Skip to main content
Aging and Disease logoLink to Aging and Disease
. 2013 Nov 30;5(4):238–255. doi: 10.14336/AD.2014.0500238

Metabolic Disturbances in Diseases with Neurological Involvement

João M N Duarte 1, Patrícia F Schuck 2, Gary L Wenk 3, Gustavo C Ferreira 2,*
PMCID: PMC4113514  PMID: 25110608

Abstract

Degeneration of specific neuronal populations and progressive nervous system dysfunction characterize neurodegenerative diseases, including Alzheimer’s disease and Parkinson’s disease. These findings are also reported in inherited diseases such as phenylketonuria and glutaric aciduria type I. The involvement of mitochondrial dysfunction in these diseases was reported, elicited by genetic alterations, exogenous toxins or buildup of toxic metabolites. In this review we shall discuss some metabolic alterations related to the pathophysiology of diseases with neurological involvement and aging process. These findings may help identifying early disease biomarkers and lead to more effective therapies to improve the quality of life of the patients affected by these devastating illnesses.

Keywords: neurodegenerative diseases, inherited diseases, brain metabolism


Neurodegenerative disorders are progressive diseases characterized by the loss of specific neuronal populations resulting in different clinical phenotypes. These diseases are also recognized as conformational diseases or disorders of protein aggregation or catabolism and are classified according to the main pathophysiological processes involved. Examples include Alzheimer’s disease (AD), Parkinson’s disease (PD), and others [1]. In most of these disorders, the cause of protein aggregation remains unknown. However, it might be attributed to autophagic dysfunction, since the autophagy consists of a proteolytic system in which cytosolic components are degraded in lysosomes and is pivotal to cellular survival [2].

The main risk factor for neurodegenerative diseases is aging. Many of these diseases share chronic pro-inflammatory environment as a common feature. The trigger of chronic inflammation in most of cases remains unclear, although it may be the interplay between innate and adaptive responses against certain stimuli, such as aggregated proteins, heat shock proteins, and other local danger-associated molecular patterns that may contribute to a final pathogenic pathway [3]. The selective expression of mutant proteins associated with neurodegenerative diseases in astrocytes [4], along with microglia [5], may contribute to the chronic neuroinflammation.

Mitochondrial alterations also play a role in these diseases, elicited by either genetic alterations (mitochondrial or nuclear DNA) or exogenous toxins affecting mitochondrial functioning [6]. In this scenario, mitochondrial DNA mutations cause enzymatic defects of the respiratory chain complexes, which render cells devoid of energy; these changes tend to affect systems with high energy demands, including brain, skeletal muscle and heart [7,8]. Furthermore, disruption of key steps of energy metabolism was previously shown in neurodegenerative diseases, particularly for respiratory chain complexes [913]. It should, however, emphasized that the mitochondrial involvement linked to the pathophysiology of neurodegenderative diseases is not restricted to defects in energy metabolism and DNA mutations, but also oxidative stress [1416] and alterations in mitochondrial shape and distribution [1719], as well as metabolic communication with other organelles [20].

Neurodegeneration and mitochondrial dysfunction are also found in many genetic disorders, including inborn errors of metabolism such as glutaric aciduria type I (GA I), phenylketonuria (PKU), methylmalonic acidemia (MMAemia), and Niemann-Pick type C (NPC) disease. Data from patients and animal models evidenced that energy dysfunction may play a role in the pathophysiology of the characteristic brain damage found in these diseases. The evidence points out bioenergetic impairments as one of the most important biochemical events that lead to neurodegeneration in these metabolic disorders.

This review will discuss some metabolic alterations related to the pathophysiology of certain diseases with neurological involvement and aging process.

Bioenergetic alterations in inherited metabolic diseases with CNS involvement

Glutaric acidemia type I

GA I is a severe disease caused by the inherited deficiency of glutaryl-CoA dehydrogenase activity, a mitochondrial enzyme that participates in the lysine and tryptophan catabolism pathways [21]. Patients affected by GA I present increased levels of glutaric (GA), 3-hydroxyglutaric (3HGA) and trans-glutaconic (tGA) acids in tissues and body fluids [21,22]. The main clinical presentation includes progressive macrocephaly and striatal necrosis (medium-spiny neurons) following episodes of metabolic decompensation, basal ganglia degeneration, frontotemporal hypoplasia, and delayed myelination/hypomyelination [2324]. Currently, excitotoxicity, oxidative stress and alterations of bioenergetics are suggested to play a role in the pathophysiology of the characteristic neurodegeneration observed in GA I patients.

Patients affected by this disease excrete increased levels of lactate and dicarboxylic acids, indicating a possible role of mitochondrial dysfunction in GA I patients [2526]. Evidence from animal models of GA I demonstrate that brain energy metabolism is severely affected. Koeller and colleagues [27] showed alterations of the expression of genes involved in mitochondrial energy metabolism and transport in cerebral cortex from a knock out mouse model. Many studies reported that GA, and 3-HGA elicit bioenergetics dysfunction both in vivo and in vitro, while tGA did not affect mitochondrial metabolism [28]. GA and 3HGA showed an inhibitory effect on complexes of mitochondrial respiratory chain in rat brain and chick embryonic neurons [2931]. This impairment of mitochondrial function causes energy failure in the brain, which may lead to secondary excitotoxicity and oxidative stress [3235]. These organic acids were also shown to decrease cellular phosphocreatine levels and creatine kinase (CK) activity, as well as glucose and pyruvate oxidation in cerebral cortex, striatum, midbrain and hippocampus in rats [29,31,36,37], similar to Alzheimer’s, Parkinson’s and Huntington’s diseases brain areas affected (see bellow), reinforcing the idea that energy metabolism impairment may be an important mechanism involved in the characteristic striatal degeneration found in GA I patients.

Phenylketonuria

PKU is a genetic disease caused by an abnormal phenylalanine (Phe) metabolism due to a phenylalanine hydroxylase deficiency. This enzymatic deficiency leads to increased levels of Phe, phenyllactate (PLA), phenylpyruvate (PPA), and phenylacetate (PAA) in tissues, plasma and urine of affected patients [38]. Severe neurological impairment is the main characteristic of phenylketonuric patients, which appears to be a direct consequence of neuronal cell loss, white matter abnormalities, and reduction of synaptic density [3841].

Many efforts have been made in order to identify the underlying mechanisms of the pathophysiology of the severe brain damage found in phenylketonuric patients. Studies in patients and animal models suggest that the accumulating metabolites exert neurotoxic effects, particularly Phe. Phe and its metabolites cause DNA damage in patient blood and brain of animals submitted to an animal model of PKU [42,43] and cell apoptosis in cultured neurons [4446]. It has been suggested that one of the main mechanisms responsible for this neuronal damage may be an energy metabolism impairment. In this context, Wasserstein and colleagues [47] demonstrated by positron emission tomography (PET) that phenylketonuric patients had decreased relative glucose metabolic rates in cortical regions. On the other hand, the same authors showed an increased activity in subcortical regions including the striatum and limbic system. Moreover, a MRS study in vivo with adult patients demonstrated subtle abnormalities of cerebral energy metabolism in PKU steady-state conditions, such as increased ADP and decreased inorganic phosphate content. Furthermore, these abnormalities were accentuated by an increased dietary Phe load [48].

Studies using animal models also described deleterious effects of Phe, PLA, PPA and PAA on important markers of bioenergetics activity. Costabeber and colleagues [49] demonstrated that Phe significantly inhibited CK activity in vitro and in vivo in brain of young rats. In addition, Phe decreased pyruvate kinase (PK) activity, glucose utilization, and lactate release, and increased ADP brain levels [50,51]. Moreover, chronic Phe administration inhibited respiratory chain complexes I-III and succinate dehydrogenase (SDH) activities in cerebral cortex of rats [52]. Taken together, these studies suggest that Phe is the main toxic metabolite in PKU and that bioenergetics imbalance contributes to its toxicity.

Methylmalonic Acidemia

MMAemia is an organic acidemia caused by the deficiency of L-methylmalonil-CoA mutase activity, a mitochondrial enzyme involved in the catabolism of the amino acids isoleucine, valine, methionine and threonine, as well as thymine, odd carbon number fatty acids, and cholesterol [53]. The lack of L-methylmalonil-CoA mutase activity increases methylmalonic acid (MMA) levels in tissues and body fluids of patients, especially in the brain. Patients affected by MMA present predominantly with neurological signs and symptoms, with marked bilateral destruction of the globus pallidus [5355].

Observations in MMAemia-affected patients and animal models of the disease suggest that mitochondrial dysfunction is an important process participating in the neurodegeneration found in this condition. In this scenario, lactic acid levels are elevated in the globus pallidus, cerebrospinal fluid (CSF), and urine of patients [56]; citric acid cycle intermediates were also elevated in the blood [53], suggesting impairment of mitochondrial metabolism. Indeed, mitochondrial abnormalities and inhibition of respiratory chain complexes were observed in various tissues from MMAemia-affected patients [5760].

Studies in vitro demonstrated that MMA inhibited succinate-supported oxygen consumption in intact isolated mitochondria by inhibiting mitochondrial succinate transport through dicarboxylate carrier [61]. In the same way, malate transport by the same transporter is also impaired by MMA [62]. Furthermore, important enzyme activities involved in energy metabolism are affected by metabolites accumulating in MMAemia, including respiratory chain complexes in brain homogenates [63,64]. Utter and colleagues [65] observed that MMA is a competitive inhibitor of pyruvate carboxylase activity. This organic acid also inhibits brain mitochondrial utilization of ketone bodies and lactate through competitive inhibition of β-hydroxybutyrate [66,67] and lactate dehydrogenases [68]. Moreover, Schuck and colleagues [69] reported decreased CK activity by MMA in rat brain homogenates. In addition, MMA blocks aerobic glucose oxidation due to a reduction of CO2 production from [2-14C] glucose and [U-14C] acetate and increased lactate synthesis in rat brain slices [70]. MMA also induces cell death by inhibiting mitochondrial permeability transition pore [7172], a structure formed by proteins as ATP synthase and adenine nucleotide traslocase and whose activity control cell death by apoptosis and superoxide generation. Furthermore, Sauer and colleagues [73] demonstrated that methylmalonyl-CoA (precursor of MMA) decreased pyruvate and α-ketoglutarate dehydrogenase activities.

These experimental evidences suggest that MMA, the main metabolite accumulating in MMAemia, impairs brain bioenergetics and this impairment may collaborate to neurodegeneration found in MMAemia patients.

Niemann-Pick type C disease (NPC)

NPC is a fatal lipid storage disorder characterized by cholesterol accumulation in endosomal/ lysosomal compartments [74]. The disease is caused by mutations in the Npc1 or Npc2 genes [75]. Clinical findings include difficulty with speaking and swallowing, cerebellar ataxia, and progressive dementia [74,7678]. Similarly to other neurodegenerative diseases, NPC presents altered metabolism of amyloid precursor protein [79] and levels of total-tau (T-tau) in cerebrospinal fluid of patient [80], toxic proteins involved in axonal degeneration, oxidative stress induction and bioenergetics impairment. Furthermore, extensive and progressive neuronal death is found, especially in the cerebellum [81].

Evidence obtained from serum of NPC patients and animal models demonstrated diminished antioxidant defenses and Coenzyme Q10 concentrations, an important component of mitochondrial respiratory chain and mitochondrial glutathione content depletion, suggesting a role for oxidative stress in the pathophysiology of NPC [82,83]. In this context, Yu and colleagues [84] demonstrated that mitochondrial metabolism impairment and subsequent ATP deficiency might be responsible for the neuronal impairment in NPC disease. The same authors also showed that the brains of NPC animal model contained smaller mitochondria, decreased ATP levels and higher cholesterol content in mitochondria [8486]. Considering the link between oxidative stress and mitochondrial dysfunction, it is tempting to speculate that bioenergetics impairment could also participate in the neurodegeneration found in NPC patients.

Metabolic alterations in aging and neurodegenerative diseases

Aging

According to the Free Radical Theory to Aging, age-related neurodegenerative diseases are mainly attributed to oxidation of cellular components by free radicals, catalyzed by oxidative enzymes and traces of metal ions [87].

Bioenergetic changes, mainly mitochondria-related alterations, comprise a more recent approach that may help to understand the aging process. Mitochondria are the organelles responsible by ATP generation, Ca2+ uptake and storage, and generation and detoxification of reactive oxygen species (ROS). Such functions are influenced by the mitochondrial membrane potential (Δψm), while many apoptotic signals are triggered by cytochrome c release into the cytoplasm [88]. On the basis of putative ROS generation by mitochondria, the mitochondrial theory of aging was developed. This theory considers somatic mutations of mitochondrial DNA (mtDNA) induced by reactive species as the primary cause of age-related energy dysfunction. The complex I is thought to be mostly affected, becoming strongly rate limiting for electron transfer [89], which leads to impaired energy metabolism.

