Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 Alzheimer’s disease (AD) is a multifactorial, complex brain disease defined by progressive cognitive decline, heterogeneity of behavioral presentations, and dementia in older people [1-3]. + 9302 word(s) 9302 2020-10-20 08:01:57 |
2 format correct -6004 word(s) 3298 2020-11-05 10:03:34 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Ejaz, H.W.; Wang, W.; Lang, M. Copper Toxicity and Alzheimer’s Disease. Encyclopedia. Available online: https://encyclopedia.pub/entry/2902 (accessed on 29 March 2024).
Ejaz HW, Wang W, Lang M. Copper Toxicity and Alzheimer’s Disease. Encyclopedia. Available at: https://encyclopedia.pub/entry/2902. Accessed March 29, 2024.
Ejaz, Hafza Wajeeha, Wei Wang, Minglin Lang. "Copper Toxicity and Alzheimer’s Disease" Encyclopedia, https://encyclopedia.pub/entry/2902 (accessed March 29, 2024).
Ejaz, H.W., Wang, W., & Lang, M. (2020, November 05). Copper Toxicity and Alzheimer’s Disease. In Encyclopedia. https://encyclopedia.pub/entry/2902
Ejaz, Hafza Wajeeha, et al. "Copper Toxicity and Alzheimer’s Disease." Encyclopedia. Web. 05 November, 2020.
Copper Toxicity and Alzheimer’s Disease
Edit

Alzheimer’s disease (AD) is an irreversible, age-related progressive neurological disorder, and the most common type of dementia in aged people. Neuropathological lesions of AD are neurofibrillary tangles (NFTs), and senile plaques comprise the accumulated amyloid-beta (Aβ), loaded with metal ions including Cu, Fe, or Zn. Some reports have identified metal dyshomeostasis as a neurotoxic factor of AD, among which Cu ions seem to be a central cationic metal in the formation of plaque and soluble oligomers, and have an essential role in the AD pathology. Cu-Aβ complex catalyzes the generation of reactive oxygen species (ROS) and results in oxidative damage. Several studies have indicated that oxidative stress plays a crucial role in the pathogenesis of AD. The connection of copper levels in AD is still ambiguous, as some researches indicate a Cu deficiency, while others show its higher content in AD, and therefore there is a need to increase and decrease its levels in animal models, respectively, to study which one is the cause. For more than twenty years, many in vitro studies have been devoted to identifying metals’ roles in Aβ accumulation, oxidative damage, and neurotoxicity. Towards the end, a short review of the modern therapeutic approach in chelation therapy, with the main focus on Cu ions, is discussed. However, some recent reports of genetic-regulating copper transporters in AD models have shed light on treating this refractory disease. This study aims to succinctly present a better understanding of Cu ions’ current status in several AD features, and some conflicting reports are present herein.

Alzheimer’s disease amyloid plaques copper oxidative damages protein modification neurodegeneration

1. Introduction

In 1907, Alois Alzheimer was the first to identify a mental decline with amyloid plaques and neurofibrillary tangles found in most dementia symptoms [1][2]. This disorder’s main risk factor is old age, because the elderly are more prone to diseases, affecting 10% of people aged 65, and this proportion rises by about three times for people aged 85 and older [3][4] AD typically destroys neurons, and their connection with the brain regions such as the entorhinal cortex and hippocampus area, the parts of the brain essential in forming memories [5]. This disorder disrupts processes necessary for healthy neurons, such as communication, metabolism, and repair [6][7]. Ultimately, the disease is fatal. It is one of the leading causes of death [8] that we are currently unable to stop or cure because the underlying etiology is poorly understood at present [8][9].

Unfortunately, the treatment of AD has often been delayed in general because it is diagnosed only after prominent signs of cognitive deterioration [10], and this is all due to the lack of awareness of cognitive problems on the part of patients and patients’ families [11]. Clinical detection of this disorder is only possible when the symptoms are advanced enough to show visible behavior or cognitive changes [12]. There could be enough time to halt or slow this disorder’s development with early AD identification before complete onset [12]. Indeed, currently, there is no such treatment for AD [13], and approved drugs that have insignificant effects at altering the pathophysiological course of this disorder [14][15], due to the disease developing from a combination of lifestyle, environment, and genetic risk factors that affect the brain over time [16][17].

One of the most common neuropathological hallmarks of AD is the misfolding and aggregation of amyloid plaques-extracellular insoluble deposits of the β-amyloid peptides [18], and the intracellular formed NFTs (neurofibrillary tangles) [19], leading to the loss of communication between nerve cells, causes brain damage and shrinkage [20]. Posterior cingulated cortex (PCC) [21], entorhinal cortex (EC) [22], hippocampus (HIP) [23] (the first part to be affected by AD), middle temporal gyrus (MTG) (role in cognitive functions such as language processing), and superior frontal gyrus (SFG) (helps in memory) [24][25] are the regions affected in this multifactorial neurological disorder. Some studies have identified the impaired function of the middle temporal gyrus [26] and superior frontal gyrus in AD [27].

Extracellular deposits of Aβ peptides in Alzheimer’s are the main pathological events in AD [28][29][30][31]. Senile plaques or amyloid plaques mainly consist of small amyloid beta-peptides (Aβ) (up to 42 or 43 amino acids long) [32]. These are β amyloid precursor protein (APP) metabolites, derived by proteolytic sequential cleavage, first through β-secretase and then with γ-secretase, in the amyloidogenic pathway of producing peptides (Aβ), which contain 39 to 43 amino acids [33]. The APP (main isoforms, APP(695), APP(751), and APP(770)) is a type 1 transmembrane glycoprotein, which is essential for neurogenesis, neurite outgrowth, neuronal guidance, synapse formation, and repair [34][35][36]. The reason for neuritic plaques (senile plaques) forming in AD is due to irregularity between the production and removal of the beta-amyloid protein that accumulates [4]. Hence, the amyloid cascade hypothesis postulates that aggregation and accumulation of Aβ is the first pathological event in AD onset and initiates a cycle of adverse physiological changes that lead to neurodegeneration.

Another study has investigated Aβ aggregations in the senile plaques and co-localization of adenosine receptors in the AD [37]. Recently, some investigations have been done on adenosine, a purine ribonucleoside, because of its neuromodulator and neuroprotection function in neurological disorders [38][39]. It is present in all cells containing glia and neurons, initiates its biological process by four G-protein coupled receptors (GPCRs), namely, the A1, A2A,…A2BAR [40][41]. It has a role in regulating and integrating neuronal excitability, affecting many essential brain activities like sleep, memory, and neural plasticity [42][43][44]. Much research has analyzed adenosine effects via its receptors A1 and A2A in AD [45]. Nonselective blockage or modulation of these two receptors could protect cognitive impairment, making them innovative feasible therapeutic agents for AD [46]. Hippocampus, a brain region important for memory, learning, and neurogenesis [47][48][49], is one of the earliest affected brain regions that tends to exhibit the most rapid volume loss in the disease progression, and its pathology was found to be central to AD [47][50][51].

The hippocampus is a sensitive part of the brain to the dysfunctional homeostasis of transition metals, more so than any other brain region. Much research has also identified another brain part, the cortex, which is damaged by AD [52][53][54], linked with motor function, planning, organization, argumentation, feeling, and language processing [55]. NFTs are mainly composed of the microtubule-associated protein tau, predominantly expressed in the neurons under physiological conditions. This protein is mis-sorted into the somatodendritic compartment due to the tau sorting process’s failure, which is another essential factor that aggregates in AD [56]. Microtubules are essential components of a neuron’s cytoskeletal system, required for several fundamental cellular and dendritic processes, such as neuronal migration, polarity, axonal production, and differentiation [57][58]. Abnormal Aβ production might lead to the activation of tau mis-sorting, inducing tau pathology [59][60].

Multivalent metal ions such as copper (Cu) [61][62][63], zinc (Zn) [64][65], and iron (Fe) [66][67] are reported to be at higher levels in Alzheimer’s senile plaques [68][69]; while the connection of these metal ions with Aβ aggregation is still not well known. Indeed, some evidence from transgenic animal studies shows that Cu accumulates in senile plaques in the brains of 5 × FAD and Tg-SwDI/NOS2−/− mice models with neurodegeneration, as compared to PSAPP, where no Cu deposition has been seen among the mice with less neurodegeneration [70]. Much research has accumulated on Zn and Cu ions’ altered homeostasis as the central pathological hallmark [71][72][73] and shows the link of proteins related to Cu metabolism with this multifactorial AD [74].

Considerable research has suggested that Cu dyshomeostasis contributes to the onset of the most common neurodegenerative disorders besides AD, including Parkinson’s disease, prion-mediated encephalopathies, Huntington’s disease (HD), and amyotrophic lateral sclerosis (ALS) [75][76][77][78]. Hence, circumstances leading to a higher or lower copper concentration can be hazardous to health, such as Menkes diseases, a genetic disorder of Cu deficiency [79][80]. Furthermore, an autosomal recessive disorder, Wilson disease (WD), caused by defects of the ATP7B gene with excessive copper deposition in the body and patients’ brain examinations have shown copper concentration eight times greater than the controls [81].

Contradicting reports about the Cu concentration in AD has been reported. Some researches indicate a copper deficiency [82][83][84][85], while the majority show its higher level in AD, and therefore, reducing its level is required [86][87][88][89][90][91][92]. Investigations have grown exponentially in the neurodegenerative disorder fields over the past two decades. However, AD’s exact etiology is still not well understood, and as such, there is no successful therapeutic option available for this disorder to date [93][94].This literature review aims to present current knowledge regarding Cu’s role in AD. Towards the end, a short review of feasible therapeutics/strategies recommended for solving the problems associated with the metal’s implication in AD has also been discussed.

