Open access peer-reviewed chapter

Neurodegenerative Diseases and Their Therapeutic Approaches

Written By

Farhin Patel and Palash Mandal

Submitted: 13 October 2018 Reviewed: 19 October 2018 Published: 20 February 2019

DOI: 10.5772/intechopen.82129

From the Edited Volume

Neurons - Dendrites and Axons

Edited by Gonzalo Emiliano Aranda Abreu and María Elena Hernández Aguilar

Chapter metrics overview

1,659 Chapter Downloads

View Full Metrics

Abstract

Alzheimer’s disease and Parkinson’s disease are characterized as a chronic and progressive neurodegenerative disorder and are manifested by the loss of neurons within the brain and/or spinal cord. In the present chapter, we would like to summarize the molecular mechanism focusing on metabolic modification associated with neurodegenerative diseases or heritable genetic disorders. The identification of the exact molecular mechanisms involved in these diseases would facilitate the discovery of earlier pathophysiological markers along with substantial therapies, which may consist (of) mitochondria-targeted antioxidant therapy, mitochondrial dynamics modulators, epigenetic modulators, and neural stem cell therapy. Therefore, all these therapies may hold particular assurance as influential neuroprotective therapies in the treatment of neurodegenerative diseases.

Keywords

  • neurons
  • mitochondria-targeted antioxidants
  • mitochondrial dynamics
  • epigenetic regulations
  • stem cell
  • neurodegenerative diseases

1. Introduction

1.1 What are neurons?

Neurons or nerve cells are the functional unit of the brain and nervous system, and they produce electrical signals known as action potentials. Action potentials permit them to speedily pass on the details over long distances. Their connections define who we are as a person. The creation of new neurons in the brain is known as neurogenesis [1].

1.2 Anatomy of a neuron

Different types of neurons may differ in a number of ways, but they all include three distinct regions with differing functions, that is, the cell body (soma), followed by the dendrites, the axons, and the connected axon terminals (Figure 1).

  1. Cell body: It is the place of biogenesis of almost all neuronal proteins and membranes. It contains a nucleus.

  2. Dendrites: The extensions of neurons that receive signals and conduct them toward the cell body (soma) are known as dendrites.

  3. Axon (nerve fiber): The extensions of neurons that conduct the signals away from the cell body to the other nerve cells or neuron are known as axons.

  4. Axon terminal (end-plate): The end part or terminal part of axons that makes a synaptic contact with other nerve cells is known as an axon terminal. It is responsible for the initiation of transmission of nerve impulse to another nerve cell [2].

Figure 1.

Anatomy of neuron.

1.3 Functions of neurons

  1. Conduction and transmission of nerve impulses

  2. Initiation and conduction of action potential

  3. Synaptic transmission [3]

1.4 How neurons transmit information throughout the body?

Neurons converse with other neurons through axons and dendrites. When a neuron receives information from another neuron, it transmits an electrical signal along the length of the respective axon, known as action potential. At the axon terminal, the electrical signal is changed into chemical signal. The axon releases chemical messengers called neurotransmitters. The neurotransmitters are released into the gap between the axon terminal and the tip of a dendrite (receptor site) of a further neuron. The space between the axon terminal and the tip of a dendrite is called a synapse. The neurotransmitters travel along the short distance through the synaptic gap to the dendrite. The dendrite receives the neurotransmitters and translates the chemical signal into electrical signal. This electrical signal travels all the way through the neuron, to be converted back into a chemical signal when it gets to adjoining neurons [4].

Advertisement

2. Neurodegenerative diseases

Etymologically, the word neurodegeneration comprises of “neuro,” which refers to neurons, and “degeneration,” which refers to the process of losing structure and/or function of either tissues or organs [5]. A neurodegenerative disease is considered as a slow, progressive failure of nerve cells within the central nervous system (CNS). This leads to deficits in particular brain functions like learning, movement, and cognition generally performed by the CNS (brain and spinal cord).

2.1 Factors associated with neurodegenerative diseases

  1. Aberrant protein dynamics with aggregation and degradation of defective protein [6]

  2. Oxidative stress and reactive oxygen species (ROS) formation

  3. Impaired bioenergetics and mitochondrial dysfunction

  4. Excessive exposure to metals and pesticides (Figure 2)

Figure 2.

Factors associated with neurodegenerative diseases.

2.2 Classification based on molecular defects

  1. Trinucleotide repeat diseases: HD, spinal cerebellar atrophy, and myotonic dystrophy [7].

  2. Prion diseases: Creutzfeldt-Jakob disease, Gerstmann-Straussler-Scheinker syndrome, and fetal familial insomnia [8].

  3. Synucleinopathies: PD, progressive supranuclear palsy and diffuse Lewy body dementia [9].

  4. Tauopathies: Corticobasal degeneration, frontotemporal dementia with parkinsonism linked to chromosome 1\(FTDP-17), and pick disease [10].

Advertisement

3. Alzheimer’s disease

Alzheimer’s disease (AD) is an irreparable, progressive neurodegenerative disease that affects normal brain functioning [11]. It is mainly the general cause of dementia [12]. Dementia is a syndrome associated with memory loss and loss of abilities like thinking, reasoning, and language skills along with other mental illness [12].

3.1 History

This disease is named after Dr. Alois Alzheimer. He observed some brain tissue abnormalities in an old woman who died due to some unusual mental illness. Later, he examined her brain and found many abnormal tangled bundles of fibers (called as tau tangles, neurofibrillary) and clumps (called as amyloid plaques). That is how he found the cause of AD [13].

3.2 Causes

The cause of AD is not clearly understood.

  1. Genetic: Nearly, 70% of the cases are related to genetic factors with the involvement of many specific genes [14].

    1. Autosomal dominant inheritance: Also known as early-onset familial AD [15], it occurs due to the mutation in one of the three genes: Presenilin 1, presenilin 2, or amyloid precursor protein (APP) [16].

      Aβ42: A protein that is the main component of senile plaques, and the levels are increased due to mutation in APP and presenilin genes [17].

    2. Sporadic Alzheimer’s disease: In this type of AD, genetic and environmental factors play a major role.

      Example: Inheritance of the epsilon 4 allele of the apolipoprotein E (APOE) [18, 19].

