U.S. flag

An official website of the United States government

NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.

Alzate O, editor. Neuroproteomics. Boca Raton (FL): CRC Press/Taylor & Francis; 2010.

Cover of Neuroproteomics

Neuroproteomics.

Show details

Chapter 10Redox Proteomics of Oxidatively Modified Brain Proteins in Mild Cognitive Impairment

, , and .

10.1. MILD COGNITIVE IMPAIRMENT

Mild cognitive impairment (MCI) is generally referred to as the transitional zone between normal cognitive aging and early dementia or clinically probable Alzheimer’s disease (AD) (1), although not all AD patients pass through an MCI stage. The term was first coined by Petersen (2). Most individuals with MCI eventually develop AD, which suggests MCI may be the earliest phase of the AD (36). MCI can be divided into two broad subtypes: amnestic (memory-affecting) MCI or non-amnestic MCI (2,7). Other functions, such as language, attention, and visuospatial skills, may be impaired in either type. Amnestic mild MCI patients characteristically have subtle but measurable memory disorder not associated with dementia. Individuals with MCI are at an increased risk of developing AD, or another form of dementia with a rate of progression between 10% and 15% per year (8,9), although there have been cases where patients have reverted to normal (1,10,11).

Criteria for MCI include (a) a memory complaint corroborated by an informant; (b) objective memory test impairment (age and education adjusted); (c) general normal global intellectual function; (d) activities of daily living not disturbed; (e) clinical dementia rating (CDR) score of 0.0 to 0.5; (f) no dementia; and (g) a clinical evaluation that revealed no other cause for memory decline (12). Moreover, neuroim-aging studies by magnetic resonance imaging (MRI) demonstrate the atrophy of the hippocampus or entorhinal cortex in MCI patients, indicating the relationships with transition of normal aging to MCI, then later to clinical AD (13). Pathologically, MCI brain shows mild degradation of the hippocampus, entorhinal cortex, sulci, and gyri using MRI (14,15). These aforementioned areas undergo considerable degradation in AD (1620). Since the hippocampus is the region of the brain primarily responsible for processing memory, it is clearly understandable why those persons with AD and MCI have memory loss.

10.2. OXIDATIVE STRESS

Under normal physiological conditions, there is equilibrium between the level of antioxidants and pro-oxidants in a cell. However, when environmental factors, stressors, or disease occur, this homeostasis can become imbalanced in favor of pro-oxidants, resulting in a phenomenon known as oxidative stress (21). Oxidative stress can also transpire through an antioxidant deficiency (22) or excess reactive oxygen species/ reactive nitrogen species (ROS/RNS) production. Moreover, oxidatively damaged proteins are often removed by the 20S proteosome. Defects in the proteosome system would lead to elevated levels of oxidatively modified proteins and neurotoxicity (21). Oxidative stress plays a significant role in neurodegenerative disease (2327). If oxidative stress is involved in the progression of AD and is not a consequence of this disease, there should be evidence of elevated oxidative stress in the beginning stages of the disease (28). Manifested by elevated levels of nucleic acid oxidation, protein oxidation, and lipid peroxidation, oxidative damage is most severe in the hippocampus, a brain region that is responsible for memory processing and cognitive function (23,24). Moreover, oxidative stress-mediating entities per se induce neuronal death in vitro, and protein oxidation and lipid peroxidation in the superior and middle temporal gyri (SMTG) of MCI patients are increased (29). All of these studies strongly suggest that oxidative stress is a primary event in the development of AD.

10.3. POST-TRANSLATIONAL MODIFICATIONS

Post-translational modifications (PMTs), such as protein oxidation, are seen in a plethora of neurodegenerative diseases (23,30). A few examples of PMTs are phosphorylation, nitration, carbonylation, 4-hydroxy-2-nonenal (HNE) modification, gly-cosylation, and S-glutathionylation. Proteins are sensitive to oxidation by ROS and RNS, and oxidative damage is a result from such interaction. Protein oxidation leads to loss of protein function and often cell death via necrotic or apoptotic processes (30). Protein oxidation is increased in MCI (31,32) and will be discussed thoroughly in the next section. A complete list of oxidatively modified proteins in MCI brain determined by redox proteomics can be seen in Table 10.1.

TABLE 10.1. Proteins Oxidatively Modified in MCI.

TABLE 10.1

Proteins Oxidatively Modified in MCI.

10.4. PROTEIN CARBONYLATION

Protein carbonyl levels have been recognized as the most widely used indicator of oxidative stress (3335). Protein carbonyl levels increase in aging and age-associated neurodegenerative diseases (36,37). Proteins can be oxidized directly by reactive oxygen species to generate protein carbonyls. This happens by several different mechanisms including amino acid side chain (Lys, Thr, Pro, and Arg) and metal-assisted oxidation (38) (Figure 10.1). Beta scission from the peptide backbone is the second method by which protein carbonyls are created (Figure 10.2). Third, the adduction of carbonyl-containing reactive aldehydes (i.e., acrolein, 4-hydroxynonenal) to amino acids leads to carbonyls (Figure 10.3). Advanced glycation end products resulting from Amadori chemistry can also lead to protein oxidation (39,40). Indexing protein carbonyl content is a fundamental key to understanding protein oxidation. Protein carbonyl levels most often are analyzed by immunochemical detection of the hydrazone formed by reaction of 2,4-dinitrophenylhydrazine with protein carbonyls (30,38).

FIGURE 10.1. Beta scission from peptide backbone.

FIGURE 10.1

Beta scission from peptide backbone.

FIGURE 10.2. Amino acid side chain oxidation.

FIGURE 10.2

Amino acid side chain oxidation.

FIGURE 10.3. Covalent modification of amino acids by Michael addition.

FIGURE 10.3

Covalent modification of amino acids by Michael addition.

10.5. HNE MODIFICATION

Lipid peroxidation is a complex process involving the interaction of oxygen-derived free radicals with polyunsaturated fatty acids, resulting in a variety of highly reactive electrophilic aldehydes that are capable of easily attaching covalently to proteins by forming adducts with cysteine, lysine, or histidine residues (41) through Michael addition (30). Among the aldehydes formed, malondialdehyde (MDA) and HNE represent the major products of lipid peroxidation (41). The brain is particularly vulnerable to lipid peroxidation (42) due to its richness in polyunsaturated fatty acids, high oxygen consumption, and abundant quantities of redox transition metals (43,44).

Lipid peroxidation is highly evident in neurodegenerative disease (24,45,46). Lipid peroxidation occurs through ongoing free radical chain reactions until termination occurs (Figure 10.4). Free radicals attack an allylic hydrogen atom to form a carbon centered radical (step 1). This radical reacts with O2 to produce peroxyl radicals (step 2). These peroxyl radicals can react with adjacent lipids forming a lipid hydroperoxide repeating the cycle (step 3). The lipid hydroperoxide can decompose to produce multiple reactive products such as acrolein, malondialdehyde, and HNE. In MCI, acrolein and HNE have been found to be significantly elevated (31,47,48). Lipid peroxidation can be terminated by two radicals reacting forming a nonradical and oxygen (step 4). HNE is a major product of lipid peroxidation and causes cell toxicity. Lipids are particularly vulnerable to oxidation due to the fact that polyunsaturated fatty acids are abundant in brain and the presence of oxygen in the lipid bilayer is at millimolar levels. Glutathione has been shown to detoxify HNE in cells (49). Increased levels of HNE cause disruption of Ca2+ homeostasis, membrane damage, and cell death (41). Vitamin E (alpha-tocopherol) is a “chain breaking” antioxidant and can terminate propagation steps of lipid peroxidation. When the hydrogen is abstracted in step 1, an alpha-tocopherol radical forms that can be reverted back to vitamin E by ascorbic acid (vitamin C) or glutathione (GSH), both potent antioxidants.

FIGURE 10.4. Lipid peroxidation reaction summary.

FIGURE 10.4

Lipid peroxidation reaction summary.

HNE is an alpha, beta-unsaturated alkenal product of omega-6 polyunsaturated fatty acids and is a major cytotoxic end product of lipid peroxidation that mediates oxidative stress-induced death in many cell types (41,50,51). HNE accumulates in membranes at concentrations of 10 μM to 5 mM in response to oxidative insults (41) and invokes a wide range of biological activities, including inhibition of protein and DNA synthesis (5256), disruption of Ca2+ homeostasis, membrane damage, cell death (41), and activation of stress signaling pathways (50,57).

Several publications report that the brains of MCI patients are under oxidative stress. Increased levels of thiobarbituric acid reactive substance (TBARS), malon-dialdehyde (MDA), F2 isoprostanes and F4 neuroprostanes, as well as soluble and protein-bound HNE, specific markers of in vivo lipid peroxidation (30), were significantly elevated in cerebrospinal fluid (CSF), plasma, urine, and brains of MCI patients compared with controls (29,31,46,58), suggesting that lipid peroxidation may be an early event in the pathogenesis of the disease.

10.6. PROTEIN NITRATION

In conjunction with the enzyme nitric oxide synthase, arginine is converted into nitric oxide and L-citrulline (Figure 10.5). Nitric oxide can react with superoxide radical anion (O2-) forming the strong oxidant peroxynitrite (Figure 10.6). Peroxynitrite can oxidize tyrosine, methionine, tryptophan, and cysteine residues to promote protein nitration (59,60). In the presence of CO2, peroxynitrite can exist as an anion (ONOO) or other reactive intermediates. A nitrosoperoxyl intermediate is formed from the combination of peroxynitrite and carbon dioxide, which rearranges to form nitrocarbonate. This species can be cleaved to form carbonate and NO2 radicals (Figure 10.6); the latter reacts with a tyrosyl free radical in the 3-position to form 3-nitrotyrosine (Figure 10.7).

FIGURE 10.5. Production of nitric oxide from L-citrulline.

FIGURE 10.5

Production of nitric oxide from L-citrulline.

FIGURE 10.6. Combination reaction of peroxynitrate and carbon dioxide.

FIGURE 10.6

Combination reaction of peroxynitrate and carbon dioxide.

FIGURE 10.7. Formation of 3-nitrotyrosine.

FIGURE 10.7

Formation of 3-nitrotyrosine.

Previous research has shown that peroxynitrite can interact with proteins (61,62), lipids (63), DNA (64,65), and RNA (66) to promote damage in these biological molecules. Tyrosine nitration is associated with Alzheimer’s disease (AD) (67) as well as Parkinson’s disease (68). Gamma-glutamylcysteinylethyl ester (GCEE), a derivative of gamma-glutamylcysteine, can cross the blood-brain barrier (BBB) and has been proven to prevent peroxynitrite-induced damage by upregulating glutathione production (69). Lipoic acid has also been used to reduce protein nitration (70) as has gamma-tocopherol (71). Increased protein nitration also can lead to an elevated release of RNS. Nitration of proteins results in the inactivation of several important mammalian proteins such as MnSOD (7275), glyceraldehyde 3-phosphate dehydrogenase (76,77), actin (78,79), synaptic proteins (80), and tyrosine hydroxylase (81,82), among others.

10.7. OXIDATIVE DAMAGE IN MCI BRAIN

Protein oxidation has been observed in brains of patients with MCI (32). Increased protein-bound HNE immunoreactivity is observed in the MCI hippocampus and MCI inferior parietal lobule (IPL) (31) as well as elevated levels of protein carbonyls (83) and 3-nitrotyrosine (32). Since MCI is arguably the earliest form of AD and there is supporting evidence of oxidative damage in the AD brain, these data suggest that elevated levels of oxidative stress in MCI can contribute to the progression of this neurodegenerative disease.

