Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2011 Aug 1.
Published in final edited form as: Biochim Biophys Acta. 2010 Feb 20;1801(8):924–929. doi: 10.1016/j.bbalip.2010.02.005

Involvements of the Lipid Peroxidation Product, HNE, in the Pathogenesis and Progression of Alzheimer’s DiseaseA

D Allan Butterfield 1,*, Miranda L Bader Lange 1, Rukhsana Sultana 1
PMCID: PMC2886168  NIHMSID: NIHMS182231  PMID: 20176130

Abstract

Alzheimer’s disease (AD) is an age-related neurodegenerative disorder. A number of hypotheses have been proposed to explain AD pathogenesis. One such hypothesis proposed to explain AD pathogenesis is the oxidative stress hypothesis. Increased levels of oxidative stress markers including the markers of lipid peroxidation such as acrolein, 4-hydroxy-2-trans-nonenal (HNE), malondialdehyde, etc. are found in brains of AD subjects. In this review we focus principally on research conducted in the area of HNE in the central nervous system (CNS) of AD and mild cognitive impairment (MCI), and further we discuss likely consequences of lipid peroxidation with respect to AD pathogenesis and progression. Based on the research conducted so far in the area of lipid peroxidation, it is suggested that lipid accessible antioxidant molecules could be a promising therapeutic approach to treat or slow progression of MCI and AD.

Keywords: lipid peroxidation, Alzheimer disease, HNE, amyloid beta-peptide, proteomics, oxidatively modified proteins


Alzheimer’s disease (AD) is an age-related neurodegenerative disorder characterized histopathologically by the presence of senile plaques (SP), neurofibrillary tangles (NFT), and synapse loss [1]. Though AD has been known for more than a century, the exact mechanism(s) of its pathogenesis still largely remains unknown. A number of hypotheses such as amyloid cascade, excitotoxicity, oxidative stress, and inflammation hypotheses, etc., have been proposed for AD pathogenesis; however, none of these hypotheses clearly account totally for all aspects of the disease. One such hypothesis proposed to explain AD pathogenesis is the oxidative stress hypothesis [2-6] . Oxidative stress is defined as an imbalance in the levels of oxidant (reactive oxygen species (ROS)/ reactive nitrogen species (RNS)) and antioxidant defense systems. This increase in the ROS/RNS may further lead to the damage of biomolecules leading to loss of function and consequently to cell loss, one of the key observations in AD brain.

The presence of high levels of unsaturated lipid content coupled to high oxygen utilization, high level of redox metal ions, and relatively poor antioxidant systems makes brain particularly vulnerable to oxidative damage. In the case of AD, amyloid β-peptide (40-42 amino acids) [Aβ(1-40) or Aβ(1-42)], a main component of SP, is generated by the proteolytic cleavage of amyloid precursor protein by the action beta- and gamma-secretases. Aβ(1-42) has been shown to induce oxidative stress in both in vitro and in vivo studies [7, 8]. Aβ(1-42) exists in various aggregated states such as monomers, oligomers, protofibrils, and fibrils, among which the oligomeric form of Aβ(1-42) is considered as highly toxic [9]. The increased toxicity associated with oligomeric Aβ(1-42) may be attributable to its ability to reside in the lipid bilayer, where lipid peroxidation can occur. Indeed, Mattson and co-workers showed that addition of Aβ(1-42) to neurons directly lead to formation of HNE [10], as did our laboratory [12].

Extensive experimental evidence from our laboratory and others showed that methionine at residue 35 of Aβ(1-42) is particularly important for its oxidative role [4, 11, 12]. Methionine can undergo one-electron oxidation to form a sulfuranyl radical cation, which has the ability to abstract an allylic H-atom from the unsaturated acyl chains of lipid molecules, thereby leading to the initiation of lipid peroxidation processes (Figure 1) [13, 14]. Methione 35 in Aβ(1-42)-mediated lipid peroxidation requires a helical secondary structure of the peptide. This secondary structure, of course, is the case for most proteins that are bilayer-resident. Indeed, NMR studies showed that Aβ(1-42) is helical when solubilized in micelles (reviewed in [9]). Like any alpha-helix, every fourth amino acid interacts with each other (i+4 rule of helices). In particular, the backbone carbonyl of Ile-31 is located within a van der Waals distance of the S-atom of Met-35 in Aβ(1-42). Since O is more electronegative than S, electron density of the lone pair of electrons on S in Met are drawn away from the S and toward the O, making these electrons more vulnerable to a one-electron oxidation to form the sulfuranyl radical cation. This radical can abstract a labile allylic H atom from an unsaturated acyl chain of lipids forming a carbon-centered free radical. The latter can immediately bind paramagnetic and non-polar oxygen to form the peroxyl free radical, which, in turn, can abstract another acyl chain-resident labile allylic H-atom, continuing the chain reaction. Note that there is a large amplification effect of a free radical on Aβ(1-42) that is mediated by the chain reaction within the lipid phase of the membrane. Note also (Figure 2) that the lipid acyl hydroperoxide formed by these reactions can lead directly to HNE. The acid formed on the sulfuranyl radical by abstraction of a labile allylic H-atom from lipid acyl chains has a pKa of -5; hence, any base, including water, can remove this H+, resulting in reduced Met again. That is, the Met acts as a catalyst for lipid peroxidation. This chemistry is discussed in greater detail in [9].

Figure 1. Involvement of Met-35 of Aβ(1-42) in lipid peroxidation.

Figure 1

The S-atom of Met-35 of the Aβ(1-42) peptide can undergo one-electron oxidation to form a sulfuranyl radical cation within the bilayer, which has the ability to abstract a labile, allylic H-atom from the unsaturated acyl chains of lipid molecules, leading to initiation of lipid peroxidation processes [13, 14]. Like most integral membrane proteins, α-helical Aβ(1-42) adheres to the i+4 rule, causing the backbone carbonyl oxygen of Ile-31 on Aβ(1-42) to draw the electron density of the Met-35 S-atom toward itself, making the S-atom more vulnerable to oxidation and subsequent formation of the sulfuranyl radical cation. The sulfuranyl radical can, in turn, abstract allylic H-atoms from neighboring fatty acyl chains within the bilayer, forming a fatty acid carbon-centered free radical that can immediately bind paramagnetic, non-polar oxygen (O2) to form a peroxyl free radical. The peroxyl radical then abstracts another labile H-atom from nearby fatty acyl chains, perpetuating the catalytic chain reaction initiated by Met-35 of the Aβ(1-42) peptide. Because Met-35 is inevitably reduced back to its starting state, this reaction can begin again, amplifying the neurotoxic affects of the Aβ(1-42) peptide within the cell.