  2. Cholinergic hypothesis: The cholinergic hypothesis states that AD is caused by the reduced synthesis of neurotransmitter acetylcholine [20].

  3. Amyloid hypothesis: The amyloid hypothesis states that AD is caused by the deposits of extracellular amyloid beta (Aβ) [21].

  4. Tau hypothesis: The tau hypothesis states that AD is caused due to abnormalities in tau protein, leading to the disintegration of microtubules in nerve cells [22, 23].

3.3 Molecular mechanism

(a) Proteopathy: AD has been recognized by plaque formation occurring due to abnormal folding of amyloid beta (Aβ) protein and tau protein in the nerve cells (brain) leading to the degeneration of nerve cells [24]. The amyloid precursor protein (APP) leads to the formation of Aβ. APP plays an important role in neuron-like developments and post-injury repair mechanism and survival [25, 26]. In AD, secreting enzymes like β-secretase and γ-secretase together will break down APP into small fragments that penetrate through the neuron membrane [27]. This leads to the formation of Aβ fibrils that later cluster together to form senile plaques and deposits in the outer side of neurons [28, 29]. Aggregated amyloid fibrils accumulation leads to the disruption of cell’s calcium ion homeostasis, which results in apoptosis [30] (Figure 3).

Figure 3.

Molecular mechanism of AD.

(b) Tauopathy: In AD, there is an abnormal accumulation of tau protein. Upon phosphorylation, tau protein stabilizes the microtubules, and it is known as microtubule-associated protein. Tau protein undergoes certain chemical changes, and becomes hyperphosphorylated. This leads to the formation of neurofibrillary tangles upon aggregation with other threads, which results in decaying the neuro-transport system [31].

3.4 Therapeutic approaches

3.4.1 Mitochondrial-directed therapies

Decline of N-acetyl aspartate and creatine is associated with dementia [32]. Supplementation of creatine was found to protect neurons in AD [33]. In hippocampal neurons, administration of creatine defends against glutamate and Aβ toxicity in rats [34].

In AD patients, administration of lipoic acid (600 mg/day) [LA - an antioxidant; coenzyme for pyruvate dehydrogenase and α-ketoglutarate dehydrogenase] stabilizes the cognitive measures [35, 36]. Decreased oxidative stress of mitochondria in fibroblasts was found in AD patients due to LA and/or N-acetyl cysteine (antioxidant and glutathione precursor) administration [37].

CoQ10 (an antioxidant and cofactor of the electron transport chain) blocks apoptosis by inhibiting the permeability transition pore (PTP) of mitochondria [38]. Treatment of CoQ10 neutralizes the brain mitochondrial alterations made by amyloid-β1–40 [39]. CoQ10 was shown to protect paraquat and rotenone-induced mitochondrial dysfunction and neuronal death in SHSY-5Y cells (human neuroblastoma cells) and primary rat mesencephalic neurons, [40, 41]. In R6/2 mice, combined treatment of CoQ10 and minocycline reduces HTT accumulation, brain atrophy, and striatal neuron atrophy [42].

MitoQ (mitochondrial coenzyme Q) reduces oxidative stress and prevents mitochondrial dysfunction [43]. Oral administration of MitoQ (1 mg/kg body weight) showed better pharmacokinetics behavior with plasma (Cmax = 33.15 ng/ml and Tmax = 1 hr.) in Phase I trial (Antipodean Pharmaceuticals Inc., San Francisco, CA).

3.4.2 Stem cell therapy

Neural stem cell therapy provides a potential to neurons derived from stem cells to integrate with existing neuronal network of the host brain [44]. In animal models, stem cell transplantation elevates the level of acetylcholine, resulting in an improved cognitive and memory function. Stem cells secrete neurotrophic factors, which modulate neuroplasticity and neurogenesis [45, 46].

Embryonic stem cells (ESCs)-derived neuron progenitor cells (NPCs) when transplanted into an amyloid-β injured in vitro model, after 2 weeks of amyloid-β injection, showed an increased escape latency when compared with phosphate-buffered saline-treated controls [47]. It has been reported that ESCs-derived NPCs improve memory impairment in AD models [48].

Human induced pluripotent stem cell (iPSC) therapy delivers a possible strategy for drug development against AD [49]. Neurons differentiated from iPSCs increase the secretion of amyloid-β42 as it is affected by γ-secretase inhibitors [50].

Bone marrow (BM)-derived mesenchymal stem cells (MSCs) play an important role in the removal of amyloid-β plaques from the hippocampus [51]. Human MSCs promoted amyloid-β clearance and enhanced autophagy and neuronal survival in an amyloid-β-treated mouse model [46]. Transplantation of adipose-derived MSCs (AMSCs) into AD brain improved the acetylcholine levels along with microglia activation and cognitive functions [52, 53]. In a transgenic mouse model, human umbilical cord-derived MSCs differentiated themselves into neuron-like cells, and these cells when transplanted into an amyloid-β precursor protein (AβPP) and PS1 (AβPP/PS1) resulted in improved cognitive function and decreased amyloid β deposition [54].

3.4.3 Epigenetic modulators

Histone deacetylases have been linked to AD. Treatment with HDACi (histone deacetylase inhibitors) induced dendrite growth, increased the number of synapses, and restored learning and memory deficits in mice with AD [55] (Table 1).

HDACiFunctionReferences
Sodium butyrateIn neuroblastoma cells, it induces phosphorylation of tau protein and programmed cell death resulting in restoring memory.[56]
Phenylbutyrate
(4-PBA)
InTg2576 mouse model, 4-PBA restores fear learning and rescues dendritic spine losses that are associated with memory shortage.[57]
Suberoylanilide hydroxamic acidIn mutant mice model, systemic treatment restores contextual memory[58]
Resveratrol
(activator of class III HDAC)
In in vivo and in vitro studies, SIRT1 reduces the amyloidogenic processing of APP[59]

Table 1.

Histone deacetylase inhibitors and their respective functions in AD.