2. Contradictory Results about Copper Level in AD

In the body cells, Cu is absorbed through a high-affinity copper transporter Ctr1, incorporating cuprous (Cu+) ions from the intestinal microvilli’s surface. Little is known about Cu2+ absorption, which is probably absorbed by divalent metal transporter 1 (DMT1) or other shared metal transporters [95]. Ctr1is responsible for the majority (~70%) of Cu import into mammalian cells, from which Cu is passed to glutathione, which carries Cu through the cytoplasm [96]. The absorbed copper ions will be targeted to Cu-binding chaperones and enzymes in different cell compartments such as cytosolic, mitochondrial, and Golgi. In the cytosol, Cu chaperone for superoxide dismutase 1 (SOD1), CCS, mediates Cu+ loading. A recent study suggested that the direct transfer of copper from Ctr1 to chaperones and then passing it to SOD1 is via forming a Ctr1-CCS-SOD1 complex [95]. Besides CCS, soluble copper chaperones such as Atox1 and Cox17 can also escort Cu+ from Ctr1 in the cytosolic pool to facilitate copper supply to their specific target compartments [97].

Consequently, in the absence of Ctr1, other pathways to absorbed Cu ions are unavailable to the organism because of the sequestration of copper in the sub-apical vesicles. This has been confirmed by making the intestinal epithelial cell-specific knockout of the Ctr1 (Ctr1int/int) mice, which manifested severe Cu deficiency, and the majority died within three weeks of post-birth [98]. A considerable portion of ingested cuprous ions are passed into circulation in enterocytes to reach different tissues by Atox1/ATPase routes. The mouse model with inactivated ATOX1/ATP7A routes showed defects in Cu distribution, which leads to pathological variations in many organs, especially the brain[99].

In the CNS, Cu deficiency has been found in the hippocampus and amygdala regions of Alzheimer’s patients, which causes severe histopathologic alterations in AD. Additionally, scientific research has put forward that the frontal cortex tissue of Alzheimer’s patients had an increased susceptibility for exchangeable copper (CuEXC), which is associated with the overproduction of free radicals (ROS) in AD [100].

In the CSF of the AD patients, there is no significant change in Cu concentration as compared to that of the healthy cases (HC) [101]. Furthermore, within peripheral fluids, abnormal homeostasis of copper ions has been intensively investigated. The relevant data point to increased [101][102], decreased [85][103], or unchanged [104] serum or plasma Cu in Alzheimer’s patients. Many other scientific analyses have also reported excessive free or diffusible copper in serum [101][103][105]. However, Rembach (2013) has suggested the possibility of decreased non-CP copper levels-copper that is not bound to ceruloplasmin in mild cognitive impairment (MCI) and AD, which leads to a decline of copper-dependent biochemical activities in AD [106], such as reducing SOD1 activity of erythrocytes [85].

Cu association for AD is ambiguous as some substantial researches showed Cu deficiency in AD and, hence, it is required to increase Cu levels [83][84][85]. In contrast, many different scientific pieces of evidence demonstrated Cu overload, and thus it is necessary to reduce it [87][88][89][90][91][92]. The main updated explanation so far is that the abnormal Cu homeostasis is due to an increment in the labile Cu ions and a reduced attachment to proteins [107][108].

Until 2012, the published contradictory scientific researches fueled the debate of copper concentrations in AD. So far, to check Cu levels in various biological specimens of AD patients, such as serum, plasma, and CSF, six meta-analyses have been done during the past six years. Studies published from 1984 to 2017 have been included in these meta-analyses [101][109][110][111][112][113], which give unambiguous results: overall and unbound Cu both are present in higher concentrations in the serum-plasma samples of the AD patients compared to that in the healthy cases [111]. According to the very recent meta-analysis, which includes a total of 35 pieces of research: eighteen report an increase, fourteen show no change, and one reports a decrease in Cu level in the serum-plasma of this disorder [113]. Subsequently, three more studies have been published, stating increased Cu2+ ions level in Alzheimer’s compared to that in the healthy controls [114][115][116].

These recent researches have contributed considerably to the explanation of the previous controversy. In blood, a higher level of free plasma Cu, which has been identified in 50–60% of Alzheimer’s patients, can explain the higher level of serum Cu in AD [108][114][117][118]. Another earlier research also observed an increased concentration of serum copper ions in a special kind of AD (Alzheimer’s disease epsilon four apolipoprotein E allele carriers) [119]. According to some scientific investigation, a genetic basis may be the reason for this particular type of AD [119][120][121][122].

3. Therapeutics to Tackle Copper Ions in AD

Despite the exponential growth of scientific literature published in the neurodegenerative disorders area, especially for AD, the exact etiology of AD is still not well understood. To date, there is no successful therapeutic option available for this disorder [93][123]. While there is no cure, there are five FDA-approved medications to cope with the symptoms of AD, which may prevent this disease from getting worse over time [124].

In vitro, removal of Cu2+ from Aβ prevents its accumulation [125][126][127], leads to its degradation, stops hydroxyl radical (•OH) production and oxidative damage, and finally reduces cell death [127]. For the effects as mentioned above, researches have suggested potential metal chelation therapy for AD [128][129][130][131][132][133][134]. Nevertheless, the challenge is to build selective and specific metal chelators, as metal ions play crucial roles in Alzheimer’s brains. The first metal chelator made for arsenic toxicity in the 1940s was 2,3-Dimercaptopropanol (BAL) [135][136]. Much later, followed by the same approach, the first-generation of metal chelator, a lipophilic small molecule clioquinol (5-chloro-7-iodo-8HQ or CQ) was introduced at the end of the 1990s [126][137]. Transgenic mouse models treated with CQ showed promise by reducing Aβ accumulation by 50%. CQ reduced Aβ aggregation during Phase II trials and improved cognitive behavior, but failed to provide sufficient evidence of a positive clinical benefit in a larger clinical trial [138][139]. Furthermore, patients exhibited some severe side effects, including neurotoxicity and mutagenicity; therefore, further clinical trials of CQ were stopped.

The most progressive chelator so far is PBT2 (5,7-dichloro-2-((dimethylamino)methyl)) [140], a second-generation of scaffold-based chelator, which has been inspired by CQ and also showed excellent antioxidant properties [141][142][143]. It is a more effective Zn/Cu ionophore than CQ, which could decrease H2O2 formation, have greater BBB (blood-brain barrier) permeability, higher solubility, and could also inhibit Cu and Zn induced Aβ accumulation in vitro [143][144]. PBT2 treatment targets metal-induced damage [145], and most importantly, it prevents the loss of necessary metal ions from the body such as the kidney, liver, lungs, and brain [123][146]. It also shifts the Alzheimer’s phenotype within days by reducing insoluble Aβ levels by ~30% [143] and alters tau and synaptophysin protein levels’ phosphorylation. Interestingly, a lowered level of insoluble total and elevated levels of the soluble total tau has been shown in the treatment with PBT2 [141]. Despite the effects mentioned above, the results of human clinical trials are not up to the mark according to some studies [123][147]. However, some scholars have denied this idea [148]. Results from the phase IIb, the randomized clinical trial, were not as promising even though phase Ib/IIa preclinical trials demonstrated significant reductions in Aβ levels and improvement in various aspects of cognitive functioning.

The research in the Tg 2576 transgenic mice model has shown parenchymal plaque [149][150], indicating that metal chelators help slow disease progression and remove Cu ions only helpful in the initial stages of the AD [151]. While PS1 and PS2 play roles in Cu2+ uptake, tissue-specific knocking down of the single presenilins ortholog (PSN) in Drosophila reduces Cu2+ levels and increases its susceptibility to oxidative insult [152]. It was observed that the silencing of PSN in flies had less sensitivity to excess dietary Cu due to the reduced copper uptake. BLOC‐1 physically interacts with ATP7A, and disruption of the Drosophila’s dysbindin/BLOC-1 complex affects copper homeostasis in both mammalian cells and Drosophila [153].

Different approaches have been used to treat the pathological hallmarks of the multifactorial AD due to Cu dyshomeostasis, including the metal chelation therapy [143][154][155][156]. Restoring the intracellular copper decreases β-amyloid production, which was found through a mechanism that depends on the activation of phosphatidylinositol 3-kinase (PI3K)/PI3K-Akt pathway, and JNK (Jun N-terminal kinase) [157][158]. Moreover, lately, studies have observed increased intracellular Cu inhibited AD-causing Aβ peptide by direct targeting of presenilin (PS1 or PS2) subunits and nicastrin (NCT) in the γ-secretase complex [61][158][159]. Hence, higher intracellular Cu levels can improve cognitive function as well, by preventing β-amyloid aggregation and tau phosphorylation [160][161][162]. There is proof of the bis(thiosemicarbazone) copper(II) complex having the immunomodulatory potential [163][164], and greater BBB permeability. It inhibits microglial as well as astrocytic inflammatory responses and also has a role in the decrement of bacterial lipopolysaccharide (LPS) induced inflammation [165]. Some researchers have suggested that excess dietary Cu intake increases AD risks, and diets with measured copper should be supported. Additionally, a study conduct with a small amount of Cu in drinking water results in rising levels of amyloid peptides in the brain [166][167], a process that appears to be linked with dysfunction of LRP1-mediated efflux of Aβ from the brain [168] in vascular smooth muscle cells [169][170]. The median intake of copper from food among children and adolescents aged 2–19 years, the recommended daily allowance (RDA) ranges from 800 to 1000 mcg/day. In adults aged 20 and older, 1400–1700 mcg/day is recommended. Despite the difficulties, balancing Cu homeostasis has numerous advantages, and can be a potential drug target for this progressive, neurological disorder. [86][171][172][173],.

4. Conclusions

Collectively, studies strongly advocate that dyshomeostasis of Cu, leads to the onset and progression of AD. Earlier researches have recognized amyloid plaques as toxic factors in AD. Though 20–40% of healthy cases have amyloid plaques, as illustrated by some studies [174]. Furthermore, cell death often leads to amyloid plaque formation in the brain. While mounting evidence implicates ROS in the AD etiology, loosely bound copper ions are very efficient catalysts for ROS generation by a copper-amyloid complex [175][176].

