Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2011 Dec 1.
Published in final edited form as: Trends Neurosci. 2010 Oct 12;33(12):541–549. doi: 10.1016/j.tins.2010.09.001

Autophagy in Neurodegenerative disorders: Pathogenic Roles and Therapeutic Implications

Rebecca Banerjee 1, M Flint Beal 1, Bobby Thomas 1,*
PMCID: PMC2981680  NIHMSID: NIHMS237330  PMID: 20947179

Abstract

Autophagy is a highly conserved intracellular pathway involved in the elimination of proteins and organelles by lysosomes. Known originally as an adaptive response to nutrient deprivation in mitotic cells, autophagy is now recognized as an arbiter of neuronal survival and death decisions amongst neurodegenerative diseases. Studies using postmortem human tissue, genetic and toxin- induced animal and cellular models indicate that many of the etiological factors associated with neurodegenerative disorders can perturb the autophagic process. The emerging data support the view that dysregulation of autophagy may play a critical role in the pathogenesis of neurodegenerative disorders. In this review, we highlight the pathophysiological roles of autophagy and its potential therapeutic implications in debilitating neurodegenerative disorders, including Amyotrophic lateral sclerosis, Alzheimer’s, Parkinson’s and Huntington’s disease.

Introduction

Autophagy or “self-eating” is a lysosome-mediated degradation process for non-essential or damaged cellular constituents. It plays an important homeostatic role in cells preserving the balance between synthesis, degradation and subsequent recycling of cellular components1. During autophagy cytoplasmic constituents (including misfolded or aggregated proteins), damaged organelles (such as mitochondria, endoplasmic reticulum (ER) and peroxisomes), as well as intracellular pathogens are sequestered into double membrane autophagosomes and their contents degraded by lysosomal hydrolases. This machinery has been implicated in multiple physiological processes including protein and organelle turnover, stress response, cellular differentiation, programmed cell death, in addition to various pathological settings2. While basal levels of autophagy ensure the physiological turnover of old and damaged organelles as a garbage removal process, significant accumulation of autophagosomes is thought to represent either an alternative pathway of cell death or an ultimate survival attempt for cells in response to stress3.

Autophagosomes in neurons have been recently observed to be abundant in a growing number of neurodegenerative disorders. Such observations have provoked controversial viewpoints about whether these structures foster neuronal cell death or render neuroprotection. It is often debated whether accumulation of autophagosomes in neurons leads to an increase in autophagic activity, or rather is a consequence of impaired autophagic degradation creating a buildup of autophagosomes. Whether autophagic induction in neurons leads to cell death (traditionally mediated via conserved canonical pathways like apoptosis and necrosis4) or if it simply occurs as a process alongside these other pathways is a controversial issue. Thus, it is crucial to address this bewilderment about the opposing views in defining its fundamental role in neuronal physiology5.

It is becoming increasingly evident that autophagy may contribute differently to cell death induction according to the type and degree of the environmental changes or stress stimuli. A cell’s response may shift gradually from elimination of damaged proteins/organelles by autophagy which leads to its recovery, to the induction of apoptotic pathways determining cellular demise. However, in the context of neurodegenerative disorders an emerging consensus is the view that induction of autophagy is a neuroprotective response and that inadequate or defective autophagy, rather than excessive autophagy, promotes neuronal cell death in most of these disorders6. In this review, we provide an overview of current understanding on the putative role of autophagic mechanisms in common neurodegenerative disorders such as Alzheimer’s disease (AD), Parkinson’s disease (PD), Huntington’s disease (HD) and Amyotrophic lateral sclerosis (ALS). Furthermore, we discuss implications for how the prevailing knowledge about this degradation machinery could be harnessed for designing tangible therapeutic approaches.

The autophagic process

Autophagy was first described by Christian De Duve in 1963, as a cellular process in which a double membrane vesicle, the autophagosome, delivers cytoplasmic constituents to lysosomes for degradation and recycling7. Our understanding towards the genetic basis of this physiological process remained ambiguous until a decade ago with the discovery of over 30 autophagy related genes (ATG) and identification of their roles in the regulation and execution of autophagy at the molecular level8. This machinery is orchestrated based on the induction and presentation of target substrates such as proteins and organelles to the lysosomes via three major distinct autophagic processes: macroautophagy, microautophagy and chaperone-mediated autophagy (CMA).

Macroautophagy

Macroautophagy is a bulk degradation pathway potentially capable of degrading large protein aggregates or damaged organelles. This process of autophagy starts with the derivation of phagophore (or isolation membrane) which is likely derived from the lipid bilayer of the plasma membrane, ER and/or trans-Golgi, endosomes or mitochondria912. This phagophore expands to engulf intracellular cargo, such as protein aggregates, organelles and ribosomes, thereby sequestering the cargo in a double membranous autophagosome1. Autophagosomes are formed randomly in the cytoplasm and are then trafficked along microtubules in a dynein-dependent fashion towards the microtubule-organizing center. Once at this center, they mature through fusion with multivesicular bodies (MVB), early and/or late endosomes, before fusing with lysosomes, promoting the degradation of autophagosomal contents by lysosomal acid proteases. Once degraded by lysosomes, the lysosomal permeases and transporters export amino acids and other by-products of degradation back out to the cytoplasm, to be recycled for building macromolecules and for metabolism1 (Figure 1). The only known mammalian protein that specifically associates with the autophagosome membrane (as opposed to other vesicles) is microtubule associated protein 1 light chain 3 (MAP1 LC3), which is post-translationally modified into cytosolic LC3-I, which conjugates with phosphatidylethanolamine upon autophagy induction to form autophagosome-associated LC3-II13.