The net ROS production by mitochondria may be the major source of oxidative damage that accumulate with aging and impair multiple mitochondrial functions. The contributing factors appear to include the intrinsic rate of proton leakage across the inner mitochondrial membrane, decreased membrane fluidity, and decreased levels and function of cardiolipin, a compound found mainly in mitochondria that supports the function of many of the proteins of the inner mitochondrial membrane. Furthermore, the enzyme activity of mitochondrial complexes I, III, and IV are decreased with aging [90], as well as mitochondrial membrane potential, respiratory control ratio and cellular O2 uptake [9192].

Certain nutritional or behavioural measures may improve brain mitochondrial function and reduce ROS production. Indeed, it was shown that calorie restriction may be an effective intervention to delay age-related cognitive decline and diseases in mammals [9394]. Calorie restriction increases cerebral mitochondrial respiratory capacity [95] and prevents age-related deficit in long-term potentiation in the hippocampus [96,97]. Moreover, a positive relation between physical activity and cognitive performance has been reported in several studies, and may reduce the risk of elderly-associated neurodegenerative disorders [reviewed in 98,99]. Physical activity was shown to increase the number of mitochondria and dendritic spine synapses in the hippocampus [100], to stimulate neurogenesis and to enhance long-term potentiation in the hippocampus [101105], which results in improved learning and memory in aging [106].

MRS studies in elderly subjects were performed at low magnetic field and focused on the most prominent resonances in 1H MR spectra, i.e. N-acetylaspartate (NAA), myo-inositol, glutamate plus glutamine (overlapped peaks at low MR field), choline and creatine [e.g. 107109]. In experimental animal models, the extended neurochemical profile detectable in vivo by high-field 1H MRS has been extensively investigated from birth to adulthood. Recently, we studied the evolution of the neurochemical profile in the mouse brain from 3 to 24 months of age in a longitudinal manner [110]. The most important findings were that aging induces general modifications of neurotransmission processes (reduced GABA and glutamate), primary energy metabolism (altered glucose, alanine and lactate) and turnover of lipid membranes (modification of choline-containing compounds and phosphorylethanolamine).

The age-induced reduction in glutamate and GABA in 2-year old mice compared to young adults [110] may be associated with the loss of synaptic efficiency. Decreased density of synaptic proteins was observed in the aged rodent brain, namely SNARE complex proteins, vesicle-mobilizing proteins, vesicular neurotransmitter transporters, post-synaptic proteins and components of the synaptic vesicle cycle [e.g. 111,112], which can contribute to brain dysfunction and decline in hippocampal-dependent memory performance [e.g. 113116]. Lower concentrations of neurotransmitters were accompanied by higher levels of myo-inositol [110] that were often related to astrogliosis in Alzheimer’s disease [discussed in 117]. These age-dependent modifications in the mouse brain resemble those in humans, namely elderly subjects displayed reduction in glutamate and increase in myo-inositol concentrations in the brain, compared to young subjects [118]. Reduced neurotransmission imposes lower energy demand and, in fact, aging is associated with reduced cerebral glucose utilization [e.g. 119120] and reduced neuronal and glial mitochondrial metabolism [118]. Lactate levels were found to decrease with age in the mouse brain [110], which suggests modification of the relative rates of cytosolic glycolysis and mitochondrial oxidative metabolism.

In the mouse brain, concentrations of phosphorylethanolamine and choline-containing phospholipids (glycerylphosphorylcholine and phosphorylcholine) were negatively and positively associated with age, respectively, likely reflecting altered membrane lipid turnover [110,121]. Altered choline levels are also patent in the human brain [reviewed in 117]. Moreover, a 31P NMR spectroscopy study showed that the glycerylated forms of such phospholipids increase with age in the human brain [122]. The inverse trajectory in the evolution of phosphorylcholine and phosphorylethanolamine levels and the relative increase in glycerated phospholipids may indeed be related to modifications in cell membrane fluidity in the aging brain [123].

Brain taurine levels were found to be highest after birth and decay during early development until adulthood [121]. During aging, while taurine content is relatively stable in the hippocampus or cortex after reaching adulthood, it continues to decrease in the striatum, where it is most concentrated [110]. Taurine is a neuromodulator and interacts with inhibitory GABAA, GABAB or glycine receptors thus displaying capability of modulating synaptic plasticity [reviewed in 124]. This inhibitory role of taurine is especially important during the period of cortical synaptogenesis, when brain taurine levels are found to be highest [121,125]. However, as only a small fraction of taurine is involved in neuromodulation, the main role of taurine may be rather related to osmoregulation, balancing modifications that occur in the concentrations of other major neurochemicals [117].

Parkinson’s Disease

Parkinson’s disease (PD) is a disorder of the central nervous system (CNS) primarily associated with a decrease in dopamine levels in the striatum caused by a degeneration of dopaminergic neurons in the substantia nigra [126]. It has been suggested that mitochondrial dysfunction is involved in the pathogenesis of the degeneration of dopaminergic neurons, particularly a decrease in NADH dehydrogenase (respiratory chain complex I) in the substantia nigra of PD patients in postmortem studies [127]. Furthermore, alterations in complex I activity may also be found in other brain regions of PD patients [128130] and also in peripheral tissues [131133]. It was also demonstrated that the administration of complex I inhibitors, including the pesticide rotenone and 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), triggers neurological changes similar to PD in rats and that genes associated with familial PD are involved in mitochondrial functioning [134135]. Finally, it was demonstrated in tissues of PD patients or animal models of PD other alterations of energy metabolism than complex I deficiency, including impaired glycolysis, oxidative decarboxylation of pyruvate and other components of oxidative phosphorylation system [136140].

Decreased NAA to creatine ratio has been extensively observed in different brain areas of patients with PD [reviewed in 117]. Other studies found that PD leads to a reduction in the signal corresponding to choline-containing compounds in cortical areas [141,142] and posterior cingulate [143], and a reduction in glutamate to creatine ratio in the anterior cingulate gyrus [144]. This later study however failed to identify the typical PD-induced reduction in NAA levels. Since mitochondrial dysfunction also contributes to neuronal degeneration in PD [145], a reduction of high-energy phosphates and increased lactate concentration was observed in several brain areas of PD patients when compared to healthy subjects [146,147]. A recent study at high magnetic field found increased GABA levels in the pons and putamen of early PD patients relative to healthy subjects, in the absence of other neurochemical modifications [148]. In this study, the authors further proposed a mechanism for the involvement of GABAergic activity in pontine-nigral-striatal pathways that could be important for developing PD. Most importantly, the detection of increased GABA levels in these areas must be regarded as a putative early biomarker of PD, before substantial degeneration occurs.

Pre-clinical MRS studies of PD have been largely limited to animal models administered with the neurotoxin MPTP or its active metabolite 1-methyl-4-phenylpyridinium (MPP+), leading to destruction of dopaminergic neurons in the substantia nigra, and triggering PD symptoms [147]. After MPTP/MPP+ intoxication, reduced NAA and increased lactate levels were observed in the striatum or substantia nigra of rodents [149152] and other mammals [reviewed in 117], consistent with neuronal loss and mitochondrial dysfunction. Rats injected with 6-hydroxydopamine and ascorbate in the substantia nigra display phenotypes of PD and reduced NAA to creatine ratio in the frontal cortex, which is associated with synaptic degeneration and dopaminergic loss [153]. GABA was suggested to be either increased [154] or reduced [152] after MPTP/MPP+ treatment. Increased concentration of both glutamate and glutamine was proposed in the striatum of animal models of PD compared to controls [154156], as consequence of excitotoxicity and/or impaired glutamatergic neurotransmission. These neurochemical modifications were suggested to be restricted to the motor part of the striatum and absent in the nucleus accumbens [157].

Alzheimer’s Disease

Alzheimer’s disease (AD) is a progressive, severe aging-related neurodegenerative disease and the most common form of dementia. This neurodegenerative disorder is characterized by extracellular deposition of amyloid aggregates and accumulation of neurofibrillary tangles of hyperphosphorylated tau protein, leading to loss of synapses and neuronal death [158]. AD pathophysiology is not yet completely understood, however mitochondrial dysfunction and oxidative stress are implied in it.

Mitochondrial genome mutations and deletions that impair mitochondrial function are associated with AD [159]. In this context, alterations in the expression of mitochondrial fission and fusion proteins were described, indicating an impairment of fusion fission process [160]. Furthermore, precursor amyloid protein (APP) interferes with important mitochondrial proteins, impairing cell bioenergetics [161]. Studies from animal models and AD patients showed that amyloid protein inhibits respiratory complexes (especially complex IV) and α-ketoglutarate dehydrogenase activities [162,163]. In addition, amyloid protein fragments decreased mitochondrial mass in the neurites of cultured neurons [164].

Mitochondrial dysfunction may also be related to AD pathophysiology due to promotion of tau phosphorylation, participating in the process of tangle formation [165]. Moreover, ATP depletion inhibits the transporter superfamily for toxic peptides (ABC transporters), including β-amyloid, which is highly dependent on ATP availability. In fact, it was recently demonstrated that the molecular alterations are pronounced near the β-amyloid plaques [166]. In this scenario, inhibition of ABC transporters impairs the clearance mechanisms for the removal of toxic peptides and its aggregates, contributing to the progression of AD [167].

The brains of patients with AD have been consistently reported to display reduced NAA and increased myo-inositol concentrations compared to cognitively healthy elderly individuals and, furthermore, these findings were often correlated with decline in cognitive performance [reviewed in 117]. While NAA indicates loss of neuronal integrity, myo-inositol has been interpreted as marker of astrocyte reactivity. However, the majority of these clinical MRS studies did not provide absolute quantification of neurochemicals but determined ratios of NAA and myo-inositol to other peaks in the 1H MR spectra, namely choline and creatine. Since some studies found altered choline [168171] or creatine [172,173] levels in certain brain areas of patients with AD compared to controls, care should be taken when interpreting metabolite ratios. Levels of choline-containing compounds are associated to phospholipid turnover and, indeed, increased glycerylphosphorylcholine concentration in post-mortem cortical gray matter measured from severely demented AD brains have been attributed to membrane breakdown [174].

The concentration of NAA in brain areas of patients with mild cognitive impairment was found to be higher than that in the brains of AD patients but also lower than that of healthy subjects [e.g. 173,175,176]. Furthermore, NAA decrease has been suggested to correlate with cognitive decline from light dementia to AD [e.g. 172,177,178]. These findings suggest a positive correlation between NAA levels and cognitive performance. Thus, in vivo 1H MRS has given the opportunity to assess metabolic and functional correlates of dementia in research and clinical settings, as well as prediction of future cognitive decline in AD [176].

Patients with AD display impaired cerebral energy metabolism [reviewed in 13]. Accordingly, in addition to these neurochemical markers, after glucose administration, Haley and colleagues [179] measured glucose levels in the brain of AD patients at higher levels than in healthy subjects. These increased brain glucose levels thus likely reflect reduced glucose metabolism, linked to impaired insulin signaling [180].

Mutations in genes associated with β-amyloid precursor protein (APP), presenillin 1 (PS1) and presenillin 2 (PS2) are involved in early-onset familial AD [181]. Mouse models of AD have thus been generated by inserting one or more of these human mutations into the mouse genome and are characterized by β-amyloid deposition in the brain. Double transgenic mice expressing human mutant APP and human mutated PS1 (APP/PS1) display decreased glutamate and NAA concentrations in the brain, particularly in the hippocampus, that were observed at the time where hippocampal volume was already reduced but amyloid plaques were not yet present [182184]. Later, upon visible amyloid deposits, these mice also presented increased brain myo-inositol to creatine ratio, when compared to wild type mice [185], accompanied by astrogliosis [186]. Increased myo-inositol and glutamine and decreased NAA and glutamate were reported to occur in the cortex of these mice. Additionally, it was found that once amyloid plaques are observed, NAA levels are inversely associated to the area of cortex occupied by plaques [182]. Dedeoglu and colleagues [187], but not Marjanska and colleagues [183], reproduced these findings in mice possessing only the mutated APP gene. Both authors reported, however, higher cerebral taurine to creatine ratio in the brain of APP transgenic mice compared to wild type mice. These results were confirmed by in vitro NMR spectroscopy of brain extracts, in which glutathione levels were additionally found to be lower than in controls [187].

Recently, high resolution 1H and 31P MRS at 14.1 T was used to investigate the brain neurochemistry of the 5xFAD transgenic mouse model of AD [188], which over-expresses APP with 3 different familial mutations and PS1 also with two mutations, resulting in early robust β-amyloid deposition. Compared to wild-type littermates, the hippocampus of 5xFAD mice was characterized by increased myo-inositol and decreased NAA and, in addition, decreased concentrations of γ-aminobutyrate (GABA). Furthermore, 5xFAD mice displayed lower brain glucose content associated with higher lactate levels, suggesting impaired energy metabolism, as occurs in AD patients [179]. However, the rate of creatine kinase and the relative concentrations of phosphorus-containing metabolites were not altered in this AD model [188].