Figure 2. Formation of 4-hydroxy-2-trans-nonenal (HNE) from arachidonic acid.

Figure 2

The catalytic conversion of the Aβ(1-42) Met-35 S-atom to the sulfuranyl radical cation leads to the abstraction of a labile, allylic H-atom from unsaturated fatty acyl chains within the bilayer. In particular, arachidonic acid is a common fatty acid within the bilayer that is readily oxidized to produce one of the highly reactive lipid peroxidation products, HNE. Reactive oxygen species (ROS) other than Aβ(1-42) can also oxidize arachidonic acid to form a reactive hydroperoxide intermediate that is quickly converted to a peroxyl radical by Fe2+ (Fenton chemistry). The highly reactive peroxyl radical causes a molecular rearrangement which cyclizes the radical, arachidonic acid intermediate. Further oxidation and Fenton chemistry results in the β-scission of the cyclized intermediate, causing eventual formation of HNE.

Recent studies reported the presence of Aβ(1-42) in mitochondrial membranes [15, 16]. Hence, Aβ(1-42) may initiate lipid peroxidation in the mitochondrial membrane by similar processes as discussed above that may not just lead to alterations in lipid components of the membrane, but affect proteins embedded in the membrane. Consequently the process of lipid peroxidation may lead to alteration in membrane fluidity and eventually lead to alterations in membrane functions. In the case of mitochondria, alteration in the membrane may lead to leakage of apoptosis-inducing molecules such as cytochrome C from the mitochondria, and also cause functional alterations of proteins involved in the electron transport system, all of which may lead to increased release and production of RNS and ROS. In addition, an in vivo study showed that increased lipid peroxidation leads to up regulation of BACE1 expression, which may lead to increased Aβ(1-42) production [17]. Lipid peroxidation products and Aβ(1-42) have been shown to induce JNK pathways, leading to neuronal apoptosis [18].

As noted above, lipid peroxidation leads to the production of HNE [Figure 2], malondialdehyde, and the α,β-unsaturated aldehyde, acrolein, which are diffusible and highly reactive with other biomolecules and, consequently, neurotoxic [19]. Two additional markers of lipid peroxidation are known, i.e., isoprostanes and neuroprostanes, that are products of arachidonic and docosahexaenoic acid oxidation, respectively, the latter fatty acid being neuronal specific [20]. The aldehydic products of lipid peroxidation are highly reactive and covalently bind to proteins through Michael addition to protein cysteines, lysines, and histidines, altering their structure [21] and function [14, 22, 23] (Figure 3). Previous studies showed that the levels of free HNE and acrolein are increased in AD brain [24, 25].

Figure 3. 4-hydroxy-2-trans-nonenal (HNE) adduction of lysine, histidine, and cysteine.

Figure 3

HNE is a membrane diffusible, highly reactive alkenal species (formed by oxidation of arachidonic acid) that can readily oxidize other biomolecules, contributing to AD neurotoxicity [19]. HNE can covalently binds to lysine, histidine, and cysteine residues of proteins via Michael addition, forming adducts that are known to change the structural conformation [21] and function of proteins [14, 22, 23]. Resultant HNE-adduct structures with Lys, His, and Cys are shown.

Lipid Peroxidation in AD Brain

Lipid peroxidation products including free HNE, acrolein, neuroprostanes, isoprostanes are elevated in AD brain [26-28]. A recent study reported increased levels of specific HNE-histidine Michael adducts in AD hippocampus compared to age-matched controls [29], confirming our earlier findings [14]. Further, Lui et al., [30] showed that HNE also can covalently modify the histidine side chains of Aβ, leading to increased aggregation of this peptide.

Increased levels of the GSH-HNE Michael adduct (HNE-GSH) were found in the AD hippocampus, and substantia innominata, entorhinal cortex, frontal and temporal cortex, as well as cerebellum [31]. In normal cells the HNE-GSH adducts are eliminated by multidrug resistant protein 1 (MRP-1); however, in AD brain, glutathione S-transferase (GST) and MRP-1 were found to be HNE- modified, which might account for loss of GST activity in AD [32, 33], and contribute to the increased levels of HNE and accumulation of HNE-protein adducts. In contrast, the expression and activity of aldehyde dehydrogenase, which converts HNE into an acid (which then abrogates its Michael addition properties), was reported to be altered in AD brain [34] . Further, AD brain demonstrated decreased activity and HNE modification of the proteasome, which in turn leads to increased accumulation of cytotoxic biomolecues, thereby increasing neuronal cytotoxicity [32, 33, 35]. In addition, neprilysin (NEP), a major protease that cleaves Aβ in vivo, also has also been shown to be HNE- modified in the brain of AD subjects [36].

The levels of F(2)-isoprostanes [F(2)-IsoP], F(4)-neuroprostane [F(4)-NP] and isoprostane 8,12-iso-iPF2(α)-VI were found to be increased in AD brain compared to controls [37, 38]. Another product of lipid peroxidation, MDA, was found to be significantly increased and colocalized with SP and NFT in AD brain [39]. Moreover, the levels of MDA correlated with the decreased activity of superoxide dismutase (SOD) [40]. Pamplona et al., (2005)[41] also found increased amounts of the direct oxidation of amino acids, glycoxidation and lipoxidation in AD brain as indicated by significantly decreased levels of docosohexanoic acid, a good substrate for lipid peroxidation, and increased concentrations of glutamic and aminoadipic semialdehydes, Nε - (carboxymethyl)-lysine, Nε -(carboxyethyl)-lysine, and Nε -(malondialdehyde)-lysine. Additional markers of protein oxidation were identified as neurofilament L, α-tubulin, glial fibrillary acidic protein, ubiquinol-cytochrome c reductase complex protein I, and the β-chain of ATP synthase, which are targets of Nε -(malondialdehyde)-lysine formation.

