search
for
 About Bioline  All Journals  Testimonials  Membership  News


Journal of Postgraduate Medicine
Medknow Publications and Staff Society of Seth GS Medical College and KEM Hospital, Mumbai, India
ISSN: 0022-3859 EISSN: 0972-2823
Vol. 53, Num. 3, 2007, pp. 207-213

Journal of Postgraduate Medicine, Vol. 53, No. 3, July-September, 2007, pp. 207-213

View Point

The oxidative hypothesis of senescence

Department of Biochemistry, Faculty of Medicine, "Carol Davila" University of Medicine and Pharmacy, 8, Eroilor Sanitari, 76241 Bucharest, Romania
Correspondence Address:Department of Biochemistry, Faculty of Medicine, "Carol Davila" University of Medicine and Pharmacy, 8, Eroilor Sanitari, 76241 Bucharest, Romania marilenagilca@gmail.com

Date of Submission: 15-Dec-2006
Date of Decision: 24-Jan-2007
Date of Acceptance: 08-Apr-2007

Code Number: jp07074

Abstract

The oxidative hypothesis of senescence, since its origin in 1956, has garnered significant evidence and growing support among scientists for the notion that free radicals play an important role in ageing, either as "damaging" molecules or as signaling molecules. Age-increasing oxidative injuries induced by free radicals, higher susceptibility to oxidative stress in short-lived organisms, genetic manipulations that alter both oxidative resistance and longevity and the anti-ageing effect of caloric restriction and intermittent fasting are a few examples of accepted scientific facts that support the oxidative theory of senescence. Though not completely understood due to the complex "network" of redox regulatory systems, the implication of oxidative stress in the ageing process is now well documented. Moreover, it is compatible with other current ageing theories (e.g., those implicating the mitochondrial damage/mitochondrial-lysosomal axis, stress-induced premature senescence, biological "garbage" accumulation, etc). This review is intended to summarize and critically discuss the redox mechanisms involved during the ageing process: sources of oxidant agents in ageing (mitochondrial -electron transport chain, nitric oxide synthase reaction- and non-mitochondrial- Fenton reaction, microsomal cytochrome P450 enzymes, peroxisomal β -oxidation and respiratory burst of phagocytic cells), antioxidant changes in ageing (enzymatic- superoxide dismutase, glutathione-reductase, glutathion peroxidase, catalase- and non-enzymatic glutathione, ascorbate, urate, bilirubine, melatonin, tocopherols, carotenoids, ubiquinol), alteration of oxidative damage repairing mechanisms and the role of free radicals as signaling molecules in ageing.

Keywords: Antioxidants, oxidants, senescence

Two principal types of ageing theories have been developed: theories of "accidental" ageing produced by "errors" represented by random deleterious mechanisms that induce progressive damage of various levels; and theories of "programmed" ageing induced by the collection of by-products of gene action selected to enhance reproductive fitness. [1] Even if there are no genes that specifically evolved to induce senescence, scientists estimated that allelic variation or mutation at up to 7000 relevant genes might modulate patterns of ageing in man. [2] Certain polymorphisms (antagonistic pleiotropy) might underlie common or "public" mechanisms of ageing, while rare mutations lead to uncommon or "private" mechanisms of ageing. [3],[4] These two theories are not mutually exclusive, especially when oxidative stress is considered. The oxidative theory of aging was first advanced in 1956 by Harman: free radicals, normally produced in the organisms, react with cellular constituents and initiate the age-associated changes. [5],[6]

A Medline search (years 1966 through December 2006, using search terms senescence, ageing, oxidative stress and free radicals), supplemented by subsequent reference searches of retrieved articles and hand search of Free Radicals Biology and Medicine 1989-2006 identified studies relevant to the oxidative hypothesis of ageing.

Age-increasing oxidative damages induced by free radicals or higher susceptibility to oxidative stress in short-lived organisms are mentioned by supporters of both types of theories:

a. Oxygen-reactive species play a key role in age-related sporadic degenerative diseases (e.g., Alzheimer′s disease, atherosclerosis, diabetes) or various specific components of what scientists refer to as the senescent phenotype (e.g., tissue atrophies, etc.). [7]

b. Resistance to oxidative stress is a common trait of long-lived genetic variants of non-mammalian and mammalian organisms. [8],[9] Genetic variations in antioxidant defense genes (e.g. prion protein gene PRNP) were found to have important influences on the trajectory of normal ageing. [10] This fits with the oxidative hypothesis of senescence.

Unfortunately, extensive research on the relationship between polymorphisms likely to be responsible for the common mechanism of ageing and resistance to oxidative stress has been neglected. Therefore, the paucity of data does not allow us to yet conclude that the oxidative theory supports the theory of programmed ageing.

However, the most recent studies strongly support the idea that oxidative stress is a significant marker of senescence, being established in different species. This age-related oxidative stress is generated by a combination of increased production of free radicals and other oxidant agents, decreased antioxidant levels and impaired repair of oxidative damages. Oxidant agents, including reactive oxygen species (ROS) and reactive nitrogen species (RNS) (e.g., nitric oxide, NO) are recognized to play a dual role as both malefic and beneficial species, being sometimes compared with fire, which is dangerous, but nonetheless useful to humans. [11] Reactive oxygen species and RNS have many crucial biological functions as signaling molecules in growth, apoptosis, neurotransmission, etc. To control these reactive species with "two faces", cells evolved complex and critical regulatory mechanisms which become disrupted with age. Senescence is just one example of pathophysiological implications of redox dysregulation. [12] The initiation of ageing is marked by a shift from redox regulation to redox dysregulation. [13] Why this shift takes place is not yet clear.

This review focuses on how alterations in oxidants, protective agents and repair machinery during the lifespan may offer insights into the possible scenarios of biological ageing.

