Mitochondrial dysfunction how and its consequences

Mitochondria and ROS and conditions


Home | Mitochondria and ROS and conditions | Fructose induces Mitochondrial dysfunction | Drugs causing mitochondrial dysfunctionn

The target for therapy is through avoid what harms the mitochondria.  First above all is fructose, and second is the vegetable oils unsaturated fatty acids which become rancid through a chain reaction.  Third would be taking antioxidants in significant amounts.  I take CoQ10, sodium ascorbate 2 grams, and vitamins A and E.  The salt of vitamin C is far easier on the digestive system. 

Boon, Pamela, Li Pun et al, Pathological Significance of Mitochondrial Glycation

https://www.hindawi.com/journals/ijcb/2012/843505/abs/   May 2013  Pamela Boon, Li Pun et al,

Nonetheless, the prospect of mitochondrial glycation contributing as a common damaging agent across a broad spectrum of diseases is an intriguing possibility and is also a novel potential therapeutic target.”
Pathological Significance of Mitochondrial Glycation

Glycation, the nonenzymatic glycosylation of biomolecules, is commonly observed in diabetes and ageing. Reactive dicarbonyl species such as methylglyoxal and glyoxal are thought to be major physiological precursors of glycation. Because these dicarbonyls tend to be formed intracellularly, the levels of advanced glycation end products on cellular proteins are higher than on extracellular ones. The formation of glycation adducts within cells can have severe functional consequences such as inhibition of protein activity and promotion of DNA mutations. Although several lines of evidence suggest that there are specific mitochondrial targets of glycation, and mitochondrial dysfunction itself has been implicated in disease and ageing, it is unclear if glycation of biomolecules specifically within mitochondria induces dysfunction and contributes to disease pathology. We discuss here the possibility that mitochondrial glycation contributes to disease, focussing on diabetes, ageing, cancer, and neurodegeneration, and highlight the current limitations in our understanding of the pathological significance of mitochondrial glycation.




1. Introduction


Glycation is a common feature of diabetic complications [12] and ageing pathologies [34]. This nonenzymatic glycosylation process stems from multiple reactions between reducing sugars or sugar derivatives and amino groups on proteins, lipids and nucleotides. Glycation involves multiple oxidative and nonoxidative reactions, collectively termed the Maillard reaction, eventually generating advanced glycation end products (AGEs) [5]. In clinical and experimental models of diabetes and ageing, the levels of intra- and extracellular AGEs have been found to increase relative to healthy or young controls [3610]. Such correlative data suggests that glycation can potentially contribute to disease progression and ageing pathology. However, whether there is a causal relationship between glycation and pathology is uncertain and the mechanistic details are unclear.


That the formation of AGEs can have significant functional consequences further supports this hypothesis. Glycation affects all major classes of biomolecules (Figure 1), with damage levels estimated at 0.1–1% of lysine and arginine residues on proteins, 1 in 107 nucleotides on DNA and 0.1% of basic phospholipids [11]. The consequences of such glycation damage can be severe. The formation and accumulation of protein AGEs, for instance, can result in biochemical dysfunction. For example, some AGEs like MOLD (methylglyoxal-derived lysine dimers) and pentosidine form protein crosslinks, altering protein structure, generally causing proteins to become more resistant to proteolysis [12]. Protein structure may also be modified as a result of charge neutralization when arginine and lysine residues are glycated [13]. Consequently, structural integrity of proteins becomes compromised. For instance, glycated collagen is stiff and nonelastic relative to the non-glycated protein [1415]. The change in charge distribution could also promote protein aggregation, for instance, of lens crystallins, leading to cataract formation in diabetes and in old age [1617]. Changes in protein conformation could also influence its function, as would glycation of amino acid residues at sites for substrate binding and allosteric regulation on enzymes. It has been shown, for instance, that methylglyoxal-induced glycation of lys 126 and arg 463 in glutamate dehydrogenase isolated from bovine liver affects its ability to bind its substrate and allosteric activator (adenosine diphosphate (ADP)), respectively, resulting in a decrease in glutamate dehydrogenase activity [18]. Similarly, glycation of extracellular matrix (ECM) proteins affects cell-ECM interactions. The reaction of methylglyoxal with arginine residues on the RGD and GFOGER motifs in the integrin-binding sites of collagen, for instance, resulted in endothelial cell detachment by interfering with cell-ECM interactions [7].


843505.fig.001


Figure 1: Reactive α,β-dicarbonyl-induced glycation damage in cells. Glycation by α,β-dicarbonyls like methylglyoxal and glyoxal affects all classes of biological macromolecules. Proteins, in particular arginine, lysine, and cysteine residues, are highly susceptible to glycation, forming protein cross-links and various AGE adducts, examples of which are shown here. Of these, the most important protein-AGEs quantitatively are arginine-derived hydroimidazolones, especially MG-H1. Formation of protein AGEs alters protein structure and function, leading to biochemical dysfunction. Nucleotides and lipids may also undergo glycation, with deoxyguanosine and basic phospholipids being particularly vulnerable. Consequences of nucleotide and lipid glycation damage are increased DNA mutations and compromised lipid membrane integrity, respectively. Within mitochondria, it is expected that glycation affects matrix and membrane proteins, phospholipids on the outer and inner mitochondrial membranes and mtDNA.


Membrane interactions [19] may also be affected by the formation of lipid glycation adducts which increase membrane fluidity [20]. Lipids possessing a free amino group such as phosphatidylethanolamine are susceptible to glycation, whereas those without, for example phosphatidylcholine, are not glycated [21]. It is thought that lipid glycation may promote lipid peroxidation, resulting in oxidative damage [2224]. For example, in vitro incubation of human low-density lipoprotein (LDL) with 200 mM glucose for up to 12 days increased levels of both glycation and lipid oxidation products, which appears to support the hypothesis that lipid glycation causes, or at least enhances, lipid peroxidation [22]. However, since lipid glycation and peroxidation can both occur in parallel, it is difficult to distinguish one from the other as being the primary initiating event versus a secondary downstream event, particularly in vivo [22].


Likewise, glycation of DNA can have multiple effects such as strand breaks, unwinding of the double helix, mutations and formation of DNA-protein and nucleotide-nucleotide cross-links [2530]. The extent of modification appears to be dependent on the glycating agent used. For example, methylglyoxal induced 10-fold more DNA-protein crosslinks than did glyoxal when incubated in vitro with DNA and DNA polymerase I derived from Escherichia coli [31]. Such DNA-DNA polymerase crosslinks could stall replication and subsequently promote frameshift mutations [31]. The steric hindrance imposed by DNA glycation adducts may also impair transcription by preventing transcription factors from binding [32]. That is, DNA glycation may not only influence genome integrity but could also alter gene expression.


Besides direct glycation damage to biomolecules, AGEs, especially extracellular AGEs, could also contribute to disease pathology by binding to cell surface receptors such as the receptor for AGEs (RAGE), thereby activating the proinflammatory NF-κB pathway and downstream signalling molecules such as p21, MAP kinases, and JNK [3334].


The formation of AGEs arises from reactions between reducing sugars (e.g., glucose) or sugar derivatives (e.g. α,β-dicarbonyls) and amino groups on biological macromolecules (Figure 2). Glycation reactions can be broadly classified as early and late stages of glycation. During early glycation, the carbonyl group of acyclic glucose reacts with amino groups, then undergoes dehydration to form a Schiff base, which subsequently undergoes further rearrangement to form a more stable Amadori product, fructosamine [35] (Figure 2). Alternatively, the Schiff base may undergo spontaneous degradation to form reactive α,β-dicarbonyl species such as methylglyoxal and glyoxal [36] (Figure 3(a)). These dicarbonyls may also be formed from glucose breakdown via the Wolff pathway [37] (Figure 3(a)), triose phosphate fragmentation [38] (Figure 3(b)), acetone [39] and threonine metabolism [40] (Figures 3(c) and 3(d)), lipid peroxidation [41], and fructose-3-phosphate decomposition [42]. Downstream of Amadori product formation, further oxidative and nonoxidative modifications may occur, generating AGEs [5] (Figure 2). Where oxidative processes are involved, the term glycoxidation has also been used [43]. AGEs can also be formed by the direct modification of amino groups by α,β-dicarbonyls (Figure 2).


843505.fig.002


Figure 2: Formation of AGEs in physiological systems. In early glycation, glucose reacts reversibly with amine groups on macromolecules to form a Schiff base and subsequently an Amadori product. This can undergo further oxidative and nonoxidative modifications to form AGEs. Alternatively, AGEs may also be formed by the direct reaction of reactive α,β-dicarbonyls with macromolecules.


fig3


Figure 3: Formation of methylglyoxal and glyoxal in biological systems. (a) The acyclic form of glucose may undergo autooxidation via the Wolff pathway to form glyoxal or break down to 3-deoxyglucosone which in turn degrades to glyceraldehyde and methylglyoxal. Glucose may also react with amine groups on proteins to form a Schiff base which can undergo similar breakdown via the Namiki pathway to also generate glyoxal and methylglyoxal. (b) The major pathway by which methylglyoxal is formed in cells involves the breakdown of the triose phosphates, glyceraldehyde-3-phosphate, and dihydroxyacetone phosphate, via phosphate elimination from an ene-diol intermediate. (c) Methylglyoxal may also be generated during threonine catabolism. This involves oxidation of threonine by threonine dehydrogenase to 2-aminoacetoacetate, followed by spontaneous decarboxylation to aminoacetone. Monoamine oxidase then catalyses the conversion of aminoacetone to methylglyoxal. (d) Another pathway by which methylglyoxal is produced is from acetoacetate metabolism. This reaction proceeds via acetone and acetol and is catalysed by acetone monooxygenase and acetol monooxygenase.


Although glycation is traditionally thought of as a reaction between sugars and amino groups, other functional groups on macromolecules, for example, protein thiols (-SH), may also react in analogous reactions [44]. Similar to glycation of lysine and arginine residues on proteins, glycation of thiol groups can modulate enzyme activity. For example, glyceraldehyde-3-phosphate dehydrogenase can be inhibited by methylglyoxal-mediated loss of thiol groups at its active site [44]. Glycation of creatine kinase by glyoxal also led to enzyme inactivation [45]. Although thiol modification by α,β-dicarbonyls could be damaging to protein function, low molecular weight thiols such as N-acetylcysteine and GSH have been suggested as therapeutic agents for the removal of α,β-dicarbonyls before they can react with proteins [4546].