In summary, compared to healthy subjects, AD patients display lower NAA and increased myo-inositol in the brain, reflecting disease progression. These observations are also mimicked in animal models of AD. However, contradictory results have appeared regarding some metabolic observations and their regional distribution.

Huntington’s Disease

Huntington’s disease (HD) is an inherited neurodegenerative disease caused by alterations in the polyglutamine region of the huntingtin protein (Htt), leading to the accumulation of intracellular Htt aggregates, affecting especially GABAergic medium-sized spiny neurons in the striatum [189]. Even though Htt neurotoxicity has been widely studied, the pathophysiological mechanisms of HD have not yet been fully understood.

Mitochondrial dysfunction contributes to the pathogenesis of HD [145,190,191]. Indeed, respiratory chain complex II activity inhibitors administration such as malonate and 3-nitropropionic acid mimics striatal neuronal degeneration found in HD patients [192,193]. Patients and animal models develop several mitochondrial defects, ranging from altered calcium metabolism [194], impaired bioenergetics [195], increased oxidative stress and abnormal trafficking [196] and mitochondrial dynamics [197,198]. Furthermore, 1H MRS of basal ganglia and cerebral cortex of HD patients showed elevated lactate production, suggesting defects in mitochondrial pyruvate utilization [199]. Additional studies revealed inhibitions of respiratory chain complexes II, III and IV and Krebs cycle enzymes activities in basal ganglia [200202].

Either chemically induced by 3-nitropropionic or malonate administration or mutant Htt animal models also demonstrated a role for oxidative stress related to mitochondrial dysfunction. In this context, mutant cells presented increased mitochondrial ROS generation and decreased NADPH oxidase and xanthine oxidase activities [203], indicating a redox imbalance, leading to mitochondrial-related apoptosis and excitotoxicity [204].

In addition, data from the literature demonstrated the involvement of impaired function of peroxisome proliferator-activated receptor (PPAR)-γ coactivator-1α (PGC-1α), a transcriptional master co-regulator of mitochondrial biogenesis, metabolism, and antioxidant defenses, in mitochondrial dysfunction characteristic of HD [205]. In this context, it has been reported that mutant Htt increases Drp1 (an important protein implicated in mitochondrial fission) activity, suggesting that Htt may contribute to mitochondrial fragmentation and dysfunction found in HD patients and animal models [198]. These alterations in mitochondria morphology may enhance cellular susceptibility to apoptosis in HD [197,198].

Reduced NAA and increased myo-inositol are known to occur in the brain of HD patients, relative to healthy subjects. In particular, levels of NAA and myo-Inositol in the putamen were found to be correlated with measures of HD severity [206]. Association was also found between cognitive dysfunction and both NAA and glutamate levels in the posterior cingulate [207].

Different transgenic mouse models with mutated huntingtin have been used to investigate the mechanisms of HD. In the R6/2 transgenic mouse model of HD, striatal concentrations of creatine, glycerophosphorylcholine, glutamine and glutathione were found to be increased and NAA levels decreased [208,209]. Further development of the disease leads to additional modifications, namely increased concentrations of phosphocreatine, taurine, ascorbate, glutamate, lactate and myo-inositol and decreased phosphorylethanolamine [208,210]. From all these alterations, reduced NAA concentration was consistently found to correlate with neuron dysfunction and to be accentuated by the length of polyglutamine expansions and the progression of the neuropathological phenotype [211]. Interestingly, these neurochemical modifications in the R6/2 mouse were found not only in the striatum but also in cortical areas, and preceded significant shrinkage of cortical and striatal volume, as measured in vivo from MR images [210].

In another model of HD, the zQ175 knock-in mouse, concentrations of the neurotransmitters glutamate and GABA were found reduced in the striatum between 4 and 8 months of age, but recover to control levels at one year of age [212]. At this age, glutamine, creatine and taurine were also increased relative to wild-type mice. Levels of NAA in the striatum were always lower in zQ175 mice, probably reflecting the reduced number and density of neurons that also resulted in reduced striatum volume [212].

Glutamine, glutamate and GABA alterations suggest the impaired neurotransmission and glutamate/GABA-glutamine cycles, while increased creatine and phosphocreatine are indicative of impaired mitochondrial bioenergetics and reduced oxidative phosphorylation [191]. Changes in membrane properties of medium-sized spiny neurons that compromise integrity of the membrane potential could be the cause for the observed alterations in phospholipid levels [213]. Increased GSH, ascorbate and taurine were suggested to be a counter-regulatory response to HD associated oxidative stress [214].

Inhibitors of the complex II of mitochondrial respiratory chain, namely 3-nitropropionic acid and malonate, effectively trigger behavioral changes and selective striatal lesions in animals, thus mimicking symptoms of HD. In fact, like in the striatum of R6/2 or zQ175 transgenic mice, rats treated with 3-nitropropionic acid display reduced striatal NAA levels [reviewed in 215]. In this case, the mitochondrial dysfunction caused by chemical inhibition of the respiratory chain leads to increased lactate concentration [e.g. 151,199,215,216]. Table 1 summarizes the main metabolic modifications in aging and neurodegenerative disorders measured by MRS in affected brain areas of patients and/or animal models.

Table 1.

Summary of main metabolic modifications in aging and neurodegenerative disorders measured by MRS in affected brain areas of patients and/or animal models.

Metabolite Aging AD PD HD
NAA
myo-Inositol
Glutamate
Glutamine
GABA
Choline-containing phospholipids
Phosphorylethanolamine
Taurine
Lactate

NAA= N-acetylaspartate; GABA= γ-aminobutyrate; AD= Alzheimer’s disease; PD= Parkinson’s disease; HD= Huntington’s disease.

Direction of the arrow indicates either increase ( ) or decrease ( ) of metabolite concentration.

Summary

Extensive research efforts are underway aiming at unraveling the molecular basis of metabolic alterations in the pathophysiology of neurodegenerative diseases, both common (such as Alzheimer, Parkinson and Huntington diseases) and inherited metabolic diseases. The identification of the exact mechanisms involved in these diseases shall help to identify early disease biomarkers as well as more effective therapies (which may include energy compounds, antioxidant therapy particularly targeted to mitochondria and modulators of mitochondrial dynamics), preventing or slowing the progression of neurodegeneration and improving the quality of life of the patients affected by these devastating illnesses.

Acknowledgments

JMND was supported by Swiss National Science Foundation (grant PZ00P3_148250) and by the Centre d’ImagerieBioMedicale (CIBM) of the UNIL, UNIGE, HUG, CHUV, EPFL and the Leenaards and Jeantet Foundations. PFS and GCF are supported by the National Council for Scientific and Technological Development (CNPq). GLW is supported by U.S. Public Health Service, RO1 AG030331 & RO1 AG037320.