Liu et al (2006) [42] reported increased levels of HNE bound to 2′- deoxyguanosine (HNE-dG) in AD brain compared to controls. In contrast, a previous study by Gotz et al., [43] reported no difference in the levels 1,N2-propanodeoxyguanosine adducts of HNE (HNE-dGp) in the AD brain compared to controls.

A CSF study by Pratico et al., [44] demonstrated increased levels of the isoprostane 8,12-iso-iPF2(α)-VI in cerebral spinal fluid (CSF) in AD. These researchers also showed that AD patients with a ventriculoperitoneal (VP) shunt had a 51% decrease in lipid peroxidation products after a year, a finding that led these researchers to suggest that improving CSF drainage may remove the end products of lipid peroxidation from the CSF and lead to decreased damage to brain lipids. In another study, the levels of CSF F2-isoprostanes (IsoPs), which were reported to be reduced upon treatment of AD patients with α-tocopherol and vitamin C, suggesting that antioxidants may be a promising therapeutic approach to treat AD [45].

Lipid Peroxidation in MCI

Mild cognitive impairment (MCI) is arguably the earliest form of AD. Brain from MCI subjects showed increase levels of TBARs, MDA, free HNE and protein-bound HNE [46-48]. Other makers of lipid peroxidation, i.e., F(2)-IsoP, F(4)-NP and acrolein, were also found to be significantly increased in MCI brain [37, 47]. Further, in MCI, Pratico et al., [49] reported elevated levels of the F2-isoprostane 8,12-iso-iPF2(α)-VI in CSF, plasma, and urine and suggested that this isoprostane could be used as a potential marker to identify MCI individuals who are at higher risk of progressing to AD.

HNE-Modified Proteins in AD and MCI Brain

Proteomics studies from our laboratory have reported a large number of proteins in AD brain that showed increased levels of protein-bound HNE, including: ATP synthase, α-enolase, aconitase, aldolase, glutamine synthetase (GS), MnSOD, peroxiredoxin 6, dihydropyriminidase related protein-2 (DRP-2), and α-tubulin [50]. These proteins play important roles in regulating various cellular functions such as glucose metabolism, glutamate levels, antioxidant defense systems, axonal growth, and structural functions, all of which are reported to be altered in AD brain. Some of these HNE-bound proteins identified by proteomics in AD brain were previously found to be either nitrated or carbonylated in AD [51-55]. Further, proteomics studies in MCI identified increased levels of protein-bound HNE for neuropolypeptide h3, carbonyl reductase (NADPH), α-enolase, lactate dehydrogenase B, phosphoglycerate kinase, heat shock protein 70, ATP synthase α-chain, pyruvate kinase, actin, elongation factor Tu, and translation initiation factor α [56]. The MCI HNE-modified proteins regulate various cellular functions such as glucose metabolism, protein synthesis, protein structural changes, lipid metabolism, antioxidant defense systems, and axonal growth. Some of the proteins that showed increased HNE modification in MCI are also found to be HNE modified in AD brain, consistent with the notion that these proteins may contribute to the progression of MCI to AD. Further, by using immunoprecipitation techniques, HNE-bound p53 levels were reported to be elevated in AD brain, while MCI brain showed a trend toward increase HNE levels [57]. HNE modification of proteins has been shown to lead to loss of protein function. The appearance of common targets of HNE-modified proteins in AD and MCI brain suggests that these proteins may be key players in AD pathogenesis.

Phospholipid Asymmetry and AD and MCI

Lipid peroxidation within the cell bilayer can also affect the distribution of membrane phospholipids, as previous studies have reported an abnormal composition of aminophospholipids in AD brain [58]. In particular, the asymmetric distribution of the anionic aminophospholipid phosphatidylserine (PtdSer) has been shown to be significantly altered by lipid peroxidation products, such as HNE and acrolein, produced in the bilayer [13, 24, 47, 59-66] as a result of the highly oxidative environment that is a hallmark of MCI and AD pathology. Normally PtdSer can be found sequestered to the cytosolic, inner-leaflet of the lipid bilayer, an asymmetric distribution that is selectively regulated by the ATP-dependent, membrane-bound, aminophospholipid translocase, flippase, which unidirectionally transports PtdSer inward against its concentration gradient [67]. Collapse of PtdSer asymmetry results in the outer-leaflet exposure of this phospholipid, which signals induction of early apoptosis and is crucial for selective recognition and mononuclear phagocytosis of target cells by macrophages and fibroblasts in the periphery or microglia in the brain [59-63, 68]. Moreover, exposure of PtdSer to the outer-leaflet has been shown to affect activity of membrane receptors and transport proteins, as well as signal transduction and cellular morphology [69-71]. Therefore, asymmetric distribution of phospholipids, like PtdSer, within the bilayer is critical to the maintenance of cellular homeostasis.

Because the oxidative modification of proteins and lipids by ROS and/or RNS during disease progression ultimately results in diminished and/or complete loss of protein function [2], it is likely that oxidative modification of flippase and/or PtdSer by HNE or acrolein in the bilayer induces PtdSer asymmetric collapse [13, 24, 47, 59-66]. By diffusing from their formation sites, these reactive alkenals could react via Michael addition with flippase, covalently binding a critical cysteine residue of its primary structure, disrupting its translocase activity [24, 47, 59, 65, 67]. Furthermore, considering that externalization of PtdSer signals early apoptosis, exposure of PtdSer to the external bilayer leaflet may also result from the activation of pro-apoptotic proteins in MCI and AD. Indeed, reports demonstrate loss of flippase activity accompanied by the subsequent outer-leaflet exposure of PtdSer occurs downstream of caspase-3 activation, which is found to be increased in MCI and AD brain [72-78]. On the whole, however, it is evident that PtdSer asymmetric collapse is an important aspect of neurodegeneration of MCI and AD brain, regardless of exposure route. Considering PtdSer asymmetric collapse signals the initiation of early apoptotic events within the cell, it appears that PtdSer externalization could be a potential link to those patients with MCI that may eventually develop AD.