Figure 1. Steps involved in the autophagic process and their role in common neurodegenerative disorders.

Figure 1

The autophagic process involves- 1) Formation of pre-autophagosomal structures- a membrane source provides lipid bilayers for formation of a phagophore by a process known as “nucleation”. 2) Phagophore/Isolation membrane formation - here a double membranous phagophore or isolation membrane derived from pre-autophagosomal structures sequester portions of cytosol, including organelles that leads to 3) the formation of autophagosomes. 4) Maturation phase- the completed autophagosomes during this step undergo maturation, which involves steps such as fusion with multivesicular bodies or endosomes to form an amphisome. 5) Docking and fusion- during this step the inner membrane compartment fuses with a lysosome and its contents are degraded by lysosomal hydrolases. The three major forms of autophagy prevalent amongst common neurodegenerative disorders are macroautophagy, microautophagy and chaperone-mediated autophagy (see text for the detailed description of the three types of autophagy and how they are associated with pathogenesis of common neurodegenerative disorders).

Microautophagy

Microautophagy is somewhat similar to macroautophagy, but far less is understood about it in mammalian systems compared to yeast. Here, the lysosomes directly engulf cytoplasm bypassing the need for autophagosome and autophagolysosome formation. It involves the pinocytosis of small quantities of cytosol directly by lysosomes14. Microautophagy is active in the resting state and is responsible for the removal of selective organelles and continuous turnover of intracellular constituents. It is not activated by nutritional deprivation or stress. Microautophagy and macroautophagy are both able to engulf large structures through selective and non-selective mechanisms (Figure 1).

Chaperone-mediated autophagy

Chaperone-mediated autophagy (CMA) is quite distinct from macroautophagy and microautophagy processes, in terms of selectivity and mechanism of degradation. CMA is responsible for the selective degradation of 30% of all soluble cytosolic proteins harboring a unique pentapeptide motif (KFERQ). This amino acid sequence is specifically recognized by a cytosolic chaperone, heat-shock cognate 70 (hsc70) and its co-chaperones, which selectively target these proteins to lysosomes for degradation. The substrate/chaperone complex moves to the lysosomes, where the CMA receptor lysosome-associated membrane protein type-2A (LAMP-2A) recognizes the protein, which is then unfolded and translocated across the lysosome membrane assisted by the lysosomal hsc70 inside the lysosomes for degradation15, 16 (Figure 1). CMA differs from macroautophagy and microautophagy since the substrates are transported across the lysosomal membrane on a one-by-one basis, where as in the macroautophagy and microautophagy, the substrates are engulfed or sequestered in bulk. CMA is active under basal conditions but stress can enhance this pathway. While macroautophagy can be induced due to nutritional scarcity for a brief period, induction of CMA may happen over prolonged periods of nutrient deprivation15, 16.

Regulation of autophagy in common neurodegenerative disorders

Cytoplasmic, nuclear and extracellular inclusions composed of aggregated and ubiquitinated proteins comprise key pathological hallmarks of numerous neurodegenerative diseases. Since protein aggregation is believed to contribute to organelle damage, synaptic dysfunction and neuronal degeneration, intracellular protein aggregate clearance pathways have been suggested as a potential therapeutic approach for such disorders. It is now apparent that aggregate-prone proteins and damaged organelles seen among a variety of neurodegenerative disorders are eliminated more efficiently via the autophagy-lysosome pathways in contrast to the ubiquitin-proteasome machinery. Here, we provide defining characteristics of autophagic pathways and their contributions to the underlying mechanisms of disease pathogenesis amongst common neurodegenerative disorders (Table 1) and describe potential targets for therapeutic modulation (Table 2).

Table 1.

Proteins involved in autophagy and their disease modifying role in common neurodegenerative disorders.

Autophagy Signaling Factors Cellular function(s) Clinical Relevance Disease Model
Beclin-1. Involved in autophagosome formation Reduced beclin-1 expression in postmortem AD brains22.
Beclin-1 sequestration into neuronal intranuclear inclusions in HD patients- reduced beclin-1 function lead to impaired autophagic clearance of mutant htt62.
Beclin-1 induction in sporadic ALS spinal cord – believed to promote autophagy66
Beclin-1 deficiency in APP transgenic AD mice leads to Aβ accumulation due to reduced autophagy. Beclin-1 over- expression attenuates this pathological phenotype22.
Beclin-1 gene transfer activated autophagy and attenuated neuropathology in α-synuclein transgenic PD mice74.
Induction of Beclin-1 in SOD1 ALS mice66.
mTOR (Phosphatidyl kinase-related kinase) Negatively regulates autophagy mTOR levels are dramatically increased in AD brains correlating with tau pathology30. mTOR inhibition induced autophagy with neuroprotective effects in AD mice and in HD flies and mice30, 31, 58.
Elevated mTOR levels are observed in α-synuclein transgenic mice75.
LC3 Conversion of LC3I to LC3II is indicative of autophagosome formation. Marked induction of LC3 in sporadic and familial ALS spinal cords – suggestive of enhanced autophagy67. LC3-II levels increased in symptomatic SOD1 ALS mice67.
p62/Sequestosome 1 Autophagic adaptor that interacts with LC3. p62 immunoreactivity observed in neurofibrially tangles of postmortem AD brains76. p62-dependent PINK1/parkin-mediated autophagy in non neuronal and neuronal cells77.
p62-mediated recognition and targeting of mutant SOD1 for autophagic degradation78.
Dynein Autophagosome-lysosome fusion. Dynein mutations – familial ALS79. Dynein-loss-of function caused premature mutant huntingtin aggregate formation in HD flies and mice, increased LC3II in cell and mice models80.
Parkin An E3 ligase that facilitates mitophagy. Parkin loss-of-function mutations – Autosomal recessive juvenile PD81. Parkin mutations fail to ubiquitinate defective mitochondria causing mitophagic deficits44, 45.
PINK1 A mitochondrial serine-threonine kinase involved in mitochondrial fission, mitochondrial quality control, and mitophagy. PINK1 loss-of-function mutations – Autosomal recessive PD82. PINK1 loss-of-function promoted mitophagy as a compensatory response47. PINK1 mutants impaired parkin-mediated ubiquitin signaling and recruitment to mitochondria for mitophagy48, 49, 83, 84.