Despite the common association of glycation with glucose, this reducing sugar is not itself highly reactive. Rather, it is reactive α,β-dicarbonyl species such as methylglyoxal and glyoxal, which are up to 50,000-fold more reactive than glucose, that are thought to be the major physiological precursors of glycation damage [47]. Because these dicarbonyls are mostly produced intracellularly through processes such as triose phosphate fragmentation [38] and lipid peroxidation [41], there is a greater likelihood for cellular proteins to be glycated relative to plasma proteins. Therefore, levels of AGEs tend to be higher in cellular proteins than in plasma proteins. For instance, levels of the methylglyoxal/arginine-derived hydroimidazolone, MG-H1 [Nδ-(5-hydro-5-methyl-4-imidazolon-2-yl)ornithine], were found to be 1.22 mmol/mol arginine in human blood cells and 0.92 mmol/mol arginine in plasma proteins [48]. Similarly, levels of another AGE, CML (Nε-carboxymethyl-lysine), was at least three-fold higher in cellular proteins (0.068–0.233 mmol/mol lysine) than in plasma proteins (0.021 mmol/mol lysine) [48].


To minimize glycation damage by reactive α,β-dicarbonyls, there exist various enzyme systems that catalytically remove these species (Figure 4). One of the most important of these defences is the glyoxalase enzyme system which functions primarily to remove methylglyoxal and, to a lesser extent, other compounds such as glyoxal [4950]. Using glutathione (GSH) as a cofactor, glyoxalase I catalyzes the formation of S-2-hydroxyacylglutathione from methylglyoxal or glyoxal. Glyoxalase II then converts this intermediate compound into an α-hydroxyacid (either D-lactate from methylglyoxal or glycolate from glyoxal), regenerating GSH in the process. Other enzymes involved in the detoxification of α,β-dicarbonyls include aldehyde dehydrogenases which oxidize methylglyoxal and glyoxal to pyruvate and glyoxylate, respectively [51], and aldo-keto reductases/aldose reductases which reduce them to form alcohols (e.g., methylglyoxal to acetol and lactaldehyde, glyoxal to glycoaldehyde and ethylene glycol) [52].


843505.fig.004


Figure 4: Physiological defences against α,β-dicarbonyl precursors of glycation. α,β-dicarbonyls like methylglyoxal (R:CH3) and glyoxal (R:H) can be removed by aldose reductase-catalysed reduction to the corresponding alcohol or by oxidation to pyruvate and glycolate. respectively, catalysed by aldehyde dehydrogenase. Reactive α,β-dicarbonyls may also combine reversibly with reduced glutathione to form a hemithioacetal which can then be reduced by aldose reductases, forming lactaldehyde and glycoaldehyde from methylglyoxal and glyoxal, respectively. Alternatively, the hemithioacetal may undergo a two-step conversion to an α-hydroxyacid, catalysed by glyoxalase I and glyoxalase II. The glyoxalase enzyme system represents the major pathway by which methylglyoxal is removed.


Under normal physiological conditions, these anti-glycation defences are sufficient to prevent significant glycation damage. Any damage that does occur is dealt with by damage repair and removal systems. For example, proteasomes and lysosomes prevent the accumulation of glycated proteins [53]; the nucleotide excision repair system removes glycated nucleotides [54]; lipid turnover helps clear glycated lipids [19]. When there is an imbalance between the production of glycation precursors such as α,β-dicarbonyls and the removal of these species such that the former is favoured, carbonyl stress occurs [55] and glycation damage accumulates.


Because α,β-dicarbonyls are the major precursors of AGEs and their production is largely intracellular, it is likely that there would be significant functional consequences of AGEs formation and accumulation at the organellar level. In this review, we focus on glycation in mitochondria and its association with mitochondrial dysfunction in disease and ageing.


2. Glycation in Mitochondria: Involvement in Diabetes and Ageing


The levels of both intra- and extracellular AGEs have been found to increase in diabetic or old animals relative to healthy or young controls [3610]. Indeed, a glycation adduct of glucose to haemoglobin, HbA1c, is used clinically as an indicator of glycaemic control [5657], and levels of glycation adducts like CML and pentosidine have been suggested as prognostic tools in diabetics [58]. The accumulation of AGEs arises from an increase in their production and/or a decrease in their removal. For instance, low protein turnover is responsible for the accumulation of extracellular AGEs on collagen with age [59]; in diabetics, plasma levels of HbA1c increase as a result of high levels of blood glucose which forms an adduct to the N-terminal valine on the β-chain of haemoglobin [5657].


Within cells, the levels of AGEs also increase with age or in diabetes, and this is likely to reflect changes in intracellular α,β-dicarbonyl levels (Figure 5). In the case of diabetes, the levels of intracellular AGEs rise extremely quickly compared to that of extracellular AGEs and are thought to be a result of the intracellular formation of the reactive dicarbonyl species, methylglyoxal and glyoxal [60]. According to the unifying hypothesis proposed by Brownlee [1], mitochondrial superoxide production increases during hyperglycaemia, resulting in DNA damage and poly(ADP-ribose) polymerase (PARP) activation. ADP-ribose polymers generated by PARP may then bind glyceraldehyde-3-phosphate (G3P) dehydrogenase, inhibiting it. Consequently, there is a build-up of glycolytic intermediates upstream of this enzyme. Among these intermediates are the triose phosphates, G3P and dihydroxyacetone phosphate (DHAP), which break down to form methylglyoxal [38]. In addition, oxidative stress resulting from hyperglycaemia promotes lipid peroxidation [6162], another source of a major α,β-dicarbonyl species, glyoxal [41]. As such, there is increased formation of reactive dicarbonyls in diabetes. With respect to ageing, decreases in glyoxalase I activity have been reported in Caenorhabditis elegans [3] and rats [63]. The less efficient removal of α,β-dicarbonyls consequently leads to elevation of the levels of AGEs in older animals [3].


843505.fig.005


Figure 5: Intra- versus extracellular glycation—involvement in diabetes and ageing. Glycation can take place in the extracellular environment or within cells in the cytosol and in organelles like mitochondria. Extracellular AGEs may arise from oxidative and nonoxidative modifications of the Amadori product or from direct reaction of α,β-dicarbonyls with proteins. Extracellular AGEs may bind to cell surface receptors such as RAGE, thereby activating cell signalling pathways. The formation of lipid-AGEs on the cell surface membrane can further generate reactive α,β-dicarbonyls such as glyoxal. These reactive carbonyl species can diffuse through lipid membranes and enter cells where they react with cellular biomolecules to form intracellular AGEs. They can also diffuse further into mitochondria and similarly cause glycation damage within these organelles. Intracellular glycation may also arise from α,β-dicarbonyls produced during the breakdown of triose phosphates generated during glycolysis. Normally, glycation damage is kept under control by defences such as the glyoxalase enzyme system and aldehyde dehydrogenases. However, in diabetes, glycation damage increases due to elevated formation of α,β-dicarbonyls, arising from high glucose levels and consequent inhibition of the glycolytic enzyme, G3P dehydrogenase, both of which lead to a rise in triose phosphate levels. These in turn break down nonenzymatically to form methylglyoxal, a major precursor of glycation. In addition, activation of RAGE is associated with decreased glyoxalase I expression, which is expected to further raise methylglyoxal levels by preventing its removal. During ageing, glycation damage also increases, but mainly as a result of a loss of glyoxalase activity with age.


Once formed within the cell, these reactive dicarbonyl species may then diffuse across membranes [4764] and access mitochondrial targets (Figure 5). It is estimated that there is 1-2 mmol CML/mol lysine in rat heart mitochondria [65], 0.5–1 mmol CML/mol lysine in rat liver mitochondria [66] and 0.7–1 mmol CML/mol lysine in rat brain mitochondria [67]. Corresponding values for carboxyethyl-lysine (CEL) are 0.5, 0.2–0.5, and 0.1–0.2 mmol CEL/mol lysine [6567]. It is striking that levels of CML and CEL in mitochondria from the three tissues (heart, liver, and brain) are kept within narrow range (<2 mmol AGE/mol lysine). This suggests that excessive glycation of mitochondrial proteins is highly damaging and that levels of mitochondrial AGEs have to be minimized to prevent mitochondrial dysfunction. Likewise, mitochondrial DNA (mtDNA) [68] and lipids [69] are also targets for glycation. Indeed, measurements of CEdG [N2-(1-carboxyethylguanine], a nucleotide glycation adduct, were found up to three-fold more CEdG in mtDNA of cultured fibroblasts than in nuclear DNA [68], suggesting that mtDNA is more vulnerable to glycation than nuclear DNA due to the absence of protective histones in the former [70]. The presence of glycated phosphatidylethanolamine in mitochondrial membranes has also prompted suggestions that mitochondrial lipid glycation could affect electron transport chain activity and cause mitochondrial dysfunction [69].


Intriguingly, specific mitochondrial protein targets of glycation have been identified in kidneys of streptozotocin-induced diabetic rats [71] and in livers of aged rats [18]. Increasing the removal of α,β-dicarbonyls either by administering the prototypical dicarbonyl scavenger, aminoguanidine, or by overexpression of glyoxalase I decreased both glycation and oxidative damage, restored complex III activity, and improved respiration in experimental models of diabetes [7172]. Likewise, increasing glyoxalase I expression in nematodes decreased mitochondrial levels of methylglyoxal and AGEs, and extended lifespan, while inhibiting the enzyme increased methylglyoxal levels and reduced lifespan [3]. These results suggest that glycation of mitochondrial proteins could account, at least in part, for the mitochondrial dysfunction and oxidative damage observed in hyperglycaemia and ageing. Separately, changes in the expression of mitochondrial proteins have been observed in Schwann cells grown under hyperglycaemic conditions [73] and in Akita mice, a mouse model of type 1 diabetes [74]. While the molecular mechanisms inducing these changes have not been elucidated, one study in cultured endothelial cells found that glycation could cause epigenetic changes in expression of nuclear-encoded genes [75], highlighting the potential for glycation to do likewise in mitochondrial genes, thereby prompting the “remodelling of the mitochondrial proteome” as is observed in Akita mice [74].