Conclusions

A large body of evidence, as discussed above, demonstrates the specific role of HNE modification of proteins in AD progression and pathogenesis and suggests that lipid accessible antioxidant molecules could be a promising therapeutic approach used to treat MCI and AD. However, Vitamin E studies in clinical trials of AD and MCI have failed, possibly explained based on the lack of reducing equivalents given with Vitamin E (e.g., GSH or Vitamin C). Alternatively, the present antioxidant status of subjects treated with antioxidants generally was not assessed. One could imagine that subjects with high antioxidants status might be non-responders to farther antioxidant treatment. Lastly, it is conceivable that antioxidant trials require a significantly long treatment to show efficacy. [79-81]. The lack of protection of antioxidants such as Vitamin E may also be an indication that lipid peroxidation is not a key process in the progression of disease. However, given that lipid peroxidation is prevalent in brain and CSF of subjects with amnestic MCI (which has no dementia), as discussed above, such a conclusion seems unlikely to be correct. Further studies are in progress to better understand the role of lipid peroxidation in the pathogenesis and progression of AD.

Acknowledgement

This work was supported in part by grants from the National Institutes of Health AG-05119; AG-10836; AG-029839) to D.A.B.

Footnotes

A

This paper is dedicated to the life and legacy of William R. Markesbery, MD, who died January 30, 2010, and who contributed enormously to the concept of brain oxidative stress in Alzheimer disease.