10.8. REDOX PROTEOMICS

The proteome is the complete set of proteins expressed in a cell or organism. This term was first coined by Wilkins et al. in 1996 (84). Proteomics is the analysis of the proteome. Proteomics describes a protein in terms of its amino acid sequence, mass, isoelectric point (pI), and structure. This technique is helpful in determining protein identification, modification, and function. Since the advent of proteomics, this methodology is one of the fastest growing areas of biomedical science. Proteomics is widely used to determine potential biomarkers in neurodegenerative diseases, specifically AD (8587). Technology in the field of proteomics is rapidly expanding and evolving. Proteomics is now being developed for clinical diagnosis (88,89) and in establishing biomarkers for disease including MCI (87,90,91) and AD (86,92).

10.8.1. 2D Gel Electrophoresis Approaches for Redox Proteomics

Proteomics traditionally has employed a two-dimensional protein separation (see Chapter 3 for a discussion of protein separation techniques, and Chapter 4 for a discussion of 2D electrophoresis approaches). There are three basic steps to any proteomics experiment: separation, quantification, and identification. In 2D electrophoresis, proteins are separated on the basis of their isoelectric point (pI) in the first dimension. Protein samples are placed on immobilized pH gradient (IPG) strips composed of a Polyacrylamide matrix. Isoelectric focusing (IEF) takes advantage of this principle by applying a potential to the proteins embedded on the IPG strip. The proteins migrate from the cathode to the anode, and at the pH where the protein reaches its specific pI they cease to move.

In the second dimension, proteins are separated according to their migration rate, which is a function of protein mass and protein shape. IPG protein strips are equilibrated in a dithiothreitol (DTT)-containing buffer. DTT reduces the disulfide bonds to thiol groups. After reduction and removal of DTT solution, an iodoacetamide (IA) solution is added to the IPG strip to alkylate the thiol groups and prevent them from recombining. This reaction is performed in the dark as IA is photosensitive. The IPG strip is loaded onto the gel and voltage is applied. Low molecular weight proteins travel faster through the Polyacrylamide gel and are located closer to the bottom of the gel. High molecular weight proteins travel slower through the gel and are found toward the upper portion of the gel.

After electrophoresis, 2D gels are fixed and then stained with an appropriate protein stain and subsequently prepared for visualization. SYPRO Ruby is a common fluorescent stain used to detect proteins on a gel. Photons must be excited (λex = 470 nm) so they can emit (λem = 618 nm) and be detected. SYPRO Ruby is a sensitive stain as it can detect 1 ng of protein/spot. Other methods of staining for protein visualization include staining with Coomassie blue and silver staining. Although there are various other staining methods, SYPRO Ruby staining is a mass spectrometry-compatible, quick method with shorter preparation time compared to others. After protein visualization, all gels are scanned and images are imported into a software analysis package for comparison studies.

An additional, identical gel is run, and left unstained. This unstained gel is soaked in a transfer buffer, and then sandwiched between filter paper, a nitrocellulose membrane, and filter paper in order for the proteins to transfer onto the nitrocellulose membrane. These components are properly arranged and a thin glass rod is rolled over the top filter paper to remove any possible bubbles that may have formed between the layers. A weighted plate is added to the top of the filter paper and current is applied to allow proper protein transfer.

Typically, Western blots are blocked with BSA to prevent any nonspecific binding. A primary antibody that is specific for the proteins of interest is added to the blot. The antibody will bind to the protein of interest (antigen); the membrane is washed to remove any unbound molecules. A second antibody (attached to an enzyme) is added that will bind to the primary antibody-antigen complex. The protein of interest will be identified through a chemiluminescent or chromophoric reaction, depending on the nature of the secondary antibody. The blots are then dried, scanned, saved as images, and exported into the software analysis program. The Western blots in a particular experiment are compared to each other through computer algorithms based on immunoreactivity levels.

For protein carbonylated proteins, post derivatization after immunoprecipitation can be used as well. In this method, the nitrocellulose membranes are equilibrated in 20% methanol for five minutes. Membranes are then incubated in 2N HCl for five minutes and then in 0.5 mM DNPH solution for exactly five minutes. The membranes are washed three times in 2N HCl and five times in 50% methanol (five minutes each wash), dried, scanned, and saved as images.

10.8.2. Immunochemical Detection

To confirm the correct identification of the proteins detected by mass spectrometry, a representative protein identified by redox proteomics is immunoprecipitated. Brain samples are first pre-cleared using protein A/G-agarose beads for one hour at 4°C and then incubated overnight with an appropriate antibody for the protein of interest. The protein A/G mixture is ideal because protein A-conjugated agarose beads respond to rabbit-raised antibodies, while protein G-conjugated agarose beads are used if the antibodies were raised in goats or mice. The following day, the samples are incubated for one hour with protein A/G-agarose then washed three times with RIPA buffer (93). Proteins are then solubilized in IEF rehydration buffer followed by 2D electrophoresis and 2D Western blot.

10.8.3. Image Analysis

The gels and nitrocellulose blots are scanned and saved as images using a transiluminator and scanner, respectively. Several commercial software packages including PDQuest, Phoretix, and Progenesis can be used for matching and analysis of visualized protein spots among different gels and blots. The principles of measuring intensity values by 2D analysis software are similar to those of densitometric measurement. Although for 2D gel comparison, two-dimensional difference gel electrophoresis (2D-DIGE) could be used since gel-to-gel matching sometimes is not reproducible using other approaches. Spot matching is performed to ascertain the average normalized intensity of gels in an experimental sample to that of spots in a control sample. Only those spots that are considered statistically significant by Student’s t-test (p < 0.05) are selected for identification. Sophisticated statistical analyses for microarray data are not applicable for proteomics studies (94,95).

10.8.4. Trypsin Digestion

The selected spots are excised from the 2D gel, transferred into microcentrifuge tubes, and washed with ammonium bicarbonate (NH4HCO3) to swell the Polyacrylamide gel and push the protein into the gel matrix (for a detailed discussion of protein identification by mass spectrometry, refer to Chapter 5). Acetonitrile is then added to shift equilibrium away from the protein and shrinks the gel matrix. The solvent is removed, and the gel pieces are dried in a flow hood. The protein’s disulfide bonds are reduced with DTT at 56°C then alkylated with IA. Acetonitrile and ammonium bicarbonate steps are repeated and protein spots are allowed to dry. The gel pieces are rehydrated with 20 ng/μL modified trypsin in NH4HCO3 with the minimal volume to cover the gel pieces. The gel pieces then are chopped into smaller pieces and incubated with shaking overnight at 37°C with trypsin. This technique is termed “in-gel” trypsin digestion and has several advantages including higher recovery because the protein is cleaved into many smaller peptides that result from sequence-specific proteolysis, which is an important means of identifying the protein of interest (96).

10.8.5. Mass Spectrometry

These tryptic digests (smaller peptides) constitute mass fingerprints, which are unique to each protein, and the molecular weight of each peptide is determined by the use of mass spectrometry (MS; see Chapters 5 and 7). The two ion sources are matrix-assisted laser desorption ionization (MALDI) and electrospray ionization (ESI). In MALDI the analyte is mixed with a matrix, and this mixture is allowed to dry. A pulsed laser hits the target containing this analyte/matrix mixture and disperses the analyte. The matrix absorbs the energy from the laser pulse, and the matrix molecules containing the analyte are transferred into the gas phase. In the gas phase, H+ transfer from the acidic matrix to the peptides occurs, thereby putting charges on the peptides in the gas phase. A time of flight (TOF) mass analyzer measures how long it takes for the ions to reach the detector. Electrospray ionization is a technique that provides the transport of ions from solution to the gas phase. In ESI, the peptide in solution flows through a narrow capillary tube into a vacuum and the MS at atmospheric pressure and 4000 V to create ions. The charges on the droplets extend the solution to form a (Taylor) cone, causing the solution to disperse as a mist of fine droplets. As the solvent evaporates, the droplet size decreases the total charges of the proteins in the droplet remain the same but the surface area of each droplet decreases to one ion per droplet. Individual ions then flow into the mass analyzer. ESI can be coupled to other separation techniques (i.e., high performance liquid chromatography [HPLC]) to make it more versatile (as discussed in Chapter 5).

10.8.6. Database Searching

Better known as peptide mass fingerprinting (PMF), this process requires the use of MS analysis to determine the experimental masses. Experimental masses are used by on-line protein databases to compare and match protein-specific mass fingerprints generated by the in silico digestion of proteins. This comparison is used to identify the protein of interest based on the quality of peptide matches (9799). PMF is used to identify proteins from tryptic peptide fragments by utilizing an appropriate search engine (Table 10.2, and Chapter 5). Database searches are conducted allowing for up to one missed trypsin cleavage and using the assumption that the peptides are monoisotopic, oxidized at methionione residues, and carbamiodomethylated at cysteine residues. Mass tolerance of 150 ppm/g is the window of error allowed for matching the peptide mass values. Probability-based MOWSE scores are estimated by comparison of search results against estimated random match population and are reported as -10 * Log10 (p), where p is the probability that the identification of the protein is incorrect. As a correlate, the higher the MOWSE score, the lower the p value, yielding a higher probability of accurate protein identification. All protein identifications must be reviewed (and in most cases validated; for this see Chapters 11 and 14) to ensure that they are in the expected size and pI range based on position in the gel.

TABLE 10.2. Listing of Online Search Engines.

TABLE 10.2

Listing of Online Search Engines.

10.9. ENZYME ASSAYS

Enzymatic activity is generally lower in oxidatively modified enzymes compared to control (31,100,101). Oxidative stress leads to oxidative modification of key proteins and enzymes (102). Oxidative modification of amino acids, particularly at the active site, can cause loss of enzymatic activity and subsequent protein dysfunction (100,103105). Therefore, the determination of enzyme activity is frequently used to help validate the identities of the proteomics-identified proteins.

10.10. ENERGY-RELATED PROTEINS

Glycolysis is a metabolic pathway that converts glucose to pyruvate and generates two ATP molecules in the process. ATP, the energy source of the cell, is extremely important at nerve terminals for normal neural communication. Decreased levels of cellular ATP at nerve terminals may lead to loss of synapses and synaptic function, and may ultimately contribute to memory loss in amnestic MCI patients. Synapse loss is an early event in AD, since it is observed in MCI and early AD (106). Energy metabolism alteration is a current major hypothesis of AD and has been reported in brain in an advanced stage of MCI (107110). Several proteins identified by proteomics as energy-related proteins in MCI brain support this notion. There is an assortment of energy-related proteins that have been identified as being oxidatively modified in MCI hippocampus and IPL in addition to previous proteomics studies of AD and cell culture models of AD, identifying an increased oxidation of enolase (111116). Alpha-enolase, a critical glycolytic protein, interconverts 2-phosphoglycerate to phosphoenolpyruvate (Figure 10.8). The enzymatic activity of alpha-enolase is reduced significantly compared to control samples (83). Reduction in enzymatic activity can lead to protein dysfunction and in this case, lowered ATP production.

FIGURE 10.8. Interconversion of 2-phosphoglycerate to phosphoenolpyruvate via enolase.

FIGURE 10.8

Interconversion of 2-phosphoglycerate to phosphoenolpyruvate via enolase.