3.4.4 Mitochondrial dynamics modulators

Two recent studies have also shown the protective effects mediated by inhibition of mitochondrial fission via Drp1 deficiency on mitochondria and neurons in tau and APP transgenic animal models for AD [60, 61].

Advertisement

4. Parkinson’s disease

Parkinson’s disease (PD) is a progressive, long-term neurodegenerative disorder that affects the motor neurons [62]. It is caused by a loss of neurons in the brain part known as substantia nigra leading to a reduction in a neurotransmitter called dopamine [62].

4.1 History

In 1817, James Parkinson (before known as Jean-Martin Charcot) published an essay named “Shaking Palsy” describing six cases of paralysis agitans showing certain characteristics of this disease [63, 64].

In 1865, William Sanders termed this disease as Parkinson’s disease [65].

4.2 Causes

The following are the causes of PD:

  1. Environmental factors: Exposure to metals, solvents, and pesticides, or any head injuries are considered to be a factor for the onset of PD [66, 67].

  2. Genetics: Few percent of cases are developing this disease due to mutation in one specific gene out of several genes related to PD (Table 2).

NameGeneReferences
Autosomal-dominant PDPARK1/PARK4SNCA (α-synuclein)[68, 69, 72]
PARK2Parkin[68, 69, 72]
PARK5UCHL[68, 69, 72]
PARK8LRRK2[68, 69, 72]
Autosomal-recessive PD
PARK6PINK1[68, 70, 71, 72]
PARK7DJ-1[68, 70, 71, 72]

Table 2.

Genes involved in PD.

4.3 Molecular mechanism

The mechanism involved in the development of PD includes various factors like the aggregations of misfolded proteins, activation of protein degradation pathways, mitochondrial damage, and oxidative stress, along with certain gene mutations [73, 74, 75].

4.3.1 Aggregation of misfolded proteins

  • Accumulation of Lewy bodies in dopamine neurons of the substantia nigra pars compacta [75]

  • Hyperphosphorylation of tau protein causes accumulation of neurofibrillary tangles [76]

4.3.2 Protein degradation pathways

  • Ubiquitin-proteasome system (UPS): It is responsible for the degradation of misfolded or damaged proteins in the cytosol, nucleus, or endoplasmic reticulum (ER) [77]. Impairment in this system leads to aggregation of misfolded amyloid proteins (Lewy bodies) [78]. Other proteins like UCH-L1 and Parkin are involved in the degradation of misfolded α-synuclein [79]

  • Chaperones (heat shock proteins/HSP): Chaperones undergo dysfunctioning in PD, as they play a vital role in cell-defense mechanism involved in protein degradation and folding of proteins. Major HSPs involved are HSP 26, HSP40, HSP 60, HSP 70, HSP 90, and HSP 100 [80]. HSPs aggregate with α-synuclein or tau protein and form insoluble structures resulting in reduced toxicity of α-synuclein or tau protein [81, 82]

  • Autophagy-lysosomal pathway (ALP): It serves to clear Lewy bodies in PD acting as an alternative clearance mechanism for proteins [83, 84]. Chaperone-mediated autophagy (CMA) helps in the degradation of α-synuclein by selectively translocating into lysosomes [83]. Therefore, dysfunctioning of CMA decreases the efficiency of α-synuclein, leading to excessive accumulation of this protein. This results in impaired neuronal activity as observed in PD [73, 85]. Failure of formation of autophagosome, its inability to bind with lysosomes due to deficiency of lysozymes, or dysfunction of HSP70 results in dysfunction of ALP in PD [73, 85] (Figure 4)

Figure 4.

Molecular mechanism of PD.

4.3.3 Damage to mitochondria and oxidative stress

  • Abnormality of complex-I in mitochondria directly interferes with ATP production in the cell, resulting in cell death [86]. Monoamines such as dopamine are cleaved by monoamine oxidase-B (MAO-B) and combined with oxygen-forming reactive oxygen species (ROS) [87]. Increased oxidative stress was observed in PD.

4.3.4 Genetic mutations

The most common genes related to PD are α-synuclein, DJ-1, PINK1, and Parkin [88] (Table 3).

GenesDysfunctionReferences
α-synucleinAggregation of misfolded amyloid proteins[89]
ParkinAggregation of misfolded amyloid proteins within SNpc[89]
DJ-1 (PARK7)Activities like transcriptional regulation, antioxidants, chaperone, and protease are dysregulated[90]
PINK1 (PARK6)Mitochondrial dysfunctioning
Degeneration of substantia nigra neuron
[91]

Table 3.

Specific gene mutations and their dysfunction involved in the development of PD.

4.4 Therapeutic approaches

4.4.1 Mitochondria-directed therapies

Administration of creatine increases tyrosine hydroxylase immunoreactive fiber density and soma size of dopaminergic neurons in mesencephalic cultures by protecting against neurotoxic insults induced by serum and glucose deprivation, MPP+, and 6-hydroxydopamine [33, 92]. It has been reported that dopamine loss was prevented by administration of creatine. In substantia nigra, creatine also reduces loss of neuron in the mice treated with 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) [93].

CoQ10 protects against iron-induced apoptosis in dopaminergic neurons [94]. In vitro, CoQ10 exerts anti-amyloidogenic effects by disrupting preformed amyloid-β fibrils [95].

SS peptides (Szeto Schiller) act as antioxidants that target mitochondria in an independent manner. In mice, reports showed that SS-20 and SS-31 provide protection against MPTP (1-methyl-4-phenyl-1, 2, 3, 6-tetrahydropyridine) neurotoxicity. SS-31 provides protection against dopamine loss in the striatum. In substantia nigra, SS-31 also provides protection against the loss of tyrosine hydroxylase immunoreactive neurons. In MPTP-treated mice, SS-20 provides potential neuronal protection on dopaminergic neurons in PD [96].

4.4.2 Stem cell therapy

In the first trial of cell-based therapy, post-mitotic dopamine neuroblasts isolated from human embryonic mesencephalic tissue have been successfully grafted in PD patients [97]. It has been confirmed through increase in 18F-dopa intake, detected through positron emission tomography (PET) [98, 99]. The grafts restore dopamine release. Disadvantages of this therapy are limited tissue availability and grafts standardization.