Definition of Ageing

Ageing is an extremely complex, multifactorial process and represents the gradual deterioration in function that occurs after maturity and leads to disability or death. A broader perspective includes all biological processes occurring within the organism from the beginning of life (fertilization) until death. A more recent and precise definition identifies ageing with the inability of the organism to respond to stress and to maintain homeostatic regulation when given a challenge, thereby decreasing the capacity of the organism to survive detrimental changes occurring with time during postmaturational life. [14]

Main Sources of Oxidant Agents in Ageing

There are two main types of oxidant sources: mitochondrial sources (which play the principal role in ageing) and non-mitochondrial sources (which play different and sometimes specific roles, especially in the pathogenesis of age-related diseases).

Mitochondrial sources

Mitochondrial sources are represented by the electron transport chain and the nitric oxide synthase reaction.

Mitochondria seem to be the principal source of endogenous oxidants implicated in ageing. The rate of respiration is responsible for the rate of generation of reactive oxygen species (ROS), this characteristic being consistent with the observation that the higher metabolic rates an organism has, the shorter maximum lifespan potential it presents. [15] One-electron reduction of O 2 to form the superoxide anion (O 2· - ) and dismutation of O 2· - to yield hydrogen peroxide (H 2 O 2 ) occurs during mitochondrial respiration.

Mitochondria are also involved in the generation of nitric oxide (NO·) via the nitric oxide synthase (NOS) reaction. O 2· - and NO· react to form another oxidant, peroxynitrite (ONOO - ), which represents a potential source for the more powerful and aggressive hydroxi radical (·OH).

Mitochondria possess superoxide dimutase (mSOD) that rapidly scavenges O 2· - . Some researchers believe that mSOD prevents the accumulation of O 2· - , while others argue that mSOD increases the rate of O 2· - generation by accelerating the removal of this radical by dismutation to H 2 O 2 . [1],[16]

Dysfunctional mitochondria accumulate during ageing due to the oxidative damage of mitochondrial macromolecules. [17] The role of mitochondrial oxidative damage in ageing and Alzheimer′s disease (AD) has important implications for therapeutics: mitochondrial antioxidant therapy has been found to be one of the most efficacious methods in reducing pathological changes in the brain tissues in AD animal model studies. [18]

The theory of mitochondrial ageing predicts that the mitochondrial DNA (mtDNA) proximity to the cell′s major source of free radicals renders it very susceptible to oxidative insults and thereby increases the rate of mtDNA mutations, leading to an aggravation of the aerobic respiration dysfunction (mtDNA encodes proteins of the respiratory chain). The consequent decrease of electron transfer leads to further production of ROS, thus establishing a vicious circle of oxidative stress and energetic decline, which is suspected to be one of the principal causes of ageing. [19]

Mitochondrial impairments lead also to the activation of nuclear genes. This signaling pathway from the mitochondrion to the nucleus, named the retrograde response, seems to influence cell division, stress resistance and, eventually, ageing rate and lifespan, at least in fungal models. [20]

Caloric restriction leads to reduced production of mitochondrial ROS and thus to a reduction in mitochondrial oxidative stress, which may be responsible for an increase in the lifespan. [21]

However, most of the available data supporting the mitochondrial theory of ageing are merely correlative and therefore do not exclude the possibility that ROS production and mtDNA mutations are effects rather than driving forces of ageing. Moreover, there is recent controversial evidence that mitochondrial mutations do not limit the lifespan of wild-type mice, specific point mutations may not accumulate with ageing in the mouse mitochondrial DNA control region and mitochondrial ROS production might not be affected in mtDNA mutator mice displaying the ageing phenotype. [22],[23],[24]

Non-mitochondrial sources

Fenton reaction

The H 2 O 2 -degrading Fenton reaction is catalyzed by free iron bivalent ions and leads to the generation of ·OH. Recent studies localized the Fenton reaction at the endoplasmic reticulum or perinuclear, but not at mitochondria or other compartments. [25] Sources of H 2 O 2 could be mitochondria (superoxide dismutase reaction), peroxisomes (acyl-CoA oxidase reaction) and amyloid b of senile plaques (superoxide dismutase-like reactions). [26] H 2 O 2 that escapes antioxidant machinery, such as glutathione peroxidase and catalase, might be converted nonenzymatically in a perinuclear-localized Fenton reaction and act as an RNA- or DNA-damaging agent. A recent study showed that the Fenton reaction is involved in the oxidation of ribosomal RNA in tissue brain obtained at autopsy from confirmed cases of Alzheimer′s disease. [27] Therefore, it may play a special role in the context of ageing, also taking into account that the body′s content of iron increases with age. [28],[29]

Microsomal cytochrome P450 enzymes

Microsomes contain the cytochrome P450 enzymes, which catalyze univalent oxidation or reduction of xenobiotic compounds (e.g., drugs); simultaneously O 2· - is generated. [30] Direct studies on drug-metabolizing capacity in elderly humans are scant. [31] However, several laboratories have concluded that ageing induces a decline in concentrations of cyt P450 monoxygenases in senescent animals, while starvation, which has a well-known lifespan-prolonging effect, increases the expression of the different proteins of the cytochromes P450 family. [32] The highest upregulated gene among the 20.000 genes analyzed before and after 24h and 48h starvation, is cytochrome P450 4A14 (Cyp4a14). [33] Beyond the secondary production of O 2· - , the members of the cyt P450 family are involved in resistance to oxidative stress, thereby increasing longevity.