References

  • [1].Barnham KJ, Masters CL, Bush AI. Neurodegenerative diseases and oxidative stress. Nat Rev Drug Discov. 2004;3:205–14. doi: 10.1038/nrd1330. [DOI] [PubMed] [Google Scholar]
  • [2].Wong E, Cuervo AM. Autophagy gone awry in neurodegenerative diseases. Nat Neurosci. 2010;13:805–11. doi: 10.1038/nn.2575. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [3].Amor S, Puentes F, Baker D, van der Valk P. Inflammation in neurodegenerative diseases. Immunology. 2010;129:154–69. doi: 10.1111/j.1365-2567.2009.03225.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [4].Maragakis NJ, Rothstein JD. Mechanisms of Disease: astrocytes in neurodegenerative disease. Nat Clin Pract Neurol. 2006;2:679–89. doi: 10.1038/ncpneuro0355. [DOI] [PubMed] [Google Scholar]
  • [5].Perry VH, Nicoll JA, Holmes C. Microglia in neurodegenerative disease. Nat Rev Neurol. 2010;6:193–201. doi: 10.1038/nrneurol.2010.17. [DOI] [PubMed] [Google Scholar]
  • [6].Schapira AH. Mitochondrial dysfunction in neurodegenerative diseases. Neurochem Res. 2008;33:2502–9. doi: 10.1007/s11064-008-9855-x. [DOI] [PubMed] [Google Scholar]
  • [7].Kwong JQ, Beal MF, Manfredi G. The role of mitochondria in inherited neurodegenerative diseases. J Neurochem. 2006;97:1659–75. doi: 10.1111/j.1471-4159.2006.03990.x. [DOI] [PubMed] [Google Scholar]
  • [8].DiMauro S, Hirano M. Mitochondrial encephalomyopathies: an update. Neuromuscul Disord. 2005;15:276–86. doi: 10.1016/j.nmd.2004.12.008. [DOI] [PubMed] [Google Scholar]
  • [9].Tabrizi SJ, Cleeter MW, Xuereb J, Taanman JW, Cooper JM, Schapira AH. Biochemical abnormalities and excitotoxicity in Huntington’s disease brain. Ann Neurol. 1999;45:25–32. doi: 10.1002/1531-8249(199901)45:1<25::aid-art6>3.0.co;2-e. [DOI] [PubMed] [Google Scholar]
  • [10].Mann VM, Cooper JM, Javoy-Agid F, Agid Y, Jenner P, Schapira AH. Mitochondrial function and parental sex effect in Huntington’s disease. Lancet. 1990;336:749. doi: 10.1016/0140-6736(90)92242-a. [DOI] [PubMed] [Google Scholar]
  • [11].DiMauro S, De Vivo DC. Genetic heterogeneity in Leigh syndrome. Ann Neurol. 1996;40:5–7. doi: 10.1002/ana.410400104. [DOI] [PubMed] [Google Scholar]
  • [12].Cardoso SM, Proença MT, Santos S, Santana I, Oliveira CR. Cytochrome c oxidase is decreased in Alzheimer’s disease platelets. Neurobiol Aging. 2004;25:105–10. doi: 10.1016/s0197-4580(03)00033-2. [DOI] [PubMed] [Google Scholar]
  • [13].Ferreira IL, Resende R, Ferreiro E, Rego AC, Pereira CF. Multiple defects in energy metabolism in Alzheimer’s disease. Curr Drug Targets. 2010;11:1193–206. doi: 10.2174/1389450111007011193. [DOI] [PubMed] [Google Scholar]
  • [14].Caldeira GL, Ferreira IL, Rego AC. Impaired transcription in Alzheimer’s disease: key role in mitochondrial dysfunction and oxidative stress. J Alzheimers Dis. 2013;34:115–31. doi: 10.3233/JAD-121444. [DOI] [PubMed] [Google Scholar]
  • [15].Johnson WM, Wilson-Delfosse AL, Mieyal JJ. Dysregulation of glutathione homeostasis in neurodegenerative diseases. Nutrients. 2012;4(10):1399–440. doi: 10.3390/nu4101399. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [16].Fatokun AA, Stone TW, Smith RA. Oxidative stress in neurodegeneration and available means of protection. Front Biosci. 2008;13:3288–311. doi: 10.2741/2926. [DOI] [PubMed] [Google Scholar]
  • [17].Guedes-Dias P, Oliveira JM. Lysine deacetylases and mitochondrial dynamics in neurodegeneration. Biochim Biophys Acta. 2013;1832:1345–59. doi: 10.1016/j.bbadis.2013.04.005. [DOI] [PubMed] [Google Scholar]
  • [18].Itoh K, Nakamura K, Iijima M, Sesaki H. Mitochondrial dynamics in neurodegeneration. Trends Cell Biol. 2013;23:64–71. doi: 10.1016/j.tcb.2012.10.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [19].Park J, Choi H, Min JS, Park SJ, Kim JH, Park HJ, Kim B, Chae JI, Yim M, Lee DS. Mitochondrial dynamics modulate the expression of pro-inflammatory mediators in microglial cells. J Neurochem. 2013;127:221–32. doi: 10.1111/jnc.12361. [DOI] [PubMed] [Google Scholar]
  • [20].Schon EA, Przedborski S. Mitochondria: the next (neurode)generation. Neuron. 2011;70:1033–53. doi: 10.1016/j.neuron.2011.06.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [21].Goodman SI, Markey SP, Moe PG, Miles BS, Teng CC. Glutaric aciduria; a “new” disorder of amino acid metabolism. Biochem Med. 1975;12:12–21. doi: 10.1016/0006-2944(75)90091-5. [DOI] [PubMed] [Google Scholar]
  • [22].Stokke O, Goodman SI, Thompson JA, Miles BS. Glutaric aciduria; presence of glutaconic and beta-hydroxyglutaric acids in urine. Biochem Med. 1975;12:386–91. doi: 10.1016/0006-2944(75)90071-x. [DOI] [PubMed] [Google Scholar]
  • [23].Brismar J, Ozand PT. CT and MR of the brain in glutaric acidemia type I: a review of 59 published cases and a report of 5 new patients. AJNR Am J Neuroradiol. 1995;16:675–83. [PMC free article] [PubMed] [Google Scholar]
  • [24].Neumaier-Probst E, Harting I, Seitz A, Ding C, Kolker S. Neuroradiological findings in glutaric aciduria type I (glutaryl-CoA dehydrogenase deficiency) J Inherit Metab Dis. 2004;27:869–76. doi: 10.1023/B:BOLI.0000045771.66300.2a. [DOI] [PubMed] [Google Scholar]
  • [25].Floret D, Divry P, Dingeon N, Monnet P. Glutaric aciduria. 1 new case. Arch Fr Pediatr. 1979;36:462–70. [PubMed] [Google Scholar]
  • [26].Gregersen N, Brandt NJ. Ketotic episodes in glutaryl-CoA dehydrogenase deficiency (glutaric aciduria) Pediatr Res. 1979;13:977–81. doi: 10.1203/00006450-197909000-00005. [DOI] [PubMed] [Google Scholar]
  • [27].Koeller DM, Woontner M, Crnic LS, Kleinschmidt-DeMasters B, Stephens J, Hunt EL, Goodman SI. Biochemical, pathologic and behavioral analysis of a mouse model of glutaric acidemia type I. Hum Mol Genet. 2002;11:347–57. doi: 10.1093/hmg/11.4.347. [DOI] [PubMed] [Google Scholar]
  • [28].Schuck PF, Busanello EN, Tonin AM, Viegas CM, Ferreira GC. Neurotoxic effects of transglutaconic acid in rats. Oxid Med Cell Longev. 20132013:607610. doi: 10.1155/2013/607610. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [29].Ullrich K, Flott-Rahmel B, Schluff P, Musshoff U, Das A, Lücke T, Steinfeld R, Christensen E, Jakobs C, Ludolph A, Neu A, Röper R. Glutaric aciduria type I: pathomechanisms of neurodegeneration. J Inherit Metab Dis. 1999;22:392–403. doi: 10.1023/a:1005595921323. [DOI] [PubMed] [Google Scholar]
  • [30].Kölker S, Köhr G, Ahlemeyer B, Okun JG, Pawlak V, Hörster F, Mayatepek E, Krieglstein J, Hoffmann GF. Ca(2+) and Na(+) dependence of 3-hydroxyglutarate-induced excitotoxicity in primary neuronal cultures from chick embryo telencephalons. Pediatr Res. 2002;52:199–206. doi: 10.1203/00006450-200208000-00011. [DOI] [PubMed] [Google Scholar]
  • [31].da C Ferreira G, Viegas CM, Schuck PF, Latini A, Dutra-Filho CS, Wyse AT, Wannmacher CM, Vargas CR, Wajner M. Glutaric acid moderately compromises energy metabolism in rat brain. Int J Dev Neurosci. 2005;23:687–93. doi: 10.1016/j.ijdevneu.2005.08.003. [DOI] [PubMed] [Google Scholar]
  • [32].Beal MF. Mechanisms of excitotoxicity in neurologic diseases. FASEB J. 1992;6:3338–44. [PubMed] [Google Scholar]
  • [33].Greene JG, Greenamyre JT. Bioenergetics and glutamate excitotoxicity. Prog Neurobiol. 1996;48:613–34. doi: 10.1016/0301-0082(96)00006-8. [DOI] [PubMed] [Google Scholar]
  • [34].Henneberry RC, Novelli A, Cox JA, Lysko PG. Neurotoxicity at the N-methyl-D-aspartate receptor in energy-compromised neurons. An hypothesis for cell death in aging and disease. Ann N Y Acad Sci. 1989;568:225–33. doi: 10.1111/j.1749-6632.1989.tb12512.x. [DOI] [PubMed] [Google Scholar]
  • [35].Ikonomidou C, Turski L. Excitotoxicity and neurodegenerative diseases. Curr Opin Neurol. 1995;8:487–97. doi: 10.1097/00019052-199512000-00017. [DOI] [PubMed] [Google Scholar]
  • [36].Das AM, Lücke T, Ullrich K. Glutaric aciduria I: creatine supplementation restores creatinephosphate levels in mixed cortex cells from rat incubated with 3-hydroxyglutarate. Mol Genet Metab. 2003;78:108–11. doi: 10.1016/s1096-7192(02)00227-5. [DOI] [PubMed] [Google Scholar]
  • [37].Ferreira GC, Tonin A, Schuck PF, Viegas CM, Ceolato PC, Latini A, Perry ML, Wyse AT, Dutra-Filho CS, Wannmacher CM, Vargas CR, Wajner M. Evidence for a synergistic action of glutaric and 3-hydroxyglutaric acids disturbing rat brain energy metabolism. Int J Dev Neurosci. 2007;25:391–398. doi: 10.1016/j.ijdevneu.2007.05.009. [DOI] [PubMed] [Google Scholar]
  • [38].Scriver CR, Kaufman S. Hyperphenylalaninemia: phenylalanine hydroxylase deficiency. In: Scriver CR, Beaudet AL, Valle D, Sly WS, editors. The Metabolic and Molecular Bases of Inherited Disease. New York: McGraw-Hill; 2001. pp. 1667–724. [Google Scholar]
  • [39].Bauman ML, Kemper TL. Morphologic and histoanatomic observations of the brain in untreated human phenylketonuria. Acta Neuropathol. 1982;58:55–63. doi: 10.1007/BF00692698. [DOI] [PubMed] [Google Scholar]
  • [40].Gazit V, Ben-Abraham R, Pick CG, Katz Y. Beta-Phenylpyruvate induces long-term neurobehavioral damage and brain necrosis in neonatal mice. Behav Brain Res. 2003;143:1–5. doi: 10.1016/s0166-4328(03)00075-5. [DOI] [PubMed] [Google Scholar]
  • [41].Hörster F, Schwab MA, Sauer SW, Pietz J, Hoffmann GF, Okun JG, Kölker S, Kins S. Phenylalanine reduces synaptic density in mixed cortical cultures from mice. Pediatr Res. 2006;59:544–8. doi: 10.1203/01.pdr.0000203091.45988.8d. [DOI] [PubMed] [Google Scholar]
  • [42].Sitta A, Manfredini V, Biasi L, Treméa R, Schwartz IV, Wajner M, Vargas CR. Evidence that DNA damage is associated to phenylalanine blood levels in leukocytes from phenylketonuric patients. Mutat Res. 2009;679:13–6. doi: 10.1016/j.mrgentox.2009.07.013. [DOI] [PubMed] [Google Scholar]
  • [43].Simon KR, Dos Santos RM, Scaini G, Leffa DD, Damiani AP, Furlanetto CB, Machado JL, Cararo JH, Macan TP, Streck EL, Ferreira GC, Andrade VM, Schuck PF. DNA damage induced by phenylalanine and its analogue pchlorophenylalanine in blood and brain of rats subjected to a model of hyperphenylalaninemia. Biochem Cell Biol. 2013;91:319–24. doi: 10.1139/bcb-2013-0023. [DOI] [PubMed] [Google Scholar]
  • [44].Gu XF, Yang XW, Chen RG. Possible mechanism of nerve damage on hyperphenylalanine in embryonic rat. Am J Hum Genet. 2000;67:S281. [Google Scholar]
  • [45].Yang XW, Gu XF, Chen RG. Toxic effects of phenylacetic acid to cultured rat cortical neurons. Chin J Neurosci. 2000;16:330–2. [Google Scholar]
  • [46].Zhang HW, Gu XF. A study of gene expression profiles of cultured embryonic rat neurons induced by phenylalanine. Metab Brain Dis. 2005;20:61–72. doi: 10.1007/s11011-005-2477-y. [DOI] [PubMed] [Google Scholar]
  • [47].Wasserstein MP, Snyderman SE, Sansaricq C, Buchsbaum MS. Cerebral glucose metabolism in adults with early treated classic phenylketonuria. Mol Genet Metab. 2006;87:272–7. doi: 10.1016/j.ymgme.2005.06.010. [DOI] [PubMed] [Google Scholar]
  • [48].Pietz J, Rupp A, Ebinger F, Rating D, Mayatepek E, Boesch C, Kreis R. Cerebral energy metabolism in phenylketonuria: findings by quantitative in vivo 31P MR spectroscopy. Pediatr Res. 2003;53:654–62. doi: 10.1203/01.PDR.0000055867.83310.9E. [DOI] [PubMed] [Google Scholar]
  • [49].Costabeber E, Kessler A, Severo Dutra-Filho C, de Souza Wyse AT, Wajner M, Wannmacher CM. Hyperphenylalaninemia reduces creatine kinase activity in the cerebral cortex of rats. Int J Dev Neurosci. 2003;21:111–6. doi: 10.1016/s0736-5748(02)00108-9. [DOI] [PubMed] [Google Scholar]