The hypothesis that glycation of mitochondrial targets profoundly influences mitochondrial function is supported by studies in which exogenous dicarbonyls were administered and mitochondrial parameters studied. For example, methylglyoxal treatment of isolated mitochondria from rat kidney [76] and from several carcinoma cell lines [7779] decreased oxygen consumption by mitochondria. Similarly, treatment of cultured cells with methylglyoxal or glyoxal decreased mitochondrial membrane potential reduced the activities of the respiratory chain complexes, reduced ATP synthesis, and increased reactive oxygen species (ROS) levels [8083]. While these studies all point towards a glycation-induced mitochondrial dysfunction, the physiological relevance of such experiments is uncertain. Some have suggested that the use of millimolar concentrations of dicarbonyls in such experiments is physiologically irrelevant [4784], especially since cellular levels of methylglyoxal have been estimated to be in the low micromolar range. Others, however, have argued that these values, being measures of steady state concentrations of dicarbonyls, do not accurately reflect the actual dicarbonyl flux within cells where α,β-dicarbonyls are produced and that millimolar concentrations are reasonable estimates of the actual flux in vivo [85]. [what about the effect in humans over decades and the synergistic effect of reduce ATP???]  That only a very small proportion of exogenous dicarbonyl becomes incorporated into cells has also been used to justify the use of high concentrations of exogenous dicarbonyls [8386]. It should also be noted that in treating cultured cells with exogenous methylglyoxal or glyoxal, extracellular, cytosolic and mitochondrial levels of these dicarbonyls and their associated AGEs are all increased. Therefore it is unclear whether any mitochondrial dysfunction observed is a direct result of the glycation of mitochondrial targets or if it is a downstream consequence following the glycation of extracellular or cytosolic targets especially since incubation of cells with exogenous AGEs can similarly induce mitochondrial dysfunction [8789].


3. Mitochondrial Glycation in Cancer and Neurodegeneration: A Hypothesis


As Experimental uncertainties notwithstanding, the issue of mitochondrial glycation and its links to disease is scientifically and clinically interesting. Should mitochondrial glycation be a major cause of disease, then a broad strategy of limiting glycation damage in mitochondria may ameliorate disease initiation and progression across multiple clinical conditions. Such a strategy would be akin to that of targeting antioxidants to mitochondria for the treatment of conditions as disparate as Parkinson’s disease, diabetes and ischaemia-reperfusion injury [90]. In this context, it is intriguing to consider whether mitochondrial glycation could contribute to conditions apart from diabetes, and ageing. We discuss here the possibility of this hypothesis in cancer and neurodegeneration.


3.1. Cancer


The Warburg phenotype in which aerobic tumour cells are largely dependent on glycolysis, rather than oxidative phosphorylation, for their energy supply has been well-described in many cancer cell lines [91]. Since triose phosphate intermediates of the glycolytic pathway are an important source of the major glycation precursor, methylglyoxal [38], the levels of methylglyoxal and consequently glycation should be increased in tumour cells. It appears that the type and distribution of such glycation damage varies between tumour types, and not all tumours necessarily display equal extents of glycation [92]. In addition, overexpression of glyoxalase I and glyoxalase II has been correlated with multidrug resistance in tumours [93]. The upregulation of the glyoxalase enzyme system is thought to allow tumour cell growth by counteracting the rise in methylglyoxal production [93]. However, tumour cells with high glyoxalase I expression have higher levels of DNA glycation adducts than those with relatively lower expression [94] suggesting that glyoxalase I expression increases to try to counteract methylglyoxal-induced cytotoxicity [94]. It is therefore not surprising that glyoxalase I inhibitors are being evaluated for use in cancer chemotherapy [95] as they render multidrug-resistant tumours vulnerable to apoptosis [96]. Glyoxalase I inhibitors may potentiate apoptosis by increasing DNA glycation that results in PARP activation which depletes cellular nicotinamide adenine dinucleotide (NAD+) and thereby inhibits G3P dehydrogenase, leading to further increases in methylglyoxal formation [97]. Methylglyoxal also induces cytotoxicity in a variety of other ways such as promoting the mitochondrial permeability transition pore, activation of protein kinase C-delta [93] and inhibition of STAT3-associated signaling [98].


As in diabetes and ageing, mitochondria are profoundly affected in cancers; the upregulation of glycolysis in tumour cells is accompanied by mitochondrial dysfunction and decreased oxidative phosphorylation [99100]. For example, a recent review [101] argues that such mitochondrial dysfunction is critical for tumour growth, citing a strong correlation between tumour progression and increased mitochondrial DNA (mtDNA) mutations. Although the role of mitochondrial glycation in cancers has not been explored, it is possible that elevated glycolysis in cancer cells leads to increased glycation and mutation of mtDNA [75]. Thus up-regulation of glycolysis in tumour cells may contribute to increased mitochondrial damage and thereby establish a vicious cycle that enhances the Warburg effect.


3.2. Neurodegeneration


As triose phosphates are a major source of methylglyoxal [38], any disturbance in triose phosphate metabolism will influence methylglyoxal formation. For example, triose phosphate isomerase (TPI) deficiency results in an inherited neurological disorder [102]. This occurs because TPI catalyzes the interconversion of DHAP and G3P, and its inhibition results in an accumulation of DHAP which breaks down to methylglyoxal [103]. Methylglyoxal-induced protein and nucleotide glycation might result in paralysis and neurodegeneration as is observed in flies expressing a mutant form of TPI [104]. Nitrotyrosinated TPI has been detected in brains from mouse models of, and human patients with, Alzheimer’s disease [105], and this nitration of TPI decreases its enzyme activity and increases methylglyoxal production [105]. Tau aggregation and neurofibrillary tangle formation can also be promoted by methylglyoxal [106107]. These suggest that methylglyoxal and glycation can contribute to the progression of neurodegeneration, and this is supported by observations that glyoxalase I expression increases in the brains of early and middle stage Alzheimer’s patients [108], indicating an attempt to scavenge excessive α,β-dicarbonyls. That type II diabetics are at 2–2.5 fold greater risk of developing dementia also lends weight to a contribution of glycation in neurodegenerative diseases [109]. It has also been proposed that intra- and extracellular AGEs contribute to neurodegeneration by two main pathways, the former by promoting protein aggregation and thus inhibiting their proper function, and the latter by accumulating on senile plaques and inducing oxidative stress and inflammation [110].


The role of mitochondrial dysfunction in neurological disorders has been extensively reviewed [111113]. While mitochondrial glycation was not specifically investigated in these studies, it may be that glycation damage of proteins involved in oxidative phosphorylation and of mtDNA may contribute to the mitochondrial dysfunction observed. Specific protein targets of carbonylation, oxidation, and nitration have been identified in mitochondria in neurodegenerative disorders [114115]. Since glycation and oxidative damage are closely correlated, with both types of damage markers often increasing in parallel with ageing and disease [2224], it is plausible that glycation of specific mitochondrial targets also occurs. A recent study in a mouse model of Alzheimer’s disease further found that mitochondrial dysfunction precedes the presentation of any neurodegenerative pathology [116]. Particularly striking was the observation in the same study of increased glycolysis and decreased oxidative phosphorylation in neurons from these mice, features reminiscent of the Warburg phenotype of tumour cells. The many similarities between the conditions discussed above—diabetes, ageing, and cancer—with neurodegeneration suggest a potential for mitochondrial glycation to contribute to these conditions.


4. Conclusion


In the four conditions discussed above—hyperglycaemia, ageing, cancer, and neurodegeneration—there is increased production of glycation precursors, namely, reactive α,β-dicarbonyls such as methylglyoxal and glyoxal, leading to elevated molecular glycation damage as evidenced by the rise in levels of AGEs. However, despite the correlation between glycation, mitochondrial dysfunction, and disease, the relative importance of mitochondrial glycation as opposed to extracellular or cytosolic glycation is still unclear. This is in part because it is the whole cell levels of α,β-dicarbonyls that are altered by experimental manipulation with aminoguanidine or glyoxalase I overexpression, and also because treatment of cell cultures with exogenous methylglyoxal or glyoxal raises both extra- and intracellular dicarbonyl levels. As such, the effects of any changes are not isolated to mitochondria. Besides, the amount of exogenous dicarbonyl that is physiologically relevant is a further point of contention. Unfortunately, it is technically difficult to accurately measure dicarbonyl levels within mitochondria in cells and in vivo as existing methods are prone to artefacts arising from variations in mitochondria isolation, sample preparation, and derivatization [84117118]. There is also a lack of experimental tools for manipulating dicarbonyl levels within mitochondria alone without modifying cytosolic and extracellular levels. Therefore, it is difficult to isolate the effects of glycation observed in mitochondria from more general consequences on the whole cell. Consequently, to divorce the contribution of mitochondrial glycation from cytosolic or extracellular glycation to disease is experimentally challenging. Nonetheless, the prospect of mitochondrial glycation contributing as a common damaging agent across a broad spectrum of diseases is an intriguing possibility and is also a novel potential therapeutic target.


Abbreviations


ADP:

Adenosine  diphosphate

AGEs:

Advanced glycation end products

ATP:

Adenosine triphosphate

CEdG:

N2-(1-carboxyethylguanine)

CEL:

Carboxyethyl-lysine

CML:

-carboxymethyl-lysine

DHAP:

Dihydroxyacetone phosphate

ECM:

Extracellular matrix

G3P:

Glyceraldehyde-3-phosphate

GSH:

Reduced glutathione

LDL:

Low-density lipoprotein

MG-H1:

[-(5-hydro-5-methyl-4-imidazolon-2-yl)ornithine]

MOLD:

Methylglyoxal-derived lysine dimers

mtDNA:

Mitochondrial DNA

:

Nicotinamide adenine dinucleotide

PARP:

Poly(ADP-ribose) polymerase

RAGE:

Receptor for AGEs

ROS:

Reactive oxygen species

TPI:

Triose phosphate isomerase.


Acknowledgments


The authors would like to thank the Medical Research Council and the Cambridge Commonwealth Trust for funding our work.