The age-induced decline of cyt P450 enzymes could contribute to the high incidence of adverse drug reactions and toxicities reported in older people, as well as decreased antioxidant protection, but cannot account for a higher ROS production during ageing. [31]

Respiratory burst of phagocytic cells

The respiratory burst of phagocytic cells is a source of O 2· - via the NADPH oxidase reaction during inflammatory or infectious conditions. According to the theory of stress-induced premature senescence (SIPS), sublethal doses of different stressor agents (H 2 O 2 , hyperoxia, tertbuthylhidroperoxide, UV) lead to the exhaustion of the replicative potential of the proliferative normal cell types and the accumulation of senescent cells, which might be responsible for the creation of a micro-inflammatory state and the activation of phagocytic cells, thereby participating in tissue ageing. [34] Franceschi has shown that in the healthy elderly and the centenarians, the proinflammatory status is elevated and IL-6 plasma levels correlate with risk of death. [35]

Short-term fasting (80h) in healthy human subjects has induced a decrease of the stimulatory index of leukocytes activated with opsonized zymosan, phytohemagglutinin P and concanavalin A. [36] Recent animal studies have also shown that short-term repeated fasting is effective in the prolongation of lifespan or protects against age-related diseases. [37],[38],[39],[40] It is a reasonable assumption that the anti-ageing effect of intermittent fasting might be also due to the reversal of the age-related proinflammatory state.

Peroxizomal β-oxidation

Peroxisomal β-oxidation of fatty acids generates H 2 O 2 . The peroxisomes also contain catalase that decomposes H 2 O 2 and thus prevents local acumulation of this toxic compound. As cells age, the ability of peroxisomes to maintain this balance between peroxisomal pro-oxidants and antioxidants is gradually compromised. Senescent peroxisomes produce an increasing amount of ROS, not by increasing the H 2 O 2 synthesis, which is, on the contrary, probably decreased, but through an inefficient import of catalase (peroxisomes import enzymes post-translationally from cytosol). [41],[42]

Peroxisome proliferators, which increase the number of peroxisomes and the activity of enzymes involved in the β-oxidation of fatty acids, also cause oxidative damage. [43] PPAR protects against the oxidative damage associated with ageing, possibly by preventing the accumulation of oxidized fatty acids.

Antioxidant Agents in Ageing

Antioxidants are classified in two main groups: enzymatic antioxidant agents and non- enzymatic antioxidant agents. Both of them are modified during the process of ageing.

Enzymatic antioxidant agents

Several studies indicate that age induces different patterns of antioxidant enzymatic activities (superoxide dismutase SOD, glutathione peroxidase GSH-Px, glutathione reductase GR, glutathione-S-transpherase GSH-S-T, MSRA- methionine sulfoxide reductase) expressions. Some discrepancy exists in the described activities of the same enzyme [Table - 1].

Most of the studies on the genetic manipulation of antioxidant enzyme genes support the oxidative theory of ageing [Table - 2].

Non-enzymatic antioxidant agents

Hydrophilic non-enzymatic antioxidants are radical scavengers (e.g. glutathione (GSH), ascorbate, urate and bilirubin). Reduced glutathione, reduced glutathione/total glutathione ratio and ascorbate decline slightly in plasma or in lymphocytes with age. [58],[59]

The altered levels of hydrophilic antioxidants have also been correlated with various age-related diseases (e.g. the decline of melatonin in aged individuals has been suggested to contribute to the development of neurodegenerative disease). [60]

There are contradictory results concerning the effects of age on serum urate, which was either increased in aged women, but not in men or unchanged in centenarian subjects when compared with healthy adults. [61],[62] Total bilirubin was significantly reduced in aged people. [62]

Lipophilic non-enzymatic antioxidants are radical scavengers such as tocopherols, carotenoids, ubiquinol and flavonoids.

In one human study, the plasma alpha and total tocopherol concentrations did not change significantly with age. However, the plasma gamma tocopherol, platelet alpha, gamma tocopherol, total tocopherol concentrations and the platelet-to-plasma ratios of tocopherol concentrations decreased significantly with age. [63] Also, decreased levels of coenzyme Q10 in rat and human tissues have been reported during ageing. [64]

Repair of Oxidative Damages in Ageing

There are several types of repairing mechanisms depending on the nature of the oxidized target. [65] These mechanisms are mainly based either on regenerating the slightly oxidized macromolecules keeping critical chemical groups in their reduced forms or on degrading defective highly oxidized macromolecules into low-molecular-mass compounds (that are then removed or re-utilized for building up new biological structures).

Some enzymes play several protective roles, simultaneously acting as an antioxidant enzyme that scavenges ROS and as a repair enzyme that eliminates damages (e.g. GSH-Px/GR, methionine sulfoxide reductase).

Many of these essential maintenance and repair systems become deficient in senescent cells, thus a high amount of biological "garbage" is accumulated (e.g. intralysosomal accumulation of lipofuscin). [66],[67] Age-related oxidative changes are most prominent in non-proliferating cells, such as neurons and cardiac myocites, because there is a lack of dilution effect of damaged structures through cell division. [68]

The DNA repair ability correlates with species-specific lifespan, being necessary but not sufficient for longevity. [69]

There is an age-related decline in proteasome peptidase activities and proteasome content in different tissues (e.g., rat liver, human epidermis), which leads to accumulation of oxidatively modified proteins. [70] The total amount of oxidatively modified proteins for an 80-year-old human is estimated to be up to 50%. [71]

Among protein-bound aminoacids, cysteine and methionine residues are particularly sensitive to oxidation to disulfide bridges and methionine derivatives sulfoxides, but these minor oxidative damages are readily repaired by the action of thioredoxin reductase and methionine sulfoxide reductase (MSRA). [72] According to several studies, MSRA seems to be a regulator of antioxidant defense, oxidative stress-response gene expression and ageing in mammals. [56],[57]

Oxidative Damages in Ageing

All the biologically active molecules are susceptible to suffering oxidative damages and thus failing to accomplish their native roles. [73] A huge diversity of negative effects results from this indiscriminate oxidation. Elevated levels of oxidized lipids, DNA, proteins and glycoxidation macromolecules are found in aged organisms. [1],[74],[75]