Recently, researchers have shed light on stem cell therapy. The production of dopamine neuroblasts from stem cells for transplantation in PD patients has been focused on. The aim was to release dopamine in a stable manner and exhibit the electrophysiological, molecular, and morphological properties of substantia nigra neurons [100, 101]. In clinical trials, it has been found that dopaminergic cells derived from embryonic stem cells can survive and reverse behavioral deficits after transplantation in PD animal models [102, 103].

4.4.3 Epigenetic modulators

In sporadic PD patients, there is an increased α-synuclein expression in dopaminergic neurons, which is linked with α-synuclein hypomethylation [104]. In familial PD patients, decreased histone acetylation is linked with increased α-synuclein levels [105]. In vitro model, mutation in α-synuclein leads to increased histone acetylation mediated through HDAC Sirt2. Treatment of Sirt2 siRNA resulted in decreased α-synuclein-mediated toxicity [106]. Administration of levodopa elevated the dopamine level, which partially showed decreased symptoms of PD. It is correlated with deacetylation of H4K5, K12, and K16 [107].

4.4.4 Mitochondrial dynamics modulators

Recombinant adeno-associated virus expressing the dominant negative Drp1 (dynamin-related protein 1) mutant or Mdivi-1, a small molecular inhibitor of Drp1, has been reported to inhibit mitochondrial fragmentation, restore dopamine release, and prevent dopamine neuron loss in PD animal models [108].

Activation of DRP1-mediated mitochondrial fission is an important contributing factor in the progression of PD. Neurons lacking PINK or Parkin accumulate DRP1, resulting in excessive mitochondrial fission, increased oxidative stress, and reduced ATP production [108, 109]. These defects can be reversed by the inhibition of mitochondrial fission with the use of mdivi-1, an inhibitor of the DRP1 pathway, or by overexpression of MFN2 (Mitofusin 2) or OPA1 (Optic atrophy protein 1) [109, 110].

In vitro models of glutamate-toxicity or OGD (oxygen-glucose deprivation) in mouse hippocampal neurons or in vivo mouse models of transient focal ischemia can be protected from enhanced mitochondrial fission and apoptosis by DRP1 knockdown or mdivi-1 inhibition [111, 112].

Advertisement

5. Conclusion

The recent advancements in the field of neurodegenerative diseases like AD and PD are based on targeting the degenerative progressions that lead to the death of neurons. The death of neurons leads to irreversible neuropathological conditions, making it difficult to be functional in humans. Because of the intricacy involved in respective neurodegenerative diseases, researchers have identified few potential biomarkers. At present, many therapeutic approaches have been suggested to treat the symptoms of both neurodegenerative diseases. Yet there exists a lacuna for the effective therapies. Hence, few therapeutic approaches like mitochondria-targeted antioxidant therapy, mitochondrial dynamics modulators, epigenetic modulators, and neural stem cell therapy may prove to have a potential in treating AD and PD.

Advertisement

Conflict of interest

Authors have no conflict of interest to declare.