Some studies indicate an increased liable pool of Cu in the brain [111][176] responsible for Cu deficiency. The reason of Cu deficiency seems to be an essential factor in AD. Copper deficiency leads to Cu enrichment in lipid rafts, so maybe an elevation in lipid raft domains could be the reason for Cu deficiency in the brain; thus, lipid raft domains could be an efficient drug target. Studies indicated that the disrupted lipid rafts (by omega-3 fatty acids) slowed the progression of AD [177][178]. Another direction for research depends on the feasibility of developing novel therapeutic approaches to work against this disease.

The proposal for direct chelation therapy of Cu ions to work in this disorder is still in discussion [147]. Support for the lowering cellular Cu levels comes from the Drosophila model of AD, where although copper chelation or genetic knockdown of copper transporters (Ctr1C) decreased the expression of Aβ degrading proteases but rescued the toxic phenotype [179]. Similar results were also observed by silencing the expression of Ctr1B, or when copper exporter DmATP7 [179] and dMTF-1 or MtnA [91] were overexpressed in the nervous system of the Aβ transgenic flies. These flies exhibited improved neurodegeneration, locomotion, longevity, and a reduction in Cu-Aβ complex-induced oxidative stress.

Furthermore, in parallel, antibody-based treatment for Aβ aggregation is now developing and providing safe results as well [180]. Research organizations should come to the same standpoint regarding the experimental requirements and procedures to be used, to avoid different and ambiguous results for such serious matters. In this context, all the struggles for a better understanding of AD pathology’s molecular mechanisms and developing innovative therapeutic approaches should be appreciated.