The hydroxil radical oxidizes DNA, leading to the formation of adducts 8-oxo-7,8 dihydro-2′-deoxyguanosine (oxo 8 dG) in susceptible 5-GC-3′. The frequency of oxidative DNA adducts increases by as much as twofold with age in different species and tissues. [76]

Mitochondrial DNA is more vulnerable to oxidative damages than nuclear DNA because it is not protected by histones and mitochondria are the primary sites of ROS generation. [77],[78] This leads to mutations of mitochondrial DNA, involving the genes coding for respiratory chain proteins and also may disturb the division of mitochondria, resulting in their enlargement. Larger mitochondria are less autophagocytosed (by lysosomes which are overloaded with lipofuscin) and undergo further oxidative damage. Briefly, the mitochondrial-lysosomal axis theory of ageing sustains that the accumulation of dysfunctional mitochondria and lysosomes leads irreversibly to cell death. [79]

Proteins isolated from aged individuals exhibit a higher carbonyl content, which estimates the overall extent of protein oxidation (e.g. "old" ceruloplasmin versus "young" ceruloplasmin contains 0.6mol versus 0.2 mol of carbonyl/mol of protein). [80] In spite of these age-induced oxidative changes, some proteins do not lose their biological function (e.g. ceruloplasmin oxidase activity is not altered in older individuals). [80]

Reactive oxygen species as mediators of cellular senescence

Reactive oxygen species also play the role of signaling molecules in cellular senescence and the age-related increase in mithocondrial-triggered apoptosis. [81],[82],[83],[84] However, the molecular mechanism of signaling remains obscure in certain cases (e.g. superoxide signaling). A new signaling hypothesis based on the frequently forgotten "super"-nucleophilic properties of superoxide anion has been recently proposed. [85] Free radicals induce alterations in gene expression (e.g. p53, HSP70, Bcl family genes). Regulation errors of the signaling cascade may also be responsible for the development of ageing. [83],[86],[87]

Seladin-1, a protein whose expression is down-regulated in Alzheimer′s disease, protects cells from oxidative stress. [88] Reactive oxygen species induce redistribution of seladin-1 from cytosol to the nucleus, where it physically binds to p53 and leads to p53 accumulation, G1 arrest and replicative senescence. [89]

Another mechanism underlying cellular senescence is telomere attrition. [90] Telomeres are located at the end of chromosomes and each division is associated with a decrease in telomere length. It is already established that a high level of oxidative stress shortens telomeres and triggers the senescence. [91]

Controversial issues concerning oxydative hypothesis of senescence

  1. "Premature" ageing without oxidative stress hypersensitivity. The oxidative hypothesis of senescence does not explain "premature" ageing. Neither accelerated nor acute oxidative stress hypersensitivity was detected in primary fibroblast or erythroblast cultures from multiple progeroid mouse models, which are mainly associated with accelerated fibroblast senescence. [92]
  2. Gene actions that alter oxidative stress resistance, but not lifespan. Genetic manipulations that increases CuZn-SOD activity and, thus, resistance to oxidative stress, have only a slight, if any, effect on maximum lifespan in several species. [93] Simultaneous overexpression of MnSOD and mitochondrial CAT in transgenic Drosophila induced a paradoxical decrease of lifespan, which might signify that only physiological levels of O 2· - /H 2 O 2 (neither excess or deficit) promote normal ageing. [94]
    No association between A16V, C47T MnSOD or C262T CAT polymorphisms and age-related mortality or phenotypes was found in humans. However, genotype AA MnSOD was associated with an immunosenescence profile and DNA damage, while TT CAT was associated with improved physical functioning. [95],[96],[97]
    In another study, P66(shc-/-) mice exhibited prolonged lifespan and increased resistance to oxidative stress, but unexpectedly, centenarians showed the highest basal levels of p66(shc) when compared with young people and the elderly. [98]
  3. Lack of predictability. Some critics of the oxidative theory claim that the failure of antioxidant interventions to stop or reverse the aging process and to quell the current pandemic of age-related diseases (e.g. cardiovascular disease) brings the oxidative hypothesis into question. [99] Nevertheless, since ageing is a complex dysregulation of many redox systems, single antioxidant administration should not necessarily be expected to influence the ageing process. Complex multiple antioxidant interventions or complete dietary changes might be more successful in this respect.
  4. Lack of direct cause-and-effect evidence. Although a growing body of evidence points towards the implication of redox dysregulation as an important determinant of ageing, a direct cause-and-effect relationship between the accumulation of oxidatively mediated damage and ageing has not been clearly established. [100]

Conclusion

In one way or another, oxidative stress is mentioned in many theoretical and experimental scenarios of ageing. The life trajectory seems to be parallel with the oxidative stress resistance in most cases. Therefore, revealing all the aspects of redox alterations during senescence may be a key for controlling the rate of ageing and the longevity potential of the organisms. Nevertheless, the main question is still without answer: which of the mechanisms (oxidative damages, garbage catastrophe, mutation accumulation, antagonistic pleiotropy, etc.) proposed by scientists to cause senescence do cause ageing in the natural population? The answer might be more complex than we expect and probably a more integrative approach should be adopted to solve the dilemma.

Acknowledgement

Viasan Grant from the Ministry of Education and Science, Romania, project "New aspects concerning redox systems involved in ageing and potential therapeutic implications". The authors would particularly like to thank Dr. Ralph Miller (Canadian Commission′s Research ex-Director) for his comments and support during the revision process.