Pyruvate kinase is also a glycolytic enzyme that catalyzes the final step in glycolysis, the conversion of phosphoenolpyruvate to pyruvate with the concomitant transfer of the high-energy phosphate group from phosphoenolpyruvate to ADP, thereby generating ATP (Figure 10.9). Under aerobic conditions, pyruvate can be transported to the mitochondria, where it enters the TCA cycle and is further broken down to produce considerably more ATP through oxidative phosphorylation. Therefore, the oxidative inactivation of this important enzyme in the hippocampi of MCI subjects could conceivably result in the reduced ATP production and alter ATP-dependent processes, such as signal transduction and cell potential maintenance, thereby leading to altered Ca2+ homeostasis and neuronal dysfunction.

FIGURE 10.9. Pyruvate kinase enzymatic reaction.

FIGURE 10.9

Pyruvate kinase enzymatic reaction.

Lactate dehydrogenase B (LDH) reduces pyruvate to lactate by NADH (Figure 10.10). Lactate is a substrate for gluconeogenesis and since glucose is the major supplier of energy to the brain, proper lactate production is crucial (117). LDH dysfunction and subsequent reduced glucose metabolism are commonly observed in the positron emission tomography (PET) scans of MCI and AD brain (118,119). Enzyme activity of lactate dehydrogenase is significantly reduced in MCI hippocampus, which further correlates protein dysfunction and enzyme activity impairment. Likewise, specific enzyme activity was also lower in MCI hippocampus. Impairment of this enzyme could initiate reduction of glucose production and creation of excess pyruvate.

FIGURE 10.10. Conversion of pyruvate to L-lactate via lactate dehydrogenase.

FIGURE 10.10

Conversion of pyruvate to L-lactate via lactate dehydrogenase.

Phosphoglycerate kinase catalyzes the reaction to convert 1,3-bisphosphoglycerate to 3-phosphoglycerate. This reaction undergoes substrate phosphorylation by phos-phoryl transfer from 1,3-bisphosphoglycerate to ADP to produce ATP (Figure 10.11). Additionally, enzyme activity is reduced, thus suggesting that oxidative modification leads to impairment of protein function. Impairment of phosphoglycerate kinase results in decreased energy production and irreversible downstream effects, such as multidrug resistance (120).

FIGURE 10.11. Phosphoglycerate kinase enzymatic reaction.

FIGURE 10.11

Phosphoglycerate kinase enzymatic reaction.

Aldolase is an essential protein that cleaves fructose 1,6-bisphosphate to dihydroxy-acetonephosphate (DHAP) and glyceraldehyde-3-phosphate (G3P) (Figure 10.12). The production of G3P is essential in continuing glycolysis in order to produce adequate amounts of ATP to maintain ion-motive ATPases, signal transduction, glu-tamate, and glucose transporters (121,122). Oxidative modification of aldolase can compromise membrane symmetry and render neurons vulnerable to excitotoxicity and apoptosis.

FIGURE 10.12. Aldolase enzymatic reaction.

FIGURE 10.12

Aldolase enzymatic reaction.

The impairment of these energy-related proteins would lead to decreased ATP production, with consequent dysfunction in electrochemical gradients, ion pumps, and voltage-gated ion channels, glucose and glutamate transporters, loss of membrane asymmetry, as well as lower efficiency on such processes as Ca2+ homeostasis, cell potential, and signal transduction. All these glycolytic protein oxidative modifications observed in the earliest stage of AD support the hypothesis of energy metabolism alteration in Alzheimer’s disease and are consistent with the known PET alterations in MCI and AD.

10.11. NEUROPLASTICITY

It is well documented that glutamine synthetase (GLUL) activity declines in AD (103,123125). GLUL catalyzes the rapid amination of glutamate to form the non-neurotoxic amino acid glutamine. This reaction maintains the optimal level of glutamate and ammonia in neurons and modulates excitotoxicity. Together with the action of glutamate receptors and glutamate transporters, this process is important to maintaining neuroplasticity (126). A decline in neuroplasticity is suggested to correlate with the progression of AD from MCI (127). Therefore, oxidative inactivation of GLUL suggests the glutamate-glutamine cycles in hippocampi of MCI subjects are impaired, which may contribute to excitoxicity and impairment of neuroplasticity during the development of AD (128).

10.12. MITOCHONDRIAL DYSFUNCTION

Mitochondrial dysfunction is associated with AD; therefore, the proteomic identification of ATP synthase in MCI IPL and hippocampus contributes to evidence suggesting a role of mitochondrial dysfunction in the progression of AD. ATP synthase alpha-chain is a mitochondrial regulating subunit of complex V, which plays a key role in energy production. Alteration of complex V of the electron transport chain, ATP synthase, results in impaired ATP production in the mitochondria, the ATP “powerhouse.” ATP synthase goes through a sequence of coordinated conformational changes of its major subunits (alpha and beta) to produce ATP. ATP synthase, alpha-subunit, is the only HNE-modified protein observed by redox proteomics in both MCI hippocampus and IPL. At an early stage, ATP synthase is tightly associated with tau aggregated proteins in neurofibrillary tangles in AD (129), making ATP synthase a potential target for AD therapeutics. The oxidation of ATP synthase leads to the inactivation of this mitochondrial complex. Failure of ATP synthase could contribute to a decrease in the activity of the entire electron transport chain and impaired ATP production, resulting in possible electron leakage from their carrier molecules to generate ROS, suggesting an alternative rationalization for the well-documented existence of oxidative stress in AD and MCI (24,26,31,32,83,130,131). The specific activity of ATP synthase was reduced in MCI hippocampus and IPL compared to age-matched controls, providing insight into enzyme efficiency in MCI (28). Additionally, malate dehydrogenase is oxidatively modified in MCI; this enzyme is responsible for oxidation of L-malate to oxaloacetate in the citric acid cycle, another ATP-producing metabolic pathway (Figure 10.13).

FIGURE 10.13. Malate dehydrogenase enzymatic reaction.

FIGURE 10.13

Malate dehydrogenase enzymatic reaction.

As chaperones, heat shock proteins assist in establishing proper protein conformation in an effort to prevent protein aggregation by repairing misfolded proteins or guiding misfolding proteins to the proteosome for degradation (132). Numerous heat shock proteins (Hsp70, Hsc71, and Hsp90 and 60, respectively) have been found to be oxidatively modified in neurodegenerative diseases including MCI (133), AD (112), and HD (134). Oxidative modification of heat shock proteins may exacerbate protein misfolding and protein aggregation, leading to reduced effective proteosomal activity and eventual proteosomal overload and dysfunction, known to occur in AD (135).

10.13. ANTIOXIDANT DEFENSE

Peptidyl-prolyl cis/trans isomerase (Pin-1) is a regulatory protein that reversibly alters the conformation of proline residues in proteins from cis to trans (136). The WW domain of Pin-1 recognizes phosphorylated Ser-Pro and phosphorylated Thr-Pro motifs in proteins, and thereby binds to many cell cycle-regulating proteins, APP, and tau protein. Pin-1 is co-localized with phosphorylated tau and also shows an inverse relationship to the expression of tau in AD brains (137139). Moreover, Pin-1 is oxidatively inhibited in AD brain (140) and is able to restore the function of tau protein in AD (141). Therefore, the oxidative inactivation of Pin-1 in the hippocampi of MCI subjects could be one of the initial events that trigger tangle formation, oxidative damage, cell cycle alterations, and eventual neuronal death in AD brains. Pin-1 specifically recognizes pSer-Pro and pThr-Pro motifs. Pin-1 is excessively carbonylated in MCI (83,142) and AD hippocampus (104). Moreover, Pin-1 binds to APP and affects production of amyloid beta peptide. Pin-1 enzymatic activity is significantly decreased in MCI and AD brain, contributing to previous research showing loss of enzymatic activity in oxidatively modified proteins (103,143). This decrease in Pin-1 activity may be responsible for the increased accumulation of both phosphorylated tau protein in AD hippocampus, which is rich in neurofibrillary tangles (NFT) and senile plaques (SP), rich in amyloid beta.

Glutathione S-transferase mu and multidrug resistance protein 3 are nitrated in MCI IPL. These two proteins play an important role in regulating cellular processes by decreasing the levels of oxidants or by removing toxic compounds that are generated in the cell. Glutathione S-transferases (GSTs) catalyze the binding of HNE to glutathione, resulting in the formation of a GSH-HNE conjugate that is exported out from the cell by the multidrug resistant protein-1 (MRP1). This process thereby plays an important role in cellular protection against oxidative stress. In AD brain, GST protein levels and activity were reported to be decreased; in addition, GST was found to be oxidatively modified by HNE (144). GSTs have a high catalytic activity against HNE and are oxidatively modified and downregulated in AD brain (144). As a corollary, overexpression of GST can combat the effects of HNE toxicity in culture (145).

Peroxiredoxin 6 (Prx VI) is found to be nitrated in MCI hippocampus. Peroxiredoxins remove toxic hydrogen peroxide from the cell, resulting in reduced ROS production. Peroxiredoxins can reduce peroxynitrite at a high catalytic rate, which may modulate protein nitration and cell damage (146). Prx VI is the only peroxiredoxin that uses glutathione as an electron donor, while all other peroxiredoxins (PrxI-PrxV) use thioredoxin. In addition, peroxiredoxins play a role in cell differentiation and apoptosis. The decrease in the activity of this enzyme may also lead to decreased phospholipase A2 activity, one of the target proteins regulated by peptidyl prolyl cis/trans isomerase (Pin-1), a protein that has been reported to be downregulated and have decreased activity in MCI and AD brain (104,141,142).

The enzyme GST forms a complex with Prx VI in order to modulate both enzyme activities. These proteins work in coordination with one another either directly or indirectly, thereby protecting the cell from toxicants. These results provide insight into how the changes of these proteins may contribute to tau hyperphosphorylation and neurofibrillary tangle formation, in addition to development of oxidative stress. This result could conceivably be related to the identification of MRP1 as a protein with elevated HNE binding in AD (144). This protein was also mapped to undergo DNA damage by oxidative stress mechanisms (147).

Carbonyl reductase is a vital enzyme that can reduce carbonyl-containing compounds to their resultant alcohols, thereby reducing protein carbonyl levels. Subsequent malfunction or downregulation of this enzyme could cause an increase in protein carbonyls, which because of the polarity of the carbonyl moiety could expose ordinarily buried hydrophobic amino acids to the solvent (i.e., disrupt conformation). Carbonyl reductase has been shown to reduce the lipid peroxidation product, HNE (148). Carbonyl reductase expression is altered in Down’s syndrome and AD patients (149). The gene for carbonyl reductase is located in close proximity to the gene for Cu/Zn superoxide dismutase (SOD1) (150). Interestingly, the genes for SOD1, carbonyl reductase, and APP are located on chromosome 21, which is a trisomy in Down’s syndrome patients (151). The link between increased amyloid deposition and decreased carbonyl reductase enzymatic activity is unclear. The current research posits a possible intriguing relationship among amyloid beta, Down’s syndrome, and carbonyl reductase in neurodegeneration.

Oxidative stress is the imbalance of pro-oxidants and antioxidants, where equilibrium is favoring the side of pro-oxidants. This can also mean antioxidant defense is low due to protein modification. Increased protein expression of peroxiredoxins (I and II), Cu/ZnSOD, carbonyl reductase, and alcohol dehydrogenase, glutathione S-transferase (GST), and multidrug resistant protein-1 (MRP1) in AD brain has been reported (149,152155). Similarly, antioxidant defense is altered in MCI (156). Several antioxidant proteins undergo oxidative modification in MCI brain (133).