References

  1. Barber, R. C. The Genetics of Alzheimer’s Disease. Scientifica 2012, 2012, 1-14.
  2. Singh, S. K.; Srivastav, S.;  Yadav, A. K.;  Srikrishna, S.; Perry, G. Overview of Alzheimer's Disease and Some Therapeutic Approaches Targeting Aβ by Using Several Synthetic and Herbal Compounds. Oxid. Med. Cell. Longev. 2016, 2016, 7361613-7361613.
  3. Guerreiro, R.; Bras, J. The age factor in Alzheimer’s disease. Genome Med. 2015, 7, 106.
  4. Wan, L.; Nie, G.;  Zhang, J.;  Luo, Y.;  Zhang, P.;  Zhang, Z.; Zhao, B. β-Amyloid peptide increases levels of iron content and oxidative stress in human cell and Caenorhabditis elegans models of Alzheimer disease. Free Radic. Biol. Med. 2011, 50, 122-129.
  5. Wang, X.; L. Michaelis, M.; K. Michaelis, E. Functional Genomics of Brain Aging and Alzheimers Disease: Focus on Selective Neuronal Vulnerability. Current Genomics 2010, 11, 618-633.
  6. Cummings, J. L.; Vinters, H. V.;  Cole, G. M.; Khachaturian, Z. S. Alzheimer's disease: Etiologies, pathophysiology, cognitive reserve, and treatment opportunities. Neurology 1998, 51, S2-S17.
  7. Sheng, M.; Sabatini, B. L.; Sudhof, T. C. Synapses and Alzheimer's Disease. Cold Spring Harb. Perspect. Biol. 2012, 4, a005777-a005777.
  8. Santos, C. Y.; Snyder, P. J.;  Wu, W.-C.;  Zhang, M.;  Echeverria, A.; Alber, J. Pathophysiologic relationship between Alzheimer's disease, cerebrovascular disease, and cardiovascular risk: A review and synthesis. Alzheimer's & Dementia: Diagnosis, Assessment & Disease Monitoring 2017, 7, 69-87.
  9. van Norden, A. G. W.; van Dijk, E. J.;  de Laat, K. F.;  Scheltens, P.;  OldeRikkert, M. G. M.; de Leeuw, F. E. Dementia: Alzheimer pathology and vascular factors: From mutually exclusive to interaction. Biochimica et Biophysica Acta (BBA) - Molecular Basis of Disease 2012, 1822, 340-349.
  10. Judge, D.; Roberts, J.;  Khandker, R.;  Ambegaonkar, B.; Black, C. M. Physician perceptions about the barriers to prompt diagnosis of mild cognitive impairment and Alzheimer’s disease. International Journal of Alzheimer’s Disease 2019, 2019.
  11. Leskovjan, A. C.; Kretlow, A.;  Lanzirotti, A.;  Barrea, R.;  Vogt, S.; Miller, L. M. Increased brain iron coincides with early plaque formation in a mouse model of Alzheimer's disease. Neuroimage 2011, 55, 32-38.
  12. Torosyan, N.; Silverman, D. H. S. Neuronuclear Imaging in the Evaluation of Dementia and Mild Decline in Cognition. Semin. Nucl. Med. 2012, 42, 415-422.
  13. Akhondzadeh, S. New Hopes for Treatment of Alzheimer's Disease. Avicenna journal of medical biotechnology 2017, 10, 1-1.
  14. Conte-Daban, A.; Day, A.;  Faller, P.; Hureau, C. How Zn can impede Cu detoxification by chelating agents in Alzheimer's disease: a proof-of-concept study. Dalton Transactions 2016, 45, 15671-15678.
  15. Pickart, L.; Vasquez-Soltero, J.; Margolina, A. The Effect of the Human Peptide GHK on Gene Expression Relevant to Nervous System Function and Cognitive Decline. Brain Sciences 2017, 7, 20-20.
  16. Mielke, M. M.; Leoutsakos, J.-M.;  Corcoran, C. D.;  Green, R. C.;  Norton, M. C.;  Welsh-Bohmer, K. A.;  Tschanz, J. T.; Lyketsos, C. G. Effects of Food and Drug Administration-approved medications for Alzheimer’s disease on clinical progression. Alzheimer's & Dementia 2012, 8, 180-187.
  17. Nabeshima, T.; Kim, H.-C. Involvement of Genetic and Environmental Factors in the Onset of Depression. Exp. Neurobiol. 2013, 22, 235-235.
  18. Ashraf, G. M.; Greig, N. H.;  Khan, T. A.;  Hassan, I.;  Tabrez, S.;  Shakil, S.;  Sheikh, I. A.;  Zaidi, S. K.;  Akram, M.;  Jabir, N. R.;  Firoz, C. K.;  Naeem, A.;  Alhazza, I. M.;  Damanhouri, G. A.; Kamal, M. A. Protein misfolding and aggregation in Alzheimer's disease and type 2 diabetes mellitus. CNS Neurol. Disord. Drug Targets 2014, 13, 1280-93.
  19. Serrano-Pozo, A.; Frosch, M. P.;  Masliah, E.; Hyman, B. T. Neuropathological Alterations in Alzheimer Disease. Cold Spring Harb. Perspect. Med. 2011, 1, a006189-a006189.
  20. Furcila, D.; Domínguez-Álvaro, M.;  DeFelipe, J.; Alonso-Nanclares, L. Subregional Density of Neurons, Neurofibrillary Tangles and Amyloid Plaques in the Hippocampus of Patients With Alzheimer’s Disease. Front. Neuroanat. 2019, 13.
  21. Leech, R.; Sharp, D. J. The role of the posterior cingulate cortex in cognition and disease. Brain : a journal of neurology 2014, 137, 12-32.
  22. Maass, A.; Schütze, H.;  Speck, O.;  Yonelinas, A.;  Tempelmann, C.;  Heinze, H.-J.;  Berron, D.;  Cardenas-Blanco, A.;  Brodersen, K. H.;  Stephan, K. E.;  Düzel, E.;  Enno Stephan, K.; Düzel, E. Laminar activity in the hippocampus and entorhinal cortex related to novelty and episodic encoding. Nature Communications 2014, 5, 5547-5547.
  23. Mu, Y.; Gage, F. H. Adult hippocampal neurogenesis and its role in Alzheimer's disease. Mol. Neurodegener. 2011, 6, 85-85.
  24. Bubbico, G.; Chiacchiaretta, P.;  Parenti, M.;  Di Marco, M.;  Panara, V.;  Sepede, G.;  Ferretti, A.; Perrucci, M. G. Effects of second language learning on the plastic aging brain: functional connectivity, cognitive decline, and reorganization. Front. Neurosci. 2019, 13, 423.
  25. Jablonska, K.; Piotrowska, M.;  Bednarek, H.;  Szymaszek, A.;  Marchewka, A.;  Wypych, M.; Szelag, E. Maintenance vs. Manipulation in Auditory Verbal Working Memory in the Elderly: New Insights Based on Temporal Dynamics of Information Processing in the Millisecond Time Range. Front. Aging Neurosci. 2020, 12, 194.
  26. Liu, X.; Chen, W.;  Tu, Y.;  Hou, H.;  Huang, X.;  Chen, X.;  Guo, Z.; Bai, G. The abnormal functional connectivity between the hypothalamus and the temporal gyrus underlying depression in Alzheimer’s disease patients. Front. Aging Neurosci. 2018, 10, 37.
  27. Vasconcelos, L. G.; Jackowski, A. P.;  Oliveira, M. O.;  Ribeiro Flor, Y. M.;  Souza, A. A.;  Bueno, O. F.; Brucki, S. M. The thickness of posterior cortical areas is related to executive dysfunction in Alzheimer's disease. Clinics 2014, 69, 28-37.
  28. Kreutzer, A. G.; Yoo, S.;  Spencer, R. K.; Nowick, J. S. Stabilization, Assembly, and Toxicity of Trimers Derived from Aβ. J. Am. Chem. Soc. 2017, 139, 966-975.
  29. Sepulcre, J.; Schultz, A. P.;  Sabuncu, M.;  Gomez-Isla, T.;  Chhatwal, J.;  Becker, A.;  Sperling, R.; Johnson, K. A. In Vivo Tau, Amyloid, and Gray Matter Profiles in the Aging Brain. J. Neurosci. 2016, 36, 7364-7374.
  30. Buendia, I.; Parada, E.;  Navarro, E.;  León, R.;  Negredo, P.;  Egea, J.; López, M. G. Subthreshold Concentrations of Melatonin and Galantamine Improves Pathological AD-Hallmarks in Hippocampal Organotypic Cultures. Mol. Neurobiol. 2016, 53, 3338-3348.
  31. Reitz, C. Alzheimer's Disease and the Amyloid Cascade Hypothesis: A Critical Review. Int. J. Alzheimers Dis. 2012, 2012, 1-11.
  32. Murphy, M. P.; LeVine, H. Alzheimer's Disease and the Amyloid-β Peptide. J. Alzheimers Dis. 2010, 19, 311-323.
  33. Multhaup, G.; Huber, O.;  Buée, L.; Galas, M.-C. Amyloid Precursor Protein (APP) Metabolites APP Intracellular Fragment (AICD), Aβ42, and Tau in Nuclear Roles. J. Biol. Chem. 2015, 290, 23515-23522.
  34. Nicolas, M.; Hassan, B. A. Amyloid precursor protein and neural development. Development 2014, 141, 2543-2548.
  35. Nalivaeva, N. N.; Turner, A. J. The amyloid precursor protein: A biochemical enigma in brain development, function and disease. FEBS Lett. 2013, 587, 2046-2054.
  36. Zhou, Z.-d.; Chan, C. H.-s.;  Ma, Q.-h.;  Xu, X.-h.;  Xiao, Z.-c.; Tan, E.-K. The roles of amyloid precursor protein (APP) in neurogenesis. Cell Adhesion & Migration 2011, 5, 280-292.
  37. Franco, R.; Rivas-Santisteban, R.;  Casanovas, M.;  Lillo, A.;  Saura, C. A.; Navarro, G. Adenosine A2A Receptor Antagonists Affects NMDA Glutamate Receptor Function. Potential to Address Neurodegeneration in Alzheimer’s Disease. Cells 2020, 9, 1075.
  38. Cortés, A.; Gracia, E.;  Moreno, E.;  Mallol, J.;  Lluís, C.;  Canela, E. I.; Casadó, V. Moonlighting Adenosine Deaminase: A Target Protein for Drug Development. Med. Res. Rev. 2015, 35, 85-125.
  39. Maiuolo, J.; Oppedisano, F.;  Gratteri, S.;  Muscoli, C.; Mollace, V. Regulation of uric acid metabolism and excretion. Int. J. Cardiol. 2016, 213, 8-14.
  40. Cunha, R. A. Adenosine as a neuromodulator and as a homeostatic regulator in the nervous system: different roles, different sources and different receptors. Neurochem. Int. 2001, 38, 107-25.
  41. Sheth, S.; Brito, R.;  Mukherjea, D.;  Rybak, L.; Ramkumar, V. Adenosine Receptors: Expression, Function and Regulation. International Journal of Molecular Sciences 2014, 15, 2024-2052.