References

1.Beckman KB, Ames BN. The free radical theory of aging matures. Physiol Rev 1998;78:547-81.  Back to cited text no. 1  [PUBMED]  [FULLTEXT]
2.Martin GM. Interaction of aging and environmental agents: The gerontological perspective. Prog Clin Bio Res 1987;228:25-80.  Back to cited text no. 2    
3.Corder EH, Saunders AM, Strittmatter WJ, Schmechel DE, Gaskell PC, Small GW, et al . Gene dose of apolipoprotein E type allele and the risk of Alzheimer's disease in late onset families. Science 1993;261:921-3.  Back to cited text no. 3  [PUBMED]  [FULLTEXT]
4.Sherrington R, Rogaev EI, Liang Y, Rogaeva EA, Levesque G, Ikeda M, et al . Cloning of a gene bearing missense mutations in early-onset familial Alzheimer's disease. Nature 1995;375:754-60.  Back to cited text no. 4  [PUBMED]  [FULLTEXT]
5.Harman D. Aging: A theory based on free radical and radiation chemistry. J Gerontol 1956;11:298-300.  Back to cited text no. 5  [PUBMED]  
6.Harman D. Free radical theory of aging: An update: Increasing the functional life span. Ann N Y Acad Sci 2006;1067:10-21.  Back to cited text no. 6  [PUBMED]  [FULLTEXT]
7.Martin GM. Genetics and the pathobiology of ageing. Philos Trans R Soc Lond B Biol Sci 1997;352:1773-80.  Back to cited text no. 7  [PUBMED]  [FULLTEXT]
8.Martin GM, Austad SN, Johnson TE. Genetic analysis of ageing: Role of oxidative damage and environmental stress. Nat Genet 1996;13:25-34.  Back to cited text no. 8  [PUBMED]  [FULLTEXT]
9.Mooijaart SP, van Heemst D, Schreuder J, van Gerwen S, Beekman M, Brandt BW et al . Variation in the SHC1 gene and longevity in humans. Exp Gerontol 2004;39:263-8.  Back to cited text no. 9    
10.Kachiwala SJ, Harris SE, Wright AF, Hayward C, Starr JM, Whalley LJ, et al . Genetic influences on oxidative stress and their association with normal cognitive ageing. Neurosci Lett 2005;386:116-20.  Back to cited text no. 10  [PUBMED]  [FULLTEXT]
11.de Magalhaes JP, Church GM. Cells discover fire: employing reactive oxygen species in development and consequences for aging. Exp Gerontol 2006;41:1-10.  Back to cited text no. 11    
12.Valko M, Leibfritz D, Moncol J, Cronin MT, Mazur M, Telser J. Free radicals and antioxidants in normal physiological functions and human disease. Int J Biochem Cell Biol 2007;39:44-84.  Back to cited text no. 12  [PUBMED]  [FULLTEXT]
13.Humphries KM, Szweda PA, Szweda LI. Aging: A shift from redox regulation to oxidative damage. Free Radic Res 2006;40:1239-43.  Back to cited text no. 13  [PUBMED]  [FULLTEXT]
14.Florez-Duquet M, McDonald RB. Cold-induced thermoregulation and biological aging. Physiol Rev 1998;78:339-58.  Back to cited text no. 14  [PUBMED]  [FULLTEXT]
15.Sohal RS. Metabolic rate and life span. In : Witler R, editor. Cellular aging: Concepts and metabolism. Basel: Karger; 1976. p. 25-40.  Back to cited text no. 15    
16.Forman HJ, Azzi A. On the virtual existence of superoxide anion in mitochondria: Thoughts regarding its role in pathophysiology. FASEB J 1997;11:374-5.  Back to cited text no. 16  [PUBMED]  [FULLTEXT]
17.Osiewacz HD, Kimpel E. Mitochondrial-nuclear interactions and lifespan control in fungi. Exp Gerontol 1999;34:901-9.  Back to cited text no. 17  [PUBMED]  [FULLTEXT]
18.Reddy PH. Mitochondrial oxidative damage in aging and Alzheimer's disease: implications for mitochondrially targeted antioxidant therapeutics. J Biomed Biotechnol 2006;2006:31372.  Back to cited text no. 18  [PUBMED]  [FULLTEXT]
19.Genova ML, Pich MM, Bernacchia A, Bianchi C, Biondi A, Bovina C, et al. The mitochondrial production of reactive oxygen species in relation to aging and pathology. Ann N Y Acad Sci 2004;1011:86-100.  Back to cited text no. 19  [PUBMED]  [FULLTEXT]
20.Jazwinski SM. Longevity, genes and aging: A review provided by genetic model systems. Exp Gerontol 1999;34:1-6.  Back to cited text no. 20  [PUBMED]  [FULLTEXT]
21.Weindruch R, Keenan KP, Carney JM, Fernandes G, Feuers RJ, Floyd RA, et al . Caloric restriction mimetics: Metabolic interventions, J Gerontol A Biol Sci Med Sci 2001;56:20-33.  Back to cited text no. 21    
22.Vermulst M, Bielas JH, Kujoth GC, Ladiges WC, Rabinovich PS, Prolla TA, et al . Mitochondrial point mutations do not limit the natural lifespan of mice. Nat Genet 2007;39:540-3.  Back to cited text no. 22    
23.Trifunovic A, Hansson A, Wredenberg A, Rovio AT, Dufour E, Khvorostov I, et al . Somatic mtDNA mutations cause aging phenotypes without affecting reactive oxygen species production. PNAS 2005;102:17993-8.  Back to cited text no. 23  [PUBMED]  [FULLTEXT]
24.Song X, Deng JH, Liu CJ, Bai Y. Specific point mutations may not accumulate with aging in the specific mitochondrial DNA control region. Gene 2005;350:193-9.  Back to cited text no. 24  [PUBMED]  [FULLTEXT]
25.Liu Q, Berchner-Pfannschmidt U, Moller U, Brecht M, Wotzlaw C, Acker H, et al . A Fenton reaction at the endoplasmic reticulum is involved in the redox control of hypoxia-inducible gene expression. Proc Natl Acad Sci USA 2004;101:4302-7.  Back to cited text no. 25    