References

  1. 1. Chudler EH. Brain Facts and Figures. Neuroscience for Kids. Retrieved: 2009-06-20
  2. 2. Al Martini Frederic Et. Anatomy and Physiology 2007 Ed. 2007 Edition. Rex Bookstore: Philippines; Inc; 2007. p. 288. ISBN 978-971-23-4807-5
  3. 3. Dudel J. Function of nerve cells. In: Schmidt RF, Thews G, editors. Human Physiology. Berlin, Heidelberg: Springer; 1983. DOI: 10.1007/978-3-642-96714-6_1
  4. 4. Max Plank Institute. How Neurons Talk to Each Other. NeuroscienceNews. September 24, 2016. Available from: http://neurosciencenews.com/neurons-synapses-neuroscience-5119/
  5. 5. Przedborski S, Vila M, Jackson-Lewis VJ. Series Introduction: Neurodegeneration: What is it and where are we? Journal of Clinical Investigation. 2003;111(1):3-10
  6. 6. Sheikh S, Safia, Haque E, Mir SS. Neurodegenerative diseases: Multifactorial conformational diseases and their therapeutic interventions. Journal of Neurodegenerative Diseases. 2013;2013:8. Article ID 563481
  7. 7. Cummings CJ, Zoghbi HY. Trinucleotide repeats: Mechanisms and pathophysiology. Annual Review of Genomics and Human Genetics. 2000;1:281-328
  8. 8. Prusiner SB. Prions. Proceedings of the National Academy of Sciences. 1998;95:13363-13383
  9. 9. Galvin JE, Lee VM, Trojanowski JQ. Synucleinopathies: Clinical and pathological implications. Archives of Neurology. 2001;58:186-190
  10. 10. Goedert M, Spillantini MG. Tau gene mutations and neurodegeneration. Biochemical Society Symposium. 2001;67:59-71
  11. 11. Burns A, Iliffe S. Alzheimer's disease. BMJ. 2009;338:b158. DOI: 10.1136/bmj.b158
  12. 12. Dementia Fact sheet. World Health Organization. December 12, 2017
  13. 13. Scott KR, Barrett AM. Dementia Syndromes: Evaluation and treatment. Expert Rev Neurother. 2007;7(4):407-422
  14. 14. Ballard C, Gauthier S, Corbett A, Brayne C, Aarsland D, Jones E. Alzheimer's disease. Lancet. 2011;377(9770):1019-1031. DOI: 10.1016/S0140-6736(10)61349-9
  15. 15. Blennow K, de Leon MJ, Zetterberg H. Alzheimer's disease. Lancet. 2006;368(9533):387-403. DOI: 10.1016/S0140-6736(06)69113-7
  16. 16. Waring SC, Rosenberg RN. Genome-wide association studies in Alzheimer disease. Archives of Neurology. 2008;65(3):329-334. DOI: 10.1001/archneur.65.3.329
  17. 17. Selkoe DJ. Translating cell biology into therapeutic advances in Alzheimer's disease. Nature. 1999;399(6738 Suppl):A23-A31. DOI: 10.1038/19866
  18. 18. Shioi J, Georgakopoulos A, Mehta P, Kouchi Z, Litterst CM, Baki L, et al. FAD mutants unable to increase neurotoxic Aβ 42 suggest that mutation effects on neurodegeneration may be independent of effects on Abeta. Journal of Neurochemistry. 2007;101(3):674-681. DOI: 10.1111/j.1471-4159.2006.04391.x
  19. 19. Strittmatter WJ, Saunders AM, Schmechel D, Pericak-Vance M, Enghild J, Salvesen GS, et al. Apolipoprotein E: high-avidity binding to beta-amyloid and increased frequency of type 4 allele in late-onset familial Alzheimer disease. Proceedings of the National Academy of Sciences of the United States of America. 1993;90(5):1977-1981. DOI: 10.1073/pnas.90.5.1977
  20. 20. Francis PT, Palmer AM, Snape M, Wilcock GK. The cholinergic hypothesis of Alzheimer's disease: A review of progress. Journal of Neurology, Neurosurgery & Psychiatry. 1999;66(2):137-147. DOI: 10.1136/jnnp.66.2.137
  21. 21. Hardy J, Allsop D. Amyloid deposition as the central event in the aetiology of Alzheimer’s disease. Trends in Pharmacological Sciences. 1991;12(10):383-388. DOI: 10.1016/0165-6147(91)90609-V
  22. 22. Mudher A, Lovestone S. Alzheimer's disease-do tauists and baptists finally shake hands? Trends in Neurosciences. 2002;25(1):22-26. DOI: 10.1016/S0166-2236(00)02031-2
  23. 23. Iqbal K et al. Tau pathology in Alzheimer disease and other tauopathies. Biochimica et Biophysica Acta. 2005;1739(2-3):198-210. DOI: 10.1016/j.bbadis.2004.09.008
  24. 24. Hashimoto M, Rockenstein E, Crews L, Masliah E. Role of protein aggregation in mitochondrial dysfunction and neurodegeneration in Alzheimer's and Parkinson's Diseases. Neuromolecular Medicine. 2003;4(1-2):21-36. DOI: 10.1385/NMM:4:1-2:21
  25. 25. Priller C, Bauer T, Mitteregger G, Krebs B, Kretzchmar HA, Herms J. Synapse formation and function is modulated by the amyloid precursor protein. The Journal of Neuroscience. 2006;26(27):7212-7221. DOI: 10.1523/JNEUROSCI.1450-06.2006
  26. 26. Turner PR, O’Connor K, Tate WP, Abraham WC. Roles of amyloid precursor protein and its fragments in regulating neural activity, Plasticity and Memory. Progress in Neurobiology. 2003;70(1):1-32. DOI: 10.1016/S0301-0082(03)00089-3
  27. 27. Hooper NM. Roles of proteolysis and lipid rafts in the processing of the amyloid precursor protein and prion protein. Biochemical Society Transactions. 2005;33(Pt 2):335-338. DOI: 10.1042/BST0330335
  28. 28. Tiraboschi P, Hansen LA, Thal LJ, Corey-Bloom J. The importance of neuritic plaques and tangles to the development and evolution of AD. Neurology. 2004;62(11):1984-1989. DOI: 10.1212/01.WNL.0000129697.01779.0A
  29. 29. Ohnishi S, Takano K. Amyloid fibrils from the viewpoint of protein folding. Cellular and Molecular Life Sciences. 2004;61(5):511-524. DOI: 10.1007/s00018-003-3264-8
  30. 30. Huang Y, Mucke L. Alzheimer mechanisms and therapeutic strategies. Cell. 2012;148(6):1204-1222. DOI: 10.1016/j.cell.2012.02.040
  31. 31. Hernández F, Avila J. Tauopathies. Cellular and Molecular Life Sciences. 2007;64(17):2219-2233. DOI: 10.1007/s00018-007-7220-x
  32. 32. Pilatus U, Lais C, Rochmont Adu M, Kratzsch T, Frolich L, Maurer K, et al. Conversion to dementia in mild cognitive impairment is associated with decline of Nactylaspartate and creatine as revealed by magnetic resonance spectroscopy. Psychiatry Research. 2009;173:1-7. DOI: 10.1016/j.pscychresns.2008.07.015