10.14. STRUCTURAL DYSFUNCTION

Dihydropyriminidase related protein 2 (DRP-2) and fascin 1 are identified as nitrated proteins in MCI hippocampus. DRP-2 is a member of the dihydropyrimidinase-related protein family that is involved in axonal outgrowth and pathfinding through transmission and modulation of extracellular signals (157,158). Additionally, DRP-2 interacts with and modulates collapsin, which aids in dendrite elongation, guidance, and growth cone collapse. DRP-2 has been reported to be associated with NFT, which may lead to decreased levels of cytosolic DRP-2 in AD (159). This, in turn, would eventually lead to abnormal neuritic and axonal growth, thus accelerating neuronal degeneration in AD (160), which is one of the characteristic hallmarks of AD pathology. Increased oxidation (41) and decreased expression of DRP-2 protein was observed in AD. In adult Down’s syndrome (DS) (159), fetal DS (161), schizophrenia, and affective disorders, DRP-2 has lower levels in brain. Since memory and learning are associated with synaptic remodeling, nitration and subsequent loss of function of this protein could conceivably be involved in the observed memory decline in MCI. Moreover, the decreased function of DRP-2 could be involved in the shortened dendritic length and synapse loss observed in AD (162). Maintaining neuronal communication is essential in learning and memory. Since memory is greatly depreciated in MCI (163,164), certain proteins involved in neuronal communication are oxidatively modified in MCI and AD (112,133,155).

Fascin 1 (FSCN1) is an actin-bundling, structural protein also known as p55 (165) and is involved in cell adhesion (166) and cell motility (167). It is a marker for dendritic functionality (168). Addition of p55 has been shown to protect cells from oxidative stress produced by an insult (169). The identification of this protein as nitrated in MCI brain is consistent with the notion that loss of function of this protein lessens protection against oxidative damage and could be an important event in the transition of MCI to AD. Fascin 1 has also been found to interact with protein kinase C alpha (PKCα), which regulates focal adhesions (170). Impairment of this protein can be related to faulty neurotransmission from the affected dendritic projections. In a beagle dog model of AD, in which beagle amyloid beta has the same sequence as human amyloid beta, fascin was a protein protected by a diet rich in antioxidants and in a program of environmental enrichment, i.e., making new synapses (171).

Actin is a principal protein playing a central role in maintaining cellular integrity, morphology, and the structure of the plasma membrane. Actin microfilaments play a role in the neuronal membrane cytoskeleton by maintaining the distribution of membrane proteins, and segregating axonal and dendritic proteins (172). In the CNS, actin is distributed widely in neurons, astrocytes, and blood vessels (173) and is particularly concentrated in presynaptic terminals, dendritic spines, and growth cones. Oxidation of actin, by HNE modification, can lead to loss of membrane cytoskeletal structure, decreased membrane fluidity, and trafficking of synaptic proteins and mitochondria. Moreover, actin is involved in the elongation of the growth cone, and loss of function of actin could play a role in synapse loss and neuronal communication, which may be associated with progressive memory loss documented in AD (174).

10.15. SIGNAL TRANSDUCTION

14–3–3-protein gamma is found to be nitrated in MCI IPL. 14–3–3 gamma is a member of the 14–3–3 protein families, which are highly expressed in the brain (175,176). These proteins are involved in a number of cellular functions including signal transduction, protein trafficking, and metabolism (175). 14–3–3 gamma is associated with neurofibrillary tangles in AD (177) and levels of 14–3–3 proteins are increased in MCI brain (155), AD brain (178,179), AD CSF (180), calorically restricted rats (181), and ICV and neuronal models of AD (116,182). The nitration of 14–3–3 gamma could change its conformation, which conceivably could lead to altered binding to two of its normal binding partners, glycogen synthase kinase 3 beta (GSK3ß) and tau. One of the isoforms of 14–3–3 can act as a scaffolding protein and simultaneously bind to tau and GSK3ß in a multiprotein tau phosphorylation complex (183). This complex may promote tau phosphorylation and polymerization (184,185), leading to the formation of tangles and further leading to neurodegeneration in AD.

Neuropolypeptide h3 is critical for modulation of the enzyme choline acetyltransferase, which is vital in signal transduction and cell communication. The loss of choline acetyltransferase leads to reduced levels of the neurotransmitter acetylcholine, causing poor neurotransmission (186). NMDA receptors activate the production of this enzyme, and alteration of the NMDA receptor mediates cholinergic deficits (187). AD has cholinergic deficits, consistent with dysregulation in acetylcholine levels and loss of cholinergic neurons (188191). Consequently, the oxidative modification of this protein in MCI and AD (113) highlights the role of cholinergism in the development of this dementing disease.

10.16. PROTEIN SYNTHESIS

EF-Tu and eIF-alpha are intimately involved in protein synthesis machinery. Oxidation, and possibly impairment, of initiation factors of protein synthesis in MCI patients were shown by redox proteomics (28). Human mitochondrial EF-Tu (EF-Tu) is a nuclear-encoded protein and functions in the translational apparatus of mitochondria (192). Like its prokaryotic homolog, the mammalian EF-Tu GTPase hydrolyzes a molecule of GTP each time an amino-acylated tRNA is accommodated on the A site of the ribosome, and its recycling depends on the exchange factor EF-Ts. Nuclear genes encode most respiratory chain subunits and all protein components necessary for maintenance and expression of mtDNA. Mitochondria play pivotal roles in eukaryotic cells in producing cellular energy and essential metabolites as well as in controlling apoptosis by integrating numerous death signals (193). Mitochondrial protein synthesis inhibition is associated with the impairment of differentiation in different cell types, including neurons (194). The coordination of mitochondrial and nuclear genetic systems in the cell is necessary for proper mitochondrial biogenesis and cellular functioning. eIF-α is an abundant protein required to bind aminoacyl-tRNA to acceptor sites of ribosomes in a GTP-dependent manner during protein synthesis (195). Recently, eIF-α has been shown to be involved in cytoskeletal organization by binding and bundling actin filaments and microtubules. eIF-α is an important determinant of cell proliferation and senescence (196), since it is regulated in aging, transformation, and growth arrest. Inhibition of eIF-α induces apoptosis (197), indicating that eIF-α activity is critical to normal cell function. Numerous studies have provided indirect evidence that suggests alterations in protein synthesis may occur in AD (53,198201) and decreased protein synthesis in AD and MCI (53,202). The dysfunction of the protein synthesis apparatus, mediated in part by oxidative stress, could compromise the ability of cells to generate the various factors needed to regulate cell homeostasis, thus contributing to impaired neuronal function and to the development of neuropathology in MCI patients.

10.17. CONCLUSIONS

In conclusion, the redox proteomics-identified brain proteins in MCI brain play important roles in different neuronal functions and are directly or indirectly linked to AD pathology. Comparative analysis of these proteins between MCI IPL and hippocampus brain regions showed alpha enolase as a common target of protein oxidation by all three modifications mentioned. This suggests that energy metabolism may be among the first cellular properties that become severely affected in MCI. A similar sensitivity to energy metabolism was observed in AD brain (112,113,116,203). Future studies using animal models of the different stages of this dementing disorder should help in further delineating the mechanisms of MCI pathogenesis and to develop effective therapies to combat conversion of MCI to AD. Several HNE-bound and nitrated proteins identified in MCI were identical to those found in not only AD, but other neurodegenerative diseases. These enzymes are involved in cholinergic processes, energy metabolism, transport, detoxification, protein synthesis, and stress response properties. Loss of function or modification to any protein can be catastrophic for neuronal communication, ATP production, and other essential cell functions. Because oxidative stress is involved in arguably the earliest phase of AD (31,32,83), and because MCI may be considered a prodromal phase of AD, the identification of specific target proteins that are oxidatively modified by products of lipid peroxidation and protein nitration may provide vital insights into the role of oxidative damage in mechanisms of neuronal death in AD. The current work forms a framework for subsequent experiments and provides potential new targets for neuroprotective therapeutic intervention in MCI.

ACKNOWLEDGMENTS

The authors thank the University of Kentucky A.D.C. Clinical and Neuropathology Cores for providing the brain specimens used for this study. This research was supported in part by grants from NIH (AG-10836; AG-05119; AG-029839).