References


1.      M. Brownlee, “Biochemistry and molecular cell biology of diabetic complications,” Nature, vol. 414, no. 6865, pp. 813–820, 2001. View at Publisher · View at Google Scholar · View at Scopus


2.      N. Rabbani and P. J. Thornalley, “Glyoxalase in diabetes, obesity and related disorders,” Seminars in Cell and Developmental Biology, vol. 22, no. 3, pp. 309–317, 2011. View at Publisher · View at Google Scholar ·View at Scopus


3.      M. Morcos, X. Du, F. Pfisterer et al., “Glyoxalase-1 prevents mitochondrial protein modification and enhances lifespan in Caenorhabditis elegans,” Aging Cell, vol. 7, no. 2, pp. 260–269, 2008. View at Publisher · View at Google Scholar · View at Scopus


4.      M. Xue, N. Rabbani, and P. J. Thornalley, “Glyoxalase in ageing,” Seminars in Cell and Developmental Biology, vol. 22, no. 3, pp. 293–301, 2011. View at Publisher · View at Google Scholar · View at Scopus


5.      A. Cerami, “Aging of proteins and nucleic acids: what is the role of glucose?” Trends in Biochemical Sciences, vol. 11, no. 8, pp. 311–314, 1986. View at Google Scholar · View at Scopus


6.      P. J. Thornalley, S. Battah, N. Ahmed et al., “Quantitative screening of advanced glycation endproducts in cellular and extracellular proteins by tandem mass spectrometry,” Biochemical Journal, vol. 375, no. 3, pp. 581–592, 2003. View at Publisher · View at Google Scholar · View at Scopus


7.      D. Dobler, N. Ahmed, L. Song, K. E. Eboigbodin, and P. J. Thornalley, “Increased dicarbonyl metabolism in endothelial cells in hyperglycemia induces anoikis and impairs angiogenesis by RGD and GFOGER motif modification,” Diabetes, vol. 55, no. 7, pp. 1961–1969, 2006. View at Publisher · View at Google Scholar · View at Scopus


8.      H. Li, S. Nakamura, S. Miyazaki et al., “N2-carboxyethyl-2′-deoxyguanosine, a DNA glycation marker, in kidneys and aortas of diabetic and uremic patients,” Kidney International, vol. 69, no. 2, pp. 388–392, 2006. View at Publisher · View at Google Scholar · View at Scopus


9.      A. J. Oudes, C. M. Herr, Y. Olsen, and J. E. Fleming, “Age-dependent accumulation of advanced glycation end-products in adult Drosophila melanogaster,” Mechanisms of Ageing and Development, vol. 100, no. 3, pp. 221–229, 1998. View at Publisher · View at Google Scholar · View at Scopus


10. J. M. Haus, J. A. Carrithers, S. W. Trappe, and T. A. Trappe, “Collagen, cross-linking, and advanced glycation end products in aging human skeletal muscle,” Journal of Applied Physiology, vol. 103, no. 6, pp. 2068–2076, 2007. View at Publisher · View at Google Scholar · View at Scopus


11. P. J. Thornalley, “The enzymatic defence against glycation in health, disease and therapeutics: a symposium to examine the concept,” Biochemical Society Transactions, vol. 31, no. 6, pp. 1341–1342, 2003. View at Google Scholar · View at Scopus


12. J. de Groot, N. Verzijl, M. J. G. Wenting-van Wijk et al., “Age-related decrease in susceptibility of human articular cartilage to matrix metalloproteinase-mediated degradation—the role of advanced glycation end products,” Arthritis & Rheumatism, vol. 44, no. 11, pp. 2562–2571, 2001. View at Google Scholar


13. M. Luthra and D. Balasubramanian, “Nonenzymatic glycation alters protein structure and stability. A study of two eye lens crystallins,” Journal of Biological Chemistry, vol. 268, no. 24, pp. 18119–18127, 1993.View at Google Scholar · View at Scopus


14. A. J. Bailey, R. G. Paul, and L. Knott, “Mechanisms of maturation and ageing of collagen,” Mechanisms of Ageing and Development, vol. 106, no. 1-2, pp. 1–56, 1998. View at Publisher · View at Google Scholar ·View at Scopus


15. P. Ulrich and A. Cerami, “Protein glycation, diabetes, and aging,” Recent Progress in Hormone Research, vol. 56, pp. 1–21, 2001. View at Publisher · View at Google Scholar · View at Scopus


16. M. J. C. Crabbe, L. R. Cooper, and D. W. Corne, “Use of essential and molecular dynamics to study γB-crystallin unfolding after non-enzymic post-translational modifications,” Computational Biology and Chemistry, vol. 27, no. 4-5, pp. 507–510, 2003. View at Publisher · View at Google Scholar · View at Scopus


17. D. R. Sell and V. M. Monnier, “Ornithine is a novel amino acid and a marker of arginine damage by oxoaldehydes in senescent proteins,” Annals of the New York Academy of Sciences, vol. 1043, pp. 118–128, 2005. View at Publisher · View at Google Scholar · View at Scopus


18. M. Hamelin, J. Mary, M. Vostry, B. Friguet, and H. Bakala, “Glycation damage targets glutamate dehydrogenase in the rat liver mitochondrial matrix during aging,” FEBS Journal, vol. 274, no. 22, pp. 5949–5961, 2007. View at Publisher · View at Google Scholar · View at Scopus


19. J. R. Requena, M. U. Ahmed, C. W. Fountain et al., “Carboxymethylethanolamine, a biomarker of phospholipid modification during the Maillard reaction in vivo,” Journal of Biological Chemistry, vol. 272, no. 28, pp. 17473–17479, 1997. View at Publisher · View at Google Scholar · View at Scopus


20. A. Ravandi, A. Kuksis, L. Marai et al., “Isolation and identification of glycated aminophospholipids from red cells and plasma of diabetic blood,” FEBS Letters, vol. 381, no. 1-2, pp. 77–81, 1996. View at Publisher· View at Google Scholar · View at Scopus


21. R. Pamplona, M. J. Bellmunt, M. Portero, D. Riba, and J. Prat, “Chromatographic evidence for Amadori product formation in rat liver aminophospholipids,” Life Sciences, vol. 57, no. 9, pp. 873–879, 1995. View at Publisher · View at Google Scholar · View at Scopus


22. R. Bucala, Z. Makita, T. Koschinsky, A. Cerami, and H. Vlassara, “Lipid advanced glycosylation: pathway for lipid oxidation in vivo,” Proceedings of the National Academy of Sciences of the United States of America, vol. 90, no. 14, pp. 6434–6438, 1993. View at Google Scholar · View at Scopus


23. C. M. Breitling-Utzmann, A. Unger, D. A. Friedl, and M. O. Lederer, “Identification and quantification of phosphatidylethanolamine-derived glucosylamines and aminoketoses from human erythrocytes—influence of glycation products on lipid peroxidation,” Archives of Biochemistry and Biophysics, vol. 391, no. 2, pp. 245–254, 2001. View at Publisher · View at Google Scholar · View at Scopus


24. K. Nakagawa, J. H. Oak, and T. Miyazawa, “Synthetically prepared Amadori-glycated phosphatidylethanolamine can trigger lipid peroxidation via free radical reactions,” FEBS Letters, vol. 481, no. 1, pp. 26–30, 2000. View at Publisher · View at Google Scholar · View at Scopus


25. H. J. C. Chen and Y. C. Chen, “Analysis of glyoxal-induced DNA cross-links by capillary liquid chromatography nanospray ionization tandem mass spectrometry,” Chemical Research in Toxicology, vol. 22, no. 7, pp. 1334–1341, 2009. View at Publisher · View at Google Scholar · View at Scopus


26. N. Murata-Kamiya, H. Kamiya, H. Kaji, and H. Kasai, “Glyoxal, a major product of DNA oxidation, induces mutations at G:C sites on a shuttle vector plasmid replicated in mammalian cells,” Nucleic Acids Research, vol. 25, no. 10, pp. 1897–1902, 1997. View at Publisher · View at Google Scholar · View at Scopus


27. H. Kasai, N. Iwamoto-Tanaka, and S. Fukada, “DNA modifications by the mutagen glyoxal: adduction to G and C, deamination of C and GC and GA cross-linking,” Carcinogenesis, vol. 19, no. 8, pp. 1459–1465, 1998. View at Publisher · View at Google Scholar · View at Scopus


28. M. Pischetsrieder, W. Seidel, G. Münch, and R. Schinzel, “N2-(1-carboxyethyl)deoxyguanosine, a nonenzymatic glycation adduct of DNA, induces single-strand breaks and increases mutation frequencies,” Biochemical and Biophysical Research Communications, vol. 264, no. 2, pp. 544–549, 1999.View at Publisher · View at Google Scholar · View at Scopus


29. G. E. Wuenschell, D. Tamae, A. Cercillieux, R. Yamanaka, C. Yu, and J. Termini, “Mutagenic potential of DNA glycation: miscoding by (R)- and (S)-N 2-(1-carboxyethyl)2′-deoxyguanosine,” Biochemistry, vol. 49, no. 9, pp. 1814–1821, 2010. View at Publisher · View at Google Scholar · View at Scopus


30. W. Seidel and M. Pischetsrieder, “DNA-glycation leads to depurination by the loss of N2-carboxyethylguanine in vitro,” Cellular and Molecular Biology, vol. 44, no. 7, pp. 1165–1170, 1998. View at Google Scholar · View at Scopus


31. N. Murata-Kamiya and H. Kamiya, “Methylglyoxal, and endogenous aldehyde, crosslinks DNA polymerase and the substrate DNA,” Nucleic Acids Research, vol. 29, no. 16, pp. 3433–3438, 2001. View at Google Scholar · View at Scopus


32. V. Breyer, M. Frischmann, C. Bidmon, A. Schemm, K. Schiebel, and M. Pischetsrieder, “Analysis and biological relevance of advanced glycation end-products of DNA in eukaryotic cells,” FEBS Journal, vol. 275, no. 5, pp. 914–925, 2008. View at Publisher · View at Google Scholar · View at Scopus


33. S. F. Yan, R. Ramasamy, Y. Naka, and A. M. Schmidt, “Glycation, inflammation, and RAGE: a scaffold for the macrovascular complications of diabetes and beyond,” Circulation Research, vol. 93, no. 12, pp. 1159–1169, 2003. View at Publisher · View at Google Scholar · View at Scopus


34. A. M. Schmidt, S. F. Yan, and R. Ramasamy, “Mechanisms of Disease: advanced glycation end-products and their receptor in inflammation and diabetes complications,” Nature Clinical Practice Endocrinology & Metabolism, vol. 4, no. 5, pp. 285–293, 2008. View at Publisher · View at Google Scholar · View at Scopus


35. J. E. Hodge, “The Amadori rearrangement,” Advances in Carbohydrate Chemistry, vol. 10, pp. 169–205, 1955. View at Publisher · View at Google Scholar · View at Scopus


36. T. Hayashi and M. Namiki, “Formation of 2-carbon sugar fragment at an early stage of the browning reaction of sugar with amine,” Agricultural and Biological Chemistry, vol. 44, no. 11, pp. 2575–2580, 1980.View at Google Scholar