  33. 33. Andres RH, Ducray AD, Perez-Bouza A, Schlattner U, Huber AW, Krebs SH, et al. Creatine supplementation improves dopaminergic cell survival and protects against MPP+ toxicity in an organotypic tissue culture system. Cell Transplantation. 2005;14:537-550. [PubMed: 16355565]
  34. 34. Brewer GJ, Wallimann TW. Protective effect of the energy precursor creatine against toxicity of glutamate and beta-amyloid in rat hippocampal neurons. Journal of Neurochemistry. 2000;74:1968-1978. [PubMed: 10800940]
  35. 35. Hager K, Marahrens A, Kenklies M, Riederer P, Munch G. Alpha-lipoic acid as a new treatment option for Azheimer type dementia. Archives of Gerontology and Geriatrics. 2001;32:275-2.82. [PubMed: 11395173]
  36. 36. Hager K, Kenklies M, McAfoose J, Engel J, Munch G. Alpha-lipoic acid as a new treatment option for Alzheimer’s disease–A 48 months follow-up analysis. Journal of Neural Transmission. 2007;(Suppl):189-193
  37. 37. Moreira PI, Harris PL, Zhu X, Santos MS, Oliveira CR, Smith MA, et al. Lipoic acid and N-acetyl cysteine decrease mitochondrial-related oxidative stress in Alzheimer disease patient fibroblasts. Journal of Alzheimer's Disease. 2007;12:195-206. [PubMed: 17917164]
  38. 38. Beal MF. Mitochondrial dysfunction and oxidative damage in Alzheimer’s and Parkinson’s diseases and coenzyme Q10 as a potential treatment. Journal of Bioenergetics and Biomembranes. 2004;36:381-386. [PubMed: 15377876]
  39. 39. Moreira PI, Santos MS, Sena C, Nunes E, Seica R, Oliveira CR. CoQ10 therapy attenuates amyloid beta-peptide toxicity in brain mitochondria isolated from aged diabetic rats. Experimental Neurology. 2005;196:112-119. [PubMed: 16126199]
  40. 40. Moon Y, Lee KH, Park JH, Geum D, Kim K. Mitochondrial membrane depolarization and the selective death of dopaminergic neurons by rotenone: protective effect of coenzyme Q10. Journal of Neurochemistry. 2005;93:1199-1208. [PubMed: 15934940]
  41. 41. McCarthy S, Somayajulu M, Sikorska M, Borowy-Borowski H, Pandey S. Paraquat induces oxidative stress and neuronal cell death; neuroprotection by water-soluble Coenzyme Q10. Toxicology and Applied Pharmacology. 2004;201:21-31. [PubMed: 15519605]
  42. 42. Stack EC, Smith KM, Ryu H, Cormier K, Chen M, Hagerty SW, et al. Combination therapy using minocycline and coenzyme Q10 in R6/2 transgenic Huntington’s disease mice. Biochimica et Biophysica Acta. 1762;2006:373-380. [PubMed: 16364609]
  43. 43. Moreira PI, Zhu X, Wang X, Lee H-g, Nunomura A, Petersen RB, et al. Mitochondria: A therapeutic target in neurodegeneration. Biochimica et Biophysica Acta. 2010;1802(1):212-220. DOI: 10.1016/j.bbadis.2009.10.007
  44. 44. Liu Y, Weick JP, Liu H, Krencik R, Zhang X, Ma L, et al. Medial ganglionic eminence-like cells derived from human embryonic stem cells correct learning and memory deficits. Nature Biotechnology. 2013;31:440-447
  45. 45. Park D, Yang YH, Bae DK, Lee SH, Yang G, Kyung J, et al. Improvement of cognitive function and physical activity of aging mice by human neural stem cells over-expressing choline acetyltransferase. Neurobiology of Aging. 2013;34:2639-2646
  46. 46. Park D, Yang G, Bae DK, Lee SH, Yang YH, Kyung J, et al. Human adipose tissue-derived mesenchymal stem cells improve cognitive function and physical activity in ageing mice. Journal of Neuroscience Research. 2013;91:660-670
  47. 47. Enciu AM, Nicolescu MI, Manole CG, Muresanu DF, Popescu LM, Popescu BO. Neuroregeneration in neurodegenerative disorders. BMC Neurology. 2011;11:75
  48. 48. Tang J, Xu H, Fan X, Li D, Rancourt D, Zhou G, et al. Embryonic stem cell-derived neural precursor cells improve memory dysfunction in Aβ(1-40) injured rats. Neuroscience Research. 2008;62:86-96
  49. 49. Yagi T, Ito D, Okada Y, Akamatsu W, Nihei Y, Yoshizaki T, et al. Modeling familial Alzheimer’s disease with induced pluripotent stem cells. Human Molecular Genetics. 2011;20:4530-4539
  50. 50. Salem AM, Ahmed HH, Atta HM, Ghazy MA, Aglan HA. Potential of bone marrow mesenchymal stem cells in management of Alzheimer’s disease in female rats. Cell Biology International. 2014. DOI: 10.1002/cbin.10331
  51. 51. Shin JY, Park HJ, Kim HN, Oh SH, Bae JS, Ha HJ, et al. Mesenchymal stem cells enhance autophagy and increase β-amyloid clearance in Alzheimer disease models. Autophagy. 2014;10:32-44
  52. 52. Ma T, Gong K, Ao Q, Yan Y, Song B, Huang H, et al. Intracerebral transplantation of adipose-derived mesenchymal stem cells alternatively activates microglia and ameliorates neuropathological deficits in Alzheimer’s disease mice. Cell Transplantation. 2013;22:S113-S126
  53. 53. Darlington D, Deng J, Giunta B, Hou H, Sanberg CD, Kuzmin-Nichols N, et al. Multiple low-dose infusions of human umbilical cord blood cells improve cognitive impairments and reduce amyloid-β-associated neuropathology in Alzheimer mice. Stem Cells and Development. 2013;22:412-421
  54. 54. Yang H, Xie Z, Wei L, Yang S, Zhu Z, Wang P, et al. Human umbilical cord mesenchymal stem cell-derived neuron-like cells rescue memory deficits and reduce amyloid-β deposition in an AβPP/PS1 transgenic mouse model. Stem Cell Research & Therapy. 2013;4:76
  55. 55. Fischer A, Sananbenesi F, Wang X, Dobbin M, Tsai LH. Recovery of learning and memory is associated with chromatin remodelling. Nature. 2007;447(7141):178-182
  56. 56. Nuydens R, Heers C, Chadarevian A, et al. Sodium butyrate induces aberrant t phosphorylation and programmed cell death in human neuroblastoma cells. Brain Research. 1995;688(1-2):86-94