Table 2.

Autophagy modulators and their disease modifying action in common neurodegenerative disorders.

Autophagy Modulators Mode of Action Disease Pathology Modifier
Rapamycin CCI-779 (rapamycin analog)

Glucose transporter (GLUT1)
Inhibit mTOR activating autophagy

Raise intracellular glucose or glucose-6- phosphate
Clearance of mutant α-synuclein in cells37.
Reduces Aβ levels and associated cognitive deficits31.
Decreased mutant aggregate-prone tau proteins in flies85.
Reduced toxicity of polyglutamine expansion in fly and mouse HD models58, 86, 87.
Small molecule enhancers (SMERs) mTOR-independent autophagic inducers Mutant htt clearance86.
Mutant A53T α-synuclein reduction86.
Lithium, Sodium valproate, carbamazepine, L-690, 330. Inositol lowering agents – inhibit inositol monophosphatase (IMPase)
mTOR-independent autophagic inducer
Increased survival in ALS mouse88.
Improved outcome in ALS clinical trial89.
Mutant Htt aggregation/toxicity reduction90.
Trehalose disaccharide, Minoxidil, Clonidine mTOR-independent autophagic inducer Increased autophagic flux, clearance of mutant htt and α-synuclein mutants91, 92.
Ameliorates dopaminergic and tau pathology93.
Combination Therapy: Rapamycin + Trehalose/Calpastatin/SMERs mTOR-dependent and mTOR-independent targets Additive effect-enhanced autophagic clearance of protein aggregates91, 92.

Alzheimer’s disease

AD is the most common neurodegenerative disorder characterized by the pathogenic accumulation of β-amyloid (Aβ) peptides, generated by the proteolytic cleavage of amyloid precursor protein (APP) via β- and γ-secretases that are localized in secretory and endocytic compartments of affected cells. Autophagosomes and endosomes within neurons actively form in synapses and along neuritic processes. Efficient trafficking of autophagosomes along microtubules is essential for autophagic degradation in neurons given the distances from neuronal processes, where endosomes and autophagosomes are formed, and the neuronal soma where the lysosomes are concentrated. In AD there is massive accumulation of autophagosomes within large swellings along dystrophic and degenerating neurites, primarily due to deficits in the maturation of autophagosomes and their retrograde transport towards the neuronal cell body17.

In general, autophagy is considered to be activated in AD primarily because of impaired clearance of autophagosomes that contain both APP and its processing enzymes, thereby increasing the propensity to generate toxic Aβ peptides18. Although most Aβ formed during autophagy is normally degraded within lysosomes via macroautophagy, Aβ also accumulates within the large pool of autophagosomes in dystrophic neurites and becomes a major intracellular reservoir of toxic peptides in AD brains. It is well established that Aβ can disrupt autophagosome membranes and trigger the release of hydrolytic enzymes into the cytoplasm. It is suggested that local accumulation of autophagosomes in dystrophic neurons may contribute to Aβ generation within plaques and that increased autophagy in the neuropil could be a major source of Aβ 17, 19. This is achieved by increased turnover of APP together with an enrichment of the γ-secretase complex inside autophagosomes that cleaves APP to Aβ. In fact, autophagosome membranes are rich in presenilin 1, a component of the APP cleaving γ-secretase complex which is encoded by the presenilin gene whose mutations lead to early onset forms of autosomal dominant AD. Presenilin 1 mutant human AD brains show accumulation of extensive lysosomal pathology, together with amyloid pathology and neurodegeneration20. It was recently demonstrated that neurons and blastocysts from presenilin 1 hypomorphic mice and fibroblasts from presenilin 1 mutant AD patients show loss of macroautophagy, resulting in increased Aβ accumulation due to impaired maturation of the V0a1 subunit of the bimodular v-type H+-ATPase proton pump that acidifies the lysosome21. Furthermore, levels of beclin 1 (or ATG6), required for the formation of autophagosomes, have been consistently found to be significantly reduced in postmortem AD brains22. In a transgenic mouse model of AD that expresses human APP, genetic reduction of beclin 1 expression stimulated formation of Aβ accumulation and neurodegeneration due to a decrease in autophagy22. Conversely, increased expression of beclin 1 in APP-transgenic mice significantly reduced amyloid pathology and neurodegeneration.