37. P. Thornalley, S. Wolff, J. Crabbe, and A. Stern, “The autoxidation of glyceraldehyde and other simple monosaccharides under physiological conditions catalysed by buffer ions,” Biochimica et Biophysica Acta, vol. 797, no. 2, pp. 276–287, 1984. View at Publisher · View at Google Scholar · View at Scopus


38. S. A. Phillips and P. J. Thornalley, “The formation of methylglyoxal from triose phosphates. Investigation using a specific assay for methylglyoxal,” European Journal of Biochemistry, vol. 212, no. 1, pp. 101–105, 1993. View at Google Scholar · View at Scopus


39. J. P. Casazza, M. E. Felver, and R. L. Veech, “The metabolism of acetone in rat,” Journal of Biological Chemistry, vol. 259, no. 1, pp. 231–236, 1984. View at Google Scholar · View at Scopus


40. M. Ray and S. Ray, “L-Threonine dehydrogenase from goat liver. Feedback inhibition by methylglyoxal,” Journal of Biological Chemistry, vol. 260, no. 10, pp. 5913–5918, 1985. View at Google Scholar · View at Scopus


41. A. Loidl-Stahlhofen and G. Spiteller, “Aλπηα-hydroxyaldehydes, products of lipid peroxidation,” Biochimica et Biophysica Acta, vol. 1211, no. 2, pp. 156–160, 1994. View at Publisher · View at Google Scholar · View at Scopus


42. P. J. Thornalley, A. Langborg, and H. S. Minhas, “Formation of glyoxal, methylglyoxal and 8-deoxyglucosone in the glycation of proteins by glucose,” Biochemical Journal, vol. 344, no. 1, pp. 109–116, 1999. View at Publisher · View at Google Scholar · View at Scopus


43. J. W. Baynes, “Role of oxidative stress in development of complications in diabetes,” Diabetes, vol. 40, no. 4, pp. 405–412, 1991. View at Google Scholar · View at Scopus


44. P. E. Morgan, R. T. Dean, and M. J. Davies, “Inactivation of cellular enzymes by carbonyls and protein-bound glycation/glycoxidation products,” Archives of Biochemistry and Biophysics, vol. 403, no. 2, pp. 259–269, 2002. View at Publisher · View at Google Scholar · View at Scopus


45. J. Zeng and M. J. Davies, “Protein and low molecular mass thiols as targets and inhibitors of glycation reactions,” Chemical Research in Toxicology, vol. 19, no. 12, pp. 1668–1676, 2006. View at Publisher ·View at Google Scholar · View at Scopus


46. J. M. Aćimović, B. D. Stanimirović, N. Todorović, V. B. Jovanović, and L. M. Mandić, “Influence of the microenvironment of thiol groups in low molecular mass thiols and serum albumin on the reaction with methylglyoxal,” Chemico-Biological Interactions, vol. 188, no. 1, pp. 21–30, 2010. View at Publisher · View at Google Scholar · View at Scopus


47. N. Rabbani and P. J. Thornalley, “Dicarbonyls linked to damage in the powerhouse: glycation of mitochondrial proteins and oxidative stress,” Biochemical Society Transactions, vol. 36, no. 5, pp. 1045–1050, 2008. View at Publisher · View at Google Scholar · View at Scopus


48. N. Ahmed and P. J. Thornalley, “Quantitative screening of protein biomarkers of early glycation, advanced glycation, oxidation and nitrosation in cellular and extracellular proteins by tandem mass spectrometry multiple reaction monitoring,” Biochemical Society Transactions, vol. 31, no. 6, pp. 1417–1422, 2003. View at Google Scholar · View at Scopus


49. P. J. Thornalley, “The glyoxalase system: new developments towards functional characterization of a metabolic pathway fundamental to biological life,” Biochemical Journal, vol. 269, no. 1, pp. 1–11, 1990.View at Google Scholar · View at Scopus


50. P. J. Thornalley, “Glyoxalase I—structure, function and a critical role in the enzymatic defence against glycation,” Biochemical Society Transactions, vol. 31, no. 6, pp. 1343–1348, 2003. View at Google Scholar ·View at Scopus


51. G. Izaguirre, A. Kikonyogo, and R. Pietruszko, “Methylglyoxal as substrate and inhibitor of human aldehyde dehydrogenase: comparison of kinetic properties among the three isozymes,” Comparative Biochemistry and Physiology, vol. 119, no. 4, pp. 747–754, 1998. View at Publisher · View at Google Scholar · View at Scopus


52. D. L. Vanderjagt, B. Robinson, K. K. Taylor, and L. A. Hunsaker, “Reduction of trioses by NADPH-dependent aldo-keto reductases. Aldose reductase, methylglyoxal, and diabetic complications,” Journal of Biological Chemistry, vol. 267, no. 7, pp. 4364–4369, 1992. View at Google Scholar · View at Scopus


53. A. Stolzing, R. Widmer, T. Jung, P. Voss, and T. Grune, “Degradation of glycated bovine serum albumin in microglial cells,” Free Radical Biology and Medicine, vol. 40, no. 6, pp. 1017–1027, 2006. View at Publisher · View at Google Scholar · View at Scopus


54. D. Tamae, P. Lim, G. E. Wuenschell, and J. Termini, “Mutagenesis and repair induced by the DNA advanced glycation end product N 2-1-(carboxyethyl)-2′-deoxyguanosine in human cells,” Biochemistry, vol. 50, no. 12, pp. 2321–2329, 2011. View at Publisher · View at Google Scholar · View at Scopus


55. N. Rabbani and P. J. Thornalley, “Glycation research in amino acids: a place to call home,” Amino Acids, vol. 42, no. 2, pp. 1087–1096, 2012. View at Publisher · View at Google Scholar · View at Scopus


56. D. B. Sacks and R. R. Little, “HbA1c: how do we measure it and what does it mean?” Current Opinion in Endocrinology, Diabetes and Obesity, vol. 16, no. 2, pp. 113–118, 2009. View at Publisher · View at Google Scholar · View at Scopus


57. S. Rahbar, “The discovery of glycated hemoglobin: a major event in the study of nonenzymatic chemistry in biological systems,” Annals of the New York Academy of Sciences, vol. 1043, pp. 9–19, 2005. View at Publisher · View at Google Scholar · View at Scopus


58. V. M. Monnier, D. R. Sell, and S. Genuth, “Glycation products as markers and predictors of the progression of diabetic complications,” Annals of the New York Academy of Sciences, vol. 1043, pp. 567–581, 2005. View at Publisher · View at Google Scholar · View at Scopus


59. N. Verzijl, J. DeGroot, S. R. Thorpe et al., “Effect of collagen turnover on the accumulation of advanced glycation end products,” Journal of Biological Chemistry, vol. 275, no. 50, pp. 39027–39031, 2000. View at Publisher · View at Google Scholar · View at Scopus


60. M. Brownlee, “Negative consequences of glycation,” Metabolism, vol. 49, no. 2, pp. 9–13, 2000. View at Google Scholar · View at Scopus


61. H. Ha, S. J. Yoon, and K. H. Kim, “High glucose can induce lipid peroxidation in the isolated rat glomeruli,” Kidney International, vol. 46, no. 6, pp. 1620–1626, 1994. View at Google Scholar · View at Scopus


62. A. Mezzetti, F. Cipollone, and F. Cuccurullo, “Oxidative stress and cardiovascular complications in diabetes: isoprostanes as new markers on an old paradigm,” Cardiovascular Research, vol. 47, no. 3, pp. 475–488, 2000. View at Publisher · View at Google Scholar · View at Scopus


63. T. Armeni, C. Pieri, M. Marra, F. Saccucci, and G. Principato, “Studies on the life prolonging effect of food restriction: glutathione levels and glyoxalase enzymes in rat liver,” Mechanisms of Ageing and Development, vol. 101, no. 1-2, pp. 101–110, 1998. View at Google Scholar


64. P. J. Thornalley, “Modification of the glyoxalase system in human red blood cells by glucose in vitro,” Biochemical Journal, vol. 254, no. 3, pp. 751–755, 1988. View at Google Scholar · View at Scopus


65. R. Pamplona, M. Portero-Otin, M. J. Bellmunt, R. Gredilla, and G. Barja, “Erratum: aging increases Nepsilon-(carboxymethyl)lysine and caloric restriction decreases Nepsilon-(carboxymethyl)lysine and Nepsilon-(malondialdehyde)lysine in rat heart mitochondrial proteins (Free Radical Research, vol. 36, no.1, pp. 47–54, 2002),” Free Radical Research, vol. 36, no. 2, 2002. View at Google Scholar · View at Scopus


66. A. J. Lambert, M. Portero-Otin, R. Pamplona, and B. J. Merry, “Effect of ageing and caloric restriction on specific markers of protein oxidative damage and membrane peroxidizability in rat liver mitochondria,” Mechanisms of Ageing and Development, vol. 125, no. 8, pp. 529–538, 2004. View at Publisher · View at Google Scholar · View at Scopus


67. J. J. Ochoa, R. Pamplona, M. C. Ramirez-Tortosa et al., “Age-related changes in brain mitochondrial DNA deletion and oxidative stress are differentially modulated by dietary fat type and coenzyme Q 10,” Free Radical Biology and Medicine, vol. 50, no. 9, pp. 1053–1064, 2011. View at Publisher · View at Google Scholar · View at Scopus


68. V. Breyer, C. M. Becker, and M. Pischetsrieder, “Intracellular glycation of nuclear DNA, mitochondrial DNA, and cytosolic proteins during senescence-like growth arrest,” DNA and Cell Biology, vol. 30, no. 9, pp. 681–689, 2011. View at Google Scholar


69. R. Pamplona, J. R. Requena, M. Portero-Otín, J. Prat, S. R. Thorpe, and M. J. Bellmunt, “Carboxymethylated phosphatidylethanolamine in mitochondrial membranes of mammals—evidence for intracellular lipid glycoxidation,” European Journal of Biochemistry, vol. 255, no. 3, pp. 685–689, 1998.View at Publisher · View at Google Scholar · View at Scopus


70. J. W. Baynes, “The Maillard hypothesis on aging: time to focus on DNA,” Annals of the New York Academy of Sciences, vol. 959, pp. 360–367, 2002. View at Google Scholar · View at Scopus