  57. 57. Ricobaraza A, Cuadrado-Tejedor M, Marco S, Perez-Otano I, Garcia-Osta A. Phenylbutyrate rescues dendritic spine loss associated with memory deficits in a mouse model of Alzheimer disease. Hippocampus. 2012;22(5):1040-1050
  58. 58. Kilgore M, Miller CA, Fass DM, et al. Inhibitors of class 1 histone deacetylases reverse contextual memory deficits in a mouse model of Alzheimer’s disease. Neuropsychopharmacology. 2010;35(4):870-880
  59. 59. Bonda DJ, Lee HG, Camins A, et al. The sirtuin pathway in ageing and Alzheimer disease: mechanistic and therapeutic considerations. Lancet Neurology. 2011;10(3):275-279
  60. 60. Kandimalla R, Manczak M, Fry D, Suneetha Y, Sesaki H, Reddy PH. Reduced dynamin-related protein 1 protects against phosphorylated Tau-induced mitochondrial dysfunction and synaptic damage in alzheimer’s disease. Human Molecular Genetics. 2016;25:4881-4897. [CrossRef] [PubMed]
  61. 61. Manczak M, Kandimalla R, Fry D, Sesaki H, Reddy PH. Protective effects of reduced dynamin-related protein 1 against amyloid beta-induced mitochondrial dysfunction and synaptic damage in Alzheimer’s disease. Human Molecular Genetics. 2016;25:5148-5166
  62. 62. “Parkinson’s Disease Information Page”. NINDS. June 30, 2016. Archived from the original on January 4, 2017. Retrieved: July 18, 2016
  63. 63. Sveinbjornsdottir S. The clinical symptoms of Parkinson's disease. Journal of Neurochemistry. 2016;139(Suppl 1):318-324
  64. 64. GBD 2015 Mortality Causes of Death Collaborators. Global, regional, and national life expectancy, all-cause mortality, and cause-specific mortality for 249 causes of death, 1980-2015: a systematic analysis for the Global Burden of Disease Study 2015. Lancet. 2016;388(10053):1459-1544. DOI: 10.1016/s0140-6736(16)31012-1
  65. 65. Barranco Quintana JL, Allam MF, Del Castillo AS, Navajas RF. Parkinson’s disease and tea: A quantitative review. Journal of the American College of Nutrition. 2009;28(1):1-6. DOI: 10.1080/07315724.2009.10719754
  66. 66. Goldman SM. Environmental toxins and Parkinson's disease. Annual Review Of Pharmacology And Toxicology;54:141-164. DOI: 10.1146/annurev-pharmtox-011613-135937
  67. 67. Gao J, Liu R, Zhao E, Huang X, Nalls MA, Singleton AB, Chen H . Head injury, potential interaction with genes, and risk for Parkinson’s disease. Parkinsonism Relat Disord. 2015;21(3):292-296
  68. 68. Lesage S, Brice A. Parkinson's disease: From monogenic forms to genetic susceptibility factors. Human Molecular Genetics. 2009;18(R1):R48-R59. DOI: 10.1093/hmg/ddp012
  69. 69. Larsen SB, Hanss Z, Krüger R. The genetic architecture of mitochondrial dysfunction in Parkinson's disease. Cell and Tissue Research. 2018;373(1):21-37. DOI: 10.1007/s00441-017-2768-8. ISSN 1432-0878
  70. 70. Davie CA. A review of Parkinson's disease. British Medical Bulletin. 2008;86(1):109-127. DOI: 10.1093/bmb/ldn013
  71. 71. Kitada T, Asakawa S, Hattori N, Matsumine H, Yamamura Y, Minoshima S, et al. Mutations in the parkin gene cause autosomal recessive juvenile parkinsonism. Nature. 1998;392(6676):605-608. DOI: 10.1038/33416. Bibcode:1998Natur.392..605K
  72. 72. Wood-Kaczmar A, Gandhi S, Wood NW. Understanding the molecular causes of Parkinson's disease. Trends in Molecular Medicine. 2006;12(11):521-528. DOI: 10.1016/j.molmed.2006.09.007
  73. 73. Dauer W, Przedborski S. Parkinson’s disease: Mechanisms and models. Neuron. 2003;39(6):889-909. DOI: 10.1016/S0896-6273(03)00568-3
  74. 74. Kim WS, Kagedal K, Halliday GM. Alpha-synuclein biology in Lewy body diseases. Alzheimer's Research & Therapy. 2014;6(5):73. DOI: 10.1186/s13195-014-0073-2
  75. 75. Cardinale A, Chiesa R, Sierks M. Protein misfolding and neurodegenerative diseases. The International Journal of Biochemistry & Cell Biology. 2014;2014:217371. DOI: 10.1155/2014/217371
  76. 76. Schraen-Maschke S, Sergeant N, Dhaenens CM, Bombois S, Deramecourt V, Caillet-Boudin ML, et al. Tau as a biomarker of neurodegenerative diseases. Biomarkers in Medicine. 2008;2(4):363-384
  77. 77. Ciechanover A, Kwon YT. Degradation of misfolded proteins in neurodegenerative diseases: Therapeutic targets and strategies. Experimental & Molecular Medicine. 2015;47:e147
  78. 78. Larsen KE, Sulzer D. Autophagy in neurons: A review. Histology and Histopathology. 2002;17(3):897-908
  79. 79. Pukass K, Richter-Landsberg C. Inhibition of UCH-L1 in oligodendroglial cells results in microtubule stabilization and prevents alpha-synuclein aggregate formation by activating the autophagic pathway: Implications for multiple system atrophy. Frontiers in Cellular Neuroscience. 2015;9:163
  80. 80. Karunanithi S, Brown IR. Heat shock response and homeostatic plasticity. Frontiers in Cellular Neuroscience. 2015;9:68
  81. 81. Bercovich B, Stancovski I, Mayer A, Blumenfeld N, Laszlo A, Schwartz AL, et al. Ubiquitin-dependent degradation of certain protein substrates in vitro requires the molecular chaperone Hsc70. The Journal of Biological Chemistry. 1997;272(14):9002-9010
  82. 82. Wyttenbach A. Role of heat shock proteins during polyglutamine neurodegeneration: Mechanisms and hypothesis. Journal of Molecular Neuroscience. 2004;23(1-2):69-96
  83. 83. Cuervo AM, Stefanis L, Fredenburg R, Lansbury PT, Sulzer D. Impaired degradation of mutant alpha-synuclein by chaperone-mediated autophagy. Science. 2004;305(5688):1292-1295
  84. 84. Webb JL, Ravikumar B, Atkins J, Skepper JN, Rubinsztein DC. AlphaSynuclein is degraded by both autophagy and the proteasome. The Journal of Biological Chemistry. 2003;278(27):25009-25013
  85. 85. Jankovic J. Motor fluctuations and dyskinesias in Parkinson's disease: Clinical manifestations. Movement Disorders. 2005;20(Suppl 11):S11-S16