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • [1].Selkoe DJ. Alzheimer’s disease: genes, proteins, and therapy. Physiol Rev. 2001;81:741–766. doi: 10.1152/physrev.2001.81.2.741. [DOI] [PubMed] [Google Scholar]
  • [2].Butterfield DA, Stadtman ER. Protein Oxidation processes in aging brain. Adv Cell Aging Gerontol. 1997;2:161–191. [Google Scholar]
  • [3].Markesbery WR. Oxidative stress hypothesis in Alzheimer’s disease. Free radical biology & medicine. 1997;23:134–147. doi: 10.1016/s0891-5849(96)00629-6. [DOI] [PubMed] [Google Scholar]
  • [4].Butterfield DA, Reed T, Newman SF, Sultana R. Roles of amyloid beta-peptide-associated oxidative stress and brain protein modifications in the pathogenesis of Alzheimer’s disease and mild cognitive impairment. Free Radic Biol Med. 2007;43:658–677. doi: 10.1016/j.freeradbiomed.2007.05.037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [5].Nunomura A, Moreira PI, Lee HG, Zhu X, Castellani RJ, Smith MA, Perry G. Neuronal death and survival under oxidative stress in Alzheimer and Parkinson diseases. CNS Neurol Disord Drug Targets. 2007;6:411–423. doi: 10.2174/187152707783399201. [DOI] [PubMed] [Google Scholar]
  • [6].Butterfield DA, Drake J, Pocernich C, Castegna A. Evidence of oxidative damage in Alzheimer’s disease brain: central role for amyloid beta-peptide. Trends Mol Med. 2001;7:548–554. doi: 10.1016/s1471-4914(01)02173-6. [DOI] [PubMed] [Google Scholar]
  • [7].Boyd-Kimball D, Mohmmad Abdul H, Reed T, Sultana R, Butterfield DA. Role of phenylalanine 20 in Alzheimer’s amyloid beta-peptide (1-42)-induced oxidative stress and neurotoxicity. Chem Res Toxicol. 2004;17:1743–1749. doi: 10.1021/tx049796w. [DOI] [PubMed] [Google Scholar]
  • [8].Boyd-Kimball D, Sultana R, Poon HF, Lynn BC, Casamenti F, Pepeu G, Klein JB, Butterfield DA. Proteomic identification of proteins specifically oxidized by intracerebral injection of amyloid beta-peptide (1-42) into rat brain: implications for Alzheimer’s disease. Neuroscience. 2005;132:313–324. doi: 10.1016/j.neuroscience.2004.12.022. [DOI] [PubMed] [Google Scholar]
  • [9].Glabe CC. Amyloid accumulation and pathogensis of Alzheimer’s disease: significance of monomeric, oligomeric and fibrillar Abeta. Subcell Biochem. 2005;38:167–177. doi: 10.1007/0-387-23226-5_8. [DOI] [PubMed] [Google Scholar]
  • [10].Mark RJ, Lovell MA, Markesbery WR, Uchida K, Mattson MP. A role for 4-hydroxynonenal, an aldehydic product of lipid peroxidation, in disruption of ion homeostasis and neuronal death induced by amyloid beta-peptide. J Neurochem. 1997;68:255–264. doi: 10.1046/j.1471-4159.1997.68010255.x. [DOI] [PubMed] [Google Scholar]
  • [11].Butterfield DA, Boyd-Kimball D. The critical role of methionine 35 in Alzheimer’s amyloid beta-peptide (1-42)-induced oxidative stress and neurotoxicity. Biochimica et biophysica acta. 2005;1703:149–156. doi: 10.1016/j.bbapap.2004.10.014. [DOI] [PubMed] [Google Scholar]
  • [12].Butterfield DA, Galvan V, Lange MB, Tang H, Sowell RA, Spilman P, Fombonne J, Gorostiza O, Zhang J, Sultana R, Bredesen DE. In vivo oxidative stress in brain of Alzheimer disease transgenic mice: Requirement for methionine 35 in amyloid beta-peptide of APP. Free radical biology & medicine. 2009 doi: 10.1016/j.freeradbiomed.2009.10.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [13].Butterfield DA, Castegna A, Lauderback CM, Drake J. Evidence that amyloid beta-peptide-induced lipid peroxidation and its sequelae in Alzheimer’s disease brain contribute to neuronal death. Neurobiol Aging. 2002;23:655–664. doi: 10.1016/s0197-4580(01)00340-2. [DOI] [PubMed] [Google Scholar]
  • [14].Lauderback CM, Hackett JM, Huang FF, Keller JN, Szweda LI, Markesbery WR, Butterfield DA. The glial glutamate transporter, GLT-1, is oxidatively modified by 4-hydroxy-2-nonenal in the Alzheimer’s disease brain: the role of Abeta1-42. J Neurochem. 2001;78:413–416. doi: 10.1046/j.1471-4159.2001.00451.x. [DOI] [PubMed] [Google Scholar]
  • [15].Reddy PH. Amyloid beta, mitochondrial structural and functional dynamics in Alzheimer’s disease. Exp Neurol. 2009;218:286–292. doi: 10.1016/j.expneurol.2009.03.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [16].Sultana R, Butterfield DA. Oxidatively modified, mitochondria-relevant brain proteins in subjects with Alzheimer disease and mild cognitive impairment. J Bioenerg Biomembr. 2009 doi: 10.1007/s10863-009-9241-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [17].Tamagno E, Parola M, Bardini P, Piccini A, Borghi R, Guglielmotto M, Santoro G, Davit A, Danni O, Smith MA, Perry G, Tabaton M. Beta-site APP cleaving enzyme up-regulation induced by 4-hydroxynonenal is mediated by stress-activated protein kinases pathways. J Neurochem. 2005;92:628–636. doi: 10.1111/j.1471-4159.2004.02895.x. [DOI] [PubMed] [Google Scholar]
  • [18].Tang SC, Lathia JD, Selvaraj PK, Jo DG, Mughal MR, Cheng A, Siler DA, Markesbery WR, Arumugam TV, Mattson MP. Toll-like receptor-4 mediates neuronal apoptosis induced by amyloid beta-peptide and the membrane lipid peroxidation product 4-hydroxynonenal. Exp Neurol. 2008;213:114–121. doi: 10.1016/j.expneurol.2008.05.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [19].Esterbauer H, Schaur RJ, Zollner H. Chemistry and biochemistry of 4-hydroxynonenal, malonaldehyde and related aldehydes. Free radical biology & medicine. 1991;11:81–128. doi: 10.1016/0891-5849(91)90192-6. [DOI] [PubMed] [Google Scholar]
  • [20].Montine TJ, Morrow JD. Fatty acid oxidation in the pathogenesis of Alzheimer’s disease. Am J Pathol. 2005;166:1283–1289. doi: 10.1016/S0002-9440(10)62347-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [21].Subramaniam R, Roediger F, Jordan B, Mattson MP, Keller JN, Waeg G, Butterfield DA. The lipid peroxidation product, 4-hydroxy-2-trans-nonenal, alters the conformation of cortical synaptosomal membrane proteins. J Neurochem. 1997;69:1161–1169. doi: 10.1046/j.1471-4159.1997.69031161.x. [DOI] [PubMed] [Google Scholar]