If reduced induction of autophagy is suggested to be the underlying mechanism for the observed higher levels of Aβ peptides and aggregates, then these results are consistent with the findings that the maturation of autophagosomes may be impaired in AD22. However, Aβ levels are known to be affected by both production and clearance23. Therefore, defects in autophagosome maturation could enhance Aβ production, whereas reduced autophagy through beclin 1 deficiency may inhibit Aβ clearance by altering APP metabolism24. Studies indicate ultrastructural alterations in mitochondrial morphology, such as reduced size, broken cristae and abnormalities in mitochondrial dynamics in AD25, 26. Sporadic AD patient fibroblasts27 and cells overexpressing the Swedish variant of APP28 demonstrate an imbalance in mitochondrial fission/fusion proteins, either via post-translational modification such as S-nitrosylation29, or by alterations in their expression26, 28, leading to an increase in mitochondrial fission. It is suggested that increased mitochondrial fission is probably an attempt to segregate and eliminate damaged mitochondria by a macroautophagic process known as “mitophagy”, which is consistent with a reduction in the level and size of mitochondria observed in human post-mortem tissue 25.

Alterations in the mammalian Target Of Rapamycin (mTOR) pathway, which is known to play a central role in signaling induced by nutrients and growth factors, is suggested to support autophagic failure in AD. Accumulation of Aβ peptides has been shown to increase signaling of the mTOR pathway, whereas decreasing mTOR signaling has been shown to reduce Aβ levels30, 31. In a mouse model of AD, pharmacological inhibition of mTOR signaling via rapamycin (an FDA-approved drug) rescues cognitive deficits and ameliorates amyloid and tau pathology by increasing autophagy31. Rapamycin-induced inhibition of mTOR led to increased neuronal autophagy in the mutant AD mice, suggesting that the reduction in Aβ levels and the improvement in cognitive function are due in part to increased autophagy via mTOR inhibition. Considering that age is an important risk factor for the development of AD and many other neurodegenerative disorders rapamycin administration in animals was recently shown to extend lifespan in multiple studies32, 33. This effect of rapamycin on lifespan due to increased autophagy further reinforces the benefits of promoting autophagy in age-related neurodegenerative disorders. Thus, autophagy induction may serve as a potential therapeutic target for blocking AD pathogenesis.

Parkinson’s disease

PD is a chronic neurodegenerative disorder characterized by progressive and relentless degeneration of dopaminergic neurons, especially within the substantia nigra34. Representing the multifactorial causes of PD, the pattern of neurodegeneration is complex, having features of necrosis, apoptosis35 and accumulation of autophagosome like structures36. The first evidence in support of a significant role of autophagy in PD came from the demonstration that α-synuclein, which is a major constituent of Lewy bodies found in PD, is degraded by macroautophagy and CMA3739. Inhibition of CMA has been shown to cause formation of high molecular weight and detergent-insoluble species of α-synuclein39, suggesting that clearance of α-synuclein by CMA is crucial for limiting the oligomerization of α-synuclein. Importantly, A53T and A30P mutants of α-synuclein, which cause autosomal dominant PD, were found to inhibit CMA through exhibiting a higher binding affinity for the lysosomal marker LAMP2A compared with wild-type α-synuclein38. However, mutant α-synuclein is poorly internalized into lysosomes, and degraded by macroautophagy instead of CMA. Accompanying the inhibition of CMA by α-synuclein mutants is a compensatory activation of macroautophagy38, although the physiological significance of this event is incompletely understood. Nonetheless, these observations suggest that A53T and A30P α-synuclein mutants may induce α-synuclein aggregation through inhibition of CMA, thereby leading to impaired clearance of these proteins.

Several posttranslational modifications, including monoubiquitination of α-synuclein, impair degradation of this protein by CMA further, but not degradation of other substrates38, 40. A modified version of α-synuclein is suggested to be responsible for neuron toxicity via a non-covalent interaction of α-synuclein and oxidized dopamine41. Dopamine-modified α-synuclein is not only poorly degraded by CMA, but also blocks degradation of other substrates, thereby increasing cellular vulnerability to stressors41. This is significant since oxidized metabolites of dopamine such as dopamine quinone derivatives are thought to play a pivotal role in the degeneration of nigrostriatal dopaminergic neurons41 and further reveal that inhibition of CMA-mediated degradation of α-synuclein may constitute an important pathogenic mechanism for PD. Interestingly, it was recently demonstrated that α-synuclein may also contribute to neuronal death in PD by inhibiting CMA-mediated degradation of MEF2D (survival factor myocyte specific enhancer factor 2D)42. Both wild type and A53T mutant α-synuclein interfered with the binding of MEF2D to the chaperone hsc70 in the CMA degradation machinery. Consistent with this observation, MEF2D levels were elevated in A53T α-synuclein transgenic mice, and also in postmortem brain tissue from PD patients42. Furthermore, over-expression of both wild type and A53T mutant forms of α-synuclein inhibited MEF2 activity, leading to neurodegeneration42. These observations reveal that in addition to α-synuclein, impaired CMA may also trigger neuronal death through inefficient degradation of other proteins. These findings collectively raise the interesting possibility that preservation of CMA functions may serve as an effective treatment strategy against PD.