26.Rottkamp CA, Raina AK, Zhu X, Gaier E, Bush AI, Atwood CS, et al . Redox-active iron mediates amyloid-beta toxicity. Free Radic Biol Med 2001;30:447-50.  Back to cited text no. 26  [PUBMED]  [FULLTEXT]
27.Honda K, Smith MA, Zhu X, Baus D, Merrick WC, Tartakoff AM, et al . Ribosomal RNA in Alzheimer disease is oxidised by bound redox-active iron. J Biol Chem 2005;280:20978-86.  Back to cited text no. 27  [PUBMED]  [FULLTEXT]
28.Koster JF, Sluiter W. Is increased tissue ferritin a risk factor for atherosclerosis and ischaemic heart disease? Br Heart J 1995;73:208-9.  Back to cited text no. 28  [PUBMED]  [FULLTEXT]
29.Vercellotti GM. A balanced budget-evaluating the iron economy. Clin Chem 1996;42:657.  Back to cited text no. 29  [PUBMED]  [FULLTEXT]
30.Halliwell BH, Gutteridge JM. Free radicals in Biology and Medicine. 2 nd ed. Oxford Univ Press: Oxford, UK; 1989.  Back to cited text no. 30    
31.Hunt CM, Westerkam WR, Stave GM. Effect of age and gender on the activity of human hepatic CYP3A. Biochem Pharmacol 1992;44:275-83.  Back to cited text no. 31    
32.Agrawal AK, Shapiro BH. Constitutive and inducible hepatic cytochrome P450 in senescent male and female rats and response to low dose Phenobarbital. Drug Metab Dispos 2003;31:612-9.  Back to cited text no. 32    
33.Bauer M, Hamm AC, Bonaus M, Jacob A, Jaekel J, Schorle H, et al . Starvation response in mouse liver shows strong correlation with life-span-prolonging processes. Physiol Genomics 2004;17:230-44.  Back to cited text no. 33    
34.Toussaint O, Remacle J. Stress and energy metabolism in age-related processes. In : Rattan SI, Toussaint O, editors. Molecular gerontology. Research status and perspectives. Plenum Press: New York; 1996. p. 87-110.  Back to cited text no. 34    
35.Franceschi C, Ottaviani E. Stress, inflammation and natural immunity in the aging process: A new theory. Aging (Milano) 1997;9:30-1.  Back to cited text no. 35    
36.Gilca M, Chirila M, Dinu V. Effect of fasting (80h) on the luminol-enhanced chemiluminescence of the polymorphonuclear leukocytes in healthy human subjects. Rom J Intern Med 2003;41:75-81.  Back to cited text no. 36    
37.Sogawa H, Kubo C. Influence of short-term repeated fasting on the longevity of female (NZB x NZW)F1 mice. Mech Ageing Dev 2000;115:61-71.  Back to cited text no. 37    
38.Mulas MF, Demuro G, Mulas C, Putzolu M, Cavallini G, Donati A, et al . Dietary restriction counteracts age-related changes in cholesterol metabolism in the rat. Mech Ageing Dev 2005;126:648-54.  Back to cited text no. 38    
39.Descamps O, Riondel J, Ducros V, Roussel AM. Mitochondrial production of reactive oxygen species and incidence of age-associated lymphoma in OF1 mice: Effect of alternate-day fasting. Mech Ageing Dev 2005;126:1185-91.  Back to cited text no. 39    
40.Martin B, Mattson MP, Maudsley S. Caloric restriction and intermittent fasting: Two potential diets for successful brain aging. Ageing Res Rev 2006;5:332-53.  Back to cited text no. 40    
41.Iemitsu M, Miyauchi T, Maeda S, Tanabe T, Takanashi M, Irukayama-Tomobe Y, et al . Aging-induced decrease in the PPAR-a level in hearts is improved by exercise training. Am J Physiol Heart Circ Physiol 2002;283:H1750-60.  Back to cited text no. 41    
42.Legakis JE, Koepke JI, Jedeszko C, Barlaskar F, Terlecky LJ, Edwards HJ, et al . Peroxisomes senescence in human fibroblasts. Mol Biol Cell 2002;13:4243-55.  Back to cited text no. 42    
43.Arnaiz SL, Travacio M, Llesuy S, Boveris A. Hydrogen peroxide metabolism during peroxisome proliferation by fenofibrate. Biochim Biophys Acta 1995;1272:175-80.  Back to cited text no. 43    
44.Andersen HR, Nielsen JB, Nielsen F, Grandjean P. Antioxidative enzyme activities in human erythrocytes. Clin Chem 1997;43:562-8.  Back to cited text no. 44    
45.Jozwiak Z, Jasnowska B. Changes in oxygen-metabolizing enzymes and lipid peroxidation in human erythrocytes as a function of age of donor. Mech Ageing Dev 1985;32:77-83.  Back to cited text no. 45    
46.Ceballos-Picot I, Trivier JM, Nicole A, Sinet PM, Thevenin M. Age-correlated modifications of copper-zinc dismutase and glutathione-related enzyme activities in human erythrocytes. Clin Chem 1992;38:66-70.  Back to cited text no. 46    
47.Constantin A, Constantinescu E, Dumitrescu M, Calin A, Popov D. Effects of ageing on carbonyl stress and antioxidant defense in RBCs of obese Type 2 diabetic patients. J Cell Mol Med 2005;9:683-91.  Back to cited text no. 47    
48.Di Massimo C, Lo Presti R, Corbacelli C, Pompei A, Scarpelli P, De Amicis D, et al . Impairment of plasma nitric oxide availability in senescent healthy individuals: Apparent involvement of extracellular superoxide dismutase activity. Clin Hemorheol Microcirc 2006;35:231-7.  Back to cited text no. 48    
49.Judge S, Jang YM, Smith A, Hagen T, Leewenburgh C. Age-associated increases in oxidative stress and antioxidant enzyme activities in cardiac interfibrillar mitochondria: Implications for the mitochondrial theory of aging. FASEB J 2005;19:419-21.  Back to cited text no. 49    