REFERENCES

1.
Winblad B., Palmer K., Kivipelto M., et al. Mild cognitive impairment—Beyond controversies, towards a consensus: Report of the International Working Group on Mild Cognitive Impairment. J Intern Med. 2004;256:240–6. [PubMed: 15324367]
2.
Petersen R. C. Mild cognitive impairment as a diagnostic entity. J Intern Med. 2004;256:183–94. [PubMed: 15324362]
3.
Almkvist O., Basun H., Backman L., et al. Mild cognitive impairment—An early stage of Alzheimer’s disease? J Neural Transm Suppl. 1998;54:21–9. [PubMed: 9850912]
4.
Flicker C., Ferris S. H., Reisberg B. Mild cognitive impairment in the elderly: Predictors of dementia. Neurology. 1991;41:1006–9. [PubMed: 2067629]
5.
Luis C. A., Loewenstein D. A., Acevedo A., Barker W. W., Duara R. Mild cognitive impairment: Directions for future research. Neurology. 2003;61:438–44. [PubMed: 12939414]
6.
Morris J. C., Storandt M., Miller J. P., et al. Mild cognitive impairment represents early-stage Alzheimer disease. Arch Neurol. 2001;58:397–405. [PubMed: 11255443]
7.
Portet F., Ousset P. J., Touchon J. What is a mild cognitive impairment? Rev Prat. 2005;55:1891–4. [PubMed: 16396229]
8.
Maioli F., Coveri M., Pagni P., et al. Conversion of mild cognitive impairment to dementia in elderly subjects: A preliminary study in a memory and cognitive disorder unit. Arch Gerontol Geriatr. 2007;44 Suppl 1:233–41. [PubMed: 17317458]
9.
Rozzini L., Chilovi B. V., Conti M., et al. Conversion of amnestic mild cognitive impairment to dementia of Alzheimer type is independent to memory deterioration. Int J Geriatr Psychiatry. 2007;22:1217–22. [PubMed: 17562522]
10.
Apostolova L. G., Dutton R. A., Dinov I. D., et al. Conversion of mild cognitive impairment to Alzheimer disease predicted by hippocampal atrophy maps. Arch Neurol. 2006;63:693–9. [PubMed: 16682538]
11.
Petersen R. C. Mild cognitive impairment: Transition between aging and Alzheimer’s disease. Neurologia. 2000;15:93–101. [PubMed: 10846869]
12.
Petersen R. C. Mild cognitive impairment clinical trials. Nat Rev Drug Discov. 2003;2:646–53. [PubMed: 12904814]
13.
de Leon M. J., DeSanti S., Zinkowski R., et al. MRI and CSF studies in the early diagnosis of Alzheimer’s disease. J Intern Med. 2004;256:205–23. [PubMed: 15324364]
14.
Devanand D. P., Pradhaban G., Liu X., et al. Hippocampal and entorhinal atrophy in mild cognitive impairment: Prediction of Alzheimer disease. Neurology. 2007;68:828–36. [PubMed: 17353470]
15.
Jack C. R. Jr., Petersen R. C., Xu Y. C., et al. Prediction of AD with MRI-based hippocampal volume in mild cognitive impairment. Neurology. 1999;52:1397–1403. [PMC free article: PMC2730146] [PubMed: 10227624]
16.
Barnes J., Godbolt A. K., Frost C., et al. Atrophy rates of the cingulate gyrus and hippocampus in AD and FTLD. Neurobiol Aging. 2007;28:20–8. [PubMed: 16406154]
17.
Du A. T., Schuff N., Amend D., et al. Magnetic resonance imaging of the entorhinal cortex and hippocampus in mild cognitive impairment and Alzheimer’s disease. J Neurol Neurosurg Psychiatry. 2001;71:441–7. [PMC free article: PMC1763497] [PubMed: 11561025]
18.
Du A. T., Schuff N., Kramer J. H., et al. Higher atrophy rate of entorhinal cortex than hippocampus in AD. Neurology. 2004;62:422–7. [PMC free article: PMC1820859] [PubMed: 14872024]
19.
Mevel K., Chetelat G., Desgranges B., Eustache F. 2006Alzheimer’s disease, hippocampus and neuroimaging Encephale 32Pt 4>S1149–54. [PubMed: 17356489]
20.
Mori E. Hippocampal atrophy and memory disturbance. No To Shinkei. 2005;57:1067–78. [PubMed: 16375192]
21.
Halliwell B. Oxidative stress and neurodegeneration: Where are we now? J Neurochem. 2006;97:1634–58. [PubMed: 16805774]
22.
Brown L. A., Harris F. L., Jones D. P. Ascorbate deficiency and oxidative stress in the alveolar type II cell. Am J Physiol. 1997;273:L782–8. [PubMed: 9357853]
23.
Butterfield D. A., Kanski J. Brain protein oxidation in age-related neurodegenerative disorders that are associated with aggregated proteins. Mech Aging Dev. 2001;122:945–62. [PubMed: 11348660]
24.
Butterfield D. A., Lauderback C. M. Lipid peroxidation and protein oxidation in Alzheimer’s disease brain: potential causes and consequences involving amyloid beta-peptide-associated free radical oxidative stress. Free Radic Biol Med. 2002;32:1050–60. [PubMed: 12031889]
25.
Giasson B. I., Ischiropoulos H., Lee V. M., Trojanowski J. Q. The relationship between oxidative/nitrative stress and pathological inclusions in Alzheimer’s and Parkinson’s diseases. Free Radic Biol Med. 2002;32:1264–75. [PubMed: 12057764]
26.
Markesbery W. R. Oxidative stress hypothesis in Alzheimer’s disease. Free Radic Biol Med. 1997;23:134–47. [PubMed: 9165306]
27.
Zhu X., Smith M. A., Perry G., Aliev G. Mitochondrial failures in Alzheimer’s disease. Am J Alzheimers Dis Other Demen. 2004;19:345–52. [PubMed: 15633943]
28.
Reed T., Perluigi M., Sultana R., et al. 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;30:107–20. [PubMed: 18325775]
29.
Keller J. N., Schmitt F. A., Scheff S. W., et al. Evidence of increased oxidative damage in subjects with mild cognitive impairment. Neurology. 2005;64:1152–6. [PubMed: 15824339]
30.
Butterfield D. A., Stadtman E. R. Protein oxidation processes in aging brain. Adv Cell Aging Gerontol. 1997;2:161–91.
31.
Butterfield D. A., Reed T., Perluigi M., et al. Elevated protein-bound levels of the lipid peroxidation product, 4-hydroxy-2-nonenal, in brain from persons with mild cognitive impairment. Neurosci Lett. 2006;397:170–3. [PubMed: 16413966]
32.
Butterfield D. A., Reed T. T., Perluigi M., et al. Elevated levels of 3-nitrotyrosine in brain from subjects with amnestic mild cognitive impairment: Implications for the role of nitration in the progression of Alzheimer’s disease. Brain Res. 2007;1148:243–8. [PMC free article: PMC1934617] [PubMed: 17395167]
33.
Dalle-Donne I., Rossi R., Colombo R., Giustarini D., Milzani A. Biomarkers of oxidative damage in human disease. Clin Chem. 2006;52:601–23. [PubMed: 16484333]
34.
Dalle-Donne I., Scaloni A., Giustarini D., et al. Proteins as biomarkers of oxidative/ nitrosative stress in diseases: the contribution of redox proteomics. Mass Spectrom Rev. 2005;24:55–99. [PubMed: 15389864]
35.
Levine R. L., Garland D., Oliver C. N., et al. Determination of carbonyl content in oxidatively modified proteins. Methods Enzymol. 1990;186:464–78. [PubMed: 1978225]
36.
Aksenov M. Y., Aksenova M. V., Butterfield D. A., Geddes J. W., Markesbery W. R. Protein oxidation in the brain in Alzheimer’s disease. Neuroscience. 2001;103:373–83. [PubMed: 11246152]
37.
Stadtman E. R. Protein oxidation in aging and age-related diseases. Ann NY Acad Sci. 2001;928:22–38. [PubMed: 11795513]
38.
Stadtman E. R., Berlett B. S. Reactive oxygen-mediated protein oxidation in aging and disease. Chem Res Toxicol. 1997;10:485–94. [PubMed: 9168245]
39.
Loske C., Neumann A., Cunningham A. M., et al. Cytotoxicity of advanced glycation endproducts is mediated by oxidative stress. J Neural Transm. 1998;105:1005–15. [PubMed: 9869332]
40.
Munch G., Schinzel R., Loske C., et al. Alzheimer’s disease—Synergistic effects of glucose deficit, oxidative stress and advanced glycation endproducts. J Neural Transm. 1998;105:439–61. [PubMed: 9720973]
41.
Esterbauer H., Schaur R. J., Zollner H. Chemistry and biochemistry of 4-hydroxynonenal, malonaldehyde and related aldehydes. Free Radic Biol Med. 1991;11:81–128. [PubMed: 1937131]
42.
Pamplona R., Dalfo E., Ayala V., et al. 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;280:21522–30. [PubMed: 15799962]
43.
O’Brien J. S., Sampson E. L. Lipid composition of the normal human brain: Gray matter, white matter, and myelin. J Lipid Res. 1965;6:537–44. [PubMed: 5865382]
44.
Skinner E. R., Watt C., Besson J. A., Best P. V. 1993Differences in the fatty acid composition of the grey and white matter of different regions of the brains of patients with Alzheimer’s disease and control subjects Brain 116(Pt 3)>717–25. [PubMed: 8513399]
45.
Montine T. J., Neely M. D., Quinn J. F., et al. Lipid peroxidation in aging brain and Alzheimer’s disease. Free Radic Biol Med. 2002;33:620–6. [PubMed: 12208348]
46.
Perluigi M., Sultana R., Cenini G., Di Domenico F., Memo M., Pierce W. M., Coccia R., Butterfield D. A. Redox Proteomics Identification of HNEModified Brain Proteins in Alzheimer’s Disease: Role of Lipid Peroxidaton in Alzheimer’s Disease Pathogenesis. Proteomics—Clin Appl. 2009;118:131–150. [PMC free article: PMC2843938] [PubMed: 20333275]
47.
Bader Lange M. L., Cenini G., Piroddi M., et al. Loss of phospholipid asymmetry and elevated brain apoptotic protein levels in subjects with amnestic mild cognitive impairment and Alzheimer disease. Neurobiol Dis. 2008;29:456–64. [PMC free article: PMC2292396] [PubMed: 18077176]
48.
Williams T. I., Lynn B. C., Markesbery W. R., Lovell M. A. 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;27:1094–9. [PubMed: 15993986]
49.
Subramaniam R., Roediger F., Jordan B., et al. The lipid peroxidation product,4-hydroxy-2-trans-nonenal, alters the conformation of cortical synaptosomal membrane proteins. J Neurochem. 1997;69:1161–9. [PubMed: 9282939]
50.
Tamagno E., Robino G., Obbili A., et al. H2O2 and 4-hydroxynonenal mediate amyloid beta-induced neuronal apoptosis by activating JNKs and p38MAPK. Exp Neurol. 2003;180:144–55. [PubMed: 12684028]
51.
Uchida K. 4-Hydroxy-2-nonenal: A product and mediator of oxidative stress. Prog Lipid Res. 2003;42:318–43. [PubMed: 12689622]
52.
Camandola S., Poli G., Mattson M. P. The lipid peroxidation product 4-hydroxy-2,3-nonenal inhibits constitutive and inducible activity of nuclear factor kappa B in neurons. Brain Res Mol Brain Res. 2000;85:53–60. [PubMed: 11146106]
53.
Ding Q., Markesbery W. R., Chen Q., Li F., Keller J. N. Ribosome dysfunction is an early event in Alzheimer’s disease. J Neurosci. 2005;25:9171–5. [PMC free article: PMC6725754] [PubMed: 16207876]
54.
Drake J., Petroze R., Castegna A., et al. 4-Hydroxynonenal oxidatively modifies histones: Implications for Alzheimer’s disease. Neurosci Lett. 2004;356:155–8. [PubMed: 15036618]
55.
Poot M., Verkerk A., Koster J. F., Esterbauer H., Jongkind J. F. Reversible inhibition of DNA and protein synthesis by cumene hydroperoxide and 4-hydroxy-nonenal. Mech Ageing Dev. 1988;43:1–9. [PubMed: 3374176]