Besides the degradation of α-synuclein, the autophagic pathway is also involved in the turnover of mitochondria in cells. Mitochondrial dysfunction represents one of the crucial pathogenic mechanisms in PD43. Mutations in genes such as parkin and PINK1 are known to cause autosomal recessive forms of PD and have been implicated in the control of mitochondrial morphology and function43. How parkin and PINK1 affects mitochondrial functionality remains poorly understood. Recently, parkin was found to facilitate macroautophagy of impaired mitochondria, by mitophagy44. Parkin is selectively targeted to functionally impaired and depolarized mitochondria, which are then eliminated by mitophagy44. The selective targeting of impaired mitochondria for mitophagy is brought about via a functional parkin-mediated ubiquitination of impaired mitochondria that serves to recruit ubiquitin-binding histone deacetylase, HDAC6, and p62, which assemble the autophagic machinery for efficient degradation of the impaired mitochondria45. It has been suggested that mitophagy could protect neurons by eliminating dysfunctional mitochondria, which may produce toxic reactive oxygen species that damage neurons46. In support of a role for autophagy in the clearance of defective mitochondria in PD, knockdown of PINK1 expression induces mitochondrial fragmentation, followed by activation of autophagy/mitophagy47. Moreover, it was recently shown that targeting of parkin to impaired mitochondria during mitophagy relies on the expression of wild type PINK1, but not PD-associated PINK1 mutants48. This occurs through the phosphorylation-dependent regulation of the ubiquitin E3 ligase activity of parkin49, suggesting a pathogenic role of PINK1 and parkin loss-of-function mutations in PD. These observations suggest that the autophagic pathway is essential for the turnover of dysfunctional mitochondria in PD. Failure to activate efficient mitophagy may serve as an important pathogenic mechanism of cell death in PD.

Huntington’s disease

HD is a progressive, autosomal dominant, neurodegenerative disorder caused by the expansion of CAG trinucleotide repeats (> 35 repeats) in the huntingtin (htt) gene, which is translated into an expanded polyglutamine tract in the N-terminus of the htt protein. Mutant htt toxicity is believed to be expressed after it is cleaved to form N-terminal fragments comprising the first 100–150 residues with the expanded polyglutamine tract. Although HD and other polyglutamine disorders are associated with intraneuronal aggregates, it is debatable whether these aggregates are toxic or protective50. Several lines of evidence have implicated the preaggregate oligomers as the most toxic species51. However, studies indicate that induction of autophagy results in the decrease of both aggregated and soluble ‘monomeric’ htt species, and results in decreased toxicity in various in vitro and in vivo models of HD52. Postmortem brains of patients with HD have endosomal and/or lysosomal organelles, and multivesicular bodies, characteristic features of autophagy53. Increased numbers of autophagosomes have been also found in lymphoblasts of HD patients in comparison to control lymphoblasts54. Mutant htt is known to induce endosomal and/or lysosomal activity55 and mouse clonal striatal cells transiently transfected with truncated and full length human wild-type and mutant htt show the presence of both normal and mutant proteins in dispersed and perinuclear vacuoles56. Strong support for the involvement of autophagy in HD is provided by examining mTOR which is sequestrated into aggregates of mutant htt in cell models, transgenic mice and human brains, leading to decreased mTOR activity fostering the induction of autophagy to clear htt aggregates 57,58. Numerous studies also suggest that pharmacological modifiers of mTOR-dependent and independent autophagic pathways can provide neuroprotection in various models of HD by inhibiting polyglutamine aggregation (Table 2).

Htt protein has been shown to modulate autophagy through its stress sensitive ER membrane association domain59. In response to ER stress, htt releases from membranes and rapidly translocates into the nucleus. However, this membrane release is inhibited when htt contains the polyglutamine expansion seen in HD, leading to perturbed ER function and an increase in autophagosomes. Notably, expression of htt with a deleted polyglutamine tract in a knockin mouse (Hdh (140Q/+)) model of HD showed upregulated autophagic markers and increased lifespan60. A recent study has also shown that despite the formation of autophagosomes at enhanced rates in HD cells there is a failure to efficiently trap cytosolic cargo in the lumen61. This phenomenon might explain how inefficient engulfment of cytosolic components by autophagosomes is responsible for the slower turnover of htt, leading to htt accumulation in HD61.

Another mechanism for increased accumulation of htt is via sequestration of beclin 1 into neuronal intranuclear inclusions. This has been demonstrated to occur in patients with HD, leading to reduced autophagic clearance of mutant huntingtin62. Expression of beclin 1 has been shown to decrease in an age-dependent fashion in human brains62. As the beclin 1 gene is haploid insufficient in regulating autophagosome function, it is proposed that an age-dependent decrease in beclin 1 expression may lead to a reduction in autophagic activity during aging, that in turn promotes accumulation of mutant htt and disease progression. The significance of autophagy in HD pathogenesis is further supported by the recent finding that age of onset of HD is modified by V471A polymorphism in ATG763. A general consensus that has emerged from these studies is that blocking autophagy will reduce cell viability and increase the number of cells bearing mutant htt aggregates, whereas stimulating autophagy promotes htt degradation. A recent report provides strong evidence that stimulating autophagy is indeed beneficial for HD. Enhancement of the CMA was achieved by expressing a fusion molecule comprising polyglutamine binding peptide 1 (QBP1) and hsc70-binding motifs in cellular and mouse models of HD. This was found to result in the selective targeting of mutant htt for degradation and ameliorated the disease phenotype in a HD mouse model64. QBP1 is known to bind only the expanded polyglutamine tract seen in mutant htt and not wild-type htt. Such novel strategies using similar adaptor molecules comprising hsc70-binding motifs fused to an appropriate structure-specific binding agent(s) may have therapeutic potential for treating diseases caused by misfolded proteins other than those with expanded polyglutamine tracts. Thus taken together, autophagy may represent an initial attempt of HD neurons to eliminate mutant htt protein, however, over the course of the disease the autophagy processes become overloaded, ineffective and dysfunctional, eventually resulting in neurodegeneration. Induction of autophagy enhances the clearance of both soluble and aggregated forms of mutant htt, and protects against toxicity caused by these mutations in several disease models of HD.