  86. 86. Parker WD Jr, Parks JK, Swerdlow RH. Complex I deficiency in Parkinson's disease frontal cortex. Brain Research. 2008;1189:215-218
  87. 87. Lotharius J, Brundin P. Pathogenesis of Parkinson’s disease: Dopamine, vesicles and alpha-synuclein. Nature Reviews. Neuroscience. 2002;3(12):932-942
  88. 88. Klein C, Westenberger A. Genetics of Parkinson’s disease. Cold Spring Harbor Perspectives in Medicine. 2012;2(1):a008888
  89. 89. Schlossmacher MG, Frosch MP, Gai WP, Medina M, Sharma N, Forno L, et al. Parkin localizes to the Lewy bodies of Parkinson disease and dementia with Lewy bodies. The American Journal of Pathology. 2002;160(5):1655-1667
  90. 90. Canet-Aviles RM, Wilson MA, Miller DW, Ahmad R, McLendon C, Bandyopadhyay S, et al. The Parkinson's disease protein DJ-1 is neuroprotective due to cysteine-sulfinic acid-driven mitochondrial localization. Proceedings of the National Academy of Sciences of the United States of America. 2004;101(24):9103-9108
  91. 91. Wu HY, Chen SF, Hsieh JY, Chou F, Wang YH, Lin WT, et al. Structural basis of antizyme-mediated regulation of polyamine homeostasis. Proceedings of the National Academy of Sciences of the United States of America. 2015;112(36):11229-11234
  92. 92. Andres RH, Huber AW, Schlattner U, Perez-Bouza A, Krebs SH, Seiler RW, et al. Effects of creatine treatment on the survival of dopaminergic neurons in cultured fetal ventral mesencephalic tissue. Neuroscience. 2005;133:701-713. [PubMed: 15890457]
  93. 93. Matthews RT, Ferrante RJ, Klivenyi P, Yang L, Klein AM, Mueller G, et al. Creatine and cyclocreatine attenuate MPTP neurotoxicity. Experimental Neurology. 1999;157:142-149. [PubMed: 10222117]
  94. 94. Kooncumchoo P, Sharma S, Porter J, Govitrapong P, Ebadi M. Coenzyme Q(10) provides neuroprotection in iron-induced apoptosis in dopaminergic neurons. Journal of Molecular Neuroscience. 2006;28:125-141. [PubMed: 16679553]
  95. 95. Ono K, Hasegawa K, Naiki H, Yamada M. Preformed beta-amyloid fibrils are destabilized by coenzyme Q10 in vitro. Biochemical and Biophysical Research Communications. 2005;330:111-116. [PubMed: 15781239]
  96. 96. Yang L, Zhao K, Calingasan NY, Luo G, Szeto HH, Beal F. Mitochondria targeted peptides protect against 1-methyl-4-phenyl-1, 2, 3, 6-tetrahydropyridine neurotoxicity. Antioxidants & Redox Signaling. 2009;11(9):2095-2104
  97. 97. Lindvall O, Kokaia Z. Prospects of stem cell therapy for replacing dopamine neurons in Parkinson’s disease. Trends in Pharmacological Sciences.2009;30(5):260-267
  98. 98. Brundin P et al. Bilateral caudate and putamen grafts of embryonic mesencephalic tissue treated with lazaroids in Parkinson’s disease. Brain. 2000;123:1380-1390
  99. 99. Wenning GK et al. Short- and long-term survival and function of unilateral intrastriatal dopaminergic grafts in Parkinson’s disease. Annals of Neurology. 1997;42:95-107
  100. 100. Mendez I et al. Cell type analysis of functional fetal dopamine cell suspension transplants in the striatum and substantia nigra of patients with Parkinson’s disease. Brain. 2005;128:1498-1510
  101. 101. Isacson O et al. Toward full restoration of synaptic and terminal function of the dopaminergic system in Parkinson’s disease by stem cells. Annals of Neurology. 2003;53(Suppl. 3):S135-S146
  102. 102. Winkler C et al. Transplantation in the rat model of Parkinson’s disease: Ectopic versus homotopic graft placement. Progress in Brain Research. 2000;127:233-265
  103. 103. Brundin P et al. Functional effects of mesencephalic and adrenal chromoffin cells grafted in the rodent striatum. In: Dunnett SB, Björklund A, editors. Functional Neural Transplantation. NewYork: Raven Press; 1994. pp. 9-46
  104. 104. Jowaed A, Schmitt I, Kaut O, Wullner U. Methylation regulates alpha-synuclein expression and is decreased in Parkinson’s disease patients’ brains. The Journal of Neuroscience. 2010;30:6355-6359
  105. 105. Voutsinas GE, Stavrou EF, Karousos G, Dasoula A, Papachatzopoulou A, Syrrou M, et al. Allelic imbalance of expression and epigenetic regulation within the alpha-synuclein wild-type and p.Ala53Thr alleles in Parkinson disease. Human Mutation. 2010;31:685-691
  106. 106. Outeiro TF, Kontopoulos E, Altmann SM, Kufareva I, Strathearn KE, Amore AM, et al. Sirtuin 2 inhibitors rescue alpha-synuclein-mediated toxicity in models of Parkinson’s disease. Science. 2007;317:516-519
  107. 107. Nicholas AP, Lubin FD, Hallett PJ, Vattem P, Ravenscroft P, Bezard E, et al. Striatal histone modifications in models of levodopa-induced dyskinesia. Journal of Neurochemistry. 2008;106:486-494
  108. 108. Rappold PM, Cui M, Grima JC, Fan RZ, de Mesy-Bentley KL, Chen L, et al. Drp1 inhibition attenuates neurotoxicity and dopamine release deficits in vivo. Nature Communications. 2014;5:5244
  109. 109. Dagda RK, Cherra SJ, Kulich SM, et al. Loss of PINK1 function promotes mitophagy through effects on oxidative stress and mitochondrial fission. The Journal of Biological Chemistry. 2009;284:13843-13855
  110. 110. Grünewald A, Gegg ME, Taanman J-W, et al. Differential effects of PINK1 nonsense and missense mutations on mitochondrial function and morphology. Experimental Neurology. 2009;219:266-273
  111. 111. Zhang N, Wang S, Li Y, et al. A selective inhibitor of Drp1, mdivi-1, acts against cerebral ischemia/reperfusion injury via an anti-apoptotic pathway in rats. Neuroscience Letters. 2013;535:104-109
  112. 112. Zhao Y-X, Cui M, Chen S-F, et al. Amelioration of ischemic mitochondrial injury and Bax-dependent outer membrane permeabilization by Mdivi-1. CNS Neuroscience & Therapeutics. 2014;20:528-538

Written By

Farhin Patel and Palash Mandal

Submitted: 13 October 2018 Reviewed: 19 October 2018 Published: 20 February 2019