71. M. G. Rosca, T. G. Mustata, M. T. Kinter et al., “Glycation of mitochondrial proteins from diabetic rat kidney is associated with excess superoxide formation,” American Journal of Physiology, vol. 289, no. 2, pp. F420–F430, 2005. View at Publisher · View at Google Scholar · View at Scopus


72. O. Brouwers, P. M. Niessen, I. Ferreira et al., “Overexpression of glyoxalase-I reduces hyperglycemiainduced levels of advanced glycation end products and oxidative stress in diabetic rats,” Journal of Biological Chemistry, vol. 286, no. 2, pp. 1374–1380, 2011. View at Publisher · View at Google Scholar · View at Scopus


73. L. Zhang, C. Yu, F. E. Vasquez et al., “Hyperglycemia alters the Schwann cell mitochondrial proteome and decreases coupled respiration in the absence of superoxide production,” Journal of Proteome Research, vol. 9, no. 1, pp. 458–471, 2010. View at Publisher · View at Google Scholar · View at Scopus


74. H. Bugger, C. Dong, C. Riehle et al., “Tissue-specific remodeling of the mitochondrial proteome in type 1 diabetic akita mice,” Diabetes, vol. 58, no. 9, pp. 1986–1997, 2009. View at Publisher · View at Google Scholar · View at Scopus


75. A. El-Osta, D. Brasacchio, D. Yao et al., “Transient high glucose causes persistent epigenetic changes and altered gene expression during subsequent normoglycemia,” Journal of Experimental Medicine, vol. 205, no. 10, pp. 2409–2417, 2008. View at Google Scholar


76. M. G. Rosca, V. M. Monnier, L. I. Szweda, and M. F. Weiss, “Alterations in renal mitochondrial respiration in response to the reactive oxoaldehyde methylglyoxal,” American Journal of Physiology, vol. 283, no. 1, pp. F52–F59, 2002. View at Google Scholar · View at Scopus


77. S. Biswas, M. Ray, S. Misra, D. P. Dutta, and S. Ray, “Selective inhibition of mitochondrial respiration and glycolysis in human leukaemic leucocytes by methylglyoxal,” Biochemical Journal, vol. 323, no. 2, pp. 343–348, 1997. View at Google Scholar · View at Scopus


78. S. Ray, S. Dutta, J. Halder, and M. Ray, “Inhibition of electron flow through complex I of the mitochondrial respiratory chain of Ehrlich ascites carcinoma cells by methylglyoxal,” Biochemical Journal, vol. 303, no. 1, pp. 69–72, 1994. View at Google Scholar · View at Scopus


79. A. Ghosh, S. Bera, S. Ray, T. Banerjee, and M. Ray, “Methylglyoxal induces mitochondria-dependent apoptosis in sarcoma,” Biochemistry, vol. 76, no. 10, pp. 1164–1171, 2011. View at Publisher · View at Google Scholar · View at Scopus


80. N. Shangari and P. J. O'Brien, “The cytotoxic mechanism of glyoxal involves oxidative stress,” Biochemical Pharmacology, vol. 68, no. 7, pp. 1433–1442, 2004. View at Publisher · View at Google Scholar · View at Scopus


81. N. Shangari, R. Mehta, and P. J. O'Brien, “Hepatocyte susceptibility to glyoxal is dependent on cell thiamin content,” Chemico-Biological Interactions, vol. 165, no. 2, pp. 146–154, 2007. View at Publisher ·View at Google Scholar · View at Scopus


82. S. G. de Arriba, G. Stuchbury, J. Yarin, J. Burnell, C. Loske, and G. Münch, “Methylglyoxal impairs glucose metabolism and leads to energy depletion in neuronal cells—protection by carbonyl scavengers,” Neurobiology of Aging, vol. 28, no. 7, pp. 1044–1050, 2007. View at Publisher · View at Google Scholar ·View at Scopus


83. H. Wang, J. Liu, and L. Wu, “Methylglyoxal-induced mitochondrial dysfunction in vascular smooth muscle cells,” Biochemical Pharmacology, vol. 77, no. 11, pp. 1709–1716, 2009. View at Publisher · View at Google Scholar · View at Scopus


84. P. J. Thornalley, “Protein and nucleotide damage by glyoxal and methylglyoxal in physiological systems—role in ageing and disease,” Drug Metabolism and Drug Interactions, vol. 23, no. 1-2, pp. 125–150, 2008.View at Google Scholar · View at Scopus


85. J. Zeng, R. A. Dunlop, K. J. Rodgers, and M. J. Davies, “Evidence for inactivation of cysteine proteases by reactive carbonyls via glycation of active site thiols,” Biochemical Journal, vol. 398, no. 2, pp. 197–206, 2006. View at Publisher · View at Google Scholar · View at Scopus


86. A. Riboulet-Chavey, A. Pierron, I. Durand, J. Murdaca, J. Giudicelli, and E. van Obberghen, “Methylglyoxal impairs the insulin signaling pathways independently of the formation of intracellular reactive oxygen species,” Diabetes, vol. 55, no. 5, pp. 1289–1299, 2006. View at Publisher · View at Google Scholar · View at Scopus


87. M. C. Lo, C. I. Lu, M. H. Chen, C. D. Chen, H. M. Lee, and S. H. Kao, “Glycoxidative stress-induced mitophagy modulates mitochondrial fates,” Annals of the New York Academy of Sciences, vol. 1201, pp. 1–7, 2010. View at Publisher · View at Google Scholar · View at Scopus


88. H. Ma, S. Y. Li, P. Xu et al., “Advanced glycation endproduct (AGE) accumulation and AGE receptor (RAGE) up-regulation contribute to the onset of diabetic cardiomyopathy,” Journal of Cellular and Molecular Medicine, vol. 13, no. 8B, pp. 1751–1764, 2009. View at Publisher · View at Google Scholar ·View at Scopus


89. G. V. Sangle, S. K. R. Chowdhury, X. Xie, G. L. Stelmack, A. J. Halayko, and G. X. Shen, “Impairment of mitochondrial respiratory chain activity in aortic endothelial cells induced by glycated low-density lipoprotein,” Free Radical Biology and Medicine, vol. 48, no. 6, pp. 781–790, 2010. View at Publisher ·View at Google Scholar · View at Scopus


90. R. A. J. Smith, V. J. Adlam, F. H. Blaikie et al., “Mitochondria-targeted antioxidants in the treatment of disease,” Annals of the New York Academy of Sciences, vol. 1147, pp. 105–111, 2008. View at Publisher ·View at Google Scholar · View at Scopus


91. P. P. Hsu and D. M. Sabatini, “Cancer cell metabolism: warburg and beyond,” Cell, vol. 134, no. 5, pp. 703–707, 2008. View at Publisher · View at Google Scholar · View at Scopus


92. J. W. J. van Heijst, H. W. M. Niessen, K. Hoekman, and C. G. Schalkwijk, “Advanced glycation end products in human cancer tissues: detection of Nε-(carboxymethyl)lysine and argpyrimidine,” Annals of the New York Academy of Sciences, vol. 1043, pp. 725–733, 2005. View at Publisher · View at Google Scholar · View at Scopus


93. P. J. Thornalley and N. Rabbani, “Glyoxalase in tumourigenesis and multidrug resistance,” Seminars in Cell and Developmental Biology, vol. 22, no. 3, pp. 318–325, 2011. View at Publisher · View at Google Scholar · View at Scopus


94. P. J. Thornalley, S. Waris, T. Fleming et al., “Imidazopurinones are markers of physiological genomic damage linked to DNA instability and glyoxalase 1-associated tumour multidrug resistance,” Nucleic Acids Research, vol. 38, no. 16, Article ID gkq306, pp. 5432–5442, 2010. View at Publisher · View at Google Scholar · View at Scopus


95. D. J. Creighton, Z. B. Zheng, R. Holewinski, D. S. Hamilton, and J. L. Eiseman, “Glyoxalase I inhibitors in cancer chemotherapy,” Biochemical Society Transactions, vol. 31, no. 6, pp. 1378–1382, 2003. View at Google Scholar · View at Scopus


96. P. J. Thornalley, L. G. Edwards, Y. Kang et al., “Antitumour activity of S-p-bromobenzylglutathione cyclopentyl diester in vitro and in vivo. Inhibition of glyoxalase I and induction of apoptosis,” Biochemical Pharmacology, vol. 51, no. 10, pp. 1365–1372, 1996. View at Publisher · View at Google Scholar · View at Scopus


97. P. J. Thornalley, “Protecting the genome: defence against nucleotide glycation and emerging role of glyoxalase I overexpression in multidrug resistance in cancer chemotherapy,” Biochemical Society Transactions, vol. 31, no. 6, pp. 1372–1377, 2003. View at Google Scholar · View at Scopus


98. H. K. Lee, I. A. Seo, D. J. Suh, H. J. Lee, and H. T. Park, “A novel mechanism of methylglyoxal cytotoxicity in neuroglial cells,” Journal of Neurochemistry, vol. 108, no. 1, pp. 273–284, 2009. View at Publisher ·View at Google Scholar · View at Scopus


99. G. Kroemer, “Mitochondria in cancer,” Oncogene, vol. 25, no. 34, pp. 4630–4632, 2006. View at Publisher· View at Google Scholar · View at Scopus


100.                     D. C. Wallace, “Mitochondria and cancer: warburg addressed,” Cold Spring Harbor Symposia on Quantitative Biology, vol. 70, pp. 363–374, 2005. View at Publisher · View at Google Scholar · View at Scopus


101.                     L. Formentini, I. Martínez-Reyes, and J. M. Cuezva, “The mitochondrial bioenergetic capacity of carcinomas,” IUBMB Life, vol. 62, no. 7, pp. 554–560, 2010. View at Publisher · View at Google Scholar ·View at Scopus


102.                     F. Orosz, J. Oláh, and J. Ovádi, “Triosephosphate isomerase deficiency: new insights into an enigmatic disease,” Biochimica et Biophysica Acta, vol. 1792, no. 12, pp. 1168–1174, 2009. View at Publisher · View at Google Scholar · View at Scopus


103.                     N. Ahmed, S. Battah, N. Karachalias et al., “Increased formation of methylglyoxal and protein glycation, oxidation and nitrosation in triosephosphate isomerase deficiency,” Biochimica et Biophysica Acta, vol. 1639, no. 2, pp. 121–132, 2003. View at Publisher · View at Google Scholar · View at Scopus