  42. Rau, A. R.; Ariwodola, O. J.; Weiner, J. L. Postsynaptic Adenosine A2A Receptors Modulate Intrinsic Excitability of Pyramidal Cells in the Rat Basolateral Amygdala. Int. J. Neuropsychopharmacol. 2015, 18.
  43. Gorgoni, M.; D'Atri, A.;  Lauri, G.;  Rossini, P. M.;  Ferlazzo, F.; De Gennaro, L. Is Sleep Essential for Neural Plasticity in Humans, and How Does It Affect Motor and Cognitive Recovery? Neural Plast. 2013, 2013, 1-13.
  44. Xiao, C.; Liu, N.;  Jacobson, K. A.;  Gavrilova, O.; Reitman, M. L. Physiology and effects of nucleosides in mice lacking all four adenosine receptors. PLoS Biol. 2019, 17, e3000161.
  45. Angulo, E.; Casadó, V.;  Mallol, J.;  Canela, E. I.;  Viñals, F.;  Ferrer, I.;  Lluis, C.; Franco, R. A1 adenosine receptors accumulate in neurodegenerative structures in Alzheimer disease and mediate both amyloid precursor protein processing and tau phosphorylation and translocation. Brain pathology (Zurich, Switzerland) 2003, 13, 440-51.
  46. Liu, Y. J.; Chen, J.;  Li, X.;  Zhou, X.;  Hu, Y. M.;  Chu, S. F.;  Peng, Y.;  Chen, N. H.; therapeutics Research progress on adenosine in central nervous system diseases. CNS neuroscience 2019, 25, 899-910.
  47. Lazarov, O.; Hollands, C. Hippocampal neurogenesis: Learning to remember. Prog. Neurobiol. 2016, 138-140, 1-18.
  48. Bengtsson, S. K. S.; Johansson, M.; Bäckström, T. Long-term continuous allopregnanolone elevation causes memory decline and hippocampus shrinkage, in female wild-type B6 mice. Horm. Behav. 2016, 78, 160-167.
  49. Ben Ahmed, O.; Benois-Pineau, J.;  Allard, M.;  Ben Amar, C.; Catheline, G. Classification of Alzheimer’s disease subjects from MRI using hippocampal visual features. Multimedia Tools and Applications 2015, 74, 1249-1266.
  50. La Joie, R.; Perrotin, A.;  de La Sayette, V.;  Egret, S.;  Doeuvre, L.;  Belliard, S.;  Eustache, F.;  Desgranges, B.; Chételat, G. Hippocampal subfield volumetry in mild cognitive impairment, Alzheimer's disease and semantic dementia. NeuroImage: Clinical 2013, 3, 155-162.
  51. Lindberg, O.; Walterfang, M.;  Looi, J. C. L.;  Malykhin, N.;  Östberg, P.;  Zandbelt, B.;  Styner, M.;  Paniagua, B.;  Velakoulis, D.;  Örndahl, E.; Wahlund, L.-O. Hippocampal Shape Analysis in Alzheimer's Disease and Frontotemporal Lobar Degeneration Subtypes. J. Alzheimers Dis. 2012, 30, 355-365.
  52. Gupta, Y.; Lee, K. H.;  Choi, K. Y.;  Lee, J. J.;  Kim, B. C.; Kwon, G.-R. Alzheimer’s Disease Diagnosis Based on Cortical and Subcortical Features. J. Healthc. Eng. 2019, 2019, 1-13.
  53. Castillo, E.; Leon, J.;  Mazzei, G.;  Abolhassani, N.;  Haruyama, N.;  Saito, T.;  Saido, T.;  Hokama, M.;  Iwaki, T.;  Ohara, T.;  Ninomiya, T.;  Kiyohara, Y.;  Sakumi, K.;  LaFerla, F. M.; Nakabeppu, Y. Comparative profiling of cortical gene expression in Alzheimer’s disease patients and mouse models demonstrates a link between amyloidosis and neuroinflammation. Sci. Rep. 2017, 7, 17762-17762.
  54. Levine, M. E.; Lu, A. T.;  Bennett, D. A.; Horvath, S. Epigenetic age of the pre-frontal cortex is associated with neuritic plaques, amyloid load, and Alzheimer’s disease related cognitive functioning. Aging 2015, 7, 1198-1211.
  55. Leisman, G.; Moustafa, A. A.; Shafir, T. Thinking, Walking, Talking: Integratory Motor and Cognitive Brain Function. Frontiers in public health 2016, 4, 94-94.
  56. Götz, J.; Xia, D.;  Leinenga, G.;  Chew, Y. L.; Nicholas, H. What Renders TAU Toxic. Front. Neurol. 2013, 4.
  57. Lasser, M.; Tiber, J.; Lowery, L. A. The Role of the Microtubule Cytoskeleton in Neurodevelopmental Disorders. Front. Cell. Neurosci. 2018, 12.
  58. Kapitein, L. C.; Hoogenraad, C. C. Building the Neuronal Microtubule Cytoskeleton. Neuron 2015, 87, 492-506.
  59. Morris, M.; Knudsen, G. M.;  Maeda, S.;  Trinidad, J. C.;  Ioanoviciu, A.;  Burlingame, A. L.; Mucke, L. Tau post-translational modifications in wild-type and human amyloid precursor protein transgenic mice. Nat. Neurosci. 2015, 18, 1183-1189.
  60. Zempel, H.; Mandelkow, E. Lost after translation: missorting of Tau protein and consequences for Alzheimer disease. Trends Neurosci. 2014, 37, 721-732.
  61. Bagheri, S.; Squitti, R.;  Haertlé, T.;  Siotto, M.; Saboury, A. A. Role of Copper in the Onset of Alzheimer’s Disease Compared to Other Metals. Front. Aging Neurosci. 2018, 9.
  62. Greenough, M. A.; Ramírez Munoz, A.;  Bush, A. I.; Opazo, C. M. Metallo-pathways to Alzheimer's disease: lessons from genetic disorders of copper trafficking. Metallomics 2016, 8, 831-839.
  63. Xu, J.; Begley, P.;  Church, S. J.;  Patassini, S.;  McHarg, S.;  Kureishy, N.;  Hollywood, K. A.;  Waldvogel, H. J.;  Liu, H.;  Zhang, S.;  Lin, W.;  Herholz, K.;  Turner, C.;  Synek, B. J.;  Curtis, M. A.;  Rivers-Auty, J.;  Lawrence, C. B.;  Kellett, K. A. B.;  Hooper, N. M.;  Vardy, E. R. L. C.;  Wu, D.;  Unwin, R. D.;  Faull, R. L. M.;  Dowsey, A. W.; Cooper, G. J. S. Elevation of brain glucose and polyol-pathway intermediates with accompanying brain-copper deficiency in patients with Alzheimer’s disease: metabolic basis for dementia. Sci. Rep. 2016, 6, 27524-27524.
  64. Mezentsev, Y. V.; Medvedev, A. E.;  Kechko, O. I.;  Makarov, A. A.;  Ivanov, A. S.;  Mantsyzov, A. B.; Kozin, S. A. Zinc-induced heterodimer formation between metal-binding domains of intact and naturally modified amyloid-beta species: implication to amyloid seeding in Alzheimer’s disease? J. Biomol. Struct. Dyn. 2016, 34, 2317-2326.
  65. Kulikova, A. A.; Makarov, A. A.; Kozin, S. A. Roles of zinc ions and structural polymorphism of β-amyloid in the development of Alzheimer’s disease. Mol. Biol. 2015, 49, 217-230.
  66. Sands, S. A.; Leung-Toung, R.;  Wang, Y.;  Connelly, J.; LeVine, S. M. Enhanced Histochemical Detection of Iron in Paraffin Sections of Mouse Central Nervous System Tissue. ASN Neuro 2016, 8, 175909141667097-175909141667097.
  67. James, S. A.; Churches, Q. I.;  de Jonge, M. D.;  Birchall, I. E.;  Streltsov, V.;  McColl, G.;  Adlard, P. A.; Hare, D. J. Iron, Copper, and Zinc Concentration in Aβ Plaques in the APP/PS1 Mouse Model of Alzheimer’s Disease Correlates with Metal Levels in the Surrounding Neuropil. ACS Chem. Neurosci. 2017, 8, 629-637.
  68. Summers, K. L.; Schilling, K. M.;  Roseman, G.;  Markham, K. A.;  Dolgova, N. V.;  Kroll, T.;  Sokaras, D.;  Millhauser, G. L.;  Pickering, I. J.; George, G. N. X-ray absorption spectroscopy investigations of copper (II) coordination in the human amyloid β peptide. Inorg. Chem. 2019, 58, 6294-6311.
  69. Ji, M.; Arbel, M.;  Zhang, L.;  Freudiger, C. W.;  Hou, S. S.;  Lin, D.;  Yang, X.;  Bacskai, B. J.; Xie, X. S. Label-free imaging of amyloid plaques in Alzheimer’s disease with stimulated Raman scattering microscopy. Science advances 2018, 4, eaat7715.
  70. Bourassa, M. W.; Leskovjan, A. C.;  Tappero, R. V.;  Farquhar, E. R.;  Colton, C. A.;  Van Nostrand, W. E.; Miller, L. M. Elevated copper in the amyloid plaques and iron in the cortex are observed in mouse models of Alzheimer's disease that exhibit neurodegeneration. Biomedical spectroscopy and imaging 2013, 2, 129-139.
  71. Arnal, N.; Castillo, O.;  de Alaniz, M. J. T.; Marra, C. A. Effects of Copper and/or Cholesterol Overload on Mitochondrial Function in a Rat Model of Incipient Neurodegeneration. Int. J. Alzheimers Dis. 2013, 2013, 1-14.
  72. Mao, X.; Ye, J.;  Zhou, S.;  Pi, R.;  Dou, J.;  Zang, L.;  Chen, X.;  Chao, X.;  Li, W.;  Liu, M.; Liu, P. The effects of chronic copper exposure on the amyloid protein metabolisim associated genes’ expression in chronic cerebral hypoperfused rats. Neurosci. Lett. 2012, 518, 14-18.
  73. Hozumi, I.; Hasegawa, T.;  Honda, A.;  Ozawa, K.;  Hayashi, Y.;  Hashimoto, K.;  Yamada, M.;  Koumura, A.;  Sakurai, T.;  Kimura, A.;  Tanaka, Y.;  Satoh, M.; Inuzuka, T. Patterns of levels of biological metals in CSF differ among neurodegenerative diseases. J. Neurol. Sci. 2011, 303, 95-99.
  74. Pal, A.; Kumar, A.; Prasad, R. Predictive association of copper metabolism proteins with Alzheimer’s disease and Parkinson’s disease: a preliminary perspective. Biometals 2014, 27, 25-31.
  75. Xiao, G.; Fan, Q.;  Wang, X.; Zhou, B. Huntington disease arises from a combinatory toxicity of polyglutamine and copper binding. Proceedings of the National Academy of Sciences 2013, 110, 14995-15000.
  76. Siggs, O. M.; Cruite, J. T.;  Du, X.;  Rutschmann, S.;  Masliah, E.;  Beutler, B.; Oldstone, M. B. A. Disruption of copper homeostasis due to a mutation of Atp7a delays the onset of prion disease. Proceedings of the National Academy of Sciences 2012, 109, 13733-13738.
  77. Rose, F.; Hodak, M.; Bernholc, J. Mechanism of copper(II)-induced misfolding of Parkinson's disease protein. Sci. Rep. 2011, 1, 11-11.