  • [50].Miller AL, Hawkins RA, Veech RL. Phenylketonuria: phenylalanine inhibits brain pyruvate kinase in vivo. Science. 1973;179:904–6. doi: 10.1126/science.179.4076.904. [DOI] [PubMed] [Google Scholar]
  • [51].Lütz Mda G, Feksa LR, Wyse AT, Dutra-Filho CS, Wajner M, Wannmacher CM. Alanine prevents the in vitro inhibition of glycolysis caused by phenylalanine in brain cortex of rats. Metab Brain Dis. 2003;18:87–94. doi: 10.1023/a:1021986820562. [DOI] [PubMed] [Google Scholar]
  • [52].Rech VC, Feksa LR, Dutra-Filho CS, Wyse AT, Wajner M, Wannmacher CM. Inhibition of the mitochondrial respiratory chain by phenylalanine in rat cerebral cortex. Neurochem Res. 2002;27:353–7. doi: 10.1023/a:1015529511664. [DOI] [PubMed] [Google Scholar]
  • [53].Fenton WA, Gravel RA, Rosenblatt DS. In: The metabolic and molecular bases of inherited disease. Scriver CR, Beaudet AL, Sly WS, Valle D, editors. New York: McGraw-Hill; 2001. pp. 2165–93. [Google Scholar]
  • [54].Heidenreich R, Natowicz M, Hainline BE, Berman P, Kelley RI, Hillman RE, Berry GT. Acute extrapyramidal syndrome in methylmalonic acidemia: “metabolic stroke” involving the globus pallidus. J Pediatr. 1988;113:1022–7. doi: 10.1016/s0022-3476(88)80574-2. [DOI] [PubMed] [Google Scholar]
  • [55].Harting I, Seitz A, Geb S, Zwickler T, Porto L, Lindner M, Kölker S, Hörster F. Looking beyond the basal ganglia: the spectrum of MRI changes in methylmalonic acidaemia. J Inherit Metab Dis. 2008;31:368–78. doi: 10.1007/s10545-008-0801-5. [DOI] [PubMed] [Google Scholar]
  • [56].Trinh BC, Melhem ER, Barker PB. Multi-slice proton MR spectroscopy and diffusion-weighted imaging in methylmalonic acidemia: report of two cases and review of the literature. AJNR Am J Neuroradiol. 2001;22:831–3. [PMC free article] [PubMed] [Google Scholar]
  • [57].Hayasaka K, Metoki K, Satoh T, Narisawa K, Tada K, Kawakami T, Matsuo N, Aoki T. Comparison of cytosolic and mitochondrial enzyme alterations in the livers of propionic or methylmalonic acidemia: a reduction of cytochrome oxidase activity. Tohoku J Exp Med. 1982;137:329–34. doi: 10.1620/tjem.137.329. [DOI] [PubMed] [Google Scholar]
  • [58].Østergaard E, Wibrand F, Ørngreen MC, Vissing J, Horn N. Impaired energy metabolism and abnormal muscle histology in mut- methylmalonic aciduria. Neurology. 2005;65:931–3. doi: 10.1212/01.wnl.0000176065.80560.26. [DOI] [PubMed] [Google Scholar]
  • [59].Chandler RJ, Zerfas PM, Shanske S, Sloan J, Hoffmann V, DiMauro S, Venditti CP. Mitochondrial dysfunction in mut methylmalonic acidemia. FASEB J. 2009;23:1252–61. doi: 10.1096/fj.08-121848. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [60].de Keyzer Y, Valayannopoulos V, Benoist JF, Batteux F, Lacaille F, Hubert L, Chrétien D, Chadefeaux-Vekemans B, Niaudet P, Touati G, Munnich A, de Lonlay P. Multiple OXPHOS deficiency in the liver, kidney, heart, and skeletal muscle of patients with methylmalonic aciduria and propionic aciduria. Pediatr Res. 2009;66:91–5. doi: 10.1203/PDR.0b013e3181a7c270. [DOI] [PubMed] [Google Scholar]
  • [61].Mirandola SR, Melo DR, Schuck PF, Ferreira GC, Wajner M, Castilho RF. Methylmalonate inhibits succinate-supported oxygen consumption by interfering with mitochondrial succinate uptake. J Inherit Metab Dis. 2008;31:44–54. doi: 10.1007/s10545-007-0798-1. [DOI] [PubMed] [Google Scholar]
  • [62].Halperin ML, Schiller CM, Fritz IB. The inhibition by methylmalonic acid of malate transport by the dicarboxylate carrier in rat liver mitochondria. A possible explantation for hypoglycemia in methylmalonic aciduria. J Clin Invest. 1971;50:2276–82. doi: 10.1172/JCI106725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [63].Brusque AM, Borba Rosa R, Schuck PF, Dalcin KB, Ribeiro CA, Silva CG, Wannmacher CM, Dutra-Filho CS, Wyse AT, Briones P, Wajner M. Inhibition of the mitochondrial respiratory chain complex activities in rat cerebral cortex by methylmalonic acid. Neurochem Int. 2002;40:593–601. doi: 10.1016/s0197-0186(01)00130-9. [DOI] [PubMed] [Google Scholar]
  • [64].Pettenuzzo LF, Ferreira Gda C, Schmidt AL, Dutra-Filho CS, Wyse AT, Wajner M. Differential inhibitory effects of methylmalonic acid on respiratory chain complex activities in rat tissues. Int J Dev Neurosci. 2006;24:45–52. doi: 10.1016/j.ijdevneu.2005.10.005. [DOI] [PubMed] [Google Scholar]
  • [65].Utter MF, Keech DB, Scrutton MC. A possible role for acetyl CoA in the control of gluconeogenesis. Adv Enzyme Regul. 1964;2:49–68. doi: 10.1016/s0065-2571(64)80005-4. [DOI] [PubMed] [Google Scholar]
  • [66].Dutra JC, Wajner M, Wannmacher CF, Dutra-Filho CS, Wannmacher CM. Effects of methylmalonate and propionate on uptake of glucose and ketone bodies in vitro by brain of developing rats. Biochem Med Metab Biol. 1991;45:56–64. doi: 10.1016/0885-4505(91)90008-9. [DOI] [PubMed] [Google Scholar]
  • [67].Dutra JC, Dutra-Filho CS, Cardozo SE, Wannmacher CM, Sarkis JJ, Wajner M. Inhibition of succinate dehydrogenase and beta-hydroxybutyrate dehydrogenase activities by methylmalonate in brain and liver of developing rats. J Inherit Metab Dis. 1993;16:147–53. doi: 10.1007/BF00711328. [DOI] [PubMed] [Google Scholar]
  • [68].Saad LO, Mirandola SR, Maciel EN, Castilho RF. Lactate dehydrogenase activity is inhibited by methylmalonate in vitro. Neurochem Res. 2006;31:541–8. doi: 10.1007/s11064-006-9054-6. [DOI] [PubMed] [Google Scholar]
  • [69].Schuck PF, Rosa RB, Pettenuzzo LF, Sitta A, Wannmacher CM, Wyse AT, Wajner M. Inhibition of mitochondrial creatine kinase activity from rat cerebral cortex by methylmalonic acid. Neurochem Int. 2004;45:661–7. doi: 10.1016/j.neuint.2004.03.006. [DOI] [PubMed] [Google Scholar]
  • [70].Wajner M, Dutra JC, Cardoso SE, Wannmacher CM, Motta ER. Effect of methylmalonate on in vitro lactate release and carbon dioxide production by brain of suckling rats. J Inherit Metab Dis. 1992;15:92–6. doi: 10.1007/BF01800350. [DOI] [PubMed] [Google Scholar]
  • [71].Maciel EN, Kowaltowski AJ, Schwalm FD, Rodrigues JM, Souza DO, Vercesi AE, Wajner M, Castilho RF. Mitochondrial permeability transition in neuronal damage promoted by Ca2+ and respiratory chain complex II inhibition. J Neurochem. 2004;90:1025–35. doi: 10.1111/j.1471-4159.2004.02565.x. [DOI] [PubMed] [Google Scholar]
  • [72].Kowaltowski AJ, Maciel EN, Fornazari M, Castilho RF. Diazoxide protects against methylmalonate-induced neuronal toxicity. Exp Neurol. 2006;201:165–71. doi: 10.1016/j.expneurol.2006.04.004. [DOI] [PubMed] [Google Scholar]
  • [73].Sauer SW, Okun JG, Hoffmann GF, Koelker S, Morath MA. Impact of short- and medium-chain organic acids, acylcarnitines, and acyl-CoAs on mitochondrial energy metabolism. Biochim Biophys Acta. 2008;1777:1276–82. doi: 10.1016/j.bbabio.2008.05.447. [DOI] [PubMed] [Google Scholar]
  • [74].Patterson MC, Vanier MT, Suzuki K, Morris JA, Carstea E, Neufeld EB, Blanchette-Mackie JE, Pentchev PG. Niemann-Pick disease type C. A lipid trafficking disorder. In: Scriver CR, Sly WS, Valle D, editors. The Metabolic and Molecular Basis of Inherited Disease. New York: Mulencer Hill; 2001. [Google Scholar]
  • [75].Vanier MT, Duthel S, Rodriguez-Lafrasse C, Pentchev P, Carstea ED. Genetic heterogeneity in Niemann-Pick C disease: a study using somatic cell hybridization and linkage analysis. Am J Hum Genet. 1996;58:118–25. [PMC free article] [PubMed] [Google Scholar]
  • [76].NP-C Guidelines Working Group. Wraith JE, Baumgartner MR, Bembi B, Covanis A, Levade T, Mengel E, Pineda M, Sedel F, Topçu M, Vanier MT, Widner H, Wijburg FA, Patterson MC. Recommendations on the diagnosis and management of Niemann-Pick disease type C. Mol Genet Metab. 2009;98:152–65. doi: 10.1016/j.ymgme.2009.06.008. [DOI] [PubMed] [Google Scholar]
  • [77].Pentchev PGV, Vanier MT, Suzuki K, Patterson MC. Niemann-Pick disease type C: a cellular cholesterol lipidosis. In: Scriver CR, Beaudet AL, Sly WS, Valle D, editors. The Metabolic and Molecular Bases of Inherited Disease. New York: McGraw Hill; 1995. pp. 2625–39. [Google Scholar]
  • [78].Ory DS. Niemann-Pick type C: a disorder of cellular cholesterol trafficking. Biochim Biophys Acta. 2000;1529:331–9. doi: 10.1016/s1388-1981(00)00158-x. [DOI] [PubMed] [Google Scholar]
  • [79].Stefulja J, Perica M, Malnar M, Kosicek M, Schweinzer C, Zivkovic J, Scholler M, Panzenboeck U, Hecimovic S. Pharmacological Activation of LXRs Decreases Amyloid-β Levels in Niemann-Pick Type C Model Cells. Curr Pharm Biotechnol. 2013 doi: 10.2174/138920101131400224. In press. [DOI] [PubMed] [Google Scholar]
  • [80].Mattsson N, Zetterberg H, Bianconi S, et al. Gamma-secretase- dependent amyloid-beta is increased in Niemann-Pick type C: a cross-sectional study. Neurology. 2011;76:366–72. doi: 10.1212/WNL.0b013e318208f4ab. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [81].Walkley SU, Suzuki K. Consequences of NPC1 and NPC2 loss of function in mammalian neurons. Biochim Biophys Acta. 2004;1685:48–62. doi: 10.1016/j.bbalip.2004.08.011. [DOI] [PubMed] [Google Scholar]
  • [82].Fu R, Yanjanin NM, Bianconi S, Pavan WJ, Porter FD. Oxidative stress in Niemann-Pick disease, type C. Mol Genet Metab. 2010;101:214–8. doi: 10.1016/j.ymgme.2010.06.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [83].Vázquez MC, Balboa E, Alvarez AR, Zanlungo S. Oxidative stress: a pathogenic mechanism for Niemann-Pick type C disease. Oxid Med Cell Longev. 20122012:205713. doi: 10.1155/2012/205713. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [84].Yu W, Gong JS, Ko M, Garver WS, Yanagisawa K, Michikawa M. Altered cholesterol metabolism in Niemann- Pick type C1 mouse brains affects mitochondrial function. J Biol Chem. 2005;280:11731–11739. doi: 10.1074/jbc.M412898200. [DOI] [PubMed] [Google Scholar]
  • [85].Lucken-Ardjomande S, Montessuit S, Martinou JC. Bax activation and stress-induced apoptosis delayed by the accumulation of cholesterol in mitochondrial membranes. Cell Death Differ. 2008;15:484–493. doi: 10.1038/sj.cdd.4402280. [DOI] [PubMed] [Google Scholar]
  • [86].Charman M, Kennedy BE, Osborne N, Karten B. MLN64 mediates egress of cholesterol from endosomes to mitochondria in the absence of functional Niemann-Pick Type C1 protein. J Lipid Res. 2010;51:1023–1034. doi: 10.1194/jlr.M002345. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [87].Harman D. Aging: a theory based on free radical and radiation chemistry. J Gerontol. 1956;11:298–300. doi: 10.1093/geronj/11.3.298. [DOI] [PubMed] [Google Scholar]
  • [88].Nicholls DG. Mitochondrial membrane potential and aging. Aging Cell. 2004;3:35–40. doi: 10.1111/j.1474-9728.2003.00079.x. [DOI] [PubMed] [Google Scholar]
  • [89].Lenaz G, D’Aurelio M, Merlo Pich M, Genova ML, Ventura B, Bovina C, Formiggini G, ParentiCastelli G. Mitochondrial bioenergetics in aging. Biochim Biophys Acta. 2000;1459:397–404. doi: 10.1016/s0005-2728(00)00177-8. [DOI] [PubMed] [Google Scholar]
  • [90].Feuers RJ. The effects of dietary restriction on mitochondrial dysfunction in aging. Ann NY Acad Sci. 1998;54:192–201. doi: 10.1111/j.1749-6632.1998.tb09902.x. [DOI] [PubMed] [Google Scholar]
  • [91].Liu J, Head E, Gharib AM, Yuan W, Ingersoll RT, Hagen TM, Cotman CW, Ames BN. Memory loss in old rats is associated with brain mitochondrial decay and RNA/DNA oxidation: partial reversal by feeding acetyl-L-carnitine and/or R-alpha -lipoic acid. Proc Natl Acad Sci U S A. 2002;99:2356–61. doi: 10.1073/pnas.261709299. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [92].Liu J, Killilea DW, Ames BN. Age-associated mitochondrial oxidative decay: improvement of carnitine acetyltransferase substrate-binding affinity and activity in brain by feeding old rats acetyl-L- carnitine and/or R-alpha -lipoic acid. Proc Natl Acad Sci U S A. 2002;99:1876–81. doi: 10.1073/pnas.261709098. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [93].Roth GS, Ingram DK, Lane MA. Caloric restriction in primates and relevance to humans. Ann N Y Acad Sci. 2001;928:305–15. doi: 10.1111/j.1749-6632.2001.tb05660.x. [DOI] [PubMed] [Google Scholar]