Amyotrophic lateral sclerosis

In ALS, motor neurons degenerate leading to respiratory failure and fatality. A potential link between autophagy and ALS was initially based on morphological findings from postmortem spinal cords of sporadic and familial ALS (fALS) patients and subsequently from various genetic models of the disease exhibiting autophagic abnormalities6567. The presence of abundant autophagosomes and associated increases in autophagy proteins suggested that autophagy activation is detrimental for the survival of motor neurons. In fact, it is quite common to find evidence of increased recruitment of the autophagic machinery to compensate for an increase in misfolded proteins and dysfunctional organelles in motor neurons68. However, other studies have found an increase in the LC3II macroautophagy marker protein and a decreased ratio of phosphorylated mTOR-positive motor neurons suggestive of defective autophagy associated with motor neuron loss65, 67. Similarly, defective autophagy is also suggested in the accumulation of ubiquitinated TAR DNA-binding protein 43kD (TDP-43) inclusions in ALS69 and in motor neuron degeneration seen in ALS due to mutations in endosomal sorting complexes required for transport subunit III (ESCRT III) and CHMP2B70, 71.

Two recent studies suggest that autophagic clearance of mutant superoxide dismutase 1 (SOD1) is beneficial for motor neuron loss in ALS. The heat-shock protein HspB8 increases mutant SOD1 clearance by promoting autophagy in an ALS mouse model72. Furthermore, the unfolded protein response transcription factor X-box binding protein 1 (XBP-1) deficiency in mice has been shown to protect against ALS by autophagic induction and the subsequent degradation of mutant SOD1 66. While the recruitment of the autophagy system in ALS has been well documented, its significance (either detrimental or beneficial) for neuronal survival has been largely speculative up till now. A plausible explanation could be that mutant SOD1 in motor neurons may impair the autophagic machinery to promote neurodegeneration whereas autophagic clearance of SOD1 aggregates could be beneficial for the survival of motor neurons. The recent realization that autophagic induction may be an important therapeutic possibility for ALS has lead to clinical trials to assess the merits of autophagy inducers like lithium and rapamycin for neuroprotection in ALS patients73. Notwithstanding this, in-depth studies of the precise role of autophagy in ALS- associated motor neuron death remain to be performed.

Concluding remarks

Protein aggregation and organellar dysfunction are characteristic features of many late-onset neurodegenerative diseases. The ensuing accumulation of damaged proteins and organelles result in a mass of toxic debris within the afflicted neurons that results in neuronal dysfunction and may ultimately cause cell death. Accumulating evidence suggests that accelerating the removal of toxic accumulation of damaged membranes, organelles, and proteins may be a tractable therapeutic strategy for neurodegenerative disorders. The ubiquitin-proteasome system and the autophagy-lysosome pathways are the major routes of clearance for toxic proteins. The ubiquitin-proteasome pathway is mainly responsible for the breakdown and degradation of short-lived proteins with low-medium molecular weights; the narrow proteasome barrel precludes the entry of oligomers and high molecular weight protein aggregates commonly seen in neurodegenerative disorders. These latter proteins and aggregates, as well as cytoplasmic contents such as damaged membranes and organelles, can undergo bulk degradation in an efficient manner via distinct types of autophagy-lysosome pathways. Hence, in neurodegenerative diseases where autophagic pathway(s) is dysregulated, therapeutic interventions that activate autophagy may be beneficial in the prevention of neuronal cell death. However, in the present scenario, the consequences of manipulating the autophagic process are likely complex. In this regard, just as down-regulation or partial inhibition of autophagy could provoke or aggravate neurodegeneration, excessive activation of autophagy may invoke a similar response such as “self-cannibalism” in neurons. Thus, it is probable that this critical biological process will require an accurate titration to ensure that its activity is carefully controlled within physiological settings.

One may suggest that a beneficial or detrimental contribution of autophagy in the pathogenesis of neurodegenerative disorders could strictly rely on precisely when (the temporal aspect), where (the spatial feature), and how much (the quantity) activation. Even if one ensures that engulfment of protein aggregates or organelles by autophagosomes during autophagy is well-controlled, another crucial factor will be the efficient transport of these autophagosomes through the long processes of neurons to the site of lysosomes (predominantly in the neuronal soma). Future avenues of research (Box 1) should focus on understanding the intricate relationships between temporal, spatial and quantitative aspects and factors regulating the transport of autophagosomes during this degradation process. One can only hope that uncovering these finer details of this complex machinery will not only yield insights to pathophysiological mechanisms underlying common neurodegenerative disorders but also help us determine unique interventional targets for therapies that can be modulated to treat these disorders that presently lack effective treatments.