56.
Uchida K., Stadtman E. R. Modification of histidine residues in proteins by reaction with 4-hydroxynonenal. Proc Natl Acad Sci U S A. 1992;89:4544–8. [PMC free article: PMC49119] [PubMed: 1584790]
57.
Okada K., Wangpoengtrakul C., Osawa T., et al. 4-Hydroxy-2-nonenal-mediated impairment of intracellular proteolysis during oxidative stress. Identification of proteasomes as target molecules. J Biol Chem. 1999;274:23787–93. [PubMed: 10446139]
58.
Pratico D., Clark C. M., Liun F., et al. Increase of brain oxidative stress in mild cognitive impairment: A possible predictor of Alzheimer disease. Arch Neurol. 2002;59:972–6. [PubMed: 12056933]
59.
Alvarez B., Radi R. Peroxynitrite reactivity with amino acids and proteins. Amino Acids. 2003;25:295–311. [PubMed: 14661092]
60.
Tien M., Berlett B. S., Levine R. L., Chock P. B., Stadtman E. R. Peroxynitrite-mediated modification of proteins at physiological carbon dioxide concentration: pH dependence of carbonyl formation, tyrosine nitration, and methionine oxidation. Proc Natl Acad Sci U S A. 1999;96:7809–14. [PMC free article: PMC22143] [PubMed: 10393903]
61.
Hazen S. L., Gaut J. P., Hsu F. F., et al. p-Hydroxyphenylacetaldehyde, the major product of L-tyrosine oxidation by the myeloperoxidase-H2O2-chloride system of phagocytes, covalently modifies epsilon-amino groups of protein lysine residues. J Biol Chem. 1997;272:16990–8. [PubMed: 9202012]
62.
Smith M. A., Richey Harris P. L., Sayre L. M., Beckman J. S., Perry G. Widespread peroxynitrite-mediated damage in Alzheimer’s disease. J Neurosci. 1997;17:2653–7. [PMC free article: PMC6573097] [PubMed: 9092586]
63.
Botti H., Trostchansky A., Batthyany C., Rubbo H. Reactivity of peroxynitrite and nitric oxide with LDL. IUBMB Life. 2005;57:407–12. [PubMed: 16012049]
64.
Niles J. C., Wishnok J. S., Tannenbaum S. R. Peroxynitrite-induced oxidation and nitration products of guanine and 8-oxoguanine: Structures and mechanisms of product formation. Nitric Oxide. 2006;14:109–21. [PubMed: 16352449]
65.
Szabo C. DNA strand breakage and activation of poly-ADP ribosyltransferase: A cytotoxic pathway triggered by peroxynitrite. Free Radic Biol Med. 1996;21:855–69. [PubMed: 8902531]
66.
Masuda M., Nishino H., Ohshima H. Formation of 8-nitroguanosine in cellular RNA as a biomarker of exposure to reactive nitrogen species. Chem Biol Interact. 2002;139:187–97. [PubMed: 11823006]
67.
Good P. F., Werner P., Hsu A., Olanow C. W., Perl D. P. Evidence of neuronal oxidative damage in Alzheimer’s disease. Am J Pathol. 1996;149:21–8. [PMC free article: PMC1865248] [PubMed: 8686745]
68.
Good P. F., Hsu A., Werner P., Perl D. P., Olanow C. W. Protein nitration in Parkinson’s disease. J Neuropathol Exp Neurol. 1998;57:338–42. [PubMed: 9600227]
69.
Drake J., Kanski J., Varadarajan S., Tsoras M., Butterfield D. A. Elevation of brain glutathione by gamma-glutamylcysteine ethyl ester protects against peroxynitrite-induced oxidative stress. J Neurosci Res. 2002;68:776–84. [PubMed: 12111838]
70.
Whiteman M., Tritschler H., Halliwell B. Protection against peroxynitrite-dependent tyrosine nitration and alpha 1-antiproteinase inactivation by oxidized and reduced lipoic acid. FEBS Lett. 1996;379:74–6. [PubMed: 8566234]
71.
Christen S., Woodall A. A., Shigenaga M. K., et al. gamma-tocopherol traps mutagenic electrophiles such as NO(X) and complements alpha-tocopherol: Physiological implications. Proc Natl Acad Sci U S A. 1997;94:3217–22. [PMC free article: PMC20349] [PubMed: 9096373]
72.
Anantharaman M., Tangpong J., Keller J. N., et al. Beta-amyloid mediated nitration of manganese superoxide dismutase: Implication for oxidative stress in a APPNLH/ NLH X PS-1P264L/P264L double knock-in mouse model of Alzheimer’s disease. Am J Pathol. 2006;168:1608–18. [PMC free article: PMC1606606] [PubMed: 16651627]
73.
Aoyama K., Matsubara K., Fujikawa Y., et al. Nitration of manganese superoxide dismutase in cerebrospinal fluids is a marker for peroxynitrite-mediated oxidative stress in neurodegenerative diseases. Ann Neurol. 2000;47:524–7. [PubMed: 10762167]
74.
Ischiropoulos H., Zhu L., Chen J., et al. Peroxynitrite-mediated tyrosine nitration catalyzed by superoxide dismutase. Arch Biochem Biophys. 1992;298:431–7. [PubMed: 1416974]
75.
MacMillan-Crow L. A., Thompson J. A. Tyrosine modifications and inactivation of active site manganese superoxide dismutase mutant (Y34F) by peroxynitrite. Arch Biochem Biophys. 1999;366:82–8. [PubMed: 10334867]
76.
Hara M. R., Cascio M. B., Sawa A. GAPDH as a sensor of NO stress. Biochim Biophys Acta. 2006;1762:502–9. [PubMed: 16574384]
77.
Souza J. M., Radi R. Glyceraldehyde-3-phosphate dehydrogenase inactivation by peroxynitrite. Arch Biochem Biophys. 1998;360:187–94. [PubMed: 9851830]
78.
Clements M. K., Siemsen D. W., Swain S. D., et al. Inhibition of actin polymerization by peroxynitrite modulates neutrophil functional responses. J Leukoc Biol. 2003;73:344–55. [PubMed: 12629148]
79.
Neumann P., Gertzberg N., Vaughan E., et al. Peroxynitrite mediates TNF- {alpha}-induced endothelial barrier dysfunction and nitration of actin. Am J Physiol Lung Cell Mol Physiol. 2006;290:L674–84. [PubMed: 16284212]
80.
Di Stasi A. M., Mallozzi C., Macchia G., et al. Peroxynitrite affects exocytosis and SNARE complex formation and induces tyrosine nitration of synaptic proteins. J Neurochem. 2002;82:420–9. [PubMed: 12124443]
81.
Blanchard-Fillion B., Souza J. M., Friel T., et al. Nitration and inactivation of tyrosine hydroxylase by peroxynitrite. J Biol Chem. 2001;276:46017–23. [PubMed: 11590168]
82.
Gow A. J., Duran D., Malcolm S., Ischiropoulos H. Effects of peroxynitrite-induced protein modifications on tyrosine phosphorylation and degradation. FEBS Lett. 1996;385:63–6. [PubMed: 8641468]
83.
Butterfield D. A., Poon H. F., St Clair D., et al. Redox proteomics identification of oxidatively modified hippocampal proteins in mild cognitive impairment: Insights into the development of Alzheimer’s disease. Neurobiol Dis. 2006;22:223–32. [PubMed: 16466929]
84.
Wilkins M. R., Sanchez J. C., Gooley A. A., et al. Progress with proteome projects: Why all proteins expressed by a genome should be identified and how to do it. Biotechnol Genet Eng Rev. 1996;13:19–50. [PubMed: 8948108]
85.
Davidsson P., Sjogren M. The use of proteomics in biomarker discovery in neurodegenerative diseases. Dis Markers. 2005;21:81–92. [PMC free article: PMC3850612] [PubMed: 15920295]
86.
Galasko D. Biomarkers for Alzheimer’s disease—Clinical needs and application. J Alzheimers Dis. 2005;8:339–46. [PubMed: 16556965]
87.
Lee J. W., Namkoong H., Kim H. K., et al. Fibrinogen gamma—A chain precursor in CSF: A candidate biomarker for Alzheimer’s disease. BMC Neurol. 2007;7:14. [PMC free article: PMC1896177] [PubMed: 17565664]
88.
Hortin G. L., Jortani S. A., Ritchie J. C. Jr., Valdes R. Jr., Chan D. W. Proteomics: A new diagnostic frontier. Clin Chem. 2006;52:1218–22. [PubMed: 16675505]
89.
Solassol J., Boulle N., Maudelonde T., Mange A. Clinical proteomics: Towards early detection of cancers. Med Sci (Paris). 2005;21:722–9. [PubMed: 16115457]
90.
Ho L., Sharma N., Blackman L., et al. From proteomics to biomarker discovery in Alzheimer’s disease. Brain Res Rev. 2005;48:360–9. [PubMed: 15850675]
91.
Simonsen A. H., McGuire J., Hansson O., et al. Novel panel of cerebrospinal fluid biomarkers for the prediction of progression to Alzheimer dementia in patients with mild cognitive impairment. Arch Neurol. 2007;64:366–70. [PubMed: 17353378]
92.
Zhang J., Goodlett D. R., Quinn J. F., et al. Quantitative proteomics of cerebrospinal fluid from patients with Alzheimer disease. J Alzheimers Dis. 2005;7:125–33. discussion 173–80. [PubMed: 15851850]
93.
Poon H. F., Vaishnav R. A., Getchell T. V., Getchell M. L., Butterfield D. A. Quantitative proteomics analysis of differential protein expression and oxidative modification of specific proteins in the brains of old mice. Neurobiol Aging. 2006;27:1010–9. [PubMed: 15979213]
94.
Boguski M. S., McIntosh M. W. Biomedical informatics for proteomics. Nature. 2003;422:233–7. [PubMed: 12634797]
95.
Maurer H. H., Peters F. T. Toward high-throughput drug screening using mass spectrometry. Ther Drug Monit. 2005;27:686–8. [PubMed: 16404794]
96.
Thongboonkerd V., McLeish K. R., Arthur J. M., Klein J. B. Proteomic analysis of normal human urinary proteins isolated by acetone precipitation or ultracentrifugation. Kidney Int. 2002;62:1461–9. [PubMed: 12234320]
97.
Aebersold R., Goodlett D. R. Mass spectrometry in proteomics. Chem Rev. 2001;101:269–95. [PubMed: 11712248]
98.
Domon B., Aebersold R. Mass spectrometry and protein analysis. Science. 2006;312:212–7. [PubMed: 16614208]
99.
Gygi S. P., Aebersold R. Mass spectrometry and proteomics. Curr Opin Chem Biol. 2000;4:489–94. [PubMed: 11006534]
100.
Aksenov M., Aksenova M., Butterfield D. A., Markesbery W. R. Oxidative modification of creatine kinase BB in Alzheimer’s disease brain. J Neurochem. 2000;74:2520–7. [PubMed: 10820214]
101.
Poon H. F., Frasier M., Shreve N., et al. Mitochondrial associated metabolic proteins are selectively oxidized in A30P alpha-synuclein transgenic mice—A model of familial Parkinson’s disease. Neurobiol Dis. 2005;18:492–8. [PubMed: 15755676]
102.
Butterfield D. A., Castegna A. Proteomic analysis of oxidatively modified proteins in Alzheimer’s disease brain: Insights into neurodegeneration. Cell Mol Biol (Noisy-le-grand). 2003;49:747–51. [PubMed: 14528911]
103.
Hensley K., Hall N., Subramaniam R., et al. Brain regional correspondence between Alzheimer’s disease histopathology and biomarkers of protein oxidation. J Neurochem. 1995;65:2146–56. [PubMed: 7595501]
104.
Sultana R., Boyd-Kimball D., Poon H. F., et al. Oxidative modification and down-regulation of Pin1 in Alzheimer’s disease hippocampus: A redox proteomics analysis. Neurobiol Aging. 2006;27:918–25. [PubMed: 15950321]
105.
Tangpong J., Cole M. P., Sultana R., et al. Adriamycin-mediated nitration of manganese superoxide dismutase in the central nervous system: insight into the mechanism of chemobrain. J Neurochem. 2007;100:191–201. [PubMed: 17227439]
106.