104.                     J. P. Gnerer, R. A. Kreber, and B. Ganetzky, “wasted away, a Drosophila mutation in triosephosphate isomerase, causes paralysis, neurodegeneration, and early death,” Proceedings of the National Academy of Sciences of the United States of America, vol. 103, no. 41, pp. 14987–14993, 2006. View at Publisher · View at Google Scholar · View at Scopus


105.                     F. X. Guix, G. Ill-Raga, R. Bravo et al., “Amyloid-dependent triosephosphate isomerase nitrotyrosination induces glycation and tau fibrillation,” Brain, vol. 132, no. 5, pp. 1335–1345, 2009. View at Publisher ·View at Google Scholar · View at Scopus


106.                     B. Kuhla, C. Haase, K. Flach, H. J. Lüth, T. Arendt, and G. Münch, “Effect of pseudophosphorylation and cross-linking by lipid peroxidation and advanced glycation end product precursors on tau aggregation and filament formation,” Journal of Biological Chemistry, vol. 282, no. 10, pp. 6984–6991, 2007. View at Publisher · View at Google Scholar · View at Scopus


107.                     S. D. Yan, S. F. Yan, X. Chen et al., “Non-enzymatically glycated tau in Alzheimer's disease induces neuronal oxidant stress resulting in cytokine gene expression and release of amyloid β-peptide,” Nature Medicine, vol. 1, no. 7, pp. 693–699, 1995. View at Google Scholar · View at Scopus


108.                     B. Kuhla, K. Boeck, A. Schmidt et al., “Age- and stage-dependent glyoxalase I expression and its activity in normal and Alzheimer's disease brains,” Neurobiology of Aging, vol. 28, no. 1, pp. 29–41, 2007. View at Publisher · View at Google Scholar · View at Scopus


109.                     R. N. Kalaria, “Neurodegenerative disease: diabetes, microvascular pathology and Alzheimer disease,” Nature Reviews Neurology, vol. 5, no. 6, pp. 305–306, 2009. View at Publisher · View at Google Scholar ·View at Scopus


110.                     G. Münch and M. Krautwald, “Advanced glycation end products as biomarkers and gerontotoxins—a basis to explore methylglyoxal-lowering agents for Alzheimer's disease?” Experimental Gerontology, vol. 45, no. 10, pp. 744–751, 2010. View at Publisher · View at Google Scholar · View at Scopus


111.                     G. Fiskum, A. N. Murphy, and M. F. Beal, “Mitochondria in neurodegeneration: acute ischemia and chronic neurodegenerative diseases,” Journal of Cerebral Blood Flow and Metabolism, vol. 19, no. 4, pp. 351–369, 1999. View at Google Scholar · View at Scopus


112.                     A. N. Murphy, G. Fiskum, and M. F. Beal, “Mitochondria in neurodegeneration: bioenergetic function in cell life and death,” Journal of Cerebral Blood Flow and Metabolism, vol. 19, no. 3, pp. 231–245, 1999.View at Google Scholar · View at Scopus


113.                     M. F. Beal, “Mitochondria take center stage in aging and neurodegeneration,” Annals of Neurology, vol. 58, no. 4, pp. 495–505, 2005. View at Publisher · View at Google Scholar · View at Scopus


114.                     R. Sultana, M. Perluigi, and D. A. Butterfield, “Oxidatively modified proteins in Alzheimer's disease (AD), mild cognitive impairment and animal models of AD: role of Abeta in pathogenesis,” Acta Neuropathologica, vol. 118, no. 1, pp. 131–150, 2009. View at Publisher · View at Google Scholar · View at Scopus


115.                     A. Martínez, M. Portero-Otin, R. Pamplona, and I. Ferrer, “Protein targets of oxidative damage in human neurodegenerative diseases with abnormal protein aggregates,” Brain Pathology, vol. 20, no. 2, pp. 281–297, 2010. View at Publisher · View at Google Scholar · View at Scopus


116.                     J. Yao, R. W. Irwin, L. Zhao, J. Nilsen, R. T. Hamilton, and R. D. Brinton, “Mitochondrial bioenergetic deficit precedes Alzheimer's pathology in female mouse model of Alzheimer's disease,” Proceedings of the National Academy of Sciences of the United States of America, vol. 106, no. 34, pp. 14670–14675, 2009.View at Publisher · View at Google Scholar · View at Scopus


117.                     A. Dhar, K. Desai, J. Liu, and L. Wu, “Methylglyoxal, protein binding and biological samples: are we getting the true measure?” Journal of Chromatography B, vol. 877, no. 11-12, pp. 1093–1100, 2009. View at Publisher · View at Google Scholar · View at Scopus


118.                     F. W. R. Chaplen, W. E. Fahl, and D. C. Cameron, “Evidence of high levels of methylglyoxal in cultured Chinese hamster ovary cells,” Proceedings of the National Academy of Sciences of the United States of America, vol. 95, no. 10, pp. 5533–5538, 1998. View at Publisher · View at Google Scholar · View at Scopus


 



Enter subhead content here

The target for therapy is through avoid what harms the mitochondria.  First above all is fructose, and second is the vegetable oils unsaturated fatty acids which become rancid through a chain reaction.  Third would be taking antioxidants in significant amounts.  I take CoQ10, sodium ascorbate 2 grams, and vitamins A and E.  The salt of vitamin C is far easier on the digestive system. 

Boon, Pamela, Li Pun et al, Pathological Significance of Mitochondrial Glycation

https://www.hindawi.com/journals/ijcb/2012/843505/abs/   May 2013  Pamela Boon, Li Pun et al,

Nonetheless, the prospect of mitochondrial glycation contributing as a common damaging agent across a broad spectrum of diseases is an intriguing possibility and is also a novel potential therapeutic target.”
Pathological Significance of Mitochondrial Glycation

Glycation, the nonenzymatic glycosylation of biomolecules, is commonly observed in diabetes and ageing. Reactive dicarbonyl species such as methylglyoxal and glyoxal are thought to be major physiological precursors of glycation. Because these dicarbonyls tend to be formed intracellularly, the levels of advanced glycation end products on cellular proteins are higher than on extracellular ones. The formation of glycation adducts within cells can have severe functional consequences such as inhibition of protein activity and promotion of DNA mutations. Although several lines of evidence suggest that there are specific mitochondrial targets of glycation, and mitochondrial dysfunction itself has been implicated in disease and ageing, it is unclear if glycation of biomolecules specifically within mitochondria induces dysfunction and contributes to disease pathology. We discuss here the possibility that mitochondrial glycation contributes to disease, focussing on diabetes, ageing, cancer, and neurodegeneration, and highlight the current limitations in our understanding of the pathological significance of mitochondrial glycation.

https://www.sciencedirect.com/science/article/pii/S0014480006001328

https://doi.org/10.1016/j.yexmp.2006.09.008 Volume 83, Issue 1August 2007, Pages 84-92  

Mitochondrial dysfunction and molecular pathways of disease

Pieczenik, SR,  John. Neustadt, August 2007, Mitochondrial dysfunction and molecular pathways of disease

Abstract

Since the first mitochondrial dysfunction was described in the 1960s, the medicine has advanced in its understanding the role mitochondria play in health, disease, and aging. A wide range of seemingly unrelated disorders, such as schizophrenia, bipolar disease, dementia, Alzheimer's diseaseepilepsymigraine headaches, strokes, neuropathic painParkinson's diseaseataxiatransient ischemic attackcardiomyopathycoronary artery disease, chronic fatigue syndromefibromyalgiaretinitis pigmentosadiabeteshepatitis C, and primary biliary cirrhosis, have underlying pathophysiological mechanisms in common, namely reactive oxygen species (ROS) production, the accumulation of mitochondrial DNA (mtDNA) damage, resulting in mitochondrial dysfunction. Antioxidant therapies hold promise for improving mitochondrial performance. Physicians seeking systematic treatments for their patients might consider testing urinary organic acids to determine how best to treat them. If in the next 50 years advances in mitochondrial treatments match the immense increase in knowledge about mitochondrial function that has occurred in the last 50 years, mitochondrial diseases and dysfunction will largely be a medical triumph.

 

Mitochondria are the powerhouses of our cells. They are responsible for generating energy as an adenosine triphosphate (ATP) and heat and are involved in the apoptosis-signaling pathway. Current theory holds that mitochondria are the descendants of aerobic bacteria that colonized an ancient prokaryote between 1 and 3 billion years ago (Spees et al., 2006DiMauro and Schon, 2003Wallace, 2005). This allowed for the evolution of the first eukaryotic cell capable of aerobic respiration, a necessary precursor to the evolution of multicellular organisms (Spees et al., 2006). Supporting this theory is the observation that mitochondria are the only other subcellular structure aside from the nucleus to contain DNA. However, unlike nuclear DNA, mitochondrial DNA (mtDNA) are not protected by histones (Croteau and Bohr, 1997). Nuclear DNA wraps around histones, which then physically shield the DNA from damaging free radicals (Milligan et al., 1993) and are also required to repair double-stranded DNA breaks (Celeste et al., 2003). Since mtDNA lacks the structural protection of histones and their repair mechanisms, they are quite susceptible to damage.

The first mitochondrial disease was described by Luft and colleagues in 1962, when a euthyroid 35-year-old female presented with myopathy, excessive perspiration, heat intolerance, polydipsia with polyuria, and a basal metabolic rate 180% of normal (Luft et al., 1962). The patient suffered from an uncoupling of oxidative phosphorylation (ox-phos). Ox-phos is the major cellular energy-producing pathway. Energy, in the form of ATP, is produced in the mitochondria through a series of reactions in which electrons liberated from the reducing substrates nicotine adenine dinucleotide (NADH) and flavin adenine dinucleotide (FADH) are delivered to O2via a chain of respiratory proton (H+) pumps (Brookes et al., 2004). The uncoupling of ox-phos leads to the generation of heat without generating ATP, which was the dysfunction underlying this patient's presentation. To compensate, her mitochondria enlarged and multiplied, which was evident in a histological examination of muscle biopsies.

Since this first documented case, mitochondrial dysfunction has been implicated in nearly all pathologic and toxicologic conditions (Aw and Jones, 1989). (These conditions are outlined in Table 1Table 2Table 3.) The conditions include sarcopenia and nonalcoholic steatohepatitis; acquired diseases such as diabetes and atherosclerosis; neurodegenerative diseases such as Parkinson's and Alzheimer's diseases; and inherited diseases, collectively called mitochondrial cytopathies.