  • [22].Pocernich CB, Butterfield DA. Acrolein inhibits NADH-linked mitochondrial enzyme activity: implications for Alzheimer’s disease. Neurotox Res. 2003;5:515–520. doi: 10.1007/BF03033161. [DOI] [PubMed] [Google Scholar]
  • [23].Drake J, Link CD, Butterfield DA. Oxidative stress precedes fibrillar deposition of Alzheimer’s disease amyloid beta-peptide (1-42) in a transgenic Caenorhabditis elegans model. Neurobiol Aging. 2003;24:415–420. doi: 10.1016/s0197-4580(02)00225-7. [DOI] [PubMed] [Google Scholar]
  • [24].Markesbery WR, Lovell MA. 4-hydroxynonenal, a product of lipid peroxidation, is increased in the brain in Alzheimer’s disease. Neurobiol. Aging. 1998;19:33–36. doi: 10.1016/s0197-4580(98)00009-8. [DOI] [PubMed] [Google Scholar]
  • [25].Montine KS, Olson SJ, Amarnath V, Whetsell WO, Jr., Graham DG, Montine TJ. Immunohistochemical detection of 4-hydroxy-2-nonenal adducts in Alzheimer’s disease is associated with inheritance of APOE4. Am J Pathol. 1997;150:437–443. [PMC free article] [PubMed] [Google Scholar]
  • [26].Sayre LM, Zelasko DA, Harris PL, Perry G, Salomon RG, Smith MA. 4-Hydroxynonenal-derived advanced lipid peroxidation end products are increased in Alzheimer’s disease. J Neurochem. 1997;68:2092–2097. doi: 10.1046/j.1471-4159.1997.68052092.x. [DOI] [PubMed] [Google Scholar]
  • [27].Lovell MA, Xie C, Markesbery WR. Acrolein is increased in Alzheimer’s disease brain and is toxic to primary hippocampal cultures. Neurobiol Aging. 2001;22:187–194. doi: 10.1016/s0197-4580(00)00235-9. [DOI] [PubMed] [Google Scholar]
  • [28].Markesbery WR, Lovell MA. Four-hydroxynonenal, a product of lipid peroxidation, is increased in the brain in Alzheimer’s disease. Neurobiol Aging. 1998;19:33–36. doi: 10.1016/s0197-4580(98)00009-8. [DOI] [PubMed] [Google Scholar]
  • [29].Fukuda M, Kanou F, Shimada N, Sawabe M, Saito Y, Murayama S, Hashimoto M, Maruyama N, Ishigami A. Elevated levels of 4-hydroxynonenal-histidine Michael adduct in the hippocampi of patients with Alzheimer’s disease. Biomed Res. 2009;30:227–233. doi: 10.2220/biomedres.30.227. [DOI] [PubMed] [Google Scholar]
  • [30].Liu L, Komatsu H, Murray IV, Axelsen PH. Promotion of amyloid beta protein misfolding and fibrillogenesis by a lipid oxidation product. J Mol Biol. 2008;377:1236–1250. doi: 10.1016/j.jmb.2008.01.057. [DOI] [PubMed] [Google Scholar]
  • [31].Volkel W, Sicilia T, Pahler A, Gsell W, Tatschner T, Jellinger K, Leblhuber F, Riederer P, Lutz WK, Gotz ME. Increased brain levels of 4-hydroxy-2-nonenal glutathione conjugates in severe Alzheimer’s disease. Neurochem Int. 2006;48:679–686. doi: 10.1016/j.neuint.2005.12.003. [DOI] [PubMed] [Google Scholar]
  • [32].Lovell MA, Xie C, Markesbery WR. Decreased glutathione transferase activity in brain and ventricular fluid in Alzheimer’s disease. Neurology. 1998;51:1562–1566. doi: 10.1212/wnl.51.6.1562. [DOI] [PubMed] [Google Scholar]
  • [33].Sultana R, Butterfield DA. Oxidatively modified GST and MRP1 in Alzheimer’s disease brain: implications for accumulation of reactive lipid peroxidation products. Neurochem Res. 2004;29:2215–2220. doi: 10.1007/s11064-004-7028-0. [DOI] [PubMed] [Google Scholar]
  • [34].Picklo MJ, Olson SJ, Markesbery WR, Montine TJ. Expression and activities of aldo-keto oxidoreductases in Alzheimer disease. J Neuropathol Exp Neurol. 2001;60:686–695. doi: 10.1093/jnen/60.7.686. [DOI] [PubMed] [Google Scholar]
  • [35].Cecarini V, Gee J, Fioretti E, Amici M, Angeletti M, Eleuteri AM, Keller JN. Protein oxidation and cellular homeostasis: Emphasis on metabolism. Biochimica et biophysica acta. 2007;1773:93–104. doi: 10.1016/j.bbamcr.2006.08.039. [DOI] [PubMed] [Google Scholar]
  • [36].Wang DS, Iwata N, Hama E, Saido TC, Dickson DW. Oxidized neprilysin in aging and Alzheimer’s disease brains. Biochem Biophys Res Commun. 2003;310:236–241. doi: 10.1016/j.bbrc.2003.09.003. [DOI] [PubMed] [Google Scholar]
  • [37].Markesbery WR, Kryscio RJ, Lovell MA, Morrow JD. Lipid peroxidation is an early event in the brain in amnestic mild cognitive impairment. Ann Neurol. 2005;58:730–735. doi: 10.1002/ana.20629. [DOI] [PubMed] [Google Scholar]
  • [38].Yao Y, Zhukareva V, Sung S, Clark CM, Rokach J, Lee VM, Trojanowski JQ, Pratico D. Enhanced brain levels of 8,12-iso-iPF2alpha-VI differentiate AD from frontotemporal dementia. Neurology. 2003;61:475–478. doi: 10.1212/01.wnl.0000070185.02546.5d. [DOI] [PubMed] [Google Scholar]
  • [39].Dei R, Takeda A, Niwa H, Li M, Nakagomi Y, Watanabe M, Inagaki T, Washimi Y, Yasuda Y, Horie K, Miyata T, Sobue G. Lipid peroxidation and advanced glycation end products in the brain in normal aging and in Alzheimer’s disease. Acta Neuropathol. 2002;104:113–122. doi: 10.1007/s00401-002-0523-y. [DOI] [PubMed] [Google Scholar]
  • [40].Casado A, Lopez-Fernandez M. Encarnacion, Casado M. Concepcion, de La Torre R. Lipid peroxidation and antioxidant enzyme activities in vascular and Alzheimer dementias. Neurochem Res. 2008;33:450–458. doi: 10.1007/s11064-007-9453-3. [DOI] [PubMed] [Google Scholar]
  • [41].Pamplona R, Dalfo E, Ayala V, Bellmunt MJ, Prat J, Ferrer I, Portero-Otin M. Proteins in human brain cortex are modified by oxidation, glycoxidation, and lipoxidation. Effects of Alzheimer disease and identification of lipoxidation targets. J Biol Chem. 2005;280:21522–21530. doi: 10.1074/jbc.M502255200. [DOI] [PubMed] [Google Scholar]
  • [42].Liu X, Lovell MA, Lynn BC. Detection and quantification of endogenous cyclic DNA adducts derived from trans-4-hydroxy-2-nonenal in human brain tissue by isotope dilution capillary liquid chromatography nanoelectrospray tandem mass spectrometry. Chem Res Toxicol. 2006;19:710–718. doi: 10.1021/tx0502903. [DOI] [PubMed] [Google Scholar]