  • [94].Weindruch, Walford . The Retardation of Aging and Disease by Dietary Restriction. Springfield, Illinois: Charles C Thomas; 1988. [Google Scholar]
  • [95].Cerqueira FM, Cunha FM, Laurindo FR, Kowaltowski AJ. Calorie restriction increases cerebral mitochondrial respiratory capacity in a NO•-mediated mechanism: impact on neuronal survival. Free Radic Biol Med. 2012;52:1236–41. doi: 10.1016/j.freeradbiomed.2012.01.011. [DOI] [PubMed] [Google Scholar]
  • [96].Eckles-Smith K, Clayton D, Bickford P, Browning MD. Caloric restriction prevents age-related deficits in LTP and in NMDA receptor expression. Brain Res Mol Brain Res. 2000;78:154–62. doi: 10.1016/s0169-328x(00)00088-7. [DOI] [PubMed] [Google Scholar]
  • [97].Okada M, Nakanishi H, Amamoto T, Urae R, Ando S, Yazawa K, Fujiwara M. How does prolonged caloric restriction ameliorate age-related impairment of long-term potentiation in the hippocampus? Brain Res Mol Brain Res. 2003;111:175–81. doi: 10.1016/s0169-328x(03)00028-7. [DOI] [PubMed] [Google Scholar]
  • [98].Hillman CH, Erickson KI, Kramer AF. Be smart, exercise your heart: exercise effects on brain and cognition. Nat Rev Neurosci. 2008;9:58–65. doi: 10.1038/nrn2298. [DOI] [PubMed] [Google Scholar]
  • [99].Kramer AF, Erickson KI, Colcombe SJ. Exercise, cognition, and the aging brain. J Appl Physiol. 2006;101:1237–42. doi: 10.1152/japplphysiol.00500.2006. [DOI] [PubMed] [Google Scholar]
  • [100].Dietrich MO, Andrews ZB, Horvath TL. Exercise-induced synaptogenesis in the hippocampus is dependent on UCP2-regulated mitochondrial adaptation. J Neurosci. 2008;28:10766–71. doi: 10.1523/JNEUROSCI.2744-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [101].Brown J, Cooper-Kuhn CM, Kempermann G, Van Praag H, Winkler J, Gage FH, Kuhn HG. Enriched environment and physical activity stimulate hippocampal but not olfactory bulb neurogenesis. Eur J Neurosci. 2003;17:2042–6. doi: 10.1046/j.1460-9568.2003.02647.x. [DOI] [PubMed] [Google Scholar]
  • [102].Van Praag H, Christie BR, Sejnowski TJ, Gage FH. Running enhances neurogenesis, learning, and long-term potentiation in mice. Proc Natl Acad Sci U S A. 1999;96:13427–31. doi: 10.1073/pnas.96.23.13427. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [103].Van Praag H, Kempermann G, Gage FH. Running increases cell proliferation and neurogenesis in the adult mouse dentate gyrus. Nat Neurosci. 1999;2:266–70. doi: 10.1038/6368. [DOI] [PubMed] [Google Scholar]
  • [104].Trejo JL, Carro E, Torres-Aleman I. Circulating insulin-like growth factor I mediates exercise-induced increases in the number of new neurons in the adult hippocampus. J Neurosci. 2001;21:1628–34. doi: 10.1523/JNEUROSCI.21-05-01628.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [105].Eadie BD, Redila VA, Christie BR. Voluntary exercise alters the cytoarchitecture of the adult dentate gyrus by increasing cellular proliferation, dendritic complexity, and spine density. J Comp Neurol. 2005;486:39–47. doi: 10.1002/cne.20493. [DOI] [PubMed] [Google Scholar]
  • [106].Van Praag H, Shubert T, Zhao C, Gage FH. Exercise enhances learning and hippocampal neurogenesis in aged mice. J Neurosci. 2005;25:8680–5. doi: 10.1523/JNEUROSCI.1731-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [107].Chang L, Ernst T, Poland RE, Jenden DJ. In vivo proton magnetic resonance spectroscopy of the normal aging human brain. Life Sci. 1996;58:2049–2056. doi: 10.1016/0024-3205(96)00197-x. [DOI] [PubMed] [Google Scholar]
  • [108].Gruber S, Pinker K, Riederer F, Chmelík M, Stadlbauer A, Bittsanský M, Mlynárik V, Frey R, Serles W, Bodamer O, Moser E. Metabolic changes in the normal ageing brain: consistent findings from short and long echo time proton spectroscopy. Eur J Radiol. 2008;68:320–327. doi: 10.1016/j.ejrad.2007.08.038. [DOI] [PubMed] [Google Scholar]
  • [109].Schuff N, Ezekiel F, Gamst AC, Amend DL, Capizzano AA, Maudsley AA, Weiner MW. Region and tissue differences of metabolites in normally aged brain using multislice 1H magnetic resonance spectroscopic imaging. Magn Reson Med. 2001;45:899–907. doi: 10.1002/mrm.1119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [110].Duarte JMN, Gruetter R. Extended neurochemical profile in the aging mouse brain detected in vivo by proton magnetic resonance spectroscopy. J Neurochem. 2013;125(Suppl 1):224. [Google Scholar]
  • [111].Canas PM, Duarte JMN, Rodrigues RJ, Köfalvi A, Cunha RA. Modification upon aging of the density of presynaptic modulation systems in the hippocampus. Neurobiology of Aging. 2009;30:1877–1884. doi: 10.1016/j.neurobiolaging.2008.01.003. [DOI] [PubMed] [Google Scholar]
  • [112].Van Guilder HD, Yan H, Farley JA, Sonntag WE, Freeman WM. Aging alters the expression of neurotransmission-regulating proteins in the hippocampal synaptoproteome. J Neurochem. 2010;113:1577–1588. doi: 10.1111/j.1471-4159.2010.06719.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [113].Allard S, Scardochio T, Cuello AC, Ribeiroda-Silva A. Correlation of cognitive performance and morphological changes in neocortical pyramidal neurons in aging. Neurobiol Aging. 2012;33:1466–1480. doi: 10.1016/j.neurobiolaging.2010.10.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [114].Gage FH, Dunnett SB, Björklund A. Spatial learning and motor deficits in aged rats. Neurobiol Aging. 1984;5:43–48. doi: 10.1016/0197-4580(84)90084-8. [DOI] [PubMed] [Google Scholar]
  • [115].Kennard JA, Woodruff-Pak DS. Age sensitivity of behavioral tests and brain substrates of normal aging in mice. Front Aging Neurosci. 2011;3:1–22. doi: 10.3389/fnagi.2011.00009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [116].Pistell PJ, Spangler EL, Kelly-Bell B, Miller MG, de Cabo R, Ingram DK. Age-associated learning and memory deficits in two mouse versions of the stone T-maze. Neurobiol Aging. 2012;33:2431–2439. doi: 10.1016/j.neurobiolaging.2011.12.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [117].Duarte JMN, Lei H, Mlynárik V, Gruetter R. The neurochemical profile quantified by in vivo 1H NMR spectroscopy. NeuroImage. 2012;61:342–362. doi: 10.1016/j.neuroimage.2011.12.038. [DOI] [PubMed] [Google Scholar]
  • [118].Boumezbeur F, Mason GF, de Graaf RA, Behar KL, Cline GW, Shulman GI, Rothman DL, Petersen KF. Altered brain mitochondrial metabolism in healthy aging as assessed by in vivo magnetic resonance spectroscopy. J Cereb Blood Flow Metab. 2010;30:211–221. doi: 10.1038/jcbfm.2009.197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [119].Gage FH, Kelly PA, Björklund A. Regional changes in brain glucose metabolism reflect cognitive impairments in aged rats. J Neurosci. 1984;4:2856–2865. doi: 10.1523/JNEUROSCI.04-11-02856.1984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [120].Tack W, Wree A, Schleicher A. Local cerebral glucose utilization in the hippocampus of old rats. Histochemistry. 1989;92:413–419. doi: 10.1007/BF00492499. [DOI] [PubMed] [Google Scholar]
  • [121].Kulak A, Duarte JMN, Do KQ, Gruetter R. Neurochemical profile of the developing mouse cortex determined by in vivo 1H NMR spectroscopy at 14.1 T and the effect of recurrent anaesthesia. J Neurochem. 2010;115:1466–1477. doi: 10.1111/j.1471-4159.2010.07051.x. [DOI] [PubMed] [Google Scholar]
  • [122].Blüml S, Seymour KJ, Ross BD. Developmental changes in choline- and ethanolamine-containing compounds measured with proton-decoupled 31P MRS in in vivo human brain. Magn Reson Med. 1999;42:643–654. doi: 10.1002/(sici)1522-2594(199910)42:4<643::aid-mrm5>3.0.co;2-n. [DOI] [PubMed] [Google Scholar]
  • [123].Calderini G, Bonetti AC, Battistella A, Crews FT, Toffano G. Biochemical changes of rat brain membranes with aging. Neurochem Res. 1983;8:483–492. doi: 10.1007/BF00965104. [DOI] [PubMed] [Google Scholar]
  • [124].Albrecht J, Schousboe A. Taurine interaction with neurotransmitter receptors in the CNS: an update. Neurochem Res. 2005;30:1615–1621. doi: 10.1007/s11064-005-8986-6. [DOI] [PubMed] [Google Scholar]
  • [125].Tkáč I, Rao R, Georgieff MK, Gruetter R. Developmental and regional changes in the neurochemical profile of the rat brain determined by in vivo1H NMR spectroscopy. Magn Reson Med. 2003;50:24–32. doi: 10.1002/mrm.10497. [DOI] [PubMed] [Google Scholar]
  • [126].Calne DB, Langston JW. Aetiology of Parkinson’s disease. Lancet. 1983;2:1457–1459. doi: 10.1016/s0140-6736(83)90802-4. [DOI] [PubMed] [Google Scholar]
  • [127].Schapira AH, Cooper JM, Dexter D, Clark JB, Jenner P, Marsden CD. Mitochondrial complex I deficiency in Parkinson’s disease. J Neurochem. 1990;54:823–827. doi: 10.1111/j.1471-4159.1990.tb02325.x. [DOI] [PubMed] [Google Scholar]
  • [128].Parker WD, Jr, Parks JK. Mitochondrial ND5 mutations in idiopathic Parkinson’s disease. Biochem Biophys Res Commun. 2005;326:667–669. doi: 10.1016/j.bbrc.2004.11.093. [DOI] [PubMed] [Google Scholar]
  • [129].Parker WD, Jr, Parks JK, Swerdlow RH. Complex I deficiency in Parkinson’s disease frontal cortex. Brain Res. 2008;1189:215–218. doi: 10.1016/j.brainres.2007.10.061. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [130].Ravid R, Ferrer I. Brain banks as key part of biochemical and molecular studies on cerebral cortex involvement in Parkinson’s disease. FEBS J. 2012;279:1167–76. doi: 10.1111/j.1742-4658.2012.08518.x. [DOI] [PubMed] [Google Scholar]
  • [131].Krige D, Carroll MT, Cooper JM, Marsden CD, Schapira AH. Platelet mitochondrial function in Parkinson’s disease. The Royal Kings and Queens Parkinson Disease Research Group. Ann Neurol. 1992;32:782–788. doi: 10.1002/ana.410320612. [DOI] [PubMed] [Google Scholar]
  • [132].Bindoff LA, Birch-Machin MA, Cartlidge NE, Parker WD, Jr, Turnbull DM. Respiratory chain abnormalities in skeletal muscle from patients with Parkinson’s disease. J Neurol Sci. 1991;104:203–208. doi: 10.1016/0022-510x(91)90311-t. [DOI] [PubMed] [Google Scholar]
  • [133].Zhu J, Chu CT. Mitochondrial dysfunction in Parkinson’s disease. J Alzheimers Dis. 2010;20:325–334. doi: 10.3233/JAD-2010-100363. [DOI] [PubMed] [Google Scholar]
  • [134].Schapira AH. Mitochondria in the aetiology and pathogenesis of Parkinson’s disease. Lancet Neurol. 2008;7:97–109. doi: 10.1016/S1474-4422(07)70327-7. [DOI] [PubMed] [Google Scholar]
  • [135].Annepu J, Ravindranath V. 1-Methyl-4-phenyl-1,2,3,6-tetrahydropyridine-induced complex I inhibition is reversed by disulfide reductant, dithiothreitol in mouse brain. Neurosci Lett. 2000;289:209–212. doi: 10.1016/s0304-3940(00)01300-8. [DOI] [PubMed] [Google Scholar]
  • [136].Mytilineou C, Werner P, Molinari S, Di Rocco A, Cohen G, Yahr MD. Impaired oxidative decarboxylation of pyruvate in fibroblasts from patients with Parkinson’s disease. J Neural Transm Park Dis Dement Sect. 1994;8:223–228. doi: 10.1007/BF02260943. [DOI] [PubMed] [Google Scholar]