~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~

Important in that this study shows a major reduction in the production of ATP, indicating defective mitochondria 94.06 ATP vs 7.5

http://science.sciencemag.org/content/300/5622/1140  Science  16 May 2003:  Vol. 300, Issue 5622, pp. 1140-1142

  •  

    REPORT

Mitochondrial Dysfunction in the Elderly: Possible Role in Insulin Resistance

 

Abstract

Insulin resistance is a major factor in the pathogenesis of type 2 diabetes in the elderly. To investigate how insulin resistance arises, we studied healthy, lean, elderly and young participants matched for lean body mass and fat mass. Elderly study participants were markedly insulin-resistant as compared with young controls, and this resistance was attributable to reduced insulin-stimulated muscle glucose metabolism. These changes were associated with increased fat accumulation in muscle and liver tissue assessed by 1Hnuclear magnetic resonance (NMR) spectroscopy, and with a 40% reduction in mitochondrial oxidative and phosphorylation activity, as assessed by in vivo 13C/31P NMR spectroscopy. These data support the hypothesis that an age-associated decline in mitochondrial function contributes to insulin resistance in the elderly.

 

Type 2 diabetes is the most common chronic metabolic disease in the elderly, affecting 30 million individuals 65 years of age or older in developed countries (1) [Estimated at 26% in the U.S.]. The estimated economic burden of diabetes in the United States is $100 billion per year [2002], of which a substantial proportion can be attributed to persons with type 2 diabetes in the elderly age group (2). Epidemiological studies have shown that the transition from the normal state to overt type 2 diabetes in aging is typically characterized by a deterioration in glucose tolerance (34) that results from impaired insulin-stimulated glucose metabolism in skeletal muscle (56). Measurements of muscle triglyceride content by biopsy (7) or intramyocellular lipid content (IMCL) by 1H nuclear magnetic resonance (NMR) spectroscopy (810) have shown a strong relationship between increased intramuscular fat content and insulin resistance in muscle. Similar correlations have been established for hepatic insulin resistance and hepatic steatosis (1113). Increases in the intracellular concentration of fatty acid metabolites have been postulated to activate a serine kinase cascade leading to defects in insulin signaling in muscle (1417) and the liver (18), which results in reduced insulin-stimulated muscle glucose transport activity (14), reduced glycogen synthesis in muscle (1920), and impaired suppression of glucose production by insulin in the liver (1113).  [This is consistent with Dr. Jason Fung’s point that the cells are stuffed with insulin due to a reduction in the production of ATP, and therefore have become resistant to more insulin, which is part of a feedback mechanism.  I hold that this is necessary to main a safe osmotic pressure in the cell, since glucose is hydroscopic.]

To examine whether insulin resistance in the elderly is associated with similar increases in intramyocellular and/or liver triglyceride content, we studied healthy elderly and young people that we matched for lean body mass (LBM) and fat mass. All study participants were nonsmoking, sedentary, lean [body mass index (BMI) < 25 m2/kg], and taking no medications. Sixteen elderly volunteers (ages 61 to 84 years, 8 male and 8 female) were screened with a 3-hour oral glucose (75 g) tolerance test and underwent dual-energy x-ray absorptiometry to assess LBM and fat mass (21). One elderly man was excluded from the study because of an abnormal glucose profile. Thirteen young volunteers (ages 18 to 39 years, 6 male and 7 female), who had no family history of diabetes or hypertension, were matched to the older participants for BMI and habitual physical activity, which was assessed by means of an activity index questionnaire (22). All participants underwent a complete medical history and physical examination, as well as blood tests to confirm that they were in excellent health (23).

Young and elderly participants had similar fat mass, percent fat mass, and LBM (Table 1) (24). The elderly participants had slightly higher plasma glucose concentrations (Fig. 1A) and significantly higher plasma insulin concentrations (Fig. 1B) during the oral glucose tolerance test, suggesting that they were relatively insulin-resistant as compared with the young controls. Basal plasma fatty acid concentrations (Fig. 1C) also tended to be higher in the elderly participants but were suppressed normally after glucose ingestion.

https://d2ufo47lrtsv5s.cloudfront.net/content/sci/300/5622/1140/F1.medium.gif

Plasma concentrations of glucose, insulin, and fatty acids before and after an oral glucose tolerance test (24) in young and elderly participants. (A) Glucose [P = 0.10 for the area under the curve (AUC) for the elderly (16,978 ± 656) as compared with the controls (14,495 ± 1,116)]. (B) Insulin [asterisks indicate P < 0.03 for AUC for the elderly (6590 ± 853) as compared with the controls (3986 ± 519)]. (C) Fatty acids (P = 0.08 for the basal concentration of fatty acids in the elderly versus the controls).

Table 1.

Body composition of study participants.

Age (years)

Body weight (kg)

Fat mass (kg)

% Fat mass (% body weight)

LBM (kg)

BMI (kg/m2)

Young (n = 13)

27 ± 2

71 ± 4

19.9 ± 2.5

28 ± 3

54 ± 5

23.8 ± 1.1

Elderly (n = 15)

70 ± 2

70 ± 3

20.1 ± 1.7

29 ± 2

49 ± 3

25.1 ± 0.5

P value

<0.0001

0.69

0.93

0.77

0.28

0.28

To determine what tissues were responsible for the insulin resistance, we performed hyperinsulinemic-euglycemic clamp studies, in combination with [6,6-2H2] glucose and [2H5] glycerol tracer infusions (24). Basal rates of glucose production were similar in the young and elderly participants (Table 2) and were suppressed completely in both groups during the hyperinsulinemic-euglycemic clamp. In contrast, the rates of glucose infusion required to maintain euglycemia during the clamp and insulin-stimulated rates of peripheral glucose uptake were 40% lower in the elderly participants (Table 2). Basal energy expenditure and respiratory quotient both tended to be lower in the elderly participants (24).

Table 2.

Metabolic rates and tissue lipid content of participants (24).

Basal rates of glucose production (mg/kg of LBM/min)

Clamp peripheral glucose metabolism rate (mg/kg of LBM/min)

Intramyocellular lipid content (%)

Intrahepatic lipid content (%)

Mitochondrial TCA flux rate (nmol/g of muscle/min)

Mitochondrial ATP synthesis rate (μmol/g of muscle/min)

Young

2.3 ± 0.1

6.2 ± 0.6

0.96 ± 0.08

0.49 ± 0.10

96 ± 10

7.50 ± 0.77

Elderly

2.4 ± 0.1

4.0 ± 0.4

1.39 ± 0.15

1.61 ± 0.38

62 ± 5

4.06 ± 0.65

Pvalue

0.34

<0.002

0.035

0.036

<0.006

<0.004

To ascertain whether lipid accumulation in muscle might be responsible for the insulin resistance in the elderly participants, we used 1H NMR spectroscopy to assess IMCL and hepatic triglyceride content (24). The IMCL content in the soleus muscle was increased by 45% in the elderly participants as compared with controls (Table 2 and fig. S1). Intrahepatic triglyceride content was also increased by 225% in the elderly participants as compared with controls, even though there was no detectable hepatic insulin resistance in the elderly participants during the clamp. It is possible that hepatic insulin resistance was not detected in the elderly participants because of the relatively high plasma insulin concentrations obtained during the clamp studies, which completely suppressed hepatic glucose production in both groups.

Because increases in intramyocellular and intrahepatic triglyceride content could occur secondarily to increased fatty acid delivery from lipolysis, we also examined this process in vivo. We assessed whole-body and subcutaneous fat lipolysis by measuring the rates of [2H5] glycerol turnover in combination with microdialysis measurements of glycerol release from subcutaneous fat. Basal rates of whole-body glycerol turnover and insulin suppression of glycerol turnover during the clamp were similar in the elderly and control participants. Consistent with this finding, the interstitial glycerol concentrations, assessed by microdialysis, decreased by a similar degree during the clamp in both groups. Taken together, these data suggest that insulin resistance was confined mostly to skeletal muscle and that increased basal rates of peripheral lipolysis, and/or defects in insulin suppression of lipolysis, do not play a major role in causing the increased intramyocellular and intrahepatic triglyceride content in the elderly.

We and others (25) have previously hypothesized that defects in mitochondrial oxidative and phosphorylation capacity might be a contributing factor to the increased triglyceride content in muscle and the liver (26). To test this hypothesis, we assessed in vivo rates of mitochondrial oxidative activity in skeletal muscle by 13C NMR and phosphorylation activity by 31P NMR (2427). Using this approach, we found that rates of mitochondrial oxidative and phosphorylation activity were both reduced by 40% in the elderly participants as compared with the young controls. These in vivo results are consistent with those of a previous in vitro study, which found decreased state III (activated) mitochondrial respiration in isolated mitochondria from elderly participants (28). However, the latter study was performed with muscle strips, from orthopedic and chronic fatigue syndrome patients, under artificial substrate concentrations that do not reflect in vivo conditions.

Our results suggest that insulin resistance in the elderly is related to increases in intramyocellular fatty acid metabolites that may be a result of an age-associated reduction in mitochondrial oxidative and phosphorylation activity (fig. S2). The similarity in mitochondrial energy coupling, assessed by the ratio between adenosine triphosphate (ATP) synthase flux and tricarboxylic acid (TCA) cycle oxidation, suggests an age-associated reduction in mitochondrial number and/or function, as opposed to an acquired defect in mitochondrial energy coupling. These possibilities are consistent with a recent study demonstrating an age-associated accumulation of mutations in control sites for mitochondrial DNA replication (29). Because mitochondrial oxidative and phosphorylation activity is the major source of energy in most organs, including the brain, our data add support to the hypothesis that a decline in mitochondrial oxidative and phosphorylation energy production may also have an important role in aging (3031). Furthermore, because mitochondrial energy metabolism plays a critical role in glucose-induced insulin secretion (32), similar age-associated reductions in pancreatic beta cell mitochondrial function, in the setting of peripheral insulin resistance, might help explain the high prevalence of diabetes in the elderly.

 

^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^^

Oxidative damage to the DNA of mitochondria is driving the reduced ATP production

http://science.sciencemag.org/content/286/5440/774  Science  22 Oct 1999:  Vol. 286, Issue 5440, pp. 774-779

~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~

Enter supporting content here

The target for therapy is through avoid what harms the mitochondria. First above all is fructose, and second is the vegetable oils unsaturated fatty acids which become rancid through a chain reaction. Third would be taking antioxidants in significant amounts. I take CoQ10, sodium ascorbate 2 grams, and vitamins A and E. The salt of vitamin C is far easier on the digestive system.