  • [43].Gotz ME, Wacker M, Luckhaus C, Wanek P, Tatschner T, Jellinger K, Leblhuber F, Ransmayr G, Riederer P, Eder E. Unaltered brain levels of 1,N2-propanodeoxyguanosine adducts of trans-4-hydroxy-2-nonenal in Alzheimer’s disease. Neurosci Lett. 2002;324:49–52. doi: 10.1016/s0304-3940(02)00163-5. [DOI] [PubMed] [Google Scholar]
  • [44].Pratico D, Sung S. Lipid peroxidation and oxidative imbalance: early functional events in Alzheimer’s disease. J Alzheimers Dis. 2004;6:171–175. doi: 10.3233/jad-2004-6209. [DOI] [PubMed] [Google Scholar]
  • [45].Quinn JF, Montine KS, Moore M, Morrow JD, Kaye JA, Montine TJ. Suppression of longitudinal increase in CSF F2-isoprostanes in Alzheimer’s disease. J Alzheimers Dis. 2004;6:93–97. doi: 10.3233/jad-2004-6110. [DOI] [PubMed] [Google Scholar]
  • [46].Keller JN, Schmitt FA, Scheff SW, Ding Q, Chen Q, Butterfield DA, Markesbery WR. Evidence of increased oxidative damage in subjects with mild cognitive impairment. Neurology. 2005;64:1152–1156. doi: 10.1212/01.WNL.0000156156.13641.BA. [DOI] [PubMed] [Google Scholar]
  • [47].Williams TI, Lynn BC, Markesbery WR, Lovell MA. Increased levels of 4-hydroxynonenal and acrolein, neurotoxic markers of lipid peroxidation, in the brain in Mild Cognitive Impairment and early Alzheimer’s disease. Neurobiol Aging. 2006;27:1094–1099. doi: 10.1016/j.neurobiolaging.2005.06.004. [DOI] [PubMed] [Google Scholar]
  • [48].Butterfield DA, Reed T, Perluigi M, De Marco C, Coccia R, Cini C, Sultana R. Elevated protein-bound levels of the lipid peroxidation product, 4-hydroxy-2-nonenal, in brain from persons with mild cognitive impairment. Neurosci Lett. 2006;397:170–173. doi: 10.1016/j.neulet.2005.12.017. [DOI] [PubMed] [Google Scholar]
  • [49].Pratico D, Clark CM, Liun F, Rokach J, Lee VY, Trojanowski JQ. Increase of brain oxidative stress in mild cognitive impairment: a possible predictor of Alzheimer disease. Arch Neurol. 2002;59:972–976. doi: 10.1001/archneur.59.6.972. [DOI] [PubMed] [Google Scholar]
  • [50].Perluigi M, Sultana R, Cenini G, Di Domenico F, Memo M, Pierce WM, Coccia R, Butterfield DA. Redox Proteomics Identification of HNE-Modified Brain Proteins in Alzheimer’s Disease: Role of Lipid Peroxidation in Alzheimer’s Disease Pathogenesis. Proteomics--Clinical Applications. 2009;13:682–693. doi: 10.1002/prca.200800161. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [51].Castegna A, Aksenov M, Aksenova M, Thongboonkerd V, Klein JB, Pierce WM, Booze R, Markesbery WR, Butterfield DA. Proteomic identification of oxidatively modified proteins in Alzheimer’s disease brain. Part I: creatine kinase BB, glutamine synthase, and ubiquitin carboxy-terminal hydrolase L-1. Free radical biology & medicine. 2002;33:562–571. doi: 10.1016/s0891-5849(02)00914-0. [DOI] [PubMed] [Google Scholar]
  • [52].Castegna A, Aksenov M, Thongboonkerd V, Klein JB, Pierce WM, Booze R, Markesbery WR, Butterfield DA. Proteomic identification of oxidatively modified proteins in Alzheimer’s disease brain. Part II: dihydropyrimidinase-related protein 2, alpha-enolase and heat shock cognate 71. J Neurochem. 2002;82:1524–1532. doi: 10.1046/j.1471-4159.2002.01103.x. [DOI] [PubMed] [Google Scholar]
  • [53].Castegna A, Thongboonkerd V, Klein JB, Lynn B, Markesbery WR, Butterfield DA. Proteomic identification of nitrated proteins in Alzheimer’s disease brain. J Neurochem. 2003;85:1394–1401. doi: 10.1046/j.1471-4159.2003.01786.x. [DOI] [PubMed] [Google Scholar]
  • [54].Sultana R, Boyd-Kimball D, Poon HF, Cai J, Pierce WM, Klein JB, Merchant M, Markesbery WR, Butterfield DA. Redox proteomics identification of oxidized proteins in Alzheimer’s disease hippocampus and cerebellum: an approach to understand pathological and biochemical alterations in AD. Neurobiol Aging. 2006;27:1564–1576. doi: 10.1016/j.neurobiolaging.2005.09.021. [DOI] [PubMed] [Google Scholar]
  • [55].Sultana R, Poon HF, Cai J, Pierce WM, Merchant M, Klein JB, Markesbery WR, Butterfield DA. Identification of nitrated proteins in Alzheimer’s disease brain using a redox proteomics approach. Neurobiol Dis. 2006;22:76–87. doi: 10.1016/j.nbd.2005.10.004. [DOI] [PubMed] [Google Scholar]
  • [56].Reed T, Perluigi M, Sultana R, Pierce WM, Klein JB, Turner DM, Coccia R, Markesbery WR, Butterfield DA. Redox proteomic identification of 4-hydroxy-2-nonenal-modified brain proteins in amnestic mild cognitive impairment: insight into the role of lipid peroxidation in the progression and pathogenesis of Alzheimer’s disease. Neurobiol Dis. 2008;30:107–120. doi: 10.1016/j.nbd.2007.12.007. [DOI] [PubMed] [Google Scholar]
  • [57].Cenini G, Sultana R, Memo M, Butterfield DA. Elevated levels of pro-apoptotic p53 and its oxidative modification by the lipid peroxidation product, HNE, in brain from subjects with amnestic mild cognitive impairment and Alzheimer’s disease. J Cell Mol Med. 2008;12:987–994. doi: 10.1111/j.1582-4934.2008.00163.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [58].Prasad MR, Lovell MA, Yatin M, Dhillon H, Markesbery WR. Regional membrane phospholipid alterations in Alzheimer’s disease. Neurochem. Res. 1998;23:81–88. doi: 10.1023/a:1022457605436. [DOI] [PubMed] [Google Scholar]
  • [59].Castegna A, Lauderback CM, Mohmmad-Abdul H, Butterfield DA. Modulation of phospholipid asymmetry in synaptosomal membranes by the lipid peroxidation products, 4-hydroxynonenal and acrolein: implications for Alzheimer’s disease. Brain Res. 2004;1004:193–197. doi: 10.1016/j.brainres.2004.01.036. [DOI] [PubMed] [Google Scholar]
  • [60].Fadok VA, Voelker DR, Campbell PA, Cohen JJ, Bratton DL, Henson PM. Exposure of phosphatidylserine on the surface of apoptotic lymphocytes triggers specific recognition and removal by macrophages. J. Immunol. 1992;148:2207–2216. [PubMed] [Google Scholar]
  • [61].Kagan VE, Borisenko GG, Serinkan BF, Tyurina YY, Tyurin VA, Jiang J, Liu SX, Shvedova AA, Fabisiak JP, Uthaisang W, Fadeel B. Appetizing rancidity of apoptotic cells for macrophages: oxidation, externalization, and recognition of phosphatidylserine. Am. J. Physiol. Lung Cell Mol. Physiol. 2003;285:L1–17. doi: 10.1152/ajplung.00365.2002. [DOI] [PubMed] [Google Scholar]