Scheff S. W., Price D. A., Schmitt F. A., Mufson E. J. Hippocampal synaptic loss in early Alzheimer’s disease and mild cognitive impairment. Neurobiol Aging. 2006;27:1372–84. [PubMed: 16289476]
107.
Cao Q., Jiang K., Zhang M., et al. Brain glucose metabolism and neuropsychological test in patients with mild cognitive impairment. Chin Med J (Engl). 2003;116:1235–8. [PubMed: 12935418]
108.
Geddes J. W., Pang Z., Wiley D. H. Hippocampal damage and cytoskeletal disruption resulting from impaired energy metabolism. Implications for Alzheimer disease. Mol Chem Neuropathol. 1996;28:65–74. [PubMed: 8871943]
109.
Messier C., Gagnon M. Glucose regulation and cognitive functions: Relation to Alzheimer’s disease and diabetes. Behav Brain Res. 1996;75:1–11. [PubMed: 8800646]
110.
Vanhanen M., Soininen H. Glucose intolerance, cognitive impairment and Alzheimer’s disease. Curr Opin Neurol. 1998;11:673–7. [PubMed: 9870136]
111.
Boyd-Kimball D., Castegna A., Sultana R., et al. Proteomic identification of proteins oxidized by Abeta(1-42) in synaptosomes: Implications for Alzheimer’s disease. Brain Res. 2005;1044:206–15. [PubMed: 15885219]
112.
Castegna A., Aksenov M., Thongboonkerd V., et al. Proteomic identification of oxidatively modified proteins in Alzheimer’s disease brain. Part II: Dihydropyrimidinaserelated protein 2, alpha-enolase and heat shock cognate 71. J Neurochem. 2002;82:1524–32. [PubMed: 12354300]
113.
Castegna A., Thongboonkerd V., Klein J. B., et al. Proteomic identification of nitrated proteins in Alzheimer’s disease brain. J Neurochem. 2003;85:1394–401. [PubMed: 12787059]
114.
Poon H. F., Castegna A., Farr S. A., et al. Quantitative proteomics analysis of specific protein expression and oxidative modification in aged senescence-acceleratedprone 8 mice brain. Neuroscience. 2004;126:915–26. [PubMed: 15207326]
115.
Sultana R., Boyd-Kimball D., Poon H. F., et al. 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;27:1564–76. [PubMed: 16271804]
116.
Sultana R., Perluigi M., Butterfield D. A. Redox proteomics identification of oxidatively modified proteins in Alzheimer’s disease brain and in vivo and in vitro models of AD centered around Abeta(1-42). J Chromatogr B Analyt Technol Biomed Life Sci. 2006;833:3–11. [PubMed: 16236561]
117.
Kida K., Nishio T., Nagai K., Matsuda H., Nakagawa H. Gluconeogenesis in the kidney in vivo in fed rats. Circadian change and substrate specificity. J Biochem (Tokyo). 1982;91:755–60. [PubMed: 7076647]
118.
Hoyer S. Glucose metabolism and insulin receptor signal transduction in Alzheimer disease. Eur J Pharmacol. 2004;490:115–25. [PubMed: 15094078]
119.
Rapoport S. I. Functional brain imaging in the resting state and during activation in Alzheimer’s disease. Implications for disease mechanisms involving oxidative phosphorylation. Ann N Y Acad Sci. 1999;893:138–53. [PubMed: 10672235]
120.
Duan Z., Lamendola D. E., Yusuf R. Z., et al. Overexpression of human phosphoglycerate kinase 1 (PGK1) induces a multidrug resistance phenotype. Anticancer Res. 2002;22:1933–41. [PubMed: 12174867]
121.
Keller J. N., Mark R. J., Bruce A. J., et al. 4-Hydroxynonenal, an aldehydic product of membrane lipid peroxidation, impairs glutamate transport and mitochondrial function in synaptosomes. Neuroscience. 1997;80:685–96. [PubMed: 9276486]
122.
Mattson M. P. Metal-catalyzed disruption of membrane protein and lipid signaling in the pathogenesis of neurodegenerative disorders. Ann N Y Acad Sci. 2004;1012:37–50. [PubMed: 15105254]
123.
Aksenov M. Y., Aksenova M. V., Butterfield D. A., et al. Glutamine synthetase-induced enhancement of beta-amyloid peptide A beta (1-40) neurotoxicity accompanied by abrogation of fibril formation and A beta fragmentation. J Neurochem. 1996;66:2050–6. [PubMed: 8780035]
124.
Butterfield D. A., Hensley K., Cole P., et al. Oxidatively induced structural alteration of glutamine synthetase assessed by analysis of spin label incorporation kinetics: Relevance to Alzheimer’s disease. J Neurochem. 1997;68:2451–7. [PubMed: 9166739]
125.
Howard B. J., Yatin S., Hensley K., et al. Prevention of hyperoxia-induced alterations in synaptosomal membrane-associated proteins by N-tert-butyl-alphaphenylnitrone and 4-hydroxy-2,2,6,6-tetramethylpiperidin-1-oxyl (Tempol). J Neurochem. 1996;67:2045–50. [PubMed: 8863512]
126.
Lamprecht R., LeDoux J. Structural plasticity and memory. Nat Rev Neurosci. 2004;5:45–54. [PubMed: 14708003]
127.
Ikonomovic M. D., Mufson E. J., Wuu J., et al. Cholinergic plasticity in hippocampus of individuals with mild cognitive impairment: Correlation with Alzheimer’s neuropathology. J Alzheimers Dis. 2003;5:39–48. [PubMed: 12590165]
128.
Lee H. G., Zhu X., Ghanbari H. A., et al. Differential regulation of glutamate receptors in Alzheimer’s disease. Neurosignals. 2002;11:282–92. [PubMed: 12566929]
129.
Sergeant N., Wattez A., Galvan-valencia M., et al. Association of ATP synthase alpha-chain with neurofibrillary degeneration in Alzheimer’s disease. Neuroscience. 2003;117:293–303. [PubMed: 12614671]
130.
Butterfield D. A., Castegna A., Lauderback C. M., 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;23:655–64. [PubMed: 12392766]
131.
Butterfield D. A., 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;7:548–54. [PubMed: 11733217]
132.
Hinault M. P., Ben-Zvi A., Goloubinoff P. Chaperones and proteases: Cellular fold-controlling factors of proteins in neurodegenerative diseases and aging. J Mol Neurosci. 2006;30:249–65. [PubMed: 17401151]
133.
Sultana R., Reed T., Perluigi M., et al. Proteomic identification of nitrated brain proteins in amnestic mild cognitive impairment: A regional study. J Cell Mol Med. 2007;11:839–51. [PMC free article: PMC3823261] [PubMed: 17760844]
134.
Perluigi M., Fai Poon H., Hensley K., et al. Proteomic analysis of 4-hydroxy-2-nonenal-modified proteins in G93A-SOD1 transgenic mice—A model of familial amyotrophic lateral sclerosis. Free Radic Biol Med. 2005;38:960–8. [PubMed: 15749392]
135.
Magrane J., Smith R. C., Walsh K., Querfurth H. W. Heat shock protein 70 participates in the neuroprotective response to intracellularly expressed beta-amyloid in neurons. J Neurosci. 2004;24:1700–6. [PMC free article: PMC6730449] [PubMed: 14973234]
136.
Schutkowski M., Bernhardt A., Zhou X. Z., et al. Role of phosphorylation in determining the backbone dynamics of the serine/threonine-proline motif and Pin1 substrate recognition. Biochemistry. 1998;37:5566–75. [PubMed: 9548941]
137.
Holzer M., Gartner U., Stobe A., et al. Inverse association of Pin1 and tau accumulation in Alzheimer’s disease hippocampus. Acta Neuropathol. 2002;104:471–81. [PubMed: 12410395]
138.
Kurt M. A., Davies D. C., Kidd M., Duff K., Howlett D. R. Hyperphosphorylated tau and paired helical filament-like structures in the brains of mice carrying mutant amyloid precursor protein and mutant presenilin-1 transgenes. Neurobiol Dis. 2003;14:89–97. [PubMed: 13678670]
139.
Ramakrishnan P., Dickson D. W., Davies P. Pin1 colocalization with phosphorylated tau in Alzheimer’s disease and other tauopathies. Neurobiol Dis. 2003;14:251–64. [PubMed: 14572447]
140.
Sultana R., Boyd-Kimball D., Poon H. F., et al. 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. 2005;27:1564–76. [PubMed: 16271804]
141.
Lu P. J., Wulf G., Zhou X. Z., Davies P., Lu K. P. The prolyl isomerase Pin1 restores the function of Alzheimer-associated phosphorylated tau protein. Nature. 1999;399:784–8. [PubMed: 10391244]
142.
Butterfield D. A., Abdul H. M., Opii W., et al. Pin1 in Alzheimer’s disease. J Neurochem. 2006;98:1697–706. [PubMed: 16945100]
143.
Lauderback C. M., Hackett J. M., Huang F. F., et al. 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;78:413–6. [PubMed: 11461977]
144.
Sultana R., Butterfield D. A. Oxidatively modified GST and MRP1 in Alzheimer’s disease brain: Implications for accumulation of reactive lipid peroxidation products. Neurochem Res. 2004;29:2215–20. [PubMed: 15672542]
145.
Xie C., Lovell M. A., Xiong S., et al. Expression of glutathione-S-transferase isozyme in the SY5Y neuroblastoma cell line increases resistance to oxidative stress. Free Radic Biol Med. 2001;31:73–81. [PubMed: 11425492]
146.
Peshenko I. V., Shichi H. Oxidation of active center cysteine of bovine 1-Cys peroxiredoxin to the cysteine sulfenic acid form by peroxide and peroxynitrite. Free Radic Biol Med. 2001;31:292–303. [PubMed: 11461766]
147.
Akman S. A., O’Connor T. R., Rodriguez H. Mapping oxidative DNA damage and mechanisms of repair. Ann N Y Acad Sci. 2000;899:88–102. [PubMed: 10863531]
148.
Doorn J. A., Maser E., Blum A., Claffey D. J., Petersen D. R. Human carbonyl reductase catalyzes reduction of 4-oxonon-2-enal. Biochemistry. 2004;43:13106–14. [PubMed: 15476404]
149.
Balcz B., Kirchner L., Cairns N., Fountoulakis M., Lubec G. Increased brain protein levels of carbonyl reductase and alcohol dehydrogenase in Down syndrome and Alzheimer’s disease. J Neural Transm Suppl. 2001;6:193–201. [PubMed: 11771743]
150.
Lemieux N., Malfoy B., Forrest G. L. Human carbonyl reductase (CBR) localized to band 21q22.1 by high-resolution fluorescence in situ hybridization displays gene dosage effects in trisomy 21 cells. Genomics. 1993;15:169–72. [PubMed: 8432528]
151.
Korenberg J. R., Bradley C., Disteche C. M. Down syndrome: Molecular mapping of the congenital heart disease and duodenal stenosis. Am J Hum Genet. 1992;50:294–302. [PMC free article: PMC1682442] [PubMed: 1531166]
152.
Kim S. H., Fountoulakis M., Cairns N., Lubec G. Protein levels of human peroxiredoxin subtypes in brains of patients with Alzheimer’s disease and Down syndrome. J Neural Transm Suppl. 2001;6:223–35. [PubMed: 11771746]
153.
Krapfenbauer K., Engidawork E., Cairns N., Fountoulakis M., Lubec G. Aberrant expression of peroxiredoxin subtypes in neurodegenerative disorders. Brain Res. 2003;967:152–60. [PubMed: 12650976]
154.
Schonberger S. J., Edgar P. F., Kydd R., Faull R. L., Cooper G. J. Proteomic analysis of the brain in Alzheimer’s disease: molecular phenotype of a complex disease process. Proteomics. 2001;1:1519–28. [PubMed: 11747211]
155.
Sultana R., Boyd-Kimball D., Cai J., et al. Proteomics analysis of the Alzheimer’s disease hippocampal proteome. J Alzheimers Dis. 2007;11:153–64. [PubMed: 17522440]
156.
