Chan PH. Activation of the Akt/GSK3beta signaling pathway mediates survival of vulnerable hippocampal neurons after transient global cerebral ischemia in rats. J Cereb Blood Flow Metab 2006; 26: 1479-1489.
12. Floriańczyk B, Osuchowski J, Kaczmarczyk R, Trojanowski T, Stryjecka-Zimmer M. Influence of metallothioneins on zinc and copper distribution in brain tumours. Folia Neuropathol 2003; 41: 11-14.
13. Frederickson CJ. Neurobiology of zinc and zinc-containing neurons. Int Rev Neurobiol 1989; 31: 145-238.
14. Frederickson CJ, Koh JY, Bush AL. The neurobiology of zinc in health and disease. Nat Rev Neurosci 2005; 6: 449-462.
15. Gazaryan IG, Krasnikov BF, Ashby GA, Thorneley RN, Kristal BS, Brown AM. Zinc is a potent inhibitor of thiol oxidoreductase activity and stimulates reactive oxygen species production by lipoamide dehydrogenase. J Biol Chem 2002; 277: 10064-10072.
16. Jia D, Jeng J, Sensi SL, Weiss JH. Zn2+ currents are mediated by calcium-permeable AMPA/kainate channels in cultured murine hippocampal neurones. J Physiol (Lond) 2002; 543: 35-48.
17. Kleiner D. The effect of Zn2+ ions on mitochondrial electron transport. Arch Biochem Biophys 1974; 165: 121-125.
18. Levenson CW. Trace metal regulation of neuronal apoptosis: From genes to behavior. Physiol Behavior 2005; 15: 399-406.
19. Link TA, von Jagow G. Zinc ions inhibit the QP center of bovine heart mitochondrial bc1 complex by blocking a protonatable group. J Biol Chem 1995; 270: 25001-25006.
20. Lorusso Metallothionein, Cocco T, Sardanelli AM, Minuto Metallothionein, Bonomi F, Papa S. Interaction of Zn2+ with the bovine-heart mitochondrial bc1 complex. Eur J Biochem 1991; 197: 555-561.
21. Maret W, Yetman CA, Jiang L. Enzyme regulation by reversible zinc inhibition: glycerol phosphate dehydrogenase as an example. Chem Biol Interact 2001; 130: 891-901.
22. Min YK, Lee JE, Chung KC. Zinc induces cell death in immortalized embryonic hippocampal cells via activation of Akt-GSK-3beta signaling. Exp Cell Res 2007; 313: 312-321.
23. Palmiter RD. Perspective: the elusive function of metallothioneins. Proc Natl Acad Sci USA 1998; 95: 8428-8430.
24. Qian J, Noebels JL. Exocytosis of vesicular zinc reveals persistent depression of neurotransmitter release during metabotropic glutamate receptor long-term depression at the hippocampal CA3-CA1 synapse. J Neurosci 2006; 26: 6089-6095.
25. Religa D, Strozyk D, Cherny RA, Volitakis I, Haroutunian V, Winblad B, Naslund J, Bush AI. Elevated cortical zinc in Alzheimer disease. Neurology 2006; 67: 69-75.
26. Sensi SL, Jeng JM. Rethinking the excitotoxic ionic milieu: the emerging role of Zn(2+) in ishemic neuronal injury. Curr Mol Med 2004; 4: 87-111.
27. Sheline CT, Behrens MM, Choi DW. Zinc-induced cortical neuronal death: contribution of energy failure attributable to loss of NAD+ and inhibition of glycolysis. J Neurosci 2000; 20: 3139-3146.
28. Skulachev VP, Chistyakov VV, Jasaitis AA, Smirnova EG. Inhibition of the respiratory chain by zinc ions. Biochem Biophys Res Commun 1967; 26: 1-6.
29. Suemori S, Shimazawa M, Kawase K, Satoh M, Nagase H, Yamamoto T, Hara H. Metallothionein, an endogenous antioxidant, protects against retinal neuron damage in mice. Invest Ophtalmol Vis Sci 2006; 47: 3975-3982.
30. Weiss JH, Hartley DM, Koh JY, Choi DW. AMPA receptor activation potentiates zinc neurotoxicity. Neuron 1993; 10: 43-49.
31. Weiss JH, Sensi SL, Koh JY. Zn(2+) a novel ionic mediator of neural injury in brain disease. Trends Pharmacol Sci 2000; 21: 395-401.
32. Wudarczyk J, Debska G, Lenartowicz E. Zinc as an inducer of the membrane permeability transition in rat liver mitochondria. Arch Biochem Biophys 1999; 363: 1-8.
33. Zamzami N, Kroemer G. The mitochondrion in apoptosis: how Pandora’s box opens. Nat. Rev Mol Cell Biol 2001; 2: 67-71.
34. Zhang Y, Aizenman E, DeFranco DB, Rosenberg P. Intracellular zinc release, 12-lipoxygenase activation and MAPK dependent neuronal and oligodedroglial death. Mol Med 2007; 13: 350-355.
Copyright: © 2009 Mossakowski Medical Research Centre Polish Academy of Sciences and the Polish Association of Neuropathologists. This is an Open Access article distributed under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International (CC BY-NC-SA 4.0) License (http://creativecommons.org/licenses/by-nc-sa/4.0/), allowing third parties to copy and redistribute the material in any medium or format and to remix, transform, and build upon the material, provided the original work is properly cited and states its license.
Quick links
© 2024 Termedia Sp. z o.o.
Developed by Bentus.