  • [62].Tyurina YY, Serinkan FB, Tyurin VA, Kini V, Yalowich JC, Schroit AJ, Fadeel B, Kagan VE. Lipid antioxidant, etoposide, inhibits phosphatidylserine externalization and macrophage clearance of apoptotic cells by preventing phosphatidylserine oxidation. J. Biol. Chem. 2004;279:6056–6064. doi: 10.1074/jbc.M309929200. [DOI] [PubMed] [Google Scholar]
  • [63].Tyurina YY, Tyurin VA, Zhao Q, Djukic M, Quinn PJ, Pitt BR, Kagan VE. Oxidation of phosphatidylserine: a mechanism for plasma membrane phospholipid scrambling during apoptosis? Biochem. Biophys. Res. Commun. 2004;324:1059–1064. doi: 10.1016/j.bbrc.2004.09.102. [DOI] [PubMed] [Google Scholar]
  • [64].Herrmann A, Devaux PF. Alteration of the aminophospholipid translocase activity during in vivo and artificial aging of human erythrocytes. Biochim. Biophys. Acta. 1990;1027:41–46. doi: 10.1016/0005-2736(90)90045-p. [DOI] [PubMed] [Google Scholar]
  • [65].Kagan VE, Fabisiak JP, Shvedova AA, Tyurina YY, Tyurin VA, Schor NF, Kawai K. Oxidative signaling pathway for externalization of plasma membrane phosphatidylserine during apoptosis. FEBS Lett. 2000;477:1–7. doi: 10.1016/s0014-5793(00)01707-5. [DOI] [PubMed] [Google Scholar]
  • [66].Mohmmad Abdul H, Butterfield DA. Protection against amyloid beta-peptide (1-42)-induced loss of phospholipid asymmetry in synaptosomal membranes by tricyclodecan-9-xanthogenate (D609) and ferulic acid ethyl ester: implications for Alzheimer’s disease. Biochimica et biophysica acta. 2005;1741:140–148. doi: 10.1016/j.bbadis.2004.12.002. [DOI] [PubMed] [Google Scholar]
  • [67].Daleke DL. Regulation of transbilayer plasma membrane phospholipid asymmetry. J. Lipid Res. 2003;44:233–242. doi: 10.1194/jlr.R200019-JLR200. [DOI] [PubMed] [Google Scholar]
  • [68].Fadok VA, de Cathelineau A, Daleke DL, Henson PM, Bratton DL. Loss of phospholipid asymmetry and surface exposure of phosphatidylserine is required for phagocytosis of apoptotic cells by macrophages and fibroblasts. J. Biol. Chem. 2001;276:1071–1077. doi: 10.1074/jbc.M003649200. [DOI] [PubMed] [Google Scholar]
  • [69].Balasubramanian K, Schroit AJ. Aminophospholipid asymmetry: A matter of life and death. Annu. Rev. Physiol. 2003;65:701–734. doi: 10.1146/annurev.physiol.65.092101.142459. [DOI] [PubMed] [Google Scholar]
  • [70].Paulusma CC, Oude Elferink RP. The type 4 subfamily of P-type ATPases, putative aminophospholipid translocases with a role in human disease. Biochim. Biophys. Acta. 2005;1741:11–24. doi: 10.1016/j.bbadis.2005.04.006. [DOI] [PubMed] [Google Scholar]
  • [71].Verkleij AJ, Post JA. Membrane phospholipid asymmetry and signal transduction. J. Membr. Biol. 2000;178:1–10. doi: 10.1007/s002320010009. [DOI] [PubMed] [Google Scholar]
  • [72].Mandal D, Mazumder A, Das P, Kundu M, Basu J. Fas-, caspase 8-, and caspase 3-dependent signaling regulates the activity of the aminophospholipid translocase and phosphatidylserine externalization in human erythrocytes. J. Biol. Chem. 2005;280:39460–39467. doi: 10.1074/jbc.M506928200. [DOI] [PubMed] [Google Scholar]
  • [73].Mandal D, Moitra PK, Saha S, Basu J. Caspase 3 regulates phosphatidylserine externalization and phagocytosis of oxidatively stressed erythrocytes. FEBS Lett. 2002;513:184–188. doi: 10.1016/s0014-5793(02)02294-9. [DOI] [PubMed] [Google Scholar]
  • [74].Martin SJ, Finucane DM, Amarante-Mendes GP, O’Brien GA, Green DR. Phosphatidylserine externalization during CD95-induced apoptosis of cells and cytoplasts requires ICE/CED-3 protease activity. J. Biol. Chem. 1996;271:28753–28756. doi: 10.1074/jbc.271.46.28753. [DOI] [PubMed] [Google Scholar]
  • [75].Nicholson DW. Caspase structure, proteolytic substrates, and function during apoptotic cell death. Cell Death Differ. 1999;6:1028–1042. doi: 10.1038/sj.cdd.4400598. [DOI] [PubMed] [Google Scholar]
  • [76].Vanags DM, Porn-Ares MI, Coppola S, Burgess DH, Orrenius S. Protease involvement in fodrin cleavage and phosphatidylserine exposure in apoptosis. J. Biol. Chem. 1996;271:31075–31085. doi: 10.1074/jbc.271.49.31075. [DOI] [PubMed] [Google Scholar]
  • [77].Bader Lange ML, Cenini G, Piroddi M, Abdul HM, Sultana R, Galli F, Memo M, Butterfield DA. Loss of phospholipid asymmetry and elevated brain apoptotic protein levels in subjects with amnestic mild cognitive impairment and Alzheimer disease. Neurobiol Dis. 2008;29:456–464. doi: 10.1016/j.nbd.2007.11.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [78].Su JH, Zhao M, Anderson AJ, Srinivasan A, Cotman CW. Activated caspase-3 expression in Alzheimer’s and aged control brain: correlation with Alzheimer pathology. Brain Res. 2001;898:350–357. doi: 10.1016/s0006-8993(01)02018-2. [DOI] [PubMed] [Google Scholar]
  • [79].Battino M, Bompadre S, Leone L, Devecchi E, Degiuli A, D’Agostino F, Cambie G, D’Agostino M, Faggi L, Colturani G, Gorini A, Villa RF, Coenzyme Q. Vitamin E and Apo-E alleles in Alzheimer Disease. Biofactors. 2003;18:277–281. doi: 10.1002/biof.5520180231. [DOI] [PubMed] [Google Scholar]
  • [80].Jack CR, Jr., Petersen RC, Grundman M, Jin S, Gamst A, Ward CP, Sencakova D, Doody RS, Thal LJ. Longitudinal MRI findings from the vitamin E and donepezil treatment study for MCI. Neurobiol Aging. 2008;29:1285–1295. doi: 10.1016/j.neurobiolaging.2007.03.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [81].Sano M, Ernesto C, Thomas R, Klauber M, Schafer K, Grundman M, Woodbury P, Growdon J, Cotman C, Pfeiffer E, Schneider L, Thal L. A controlled trial of selegiline, alpha-tocopherol, or both as treatment for Alzheimer’s disease. The Alzheimer’s Disease Cooperative Study. The New England journal of medicine. 1997;336:1216–1222. doi: 10.1056/NEJM199704243361704. [DOI] [PubMed] [Google Scholar]

RESOURCES