  • [137].Campello L, Esteve-Rudd J, Bru-Martínez R, Herrero MT, Fernández-Villalba E, Cuenca N, Martín-Nieto J. Alterations in Energy Metabolism, Neuroprotection and Visual Signal Transduction in the Retina of Parkinsonian, MPTP-Treated Monkeys. PLoS One. 2013;8:e74439. doi: 10.1371/journal.pone.0074439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [138].Bender A, Krishnan KJ, Morris CM, Taylor GA, Reeve AK, Perry RH, Jaros E, Hersheson JS, Betts J, Klopstock T, Taylor RW, Turnbull DM. High levels of mitochondrial DNA deletions in substantia nigra neurons in aging and Parkinson disease. Nat Genet. 2006;38:515–517. doi: 10.1038/ng1769. [DOI] [PubMed] [Google Scholar]
  • [139].Arthur CR, Morton SL, Dunham LD, Keeney PM, Bennett JP., Jr Parkinson’s disease brain mitochondria have impaired respirasome assembly, age-related increases in distribution of oxidative damage to mtDNA and no differences in heteroplasmic mtDNA mutation abundance. Mol Neurodegener. 2009;4:37. doi: 10.1186/1750-1326-4-37. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [140].Kraytsberg Y, Kudryavtseva E, McKee AC, Geula C, Kowall NW, Khrapko K. Mitochondrial DNA deletions are abundant and cause functional impairment in aged human substantia nigra neurons. Nat Genet. 2006;38:518–520. doi: 10.1038/ng1778. [DOI] [PubMed] [Google Scholar]
  • [141].Lucetti C, Del Dotto P, Gambaccini G, Ceravolo R, Logi C, Berti C, Rossi G, Bianchi MC, Tosetti M, Murri L, Bonuccelli U. Influences of dopaminergic treatment on motor cortex in Parkinson disease: a MRI/MRS study. Mov Disord. 2007;22:2170–2175. doi: 10.1002/mds.21576. [DOI] [PubMed] [Google Scholar]
  • [142].Taylor-Robinson SD, Turjanski N, Bhattacharya S, Seery JP, Sargentoni J, Brooks DJ, Bryant DJ, Cox IJ. A proton magnetic resonance spectroscopy study of the striatum and cerebral cortex in Parkinson’s disease. Metab Brain Dis. 1999;14:45–55. doi: 10.1023/a:1020609530444. [DOI] [PubMed] [Google Scholar]
  • [143].Nie K, Zhang Y, Huang B, Wang L, Zhao J, Huang Z, Gan R, Wang L. Marked N-acetylaspartate and choline metabolite changes in Parkinson’s disease patients with mild cognitive impairment. Parkinsonism Relat Disord. 2013;19:329–334. doi: 10.1016/j.parkreldis.2012.11.012. [DOI] [PubMed] [Google Scholar]
  • [144].Griffith HR, Okonkwo OC, O’Brien T, Hollander JA. Reduced brain glutamate in patients with Parkinson’s disease. NMR Biomed. 2008;21:381–387. doi: 10.1002/nbm.1203. [DOI] [PubMed] [Google Scholar]
  • [145].Lin MT, Beal MF. Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases. Nature. 2006;443:787–795. doi: 10.1038/nature05292. [DOI] [PubMed] [Google Scholar]
  • [146].Hattingen E, Magerkurth J, Pilatus U, Mozer A, Seifried C, Steinmetz H, Zanella F, Hilker R. Phosphorus and proton magnetic resonance spectroscopy demonstrates mitochondrial dysfunction in early and advanced Parkinson’s disease. Brain. 2009;132:3285–3297. doi: 10.1093/brain/awp293. [DOI] [PubMed] [Google Scholar]
  • [147].Henchcliffe C, Shungu DC, Mao X, Huang C, Nirenberg MJ, Jenkins BG, Beal MF. Multinuclear magnetic resonance spectroscopy for in vivo assessment of mitochondrial dysfunction in Parkinson’s disease. Ann N Y Acad Sci. 2008;1147:206–220. doi: 10.1196/annals.1427.037. [DOI] [PubMed] [Google Scholar]
  • [148].Emir UE, Tuite PJ, Öz G. Elevated pontine and putamenal GABA levels in mild-moderate Parkinson disease detected by 7 tesla proton MRS. PLoS One. 2012;7:e30918. doi: 10.1371/journal.pone.0030918. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [149].Boska MD, Lewis TB, Destache CJ, Benner EJ, Nelson JA, Uberti M, Mosley RL, Gendelman HE. Quantitative 1H magnetic resonance spectroscopic imaging determines therapeutic immunization efficacy in an animal model of Parkinson’s disease. J Neurosci. 2005;25:1691–1700. doi: 10.1523/JNEUROSCI.4364-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [150].Koga K, Mori A, Ohashi S, Kurihara N, Kitagawa H, Ishikawa M, Mitsumoto Y, Nakai M. 1H MRS identifies lactate rise in the striatum of MPTP-treated C57BL/6 mice. Eur J Neurosci. 2006;23:1077–1081. doi: 10.1111/j.1460-9568.2006.04610.x. [DOI] [PubMed] [Google Scholar]
  • [151].Jenkins BG, Brouillet E, Chen YC, Storey E, Schulz JB, Kirschner P, Beal MF, Rosen BR. Non-invasive neurochemical analysis of focal excitotoxic lesions in models of neurodegenerative illness using spectroscopic imaging. J Cereb Blood Flow Metab. 1996;16:450–461. doi: 10.1097/00004647-199605000-00011. [DOI] [PubMed] [Google Scholar]
  • [152].Storey E, Hyman BT, Jenkins B, Brouillet E, Miller JM, Rosen BR, Beal MF. 1-Methyl-4-phenylpyridinium produces excitotoxic lesions in rat striatum as a result of impairment of oxidative metabolism. J Neurochem. 1992;58:1975–1978. doi: 10.1111/j.1471-4159.1992.tb10080.x. [DOI] [PubMed] [Google Scholar]
  • [153].Hou Z, Lei H, Hong S, Sun B, Fang K, Lin X, Liu M, Yew DT, Liu S. Functional changes in the frontal cortex in Parkinson’s disease using a rat model. J Clin Neurosci. 2010;17:628–633. doi: 10.1016/j.jocn.2009.07.101. [DOI] [PubMed] [Google Scholar]
  • [154].Chassain C, Bielicki G, Durand E, Lolignier S, Essafi F, Traoré A, Durif F. Metabolic changes detected by proton magnetic resonance spectroscopy in vivo and in vitro in a murin model of Parkinson’s disease, the MPTP-intoxicated mouse. J Neurochem. 2008;105:874–882. doi: 10.1111/j.1471-4159.2007.05185.x. [DOI] [PubMed] [Google Scholar]
  • [155].Chassain C, Bielicki G, Keller C, Renou JP, Durif F. Metabolic changes detected in vivo by 1H MRS in the MPTP-intoxicated mouse. NMR Biomed. 2010;23:547–553. doi: 10.1002/nbm.1504. [DOI] [PubMed] [Google Scholar]
  • [156].Podell M, Hadjiconstantinou M, Smith MA, Neff NH. Proton magnetic resonance imaging and spectroscopy identify metabolic changes in the striatum in the MPTP feline model of parkinsonism. Exp Neurol. 2003;179:159–166. doi: 10.1016/s0014-4886(02)00015-8. [DOI] [PubMed] [Google Scholar]
  • [157].Chassain C, Bielicki G, Carcenac C, Ronsin AC, Renou JP, Savasta M, Durif F. Does MPTP intoxication in mice induce metabolite changes in the nucleus accumbens? A ¹H nuclear MRS study. NMR Biomed. 2013;26:336–47. doi: 10.1002/nbm.2853. [DOI] [PubMed] [Google Scholar]
  • [158].Huang Y, Mucke L. Alzheimer mechanisms and therapeutic strategies. Cell. 2012;148:1204–1222. doi: 10.1016/j.cell.2012.02.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [159].Santoro A, Balbi V, Balducci E, Pirazzini C, Rosini F, Tavano F, et al. Evidence for subhaplogroup h5 of mitochondrial DNA as a risk factor for late onset Alzheimer’s disease. PLoS One. 2010;5:e12037. doi: 10.1371/journal.pone.0012037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [160].Wang X, Su B, Lee HG, Li X, Perry G, Smith MA, Zhu X. Impaired balance of mitochondrial fission and fusion in Alzheimer’s disease. J Neurosci. 2009;29:9090–9103. doi: 10.1523/JNEUROSCI.1357-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [161].Palau F, Estela A, Pla-Martín D, Sánchez-Piris M. The role of mitochondrial network dynamics in the pathogenesis of Charcot-Marie-Tooth disease. Adv Exp Med Biol. 2009;652:129–137. doi: 10.1007/978-90-481-2813-6_9. [DOI] [PubMed] [Google Scholar]
  • [162].Casley CS, Canevari L, Land JM, Clark JB, Sharpe MA. Beta-amyloid inhibits inte- grated mitochondrial respiration and key enzyme activities. J Neurochem. 2002;80:91–100. doi: 10.1046/j.0022-3042.2001.00681.x. [DOI] [PubMed] [Google Scholar]
  • [163].Manczak M, Anekonda TS, Henson E, Park BS, Quinn J, Reddy PH. Mitochondria are a direct site of A beta accumulation in Alzheimer’s disease neurons: implications for free radical generation and oxidative damage in disease progression. Hum Mol Genet. 2006;15:1437–1449. doi: 10.1093/hmg/ddl066. [DOI] [PubMed] [Google Scholar]
  • [164].Cho DH, Nakamura T, Fang J, Cieplak P, Godzik A, Gu Z, Lipton SA. S-nitrosylation of Drp1 mediates beta-amyloid-related mitochondrial fission and neuronal injury. Science. 2009;324:102–5. doi: 10.1126/science.1171091. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [165].Melov S, Adlard PA, Morten K, Johnson F, Golden TR, Hinerfeld D, Schilling B, Mavros C, Masters CL, Volitakis I, Li QX, Laughton K, Hubbard A, Cherny RA, Gibson B, Bush AI. Mitochondrial oxidative stress causes hyperphosphorylation of tau. PLoS One. 2007;2:536. doi: 10.1371/journal.pone.0000536. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [166].Xie H, Guan J, Borrelli LA, Xu J, Serrano-Pozo A, Bacskai BJ. Mitochondrial Alterations near Amyloid Plaques in an Alzheimer’s Disease Mouse Model. J Neurosci. 2013;33:17042–170451. doi: 10.1523/JNEUROSCI.1836-13.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [167].Pahnke J, Fröhlich C, Krohn M, Schumacher T, Paarmann K. Impaired mitochondrial energy production and ABC transporter function-A crucial interconnection in dementing proteopathies of the brain. Mech Ageing Dev. 2013 doi: 10.1016/j.mad.2013.08.007. In press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [168].Meyerhoff DJ, MacKay S, Constans JM, Norman D, Van Dyke C, Fein G, Weiner MW. Axonal injury and membrane alterations in Alzheimer’s disease suggested by in vivo proton magnetic resonance spectroscopic imaging. Ann Neurol. 1994;36:40–7. doi: 10.1002/ana.410360110. [DOI] [PubMed] [Google Scholar]
  • [169].Pfefferbaum A, Adalsteinsson E, Spielman D, Sullivan EV, Lim KO. In vivo brain concentrations of N-acetyl compounds, creatine, and choline in Alzheimer disease. Arch Gen Psychiatry. 1999;56:185–192. doi: 10.1001/archpsyc.56.2.185. [DOI] [PubMed] [Google Scholar]
  • [170].Pfefferbaum A, Adalsteinsson E, Spielman D, Sullivan EV, Lim KO. In vivo spectroscopic quantification of the N-acetyl moiety, creatine, and choline from large volumes of brain gray and white matter: effects of normal aging. Magn Reson Med. 1999;41:276–284. doi: 10.1002/(sici)1522-2594(199902)41:2<276::aid-mrm10>3.0.co;2-8. [DOI] [PubMed] [Google Scholar]
  • [171].Kantarci K, Petersen RC, Boeve BF, Knopman DS, Tang-Wai DF, O’Brien PC, Weigand SD, Edland SD, Smith GE, Ivnik RJ, Ferman TJ, Tangalos EG, Jack CR., Jr 1H MR spectroscopy in common dementias. Neurology. 2004;63:1393–1398. doi: 10.1212/01.wnl.0000141849.21256.ac. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [172].Chantal S, Labelle M, Bouchard RW, Braun CM, Boulanger Y. Correlation of regional proton magnetic resonance spectroscopic metabolic changes with cognitive deficits in mild Alzheimer disease. Arch Neurol. 2002;59:955–962. doi: 10.1001/archneur.59.6.955. [DOI] [PubMed] [Google Scholar]
  • [173].Watanabe T, Shiino A, Akiguchi I. Absolute quantification in proton magnetic resonance spectroscopy is useful to differentiate amnesic mild cognitive impairment from Alzheimer’s disease and healthy aging. Dement Geriatr Cogn Disord. 2010;30:71–77. doi: 10.1159/000318750. [DOI] [PubMed] [Google Scholar]
  • [174].Nitsch RM, Blusztajn JK, Pittas AG, Slack BE, Growdon JH, Wurtman RJ. Evidence for a membrane defect in Alzheimer disease brain. Proc Natl Acad Sci U S A. 1992;89:1671–1675. doi: 10.1073/pnas.89.5.1671. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [175].Jessen F, Gür O, Block W, Ende G, Frölich L, Hammen T, Wiltfang J, Kucinski T, Jahn H, Heun R, Maier W, et al. A multicenter 1H-MRS study of the medial temporal lobe in AD and MCI. Neurology. 2009;72:1735–1740. doi: 10.1212/WNL.0b013e3181a60a20. [DOI] [PubMed] [Google Scholar]
  • [176].Kantarci K. 1H magnetic resonance spectroscopy in dementia. Br J Radiol. 2007;80:S146–S152. doi: 10.1259/bjr/60346217. [DOI] [PubMed] [Google Scholar]