  78. Souza, P. C.; Thallmair, S.;  Marrink, S. J.; Mera-Adasme, R. l. An Allosteric Pathway in Copper, Zinc Superoxide Dismutase Unravels the Molecular Mechanism of the G93A Amyotrophic Lateral Sclerosis-Linked Mutation. The journal of physical chemistry letters 2019, 10, 7740-7744.
  79. Delangle, P.; Mintz, E. Chelation therapy in Wilson's disease: from d-penicillamine to the design of selective bioinspired intracellular Cu(i) chelators. Dalton Transactions 2012, 41, 6359-6359.
  80. Lutsenko, S.; Bhattacharjee, A.; Hubbard, A. L. Copper handling machinery of the brain. Metallomics 2010, 2, 596-596.
  81. Squitti, R.; Ghidoni, R.;  Simonelli, I.;  Ivanova, I. D.;  Colabufo, N. A.;  Zuin, M.;  Benussi, L.;  Binetti, G.;  Cassetta, E.;  Rongioletti, M.; Biology Copper dyshomeostasis in Wilson disease and Alzheimer's disease as shown by serum and urine copper indicators. Journal of Trace Elements in Medicine 2018, 45, 181-188.
  82. Xu, J.; Church, S. J.;  Patassini, S.;  Begley, P.;  Waldvogel, H. J.;  Curtis, M. A.;  Faull, R. L. M.;  Unwin, R. D.; Cooper, G. J. S. Evidence for widespread, severe brain copper deficiency in Alzheimer's dementia. Metallomics 2017, 9, 1106-1119.
  83. Exley, C.; House, E.;  Polwart, A.; Esiri, M. M. Brain Burdens of Aluminum, Iron, and Copper and their Relationships with Amyloid-β Pathology in 60 Human Brains. J. Alzheimers Dis. 2012, 31, 725-730.
  84. Kaden, D.; Bush, A. I.;  Danzeisen, R.;  Bayer, T. A.; Multhaup, G. Disturbed Copper Bioavailability in Alzheimer's Disease. Int. J. Alzheimers Dis. 2011, 2011, 1-5.
  85. Vural, H.; Demirin, H.;  Kara, Y.;  Eren, I.; Delibas, N. Alterations of plasma magnesium, copper, zinc, iron and selenium concentrations and some related erythrocyte antioxidant enzyme activities in patients with Alzheimer’s disease. J. Trace Elem. Med. Biol. 2010, 24, 169-173.
  86. Yu, J.; Luo, X.;  Xu, H.;  Ma, Q.;  Yuan, J.;  Li, X.;  Chang, R. C.-C.;  Qu, Z.;  Huang, X.;  Zhuang, Z.;  Liu, J.; Yang, X. Identification of the Key Molecules Involved in Chronic Copper Exposure-Aggravated Memory Impairment in Transgenic Mice of Alzheimer's Disease Using Proteomic Analysis. J. Alzheimers Dis. 2015, 44, 455-469.
  87. Squitti, R.; Siotto, M.; Polimanti, R. Low-copper diet as a preventive strategy for Alzheimer's disease. Neurobiol. Aging 2014, 35, S40-S50.
  88. Brewer, G. J. Alzheimer’s disease causation by copper toxicity and treatment with zinc. Front. Aging Neurosci. 2014, 6.
  89. Ceccom, J.; Coslédan, F.;  Halley, H.;  Francès, B.;  Lassalle, J. M.; Meunier, B. Copper Chelator Induced Efficient Episodic Memory Recovery in a Non-Transgenic Alzheimer’s Mouse Model. PLoS One 2012, 7, e43105-e43105.
  90. Eskici, G.; Axelsen, P. H. Copper and Oxidative Stress in the Pathogenesis of Alzheimer’s Disease. Biochemistry 2012, 51, 6289-6311.
  91. Hua, H.; Münter, L.;  Harmeier, A.;  Georgiev, O.;  Multhaup, G.; Schaffner, W. Toxicity of Alzheimer's disease-associated Aβ peptide is ameliorated in a Drosophila model by tight control of zinc and copper availability. Biol. Chem. 2011, 392.
  92. Luo, Y.; Zhang, J.;  Liu, N.;  Luo, Y.; Zhao, B. Copper ions influence the toxicity of β-amyloid(1–42) in a concentration-dependent manner in a Caenorhabditis elegans model of Alzheimer’s disease. Science China Life Sciences 2011, 54, 527-534.
  93. Yiannopoulou, K. G.; Papageorgiou, S. G. Current and future treatments for Alzheimer’s disease. Ther. Adv. Neurol. Disord. 2013, 6, 19-33.
  94. Karthivashan, G.; Ganesan, P.;  Park, S.-Y.;  Kim, J.-S.; Choi, D.-K. Therapeutic strategies and nano-drug delivery applications in management of ageing Alzheimer’s disease. Drug delivery 2018, 25, 307-320.
  95. Skopp, A.; Boyd, S. D.;  Ullrich, M. S.;  Liu, L.; Winkler, D. D. Copper–zinc superoxide dismutase (Sod1) activation terminates interaction between its copper chaperone (Ccs) and the cytosolic metal-binding domain of the copper importer Ctr1. Biometals 2019, 32, 695-705.
  96. Ackerman, C. M.; Chang, C. J. Copper signaling in the brain and beyond. Journal of Biological Chemistry 2018, 293, 4628-4635.
  97. Witt, B.; Schaumlöffel, D.; Schwerdtle, T. Subcellular Localization of Copper—Cellular Bioimaging with Focus on Neurological Disorders. International Journal of Molecular Sciences 2020, 21, 2341.
  98. Nose, Y.; Wood, L. K.;  Kim, B.-E.;  Prohaska, J. R.;  Fry, R. S.;  Spears, J. W.; Thiele, D. J. Ctr1 is an apical copper transporter in mammalian intestinal epithelial cells in vivo that is controlled at the level of protein stability. J. Biol. Chem. 2010, 285, 32385-32392.
  99. Hodgkinson, V. L.; Dale, J. M.;  Garcia, M. L.;  Weisman, G. A.;  Lee, J.;  Gitlin, J. D.; Petris, M. J. X‐linked spinal muscular atrophy in mice caused by autonomous loss of ATP7A in the motor neuron. The Journal of pathology 2015, 236, 241-250.
  100. Deibel, M. A.; Ehmann, W. D.; Markesbery, W. R. Copper, iron, and zinc imbalances in severely degenerated brain regions in Alzheimer's disease: possible relation to oxidative stress. J. Neurol. Sci. 1996, 143, 137-142.
  101. Bucossi, S.; Ventriglia, M.;  Panetta, V.;  Salustri, C.;  Pasqualetti, P.;  Mariani, S.;  Siotto, M.;  Rossini, P. M.; Squitti, R. Copper in Alzheimer's Disease: A Meta-Analysis of Serum, Plasma, and Cerebrospinal Fluid Studies. J. Alzheimers Dis. 2011, 24, 175-185.
  102. Squitti, R.; Lupoi, D.;  Pasqualetti, P.;  Dal Forno, G.;  Vernieri, F.;  Chiovenda, P.;  Rossi, L.;  Cortesi, M.;  Cassetta, E.; Rossini, P. M. Elevation of serum copper levels in Alzheimer's disease. Neurology 2002, 59, 1153-1161.
  103. Brewer, G. J.; Kanzer, S. H.;  Zimmerman, E. A.;  Celmins, D. F.;  Heckman, S. M.; Dick, R. Copper and Ceruloplasmin Abnormalities in Alzheimer’s Disease. American Journal of Alzheimer's Disease & Other Dementiasr 2010, 25, 490-497.
  104. Ozcankaya, R.; Delibas, N. Malondialdehyde, superoxide dismutase, melatonin, iron, copper, and zinc blood concentrations in patients with Alzheimer disease: cross-sectional study. Croat. Med. J. 2002, 43, 28-32.
  105. Squitti, R.; Ghidoni, R.;  Scrascia, F.;  Benussi, L.;  Panetta, V.;  Pasqualetti, P.;  Moffa, F.;  Bernardini, S.;  Ventriglia, M.;  Binetti, G.; Rossini, P. M. Free Copper Distinguishes Mild Cognitive Impairment Subjects from Healthy Elderly Individuals. J. Alzheimers Dis. 2011, 23, 239-248.
  106. Rembach, A.; Doecke, J. D.;  Roberts, B. R.;  Watt, A. D.;  Faux, N. G.;  Volitakis, I.;  Pertile, K. K.;  Rumble, R. L.;  Trounson, B. O.;  Fowler, C. J.;  Wilson, W.;  Ellis, K. A.;  Martins, R. N.;  Rowe, C. C.;  Villemagne, V. L.;  Ames, D.;  Masters, C. L.; Bush, A. I. Longitudinal Analysis of Serum Copper and Ceruloplasmin in Alzheimer's Disease. J. Alzheimers Dis. 2013, 34, 171-182.
  107. Kepp, K. P. Alzheimer’s disease due to loss of function: A new synthesis of the available data. Prog. Neurobiol. 2016, 143, 36-60.
  108. Squitti, R.; Siotto, M.;  Arciello, M.; Rossi, L. Non-ceruloplasmin bound copper and ATP7B gene variants in Alzheimer's disease. Metallomics 2016, 8, 863-873.
  109. Wang, Z.-X.; Tan, L.;  Wang, H.-F.;  Ma, J.;  Liu, J.;  Tan, M.-S.;  Sun, J.-H.;  Zhu, X.-C.;  Jiang, T.; Yu, J.-T. Serum Iron, Zinc, and Copper Levels in Patients with Alzheimer’s Disease: A Replication Study and Meta-Analyses. J. Alzheimers Dis. 2015, 47, 565-581.
  110. Ventriglia, M.; Bucossi, S.;  Panetta, V.; Squitti, R. Copper in Alzheimer's Disease: A Meta-Analysis of Serum, Plasma, and Cerebrospinal Fluid Studies. J. Alzheimers Dis. 2012, 30, 981-984.
  111. Squitti, R.; Simonelli, I.;  Ventriglia, M.;  Siotto, M.;  Pasqualetti, P.;  Rembach, A.;  Doecke, J.; Bush, A. I. Meta-Analysis of Serum Non-Ceruloplasmin Copper in Alzheimer's Disease. J. Alzheimers Dis. 2013, 38, 809-822.
  112. Schrag, M.; Mueller, C.;  Zabel, M.;  Crofton, A.;  Kirsch, W. M.;  Ghribi, O.;  Squitti, R.; Perry, G. Oxidative stress in blood in Alzheimer's disease and mild cognitive impairment: A meta-analysis. Neurobiol. Dis. 2013, 59, 100-110.
  113. Li, D.-D.; Zhang, W.;  Wang, Z.-Y.; Zhao, P. Serum Copper, Zinc, and Iron Levels in Patients with Alzheimer's Disease: A Meta-Analysis of Case-Control Studies. Front. Aging Neurosci. 2017, 9.
  114. Talwar, P.; Grover, S.;  Sinha, J.;  Chandna, P.;  Agarwal, R.;  Kushwaha, S.; Kukreti, R. Multifactorial Analysis of a Biomarker Pool for Alzheimer Disease Risk in a North Indian Population. Dement. Geriatr. Cogn. Disord. 2017, 44, 25-34.
  115. Pu, Z.; Xu, W.;  Lin, Y.;  He, J.; Huang, M. Oxidative Stress Markers and Metal Ions are Correlated With Cognitive Function in Alzheimer’s Disease. American Journal of Alzheimer's Disease & Other Dementiasr 2017, 32, 353-359.