Box 1. Outstanding questions.

  • How are the different autophagic processes regulated in neurons both during physiological conditions and in the context of specific neurodegenerative disorders? Are these processes prevalent in all neurons that undergo degeneration and follow common regulatory mechanisms?

  • What leads to a decline in autophagy during aging and in age-related neurodegenerative disorders? Does an aging-associated decline in autophagy lead to inability of a neuron to clear undesired cytosolic components?

  • What is the precise relationship between the autophagy and the ubiquitin proteasome systems? Do these systems work in a collaborative manner? What determines this cross-talk between proteolytic systems and other signaling pathways? Are ubiquitinated proteins turned over via autophagy, or does ubiquitination occur in proteins that cannot be properly eliminated from the cytosol through other systems of elimination?

  • What factors control cargo recognition and transport of autophagosomes to sites of lysosomes in a cell during autophagy? What kinds of signals trigger mitophagy? Is it selective to damaged mitochondria? What factors tag damaged mitochondria for degradation? How do alterations in the mitochondrial fission/fusion machinery impact this process? Is mitophagy universal for all age-related neurodegenerative disorders that demonstrate mitochondrial dysfunction?

  • Can the activation of autophagic pathways be accurately titrated within the physiological range to allow this strategy to be effectively used for therapeutic intervention in neurological disorders?

Acknowledgments

This work is supported by National Institutes of Health grants NS060885, NS062165 (B.T.) and ES017295 (M.F.B.), and grants from Michael J Fox Foundation for Parkinson’s disease (B.T and M.F.B) and the Department of Defense (M.F.B.).

Footnotes

Disclosure/Conflict of Interest: The authors declare no conflict of interest.