Mecocci P. Oxidative stress in mild cognitive impairment and Alzheimer disease: A continuum. J Alzheimers Dis. 2004;6:159–63. [PubMed: 15096699]
157.
Hamajima N., Matsuda K., Sakata S., et al. A novel gene family defined by human dihydropyrimidinase and three related proteins with differential tissue distribution. Gene. 1996;180:157–63. [PubMed: 8973361]
158.
Kato Y., Hamajima N., Inagaki H., et al. Post-meiotic expression of the mouse dihydropyrimidinase-related protein 3 (DRP-3) gene during spermiogenesis. Mol Reprod Dev. 1998;51:105–11. [PubMed: 9712324]
159.
Lubec G., Nonaka M., Krapfenbauer K., et al. Expression of the dihydropyrimidinase related protein 2 (DRP-2) in Down syndrome and Alzheimer’s disease brain is downregulated at the mRNA and dysregulated at the protein level. J Neural Transm Suppl. 1999;57:161–77. [PubMed: 10666674]
160.
Yoshida H., Watanabe A., Ihara Y. Collapsin response mediator protein-2 is associated with neurofibrillary tangles in Alzheimer’s disease. J Biol Chem. 1998;273:9761–8. [PubMed: 9545313]
161.
Weitzdoerfer R., Fountoulakis M., Lubec G. Aberrant expression of dihydropyrimidinase related proteins-2,-3 and -4 in fetal Down syndrome brain. J Neural Transm Suppl. 2001;95:107. [PubMed: 11771764]
162.
Coleman P. D., Flood D. G. Neuron numbers and dendritic extent in normal aging and Alzheimer’s disease. Neurobiol Aging. 1987;8:521–45. [PubMed: 3323927]
163.
Arnaiz E., Almkvist O. Neuropsychological features of mild cognitive impairment and preclinical Alzheimer’s disease. Acta Neurol Scand Suppl. 2003;179:34–41. [PubMed: 12603249]
164.
Rami L., Molinuevo J. L., Sanchez-Valle R., Bosch B., Villar A. Screening for amnestic mild cognitive impairment and early Alzheimer’s disease with M@T (Memory Alteration Test) in the primary care population. Int J Geriatr Psychiatry. 2007;22:294–304. [PubMed: 16998781]
165.
Yamashiro S., Yamakita Y., Ono S., Matsumura F. Fascin, an actin-bundling protein, induces membrane protrusions and increases cell motility of epithelial cells. Mol Biol Cell. 1998;9:993–1006. [PMC free article: PMC25324] [PubMed: 9571235]
166.
Adams J. C. 1995Formation of stable microspikes containing actin and the 55 kDa actin bundling protein, fascin, is a consequence of cell adhesion to thrombospondin-1: Implications for the anti-adhesive activities of thrombospondin-1 J Cell Sci 108Pt 5>1977–90. [PubMed: 7657718]
167.
Adams J. C. Roles of fascin in cell adhesion and motility. Curr Opin Cell Biol. 2004;16:590–6. [PubMed: 15363811]
168.
Pinkus G. S., Lones M. A., Matsumura F., et al. Langerhans cell histiocytosis immunohistochemical expression of fascin, a dendritic cell marker. Am J Clin Pathol. 2002;118:335–43. [PubMed: 12219775]
169.
Graziewicz M. A., Day B. J., Copeland W. C. The mitochondrial DNA polymerase as a target of oxidative damage. Nucleic Acids Res. 2002;30:2817–24. [PMC free article: PMC117047] [PubMed: 12087165]
170.
Anilkumar N., Parsons M., Monk R., Ng T., Adams J. C. Interaction of fascin and protein kinase Calpha: a novel intersection in cell adhesion and motility. Embo J. 2003;22:5390–402. [PMC free article: PMC213775] [PubMed: 14532112]
171.
Opii W. O., Joshi G., Head E., et al. Proteomic identification of brain proteins in the canine model of human aging following a long-term treatment with antioxidants and a program of behavioral enrichment: Relevance to Alzheimer’s disease. Neurobiol Aging. 2008;29:51–70. [PMC free article: PMC2203613] [PubMed: 17055614]
172.
Battaini F., Pascale A., Lucchi L., Pasinetti G. M., Govoni S. Protein kinase C anchoring deficit in postmortem brains of Alzheimer’s disease patients. Exp Neurol. 1999;159:559–64. [PubMed: 10506528]
173.
Goldman J. E. Immunocytochemical studies of actin localization in the central nervous system. J Neurosci. 1983;3:1952–62. [PMC free article: PMC6564568] [PubMed: 6352870]
174.
Masliah E., Mallory M., Hansen L., et al. Synaptic and neuritic alterations during the progression of Alzheimer’s disease. Neurosci Lett. 1994;174:67–72. [PubMed: 7970158]
175.
Dougherty M. K., Morrison D. K. Unlocking the code of 14-3-3. J Cell Sci. 2004;117:1875–84. [PubMed: 15090593]
176.
Takahashi Y. The 14-3-3 proteins: Gene, gene expression, and function. Neurochem Res. 2003;28:1265–73. [PubMed: 12834267]
177.
Sugimori K., Kobayashi K., Kitamura T., Sudo S., Koshino Y. 14-3-3 protein beta isoform is associated with 3-repeat tau neurofibrillary tangles in Alzheimer’s disease. Psychiatry Clin Neurosci. 2007;61:159–67. [PubMed: 17362433]
178.
Frautschy S. A., Baird A., Cole G. M. Effects of injected Alzheimer betaamyloid cores in rat brain. Proc Natl Acad Sci U S A. 1991;88:8362–6. [PMC free article: PMC52508] [PubMed: 1924295]
179.
Layfield R., Fergusson J., Aitken A., et al. Neurofibrillary tangles of Alzheimer’s disease brains contain 14-3-3 proteins. Neurosci Lett. 1996;209:57–60. [PubMed: 8734909]
180.
Burkhard P. R., Sanchez J. C., Landis T., Hochstrasser D. F. CSF detection of the 14-3-3 protein in unselected patients with dementia. Neurology. 2001;56:1528–33. [PubMed: 11402110]
181.
Poon H. F., Shepherd H. M., Reed T. T., et al. Proteomics analysis provides insight into caloric restriction mediated oxidation and expression of brain proteins associated with age-related impaired cellular processes: Mitochondrial dysfunction, glutamate dysregulation and impaired protein synthesis. Neurobiol Aging. 2006;27:1020–34. [PubMed: 15996793]
182.
Boyd-Kimball D., Sultana R., Poon H. F., et al. 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;132:313–24. [PubMed: 15802185]
183.
Agarwal-Mawal A., Qureshi H.Y., Cafferty P. W., et al. 14-3-3 connects glycogen synthase kinase-3 beta to tau within a brain microtubule-associated tau phosphorylation complex. J Biol Chem. 2003;278:12722–8. [PubMed: 12551948]
184.
Hashiguchi M., Sobue K., Paudel H. K. 14-3-3 zeta is an effector of tau protein phosphorylation. J Biol Chem. 2000;275:25247–54. [PubMed: 10840038]
185.
Hernandez F., Cuadros R., Avila J. Zeta 14-3-3 protein favours the formation of human tau fibrillar polymers. Neurosci Lett. 2004;357:143–6. [PubMed: 15036595]
186.
Ojika K., Tsugu Y., Mitake S., Otsuka Y., Katada E. NMDA receptor activation enhances the release of a cholinergic differentiation peptide (HCNP) from hippocampal neurons in vitro. Brain Res Dev Brain Res. 1998;106:173–80. [PubMed: 9555001]
187.
Jouvenceau A., Dutar P., Billard J. M. Alteration of NMDA receptormediated synaptic responses in CA1 area of the aged rat hippocampus: contribution of GABAergic and cholinergic deficits. Hippocampus. 1998;8:627–37. [PubMed: 9882020]
188.
Davies P., Terry R. D. Cortical somatostatin-like immunoreactivity in cases of Alzheimer’s disease and senile dementia of the Alzheimer type. Neurobiol Aging. 1981;2:9–14. [PubMed: 6115327]
189.
Davis B. M., Mohs R. C., Greenwald B. S., et al. Clinical studies of the cholinergic deficit in Alzheimer’s disease. I. Neurochemical and neuroendocrine studies. J Am Geriatr Soc. 1985;33:741–8. [PubMed: 2414354]
190.
Perry E. K., Perry R. H., Smith C. J., et al. Cholinergic receptors in cognitive disorders. Can J Neurol Sci. 1986;13:521–7. [PubMed: 3791066]
191.
Rossor M. N., Iversen L. L., Johnson A. J., Mountjoy C. Q., Roth M. Cholinergic deficit in frontal cerebral cortex in Alzheimer’s disease is age dependent. Lancet. 1981;2:1422. [PubMed: 6118790]
192.
Ling M., Merante F., Chen H. S., et al. The human mitochondrial elongation factor tu (EF-Tu) gene: cDNA sequence, genomic localization, genomic structure, and identification of a pseudogene. Gene. 1997;197:325–36. [PubMed: 9332382]
193.
Orrenius S., Burgess D. H., Hampton M. B., Zhivotovsky B. Mitochondria as the focus of apoptosis research. Cell Death Differ. 1997;4:427–8. [PubMed: 16465262]
194.
Vayssiere J. L., Cordeau-Lossouarn L., Larcher J. C., et al. Participation of the mitochondrial genome in the differentiation of neuroblastoma cells. In Vitro Cell Dev Biol. 1992;28A:763–72. [PubMed: 1483966]
195.
Pestova T. V., Hellen C. U. The structure and function of initiation factors in eukaryotic protein synthesis. Cell Mol Life Sci. 2000;57:651–74. [PubMed: 11130464]
196.
Thompson J. E., Hopkins M. T., Taylor C., Wang T. W. Regulation of senescence by eukaryotic translation initiation factor 5A: Implications for plant growth and development. Trends Plant Sci. 2004;9:174–9. [PubMed: 15063867]
197.
Tome M. E., Fiser S. M., Payne C. M., Gerner E. W. Excess putrescine accumulation inhibits the formation of modified eukaryotic initiation factor 5A (eIF-5A) and induces apoptosis. Biochem J. 1997;328:847–54. (Pt 3) [PMC free article: PMC1218996] [PubMed: 9396730]
198.
Chang R. C., Wong A. K., Ng H. K., Hugon J. Phosphorylation of eukaryotic initiation factor-2 alpha (eIF2α) is associated with neuronal degeneration in Alzheimer’s disease. Neuroreport. 2002;13:2429–32. [PubMed: 12499843]
199.
Ferrer I. Differential expression of phosphorylated translation initiation factor 2 alpha in Alzheimer’s disease and Creutzfeldt-Jakob’s disease. Neuropathol Appl Neurobiol. 2002;28:441–51. [PubMed: 12445160]
200.
Li X., An W. L., Alafuzoff I., et al. Phosphorylated eukaryotic translation factor 4E is elevated in Alzheimer brain. Neuroreport. 2004;15:2237–40. [PubMed: 15371741]
201.
Sajdel-Sulkowska E. M., Marotta C.A. Alzheimer’s disease brain: Alterations in RNA levels and in a ribonuclease-inhibitor complex. Science. 1984;225:947–9. [PubMed: 6206567]
202.
Ding Q., Markesbery W. R., Cecarini V., Keller J. N. Decreased RNA, and increased RNA oxidation, in ribosomes from early Alzheimer’s disease. Neurochem Res. 2006;31:705–10. [PubMed: 16770743]
203.
Sultana R., Poon H. F., Cai J., et al. Identification of nitrated proteins in Alzheimer’s disease brain using a redox proteomics approach. Neurobiol Dis. 2006;22:76–87. [PubMed: 16378731]
Copyright © 2010 by Taylor and Francis Group, LLC.
Bookshelf ID: NBK56020PMID: 21882450

Views

  • PubReader
  • Print View
  • Cite this Page

Other titles in this collection

Related information

  • PMC
    PubMed Central citations
  • PubMed
    Links to PubMed

Similar articles in PubMed

See reviews...See all...

Recent Activity

Your browsing activity is empty.

Activity recording is turned off.

Turn recording back on

See more...