50.Tower J. Aging mechanisms in fruit files. J Bioessays 1996;18:779-807.  Back to cited text no. 50    
51.Aggarwal A, Nicholson G. Age dependent penetrance of three different superoxide dismutase 1 (sod 1) mutations. Int J Neurosci 2005;115:1119-30.  Back to cited text no. 51    
52.Van Remmen H, Ikeno Y, Hamilton M, Pahlavani M, Wolf N, Thorpe SR, et al . Life-long reduction in MnSOD activity results in increased DNA damage and higher incidence of cancer but does not accelerate aging. Physiol Genomics 2003;16:29-37.  Back to cited text no. 52    
53.Melov S, Coskun PE, Wallace DC. Mouse models of mitochondrial disease, oxidative stress and senescence. Mutat Res 1999;434:233-42.  Back to cited text no. 53    
54.Orr WC, Sohal RS. Extension of life span by over expression of superoxide dismutase and catalase in Drosophila melanogaster. Science 1994;263:1128-30.  Back to cited text no. 54    
55.Carlsson LM, Jonsson J, Edlund T, Marklund SL. Mice lacking extracellular superoxide dismutase are more sensitive to hyperoxia. Proc Natl Acad Sci USA 1995;92:6264-8.  Back to cited text no. 55    
56.Cabreiro F, Picot CR, Friguet B, Petroupoulos I. Methionine sulfoxide reductases: Relevance to aging and protection against oxidative stress. Ann NY Acad Sci 2006;1067:37-44.  Back to cited text no. 56    
57.Moskovitz J, Bar-Noy S, Williams WM, Requena J, Berlett BS, Stadtman ER. Methionine sulfoxide reductase (MsrA) is a regulator of antioxidant defense and lifespan in mammals. Proc Natl Acad Sci USA 2001;98:12920-5.  Back to cited text no. 57    
58.Michelet F, Gueguen R, Leroy P, Wellman M, Nicolas A, Siest G. Blood and plasma glutathione measured in healthy subjects by HPLC: Relation to sex, aging, biological variables and life habits. Clin Chem 1995;41:1509-17.  Back to cited text no. 58    
59.Lenton KJ, Therriault H, Cantin AM, Fulop T, Payette H, Wagner JR. Direct correlation of glutathione and ascorbate and their dependence on age season in human lymphocytes. Am J Clin Nutr 2000;71:1194-200.  Back to cited text no. 59    
60.Srinivasan V, Pandi-Perumal SR, Maestroni GJ, Esquifino AI, Hardeland R, Cardinali DP. Role of melatonin in neurodegenerative diseases. Neurotox Res 2005;7:293-318.  Back to cited text no. 60    
61.Tietz NW, Shuey DF, Wekstein DR. Laboratory values in fit aging individuals-sexagenarians through centenarians. Clin Chem 1992;38:1167-85.  Back to cited text no. 61    
62.Pinzani P, Petruzzi E orlando C, Stefanescu A, Antonini MF, Serio M, et al . Reduced serum antioxidant capacity in healthy centenarians. Clin Chem 1997;43:855-6.  Back to cited text no. 62    
63.Vatassery GT, Johnson GJ, Krezowski AM. Changes in vitamin E concentrations in human plasma and platelets with age. J Am Coll Nutr 1983;2:369-75.  Back to cited text no. 63    
64.Kalen A, Appelkvist EL, Dallner G. Age-related changes in the lipid compositions of rat and human tissues. Lipids 1989;24:579-84.  Back to cited text no. 64    
65.Pacifici RE, Davies KJ. Protein, lipid and DNA repair systems in oxidative stress: the free-radical theory of aging revisited. Gerontology 1991;37:166-80.  Back to cited text no. 65    
66.Terman A, Gustafsson B, Brunk UT. Mitochondrial damage and intralysosomal degradation in cellular aging. Mol Aspects Med 2006;27:471-82.  Back to cited text no. 66    
67.Brunk UT, Jones CB, Sohal RS. A novel hypothesis of lipofuscinogenesis and cellular aging based on interaction between oxidative stress and autophagocitosis. Mutat Res 1992;275:395-403.  Back to cited text no. 67    
68.Terman A. Garbage catastrophe theory of aging: Imperfect removal of oxidative damage? Redox Rep 2001;6:15-26.  Back to cited text no. 68    
69.Cortopassi GA, Wang E. There is substantial agreement among interspecies estimates of DNA repair activity. Mech Ageing Dev 1996;91:211-8.  Back to cited text no. 69    
70.Grune T, Reinheckel T, Davies KJ. Degradation of oxidised proteins in mammalian cells. FASEB J 1997;11:526-34.  Back to cited text no. 70    
71.Stadtman ER. Protein oxidation and aging. Science 1992;257:1220-4.  Back to cited text no. 71    
72.Hoshi T, Heinemann S. Regulation of cell function by methionine oxidation and reduction. J Physiol 2001;531:1-11.  Back to cited text no. 72    
73.Melhorn RJ. Oxidants and antioxidants in aging. In : Raton B, editors. Physiological basis of aging and geriatrics. 3 rd ed. Timiras PS, FL: CRC; 2003. p. 61-83.  Back to cited text no. 73    
74.Shringarpure R, Davies KJ. Protein turnover by the proteasome in aging and disease. Free Radic Biol Med 2002;32:1084-9.  Back to cited text no. 74    
75.Sell DR, Lane MA, Johnson WA, Masoro EJ, Mock OB, Reiser KM, et al . Longevity and the genetic determination of collagen glycoxidation kinetics in mammalian senescence. Proc Natl Acad Sci USA 1996;93:485-90.  Back to cited text no. 75    
76.Beckman KB, Ames BN. Oxidative decay of DNA. J Biol Chem 1997;272:19633-6.  Back to cited text no. 76    