  116. Guan, C.; Dang, R.;  Cui, Y.;  Liu, L.;  Chen, X.;  Wang, X.;  Zhu, J.;  Li, D.;  Li, J.; Wang, D. Characterization of plasma metal profiles in Alzheimer’s disease using multivariate statistical analysis. PLoS One 2017, 12, e0178271-e0178271.
  117. Szabo, S. T.; Harry, G. J.;  Hayden, K. M.;  Szabo, D. T.; Birnbaum, L. Comparison of Metal Levels between Postmortem Brain and Ventricular Fluid in Alzheimer’s Disease and Nondemented Elderly Controls. Toxicol. Sci. 2016, 150, 292-300.
  118. Tecchio, F.; Vecchio, F.;  Ventriglia, M.;  Porcaro, C.;  Miraglia, F.;  Siotto, M.;  M. Rossini, P.;  Rongioletti, M.; Squitti, R. Non-Ceruloplasmin Copper Distinguishes A Distinct Subtype of Alzheimer's Disease: A Study of EEG-Derived Brain Activity. Current Alzheimer Research 2016, 13, 1374-1384.
  119. Gonzalez; Martin;  Cacho;  Brenas;  Arroyo;  Garcia, B.;  Navajo; Gonzalez, B. Serum zinc, copper, insulin and lipids in Alzheimer's disease epsilon 4 apolipoprotein E allele carriers*. Eur. J. Clin. Invest. 1999, 29, 637-642.
  120. Squitti, R.; Ventriglia, M.;  Gennarelli, M.;  Colabufo, N. A.;  El Idrissi, I. G.;  Bucossi, S.;  Mariani, S.;  Rongioletti, M.;  Zanetti, O.;  Congiu, C.;  Rossini, P. M.; Bonvicini, C. Non-Ceruloplasmin Copper Distincts Subtypes in Alzheimer’s Disease: a Genetic Study of ATP7B Frequency. Mol. Neurobiol. 2017, 54, 671-681.
  121. Mercer, S. W.; Wang, J.; Burke, R. In Vivo Modeling of the Pathogenic Effect of Copper Transporter Mutations That Cause Menkes and Wilson Diseases, Motor Neuropathy, and Susceptibility to Alzheimer's Disease. J. Biol. Chem. 2017, 292, 4113-4122.
  122. Squitti, R.; Polimanti, R.;  Siotto, M.;  Bucossi, S.;  Ventriglia, M.;  Mariani, S.;  Vernieri, F.;  Scrascia, F.;  Trotta, L.; Rossini, P. M. ATP7B Variants as Modulators of Copper Dyshomeostasis in Alzheimer’s Disease. Neuromolecular Med. 2013, 15, 515-522.
  123. Hegde, M. L.; Bharathi, P.;  Suram, A.;  Venugopal, C.;  Jagannathan, R.;  Poddar, P.;  Srinivas, P.;  Sambamurti, K.;  Rao, K. J.;  Scancar, J.;  Messori, L.;  Zecca, L.; Zatta, P. Challenges Associated with Metal Chelation Therapy in Alzheimer's Disease. Journal of Alzheimer's disease : JAD 2009, 17, 457-468.
  124. Y.Y. Szeto, J.; J.G. Lewis, S. Current Treatment Options for Alzheimer's Disease and Parkinson's Disease Dementia. Curr. Neuropharmacol. 2016, 14, 326-338.
  125. Behbehani, G. R.; Barzegar, L.;  Mohebbian, M.; Saboury, A. A. A Comparative Interaction between Copper Ions with Alzheimer's β Amyloid Peptide and Human Serum Albumin. Bioinorg. Chem. Appl. 2012, 2012, 1-4.
  126. Cherny, R. A.; Atwood, C. S.;  Xilinas, M. E.;  Gray, D. N.;  Jones, W. D.;  McLean, C. A.;  Barnham, K. J.;  Volitakis, I.;  Fraser, F. W.;  Kim, Y.-S.;  Huang, X.;  Goldstein, L. E.;  Moir, R. D.;  Lim, J. T.;  Beyreuther, K.;  Zheng, H.;  Tanzi, R. E.;  Masters, C. L.; Bush, A. I. Treatment with a Copper-Zinc Chelator Markedly and Rapidly Inhibits β-Amyloid Accumulation in Alzheimer's Disease Transgenic Mice. Neuron 2001, 30, 665-676.
  127. Wu, W.-h.; Lei, P.;  Liu, Q.;  Hu, J.;  Gunn, A. P.;  Chen, M.-s.;  Rui, Y.-f.;  Su, X.-y.;  Xie, Z.-p.;  Zhao, Y.-F.;  Bush, A. I.; Li, Y.-m. Sequestration of Copper from β-Amyloid Promotes Selective Lysis by Cyclen-Hybrid Cleavage Agents. J. Biol. Chem. 2008, 283, 31657-31664.
  128. Yang, T.; Yang, L.;  Zhang, C.;  Wang, Y.;  Ma, X.;  Wang, K.;  Luo, J.;  Yao, C.;  Wang, X.; Wang, X. A copper–amyloid-β targeted fluorescent chelator as a potential theranostic agent for Alzheimer's disease. Inorganic Chemistry Frontiers 2016, 3, 1572-1581.
  129. Geng, J.; Li, M.;  Wu, L.;  Ren, J.; Qu, X. Liberation of Copper from Amyloid Plaques: Making a Risk Factor Useful for Alzheimer’s Disease Treatment. J. Med. Chem. 2012, 55, 9146-9155.
  130. Hauser-Davis, R. A.; de Freitas, L. V.;  Cukierman, D. S.;  Cruz, W. S.;  Miotto, M. C.;  Landeira-Fernandez, J.;  Valiente-Gabioud, A. A.;  Fernández, C. O.; Rey, N. A. Disruption of zinc and copper interactions with Aβ(1–40) by a non-toxic, isoniazid-derived, hydrazone: a novel biometal homeostasis restoring agent in Alzheimer's disease therapy? Metallomics 2015, 7, 743-747.
  131. Hung, V. W. S.; Bressan, L. P.;  Seo, K.; Kerman, K. Electroanalysis of Natural Compounds as Copper Chelating Agents for Alzheimer’s Disease Therapy. Electroanalysis 2015, 27, 2670-2678.
  132. Liu, G.; Garrett, M. R.;  Men, P.;  Zhu, X.;  Perry, G.; Smith, M. A. Nanoparticle and other metal chelation therapeutics in Alzheimer disease. Biochimica et Biophysica Acta (BBA) - Molecular Basis of Disease 2005, 1741, 246-252.
  133. Moret, V.; Laras, Y.;  Pietrancosta, N.;  Garino, C.;  Quéléver, G.;  Rolland, A.;  Mallet, B.;  Norreel, J.-C.; Kraus, J.-L. 1,1′-Xylyl bis-1,4,8,11-tetraaza cyclotetradecane: A new potential copper chelator agent for neuroprotection in Alzheimer’s disease. Its comparative effects with clioquinol on rat brain copper distribution. Bioorg. Med. Chem. Lett. 2006, 16, 3298-3301.
  134. Nguyen, M.; Robert, A.;  Sournia-Saquet, A.;  Vendier, L.; Meunier, B. Characterization of New Specific Copper Chelators as Potential Drugs for the Treatment of Alzheimer’s Disease. Chemistry - A European Journal 2014, 20, 6771-6785.
  135. Scott, L. E.; Orvig, C. Medicinal Inorganic Chemistry Approaches to Passivation and Removal of Aberrant Metal Ions in Disease. Chem. Rev. 2009, 109, 4885-4910.
  136. Kosnett, M. J. The Role of Chelation in the Treatment of Arsenic and Mercury Poisoning. J. Med. Toxicol. 2013, 9, 347-354.
  137. Cahoon, L. The curious case of clioquinol. Nat. Med. 2009, 15, 356-359.
  138. Barnham, K. J.; Bush, A. I. Metals in Alzheimer's and Parkinson's Diseases. Curr. Opin. Chem. Biol. 2008, 12, 222-228.
  139. Bush, A. I. Drug development based on the metals hypothesis of Alzheimer's disease. Journal of Alzheimer's disease : JAD 2008, 15, 223-40.
  140. Liang, S. H.; Southon, A. G.;  Fraser, B. H.;  Krause-Heuer, A. M.;  Zhang, B.;  Shoup, T. M.;  Lewis, R.;  Volitakis, I.;  Han, Y.;  Greguric, I.;  Bush, A. I.; Vasdev, N. Novel Fluorinated 8-Hydroxyquinoline Based Metal Ionophores for Exploring the Metal Hypothesis of Alzheimer’s Disease. ACS Med. Chem. Lett. 2015, 6, 1025-1029.
  141. Rodríguez-Rodríguez, C.; Telpoukhovskaia, M.; Orvig, C. The art of building multifunctional metal-binding agents from basic molecular scaffolds for the potential application in neurodegenerative diseases. Coord. Chem. Rev. 2012, 256, 2308-2332.
  142. Prachayasittikul, V.; Prachayasittikul, V.;  Prachayasittikul, S.; Ruchirawat, S. 8-Hydroxyquinolines: a review of their metal chelating properties and medicinal applications. Drug Des. Devel. Ther. 2013, 1157-1157.
  143. Adlard, P. A.; Cherny, R. A.;  Finkelstein, D. I.;  Gautier, E.;  Robb, E.;  Cortes, M.;  Volitakis, I.;  Liu, X.;  Smith, J. P.;  Perez, K.;  Laughton, K.;  Li, Q.-X.;  Charman, S. A.;  Nicolazzo, J. A.;  Wilkins, S.;  Deleva, K.;  Lynch, T.;  Kok, G.;  Ritchie, C. W.;  Tanzi, R. E.;  Cappai, R.;  Masters, C. L.;  Barnham, K. J.; Bush, A. I. Rapid Restoration of Cognition in Alzheimer's Transgenic Mice with 8-Hydroxy Quinoline Analogs Is Associated with Decreased Interstitial Aβ. Neuron 2008, 59, 43-55.
  144. Rana, M.; Sharma, A. K. Cu and Zn interactions with Aβ peptides: consequence of coordination on aggregation and formation of neurotoxic soluble Aβ oligomers. Metallomics 2019, 11, 64-84.
  145. Rana, M.; Cho, H.-J.;  Roy, T. K.;  Mirica, L. M.; Sharma, A. K. Azo-dyes based small bifunctional molecules for metal chelation and controlling amyloid formation. Inorganica Chimica Acta 2018, 471, 419-429.
  146. Crouch, P. J.; Savva, M. S.;  Hung, L. W.;  Donnelly, P. S.;  Mot, A. I.;  Parker, S. J.;  Greenough, M. A.;  Volitakis, I.;  Adlard, P. A.;  Cherny, R. A.;  Masters, C. L.;  Bush, A. I.;  Barnham, K. J.; White, A. R. The Alzheimer’s therapeutic PBT2 promotes amyloid-β degradation and GSK3 phosphorylation via a metal chaperone activity. J. Neurochem. 2011, 119, 220-230.
  147. Drew, S. C. The Case for Abandoning Therapeutic Chelation of Copper Ions in Alzheimer's Disease. Front. Neurosci. 2017, 11.
  148. Squitti, R.; Salustri, C.;  Rongioletti, M.; Siotto, M. Commentary: The Case for Abandoning Therapeutic Chelation of Copper Ions in Alzheimer’s Disease. Front. Neurol. 2017, 8.
  149. Elder, G. A.; Gama Sosa, M. A.; De Gasperi, R. Transgenic Mouse Models of Alzheimer's Disease. Mount Sinai Journal of Medicine: A Journal of Translational and Personalized Medicine 2010, 77, 69-81.