  • [177].Chantal S, Braun CM, Bouchard RW, Labelle M, Boulanger Y. Similar 1H magnetic resonance spectroscopic metabolic pattern in the medial temporal lobes of patients with mild cognitive impairment and Alzheimer disease. Brain Res. 2004;1003:26–35. doi: 10.1016/j.brainres.2003.11.074. [DOI] [PubMed] [Google Scholar]
  • [178].Pilatus U, Lais C, Rochmont Adu M, Kratzsch T, Frölich L, Maurer K, Zanella FE, Lanfermann H, Pantel J. Conversion to dementia in mild cognitive impairment is associated with decline of N-actylaspartate and creatine as revealed by magnetic resonance spectroscopy. Psychiatry Res. 2009;173:1–7. doi: 10.1016/j.pscychresns.2008.07.015. [DOI] [PubMed] [Google Scholar]
  • [179].Haley AP, Knight-Scott J, Simnad VI, Manning CA. Increased glucose concentration in the hippocampus in early Alzheimer’s disease following oral glucose ingestion. Magn Reson Imaging. 2006;24:715–720. doi: 10.1016/j.mri.2005.12.020. [DOI] [PubMed] [Google Scholar]
  • [180].Duarte AI, Moreira PI, Oliveira CR. Insulin in central nervous system: more than just a peripheral hormone. J Aging Res. 2012:1–21. doi: 10.1155/2012/384017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [181].Selkoe DJ. Deciphering Alzheimer’s disease: molecular genetics and cell biology yield major clues. J NIH Res. 1995;7:57–64. [Google Scholar]
  • [182].Choi JK, Jenkins BG, Carreras I, Kaymakcalan S, Cormier K, Kowall NW, Dedeoglu A. Anti-inflammatory treatment in AD mice protects against neuronal pathology. Exp Neurol. 2010;223:377–384. doi: 10.1016/j.expneurol.2009.07.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [183].Marjanska M, Curran GL, Wengenack TM, Henry PG, Bliss RL, Poduslo JF, Jack CR, Jr, Ugurbil K, Garwood M. Monitoring disease progression in transgenic mouse models of Alzheimer’s disease with proton magnetic resonance spectroscopy. Proc Natl Acad Sci U S A. 2005;102:11906–11910. doi: 10.1073/pnas.0505513102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [184].Oberg J, Spenger C, Wang FH, Andersson A, Westman E, Skoglund P, Sunnemark D, Norinder U, Klason T, Wahlund LO, Lindberg M. Age related changes in brain metabolites observed by 1H MRS in APP/PS1 mice. Neurobiol Aging. 2008;29:1423–1433. doi: 10.1016/j.neurobiolaging.2007.03.002. [DOI] [PubMed] [Google Scholar]
  • [185].Jack CR, Jr, Marjanska M, Wengenack TM, Reyes DA, Curran GL, Lin J, Preboske GM, Poduslo JF, Garwood M. Magnetic resonance imaging of Alzheimer’s pathology in the brains of living transgenic mice: a new tool in Alzheimer’s disease research. Neuroscientist. 2007;13:38–48. doi: 10.1177/1073858406295610. [DOI] [PubMed] [Google Scholar]
  • [186].Chen SQ, Wang PJ, Ten GJ, Zhan W, Li MH, Zang FC. Role of myo-inositol by magnetic resonance spectroscopy in early diagnosis of Alzheimer’s disease in APP/PS1 transgenic mice. Dement Geriatr Cogn Disord. 2009;28:558–566. doi: 10.1159/000261646. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [187].Dedeoglu A, Choi JK, Cormier K, Kowall NW, Jenkins BG. Magnetic resonance spectroscopic analysis of Alzheimer’s disease mouse brain that express mutant human APP shows altered neurochemical profile. Brain Res. 2004;1012:60–65. doi: 10.1016/j.brainres.2004.02.079. [DOI] [PubMed] [Google Scholar]
  • [188].Mlynárik V, Cacquevel M, Sun-Reimer L, Janssens S, Cudalbu C, Lei H, Schneider BL, Aebischer P, Gruetter R. Proton and phosphorus magnetic resonance spectroscopy of a mouse model of Alzheimer’s disease. J Alzheimers Dis. 2012;3:S87–99. doi: 10.3233/JAD-2012-112072. [DOI] [PubMed] [Google Scholar]
  • [189].Zheng Z, Diamond MI. Huntington disease and the huntingtin protein. Prog Mol Biol Transl Sci. 2012;107:189–214. doi: 10.1016/B978-0-12-385883-2.00010-2. [DOI] [PubMed] [Google Scholar]
  • [190].Bossy-Wetzel E, Bossy-Wetzel E, Petrilli A, Knott AB. Mutant huntingtin and mitochondrial dysfunction. Trends Neurosci. 2008;31:609–616. doi: 10.1016/j.tins.2008.09.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [191].Oliveira JM. Mitochondrial bioenergetics and dynamics in Huntington’s disease: tripartite synapses and selective striatal degeneration. J Bioenerg Biomembr. 2010;42:227–234. doi: 10.1007/s10863-010-9287-6. [DOI] [PubMed] [Google Scholar]
  • [192].Greene JG, Greenamyre JT. Characterization of the excitotoxic potential of the reversible succinate dehydrogenase inhibitor malonate. J Neurochem. 1995;64:430–436. doi: 10.1046/j.1471-4159.1995.64010430.x. [DOI] [PubMed] [Google Scholar]
  • [193].Massieu L, Del Río P, Montiel T. Neurotoxicity of glutamate uptake inhibition in vivo: correlation with succinate dehydrogenase activity and prevention by energy substrates. Neuroscience. 2001;106:669–677. doi: 10.1016/s0306-4522(01)00323-2. [DOI] [PubMed] [Google Scholar]
  • [194].Panov AV, Gutekunst CA, Leavitt BR, Hayden MR, Burke JR, Strittmatter WJ, Greenamyre JT. Early mitochondrial calcium defects in Huntington’s disease are a direct effect of polyglutamines. Nat Neurosci. 2002;5:731–736. doi: 10.1038/nn884. [DOI] [PubMed] [Google Scholar]
  • [195].Grunewald T, Beal MF. Bioenergetics in Huntington’s disease. Ann NY AcadSci. 1999;893:203–213. doi: 10.1111/j.1749-6632.1999.tb07827.x. [DOI] [PubMed] [Google Scholar]
  • [196].Trushina E, Dyer RB, Badger II JD, Ure D, Eide L, Tran DD, et al. Mutant huntingtin impairs axonal trafficking in mammalian neurons in vivo and in vitro. Mol Cell Biol. 2004;24:8195–8209. doi: 10.1128/MCB.24.18.8195-8209.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [197].Costa V, Giacomello M, Hudec R, Lopreiato R, Ermak G, Lim D, et al. Mitochondrial fission and cristae disruption increase the response of cell models of Huntington’s disease to apoptotic stimuli. EMBO Mol Med. 2010;2:490–503. doi: 10.1002/emmm.201000102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [198].Song W, Chen J, Petrilli A, Liot G, Klinglmayr E, Zhou Y, et al. Mutant huntingtin binds the mitochondrial fission GTPasedynamin-related protein-1 and increases its enzymatic activity. Nat Med. 2011;17:377–382. doi: 10.1038/nm.2313. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [199].Jenkins BG, Koroshetz WJ, Beal MF, Rosen BR. Evidence for impairment of energy metabolism in vivo in Huntington’s disease using localized 1H NMR spectroscopy. Neurology. 1993;43:2689–2695. doi: 10.1212/wnl.43.12.2689. [DOI] [PubMed] [Google Scholar]
  • [200].Brennan WA, Jr, Bird ED, Aprille JR. Regional mitochondrial respiratory activity in Huntington’s disease brain. J Neurochem. 1985;44:1948–1950. doi: 10.1111/j.1471-4159.1985.tb07192.x. [DOI] [PubMed] [Google Scholar]
  • [201].Gu M, Gash MT, Mann VM, Javoy-Agid F, Cooper JM, Schapira AH. Mitochondrial defect in Huntington’s disease caudate nucleus. Ann Neurol. 1996;39:385–389. doi: 10.1002/ana.410390317. [DOI] [PubMed] [Google Scholar]
  • [202].Browne SE, Bowling AC, MacGarvey U, Baik MJ, Berger SC, Muqit MM, Bird ED, Beal MF. Oxidative damage and metabolic dysfunction in Huntington’s disease: selective vulnerability of the basal ganglia. Ann Neurol. 1997;41:646–653. doi: 10.1002/ana.410410514. [DOI] [PubMed] [Google Scholar]
  • [203].Ribeiro M, Silva AC, Rodrigues J, Naia L, Rego AC. Oxidizing Effects of Exogenous Stressors in Huntington’s Disease Knock-in Striatal Cells--Protective Effect of Cystamine and Creatine. Toxicol Sci. 2013 doi: 10.1093/toxsci/kft199. In press. [DOI] [PubMed] [Google Scholar]
  • [204].Fan MM, Raymond LA. N-methyl-D-aspartate (NMDA) receptor function and excitotoxicity in Huntington’s disease. Prog Neurobiol. 2007;81:272–293. doi: 10.1016/j.pneurobio.2006.11.003. [DOI] [PubMed] [Google Scholar]
  • [205].Johri A, Chandra A, Beal MF.PGC-1α, mitochondrial dysfunction, and Huntington’s disease. Free RadicBiol Med. 2013;62:37–46. doi: 10.1016/j.freeradbiomed.2013.04.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [206].Sturrock A, Laule C, Decolongon J, Dar Santos R, Coleman AJ, Creighton S, Bechtel N, Reilmann R, Hayden MR, Tabrizi SJ, Mackay AL, Leavitt BR. Magnetic resonance spectroscopy biomarkers in premanifest and early Huntington disease. Neurology. 2010;75:1702–1710. doi: 10.1212/WNL.0b013e3181fc27e4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [207].Unschuld PG, Edden RA, Carass A, Liu X, Shanahan M, Wang X, Oishi K, Brandt J, Bassett SS, Redgrave GW, Margolis RL, van Zijl PC, Barker PB, Ross CA. Brain metabolite alterations and cognitive dysfunction in early Huntington’s disease. Mov Disord. 2012;27:895–902. doi: 10.1002/mds.25010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [208].Tkáč I, Dubinsky JM, Keene CD, Gruetter R, Low WC. Neurochemical changes in Huntington R6/2 mouse striatum detected by in vivo1H NMR spectroscopy. J Neurochem. 2007;100:1397–1406. doi: 10.1111/j.1471-4159.2006.04323.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [209].Jenkins BG, Klivenyi P, Kustermann E, Andreassen OA, Ferrante RJ, Rosen BR, Beal MF. Nonlinear decrease over time in N-acetyl aspartate levels in the absence of neuronal loss and increases in glutamine and glucose in transgenic Huntington’s disease mice. J Neurochem. 2000;74:2108–2119. doi: 10.1046/j.1471-4159.2000.0742108.x. [DOI] [PubMed] [Google Scholar]
  • [210].Zacharoff L, Tkac I, Song Q, Tang C, Bolan PJ, Mangia S, Henry PG, Li T, Dubinsky JM. Cortical metabolites as biomarkers in the R6/2 model of Huntington’s disease. J Cereb Blood Flow Metab. 2012;32:502–14. doi: 10.1038/jcbfm.2011.157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [211].Jenkins BG, Andreassen OA, Dedeoglu A, Leavitt B, Hayden M, Borchelt D, Ross CA, Ferrante RJ, Beal MF. Effects of CAG repeat length, HTT protein length and protein context on cerebral metabolism measured using magnetic resonance spectroscopy in transgenic mouse models of Huntington’s disease. J Neurochem. 2005;95:553–562. doi: 10.1111/j.1471-4159.2005.03411.x. [DOI] [PubMed] [Google Scholar]
  • [212].Heikkinen T, Lehtimäki K, Vartiainen N, Puoliväli J, Hendricks SJ, Glaser JR, Bradaia A, Wadel K, Touller C, Kontkanen O, Yrjänheikki JM, Buisson B, Howland D, Beaumont V, Munoz-Sanjuan I, Park LC. Characterization of neurophysiological and behavioral changes, MRI brain volumetry and 1H MRS in zQ175 knock-in mouse model of Huntington’s disease. PLoS One. 2012;7:e50717. doi: 10.1371/journal.pone.0050717. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [213].Klapstein GJ, Fisher RS, Zanjani H, Cepeda C, Jokel ES, Chesselet MF, Levine MS. Electrophysiological and morphological changes in striatal spiny neurons in R6/2 Huntington’s disease transgenic mice. J Neurophysiol. 2001;86:2667–2677. doi: 10.1152/jn.2001.86.6.2667. [DOI] [PubMed] [Google Scholar]
  • [214].Choo YS, Mao Z, Johnson GV, Lesort M. Increased glutathione levels in cortical and striatal mitochondria of the R6/2 Huntington’s disease mouse model. Neurosci Lett. 2005;386:63–68. doi: 10.1016/j.neulet.2005.05.065. [DOI] [PubMed] [Google Scholar]
  • [215].Lee WT, Chang C. Magnetic resonance imaging and spectroscopy in assessing 3-nitropropionic acid-induced brain lesions: an animal model of Huntington’s disease. Prog Neurobiol. 2004;72:87–110. doi: 10.1016/j.pneurobio.2004.02.002. [DOI] [PubMed] [Google Scholar]
  • [216].Tsai MJ, Goh CC, Wan YL, Chang C. Metabolic alterations produced by 3-nitropropionic acid in rat striata and cultured astrocytes: quantitative in vitro 1H nuclear magnetic resonance spectroscopy and biochemical characterization. Neuroscience. 1997;79:819–826. doi: 10.1016/s0306-4522(97)00015-8. [DOI] [PubMed] [Google Scholar]

Articles from Aging and Disease are provided here courtesy of JKL International LLC

RESOURCES