77.Cadenas E, Davies KJ. Mitochondrial free radical generation, oxidative stress and aging. Free Rad Biol Med 2000;29:222-30.  Back to cited text no. 77    
78.Richter C. Oxidative damage to mitochondrial DNA and is relationship to ageing. Int J Biochem Cell Biol 1995;27:647-53.  Back to cited text no. 78    
79.Brunk UT, Terman A. The mitochondrial-lysosomal axis theory of aging. Accumulation of damaged mitochondria as a result of imperfect autophagocytosis. Eur J Biochem 2002;269:1996-2002.  Back to cited text no. 79    
80.Musci G, Bonaccorsi di Patti MC, Fagiolo U, Calabrese L. Age-related changes in human ceruloplasmin. Evidence for oxidative modifications. J Biol Chem 1993;268:13388-95.  Back to cited text no. 80    
81.Colavitti R, Finkel T. Reactive oxygen species as mediators of cellular senescence. IUBMB Life 2005;57:277-81.  Back to cited text no. 81    
82.Stoian I, Oros A, Moldoveanu E. Apoptosis and free radicals. Biochem Mol Med 1996;59:93-7.  Back to cited text no. 82    
83.Afanas'ev I. Interplay between superoxide and nitric oxide in aging and diseases. Biogerontology 2004;5:267-70.  Back to cited text no. 83    
84.Pollack M, Phaneuf S, Dirks A, Leeuwenburgh C. The role of apoptosis in the normal aging brain, skeletal muscle and heart. Ann N Y Acad Sci 2002;959:93-107.  Back to cited text no. 84    
85.Afanas'ev IB. On mechanism of superoxide signaling under physiological and pathophysiological conditions. Med Hypotheses 2005;64:127-9.  Back to cited text no. 85    
86.Moldoveanu E, Stoian I, Voinea L, Marta D, Popescu LM. BCL-2-- general considerations. Haematologia (Budap) 1998;29:167-80.  Back to cited text no. 86    
87.Chung L, Ng YC. Age-related alterations in expression of apoptosis regulatory proteins and heat shock proteins in rat skeletal muscle. Biochim Biophys Acta 2006;1762:103-9.  Back to cited text no. 87    
88.Greeve I, Hermans-Borgmeyer I, Brellinger C, Kasper D, Gomez-Isla T, Behl C, et al . The human DIMINUTO/DWARF1 homolog seladin-1 confers resistance to Alzheimer's disease-associated neurodegeneration and oxidative stress. J Neurosci 2000;20:7345-52.  Back to cited text no. 88    
89.Wu C, Miloslavskaya I, Demontis S, Maestro R, Galaktiov K. Regulation of cellular response to oncogenic and oxidative stress by Seladin-1. Nature 2004;432:640-5.  Back to cited text no. 89    
90.Harley CB, Futcher AB, Greider CW. Telomeres shorten during ageing of human fibroblasts. Nature 1990;345:458-60.  Back to cited text no. 90    
91.von Zglinicki T, Saretzki G, Docke W, Lotze C. Mild hyperoxia shortens telomeres and inhibits proliferation of fibroblasts: A model for senescence? Exp Cell Res 1995;220:186-93.  Back to cited text no. 91    
92.van de ven M, Andressoo JO, Holcomb VB, von Lindern M, Jong WM, Zeeuw CI, et al . Adaptive stress response in segmental progeria resembles long-lived dwarfism and calorie restriction in mice. PLoS Genet 2006;2:e192.  Back to cited text no. 92    
93.Warner HR. Superoxide dismutase, aging and degenerative disease. Free Radic Biol Med 1994;17:249-58.  Back to cited text no. 93    
94.Bayne C, Mockett RJ, Orr WC, Sohal RS. Enhanced catabolism of mitochondrial superoxide/hydrogen peroxide and aging in transgenic Drosophila. Biochem J 2005;391:277-84.  Back to cited text no. 94    
95.Taufer M, Peres A, de Andrade VM, de Oliveira G, Sa G, do Canto ME, et al . Is the Val16Ala manganese superoxide dismutase polymorphism associated with the aging process. J Gerontol A Biol Sci Med Sci 2005;60:432-8.  Back to cited text no. 95    
96.Genkinger JM, Platz EA, Hoffman SC, Strickland P, Huang HY, Comstock GW, et al . C47T polymorphism in manganese superoxide dismutase (mnSOD), antioxidant intake and survival. Mech Ageing Dev 2006;127:371-7.  Back to cited text no. 96    
97.Christiansen L, Petersen HC, Bathum L, Frederiksen H, McGue M, Christensen K. The catalase -262C/T promoter polymorphism and aging phenotypes. J Gerontol A Biol Sci Med Sci 2004;59:B886-9.  Back to cited text no. 97    
98.Pandolfi S, Bonafe M, Di Tella L, Tiberi L, Salvioli S, Monti D, et al . p66(shc) is highly expressed in fibroblasts form centenarians. Mech Ageing Dev 2005;126:839-44.  Back to cited text no. 98    
99.Howes RM. The free radical fantasy: A panoply of paradoxes. Ann N Y Acad Sci 2006;1067:22-6.  Back to cited text no. 99    
100.Kregel KC, Zhang HJ. An integrated view of oxidative stress in aging: Basic mechanisms, functional effects and pathological considerations. Am J Physiol Regul Integr Physiol 2007;292:R18-36.  Back to cited text no. 100    

Copyright 2007 - Journal of Postgraduate Medicine


The following images related to this document are available:

Photo images

[jp07074t2.jpg] [jp07074t1.jpg]
Home Faq Resources Email Bioline
© Bioline International, 1989 - 2024, Site last up-dated on 01-Sep-2022.
Site created and maintained by the Reference Center on Environmental Information, CRIA, Brazil
System hosted by the Google Cloud Platform, GCP, Brazil