  150. Sasaguri, H.; Nilsson, P.;  Hashimoto, S.;  Nagata, K.;  Saito, T.;  De Strooper, B.;  Hardy, J.;  Vassar, R.;  Winblad, B.; Saido, T. C. APP mouse models for Alzheimer's disease preclinical studies. The EMBO Journal 2017, 36, 2473-2487.
  151. Quinn, J. F.; Harris, C. J.;  Cobb, K. E.;  Domes, C.;  Ralle, M.;  Brewer, G.; Wadsworth, T. L. A Copper-Lowering Strategy Attenuates Amyloid Pathology in a Transgenic Mouse Model of Alzheimer's Disease. J. Alzheimers Dis. 2010, 21, 903-914.
  152. Southon, A.; Greenough, M. A.;  Ganio, G.;  Bush, A. I.;  Burke, R.; Camakaris, J. Presenilin Promotes Dietary Copper Uptake. PLoS One 2013, 8, e62811-e62811.
  153. Gokhale, A.; Vrailas-Mortimer, A.;  Larimore, J.;  Comstra, H. S.;  Zlatic, S. A.;  Werner, E.;  Manvich, D. F.;  Iuvone, P. M.;  Weinshenker, D.; Faundez, V. Neuronal copper homeostasis susceptibility by genetic defects in dysbindin, a schizophrenia susceptibility factor. Hum. Mol. Genet. 2015, 24, 5512-5523.
  154. Grossi, C.; Francese, S.;  Casini, A.;  Rosi, M. C.;  Luccarini, I.;  Fiorentini, A.;  Gabbiani, C.;  Messori, L.;  Moneti, G.; Casamenti, F. Clioquinol Decreases Amyloid-β Burden and Reduces Working Memory Impairment in a Transgenic Mouse Model of Alzheimer's Disease. J. Alzheimers Dis. 2009, 17, 423-440.
  155. Matlack, K. E. S.; Tardiff, D. F.;  Narayan, P.;  Hamamichi, S.;  Caldwell, K. A.;  Caldwell, G. A.; Lindquist, S. Clioquinol promotes the degradation of metal-dependent amyloid- (A ) oligomers to restore endocytosis and ameliorate A toxicity. Proceedings of the National Academy of Sciences 2014, 111, 4013-4018.
  156. Segal-Gavish, H.; Danino, O.;  Barhum, Y.;  Ben-Zur, T.;  Shai, E.;  Varon, D.;  Offen, D.; Fischer, B. A Multifunctional Biocompatible Drug Candidate is Highly Effective in Delaying Pathological Signs of Alzheimer’s Disease in 5XFAD Mice. J. Alzheimers Dis. 2017, 58, 389-400.
  157. Donnelly, P. S.; Caragounis, A.;  Du, T.;  Laughton, K. M.;  Volitakis, I.;  Cherny, R. A.;  Sharples, R. A.;  Hill, A. F.;  Li, Q.-X.;  Masters, C. L.;  Barnham, K. J.; White, A. R. Selective Intracellular Release of Copper and Zinc Ions from Bis(thiosemicarbazonato) Complexes Reduces Levels of Alzheimer Disease Amyloid-β Peptide. J. Biol. Chem. 2008, 283, 4568-4577.
  158. Gerber, H.; Wu, F.;  Dimitrov, M.;  Garcia Osuna, G. M.; Fraering, P. C. Zinc and Copper Differentially Modulate Amyloid Precursor Protein Processing by γ-Secretase and Amyloid-β Peptide Production. J. Biol. Chem. 2017, 292, 3751-3767.
  159. Venugopal, C.; Demos, C. M.;  Rao, K. S. J.;  Pappolla, M. A.; Sambamurti, K. Beta-secretase: structure, function, and evolution. CNS Neurol. Disord. Drug Targets 2008, 7, 278-94.
  160. Crouch, P. J.; Hung, L. W.;  Adlard, P. A.;  Cortes, M.;  Lal, V.;  Filiz, G.;  Perez, K. A.;  Nurjono, M.;  Caragounis, A.;  Du, T.;  Laughton, K.;  Volitakis, I.;  Bush, A. I.;  Li, Q. X.;  Masters, C. L.;  Cappai, R.;  Cherny, R. A.;  Donnelly, P. S.;  White, A. R.; Barnham, K. J. Increasing Cu bioavailability inhibits A oligomers and tau phosphorylation. Proceedings of the National Academy of Sciences 2009, 106, 381-386.
  161. Kenche, V. B.; Barnham, K. J. Alzheimer's disease & metals: therapeutic opportunities. Br. J. Pharmacol. 2011, 163, 211-9.
  162. Ribarič, S. Peptides as Potential Therapeutics for Alzheimer’s Disease. Molecules 2018, 23, 283-283.
  163. McKenzie-Nickson, S.; I. Bush, A.; J. Barnham, K. Bis(thiosemicarbazone) Metal Complexes as Therapeutics for Neurodegenerative Diseases. Curr. Top. Med. Chem. 2016, 16, 3058-3068.
  164. Paterson, B. M.; Cullinane, C.;  Crouch, P. J.;  White, A. R.;  Barnham, K. J.;  Roselt, P. D.;  Noonan, W.;  Binns, D.;  Hicks, R. J.; Donnelly, P. S. Modification of Biodistribution and Brain Uptake of Copper Bis(thiosemicarbazonato) Complexes by the Incorporation of Amine and Polyamine Functional Groups. Inorg. Chem. 2019, 58, 4540-4552.
  165. Choo, X. Y.; Liddell, J. R.;  Huuskonen, M. T.;  Grubman, A.;  Moujalled, D.;  Roberts, J.;  Kysenius, K.;  Patten, L.;  Quek, H.;  Oikari, L. E.;  Duncan, C.;  James, S. A.;  McInnes, L. E.;  Hayne, D. J.;  Donnelly, P. S.;  Pollari, E.;  Vähätalo, S.;  Lejavová, K.;  Kettunen, M. I.;  Malm, T.;  Koistinaho, J.;  White, A. R.; Kanninen, K. M. CuII(atsm) Attenuates Neuroinflammation. Front. Neurosci. 2018, 12.
  166. Kitazawa, M.; Hsu, H.-W.; Medeiros, R. Copper Exposure Perturbs Brain Inflammatory Responses and Impairs Clearance of Amyloid-Beta. Toxicol. Sci. 2016, 152, 194-204.
  167. Sparks, D. L.; Schreurs, B. G. Trace amounts of copper in water induce -amyloid plaques and learning deficits in a rabbit model of Alzheimer's disease. Proceedings of the National Academy of Sciences 2003, 100, 11065-11069.
  168. Ramanathan, A.; Nelson, A. R.;  Sagare, A. P.; Zlokovic, B. V. Impaired vascular-mediated clearance of brain amyloid beta in Alzheimer’s disease: the role, regulation and restoration of LRP1. Front. Aging Neurosci. 2015, 7.
  169. Kanekiyo, T.; Liu, C.-C.;  Shinohara, M.;  Li, J.; Bu, G. LRP1 in brain vascular smooth muscle cells mediates local clearance of Alzheimer's amyloid-β. The Journal of neuroscience : the official journal of the Society for Neuroscience 2012, 32, 16458-65.
  170. Singh, I.; Sagare, A. P.;  Coma, M.;  Perlmutter, D.;  Gelein, R.;  Bell, R. D.;  Deane, R. J. R.;  Zhong, E.;  Parisi, M.;  Ciszewski, J.;  Kasper, R. T.; Deane, R. J. R. Low levels of copper disrupt brain amyloid-β homeostasis by altering its production and clearance. Proceedings of the National Academy of Sciences 2013, 110, 14771-14776.
  171. Kitazawa, M.; Cheng, D.; LaFerla, F. M. Chronic copper exposure exacerbates both amyloid and tau pathology and selectively dysregulates cdk5 in a mouse model of AD. J. Neurochem. 2009, 108, 1550-1560.
  172. Lasagna-Reeves, C. A.; Castillo-Carranza, D. L.;  Sengupta, U.;  Clos, A. L.;  Jackson, G. R.; Kayed, R. Tau oligomers impair memory and induce synaptic and mitochondrial dysfunction in wild-type mice. Mol. Neurodegener. 2011, 6, 39-39.
  173. Tai, H.-C.; Serrano-Pozo, A.;  Hashimoto, T.;  Frosch, M. P.;  Spires-Jones, T. L.; Hyman, B. T. The Synaptic Accumulation of Hyperphosphorylated Tau Oligomers in Alzheimer Disease Is Associated With Dysfunction of the Ubiquitin-Proteasome System. The American Journal of Pathology 2012, 181, 1426-1435.
  174. Jack, C. R.; Knopman, D. S.;  Jagust, W. J.;  Shaw, L. M.;  Aisen, P. S.;  Weiner, M. W.;  Petersen, R. C.; Trojanowski, J. Q. Hypothetical model of dynamic biomarkers of the Alzheimer's pathological cascade. The Lancet Neurology 2010, 9, 119-128.
  175. Arce-Varas, N.; Abate, G.;  Prandelli, C.;  Martínez, C.;  Cuetos, F.;  Menéndez, M.;  Marziano, M.;  Cabrera-García, D.;  Fernández-Sánchez, M. T.;  Novelli, A.;  Memo, M.; Uberti, D. Comparison of Extracellular and Intracellular Blood Compartments Highlights Redox Alterations in Alzheimer's and Mild Cognitive Impairment Patients. Current Alzheimer research 2017, 14, 112-122.
  176. James, S. A.; Volitakis, I.;  Adlard, P. A.;  Duce, J. A.;  Masters, C. L.;  Cherny, R. A.; Bush, A. I. Elevated labile Cu is associated with oxidative pathology in Alzheimer disease. Free Radic. Biol. Med. 2012, 52, 298-302.
  177. Arbor, S. C.; LaFontaine, M.; Cumbay, M. Amyloid-beta Alzheimer targets - protein processing, lipid rafts, and amyloid-beta pores. The Yale journal of biology and medicine 2016, 89, 5-21.
  178. Dannenberger, D.; Nuernberg, G.;  Renne, U.;  Nuernberg, K.;  Langhammer, M.;  Huber, K.; Breier, B. High-fat diets rich in ω-3 or ω-6 polyunsaturated fatty acids have distinct effects on lipid profiles and lipid peroxidation in mice selected for either high body weight or leanness. Nutrition 2013, 29, 765-771.
  179. Lang, M.; Fan, Q.;  Wang, L.;  Zheng, Y.;  Xiao, G.;  Wang, X.;  Wang, W.;  Zhong, Y.; Zhou, B. Inhibition of human high-affinity copper importer Ctr1 orthologous in the nervous system of Drosophila ameliorates Aβ42-induced Alzheimer's disease–like symptoms. Neurobiol. Aging 2013, 34, 2604-2612.
  180. Sevigny, J.; Chiao, P.;  Bussière, T.;  Weinreb, P. H.;  Williams, L.;  Maier, M.;  Dunstan, R.;  Salloway, S.;  Chen, T.;  Ling, Y.;  O’Gorman, J.;  Qian, F.;  Arastu, M.;  Li, M.;  Chollate, S.;  Brennan, M. S.;  Quintero-Monzon, O.;  Scannevin, R. H.;  Arnold, H. M.;  Engber, T.;  Rhodes, K.;  Ferrero, J.;  Hang, Y.;  Mikulskis, A.;  Grimm, J.;  Hock, C.;  Nitsch, R. M.; Sandrock, A. The antibody aducanumab reduces Aβ plaques in Alzheimer’s disease. Nature 2016, 537, 50-56.
More
Information
Subjects: Others
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , ,
View Times: 598
Revisions: 2 times (View History)
Update Date: 20 Apr 2021
1000/1000