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

References

  • 1.Mizushima N. Autophagy: process and function. Genes Dev. 2007;21:2861–2873. doi: 10.1101/gad.1599207. [DOI] [PubMed] [Google Scholar]
  • 2.Shintani T, Klionsky DJ. Autophagy in health and disease: a double-edged sword. Science. 2004;306:990–995. doi: 10.1126/science.1099993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Mizushima N, et al. Autophagy fights disease through cellular self-digestion. Nature. 2008;451:1069–1075. doi: 10.1038/nature06639. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Galluzzi L, et al. Cell death modalities: classification and pathophysiological implications. Cell Death Differ. 2007;14:1237–1243. doi: 10.1038/sj.cdd.4402148. [DOI] [PubMed] [Google Scholar]
  • 5.Debnath J, et al. Does autophagy contribute to cell death? Autophagy. 2005;1:66–74. doi: 10.4161/auto.1.2.1738. [DOI] [PubMed] [Google Scholar]
  • 6.Wong E, Cuervo AM. Autophagy gone awry in neurodegenerative diseases. Nat Neurosci. 2010;13:805–811. doi: 10.1038/nn.2575. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.De Duve C. The lysosome. Sci Am. 1963;208:64–72. doi: 10.1038/scientificamerican0563-64. [DOI] [PubMed] [Google Scholar]
  • 8.Klionsky DJ. Autophagy: from phenomenology to molecular understanding in less than a decade. Nat Rev Mol Cell Biol. 2007;8:931–937. doi: 10.1038/nrm2245. [DOI] [PubMed] [Google Scholar]
  • 9.Axe EL, et al. Autophagosome formation from membrane compartments enriched in phosphatidylinositol 3-phosphate and dynamically connected to the endoplasmic reticulum. J Cell Biol. 2008;182:685–701. doi: 10.1083/jcb.200803137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Simonsen A, Tooze SA. Coordination of membrane events during autophagy by multiple class III PI3-kinase complexes. J Cell Biol. 2009;186:773–782. doi: 10.1083/jcb.200907014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Hailey DW, et al. Mitochondria supply membranes for autophagosome biogenesis during starvation. Cell. 2010;141:656–667. doi: 10.1016/j.cell.2010.04.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Ravikumar B, et al. Plasma membrane contributes to the formation of pre-autophagosomal structures. Nat Cell Biol. 2010;12:747–757. doi: 10.1038/ncb2078. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Kabeya Y, et al. LC3, a mammalian homologue of yeast Apg8p, is localized in autophagosome membranes after processing. EMBO J. 2000;19:5720–5728. doi: 10.1093/emboj/19.21.5720. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Todde V, et al. Autophagy: principles and significance in health and disease. Biochim Biophys Acta. 2009;1792:3–13. doi: 10.1016/j.bbadis.2008.10.016. [DOI] [PubMed] [Google Scholar]
  • 15.Massey AC, et al. Chaperone-mediated autophagy in aging and disease. Curr Top Dev Biol. 2006;73:205–235. doi: 10.1016/S0070-2153(05)73007-6. [DOI] [PubMed] [Google Scholar]
  • 16.Kon M, Cuervo AM. Chaperone-mediated autophagy in health and disease. FEBS Lett. 2010;584:1399–1404. doi: 10.1016/j.febslet.2009.12.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Nixon RA. Autophagy, amyloidogenesis and Alzheimer disease. J Cell Sci. 2007;120:4081–4091. doi: 10.1242/jcs.019265. [DOI] [PubMed] [Google Scholar]
  • 18.Yu WH, et al. Macroautophagy--a novel Beta-amyloid peptide-generating pathway activated in Alzheimer’s disease. J Cell Biol. 2005;171:87–98. doi: 10.1083/jcb.200505082. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Nixon RA, et al. Extensive involvement of autophagy in Alzheimer disease: an immuno-electron microscopy study. J Neuropathol Exp Neurol. 2005;64:113–122. doi: 10.1093/jnen/64.2.113. [DOI] [PubMed] [Google Scholar]
  • 20.Cataldo AM, et al. Presenilin mutations in familial Alzheimer disease and transgenic mouse models accelerate neuronal lysosomal pathology. J Neuropathol Exp Neurol. 2004;63:821–830. doi: 10.1093/jnen/63.8.821. [DOI] [PubMed] [Google Scholar]
  • 21.Lee JH, et al. Lysosomal proteolysis and autophagy require presenilin 1 and are disrupted by Alzheimer-related PS1 mutations. Cell. 2010;141:1146–1158. doi: 10.1016/j.cell.2010.05.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Pickford F, et al. The autophagy-related protein beclin 1 shows reduced expression in early Alzheimer disease and regulates amyloid beta accumulation in mice. J Clin Invest. 2008;118:2190–2199. doi: 10.1172/JCI33585. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Selkoe DJ. Clearing the brain’s amyloid cobwebs. Neuron. 2001;32:177–180. doi: 10.1016/s0896-6273(01)00475-5. [DOI] [PubMed] [Google Scholar]
  • 24.Jaeger PA, et al. Regulation of amyloid precursor protein processing by the Beclin 1 complex. PLoS One. 2010;5:e11102. doi: 10.1371/journal.pone.0011102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Hirai K, et al. Mitochondrial abnormalities in Alzheimer’s disease. J Neurosci. 2001;21:3017–3023. doi: 10.1523/JNEUROSCI.21-09-03017.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Santos RX, et al. A synergistic dysfunction of mitochondrial fission/fusion dynamics and mitophagy in Alzheimer’s disease. J Alzheimers Dis. 2010;20(Suppl 2):S401–412. doi: 10.3233/JAD-2010-100666. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Wang X, et al. Dynamin-like protein 1 reduction underlies mitochondrial morphology and distribution abnormalities in fibroblasts from sporadic Alzheimer’s disease patients. Am J Pathol. 2008;173:470–482. doi: 10.2353/ajpath.2008.071208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Wang X, et al. Amyloid-beta overproduction causes abnormal mitochondrial dynamics via differential modulation of mitochondrial fission/fusion proteins. Proc Natl Acad Sci U S A. 2008;105:19318–19323. doi: 10.1073/pnas.0804871105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Cho DH, et al. S-nitrosylation of Drp1 mediates beta-amyloid-related mitochondrial fission and neuronal injury. Science. 2009;324:102–105. doi: 10.1126/science.1171091. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Caccamo A, et al. Molecular interplay between mammalian target of rapamycin (mTOR), amyloid-beta, and Tau: effects on cognitive impairments. J Biol Chem. 2010;285:13107–13120. doi: 10.1074/jbc.M110.100420. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Spilman P, et al. Inhibition of mTOR by rapamycin abolishes cognitive deficits and reduces amyloid-beta levels in a mouse model of Alzheimer’s disease. PLoS One. 2010;5:e9979. doi: 10.1371/journal.pone.0009979. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Harrison DE, et al. Rapamycin fed late in life extends lifespan in genetically heterogeneous mice. Nature. 2009;460:392–395. doi: 10.1038/nature08221. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Anisimov VN, et al. Rapamycin extends maximal lifespan in cancer-prone mice. Am J Pathol. 2010;176:2092–2097. doi: 10.2353/ajpath.2010.091050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Thomas B. Parkinson’s disease: from molecular pathways in disease to therapeutic approaches. Antioxid Redox Signal. 2009;11:2077–2082. doi: 10.1089/ars.2009.2697. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Levy OA, et al. Cell death pathways in Parkinson’s disease: proximal triggers, distal effectors, and final steps. Apoptosis. 2009;14:478–500. doi: 10.1007/s10495-008-0309-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Alvarez-Erviti L, et al. Chaperone-Mediated Autophagy Markers in Parkinson Disease Brains. Arch Neurol. 2010 Aug 9; doi: 10.1001/archneurol.2010.198. [Epub ahead of print] [DOI] [PubMed] [Google Scholar]
  • 37.Webb JL, et al. Alpha-Synuclein is degraded by both autophagy and the proteasome. J Biol Chem. 2003;278:25009–25013. doi: 10.1074/jbc.M300227200. [DOI] [PubMed] [Google Scholar]
  • 38.Cuervo AM, et al. Impaired degradation of mutant alpha-synuclein by chaperone-mediated autophagy. Science. 2004;305:1292–1295. doi: 10.1126/science.1101738. [DOI] [PubMed] [Google Scholar]
  • 39.Vogiatzi T, et al. Wild type alpha-synuclein is degraded by chaperone-mediated autophagy and macroautophagy in neuronal cells. J Biol Chem. 2008;283:23542–23556. doi: 10.1074/jbc.M801992200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Engelender S. Ubiquitination of alpha-synuclein and autophagy in Parkinson’s disease. Autophagy. 2008;4:372–374. doi: 10.4161/auto.5604. [DOI] [PubMed] [Google Scholar]
  • 41.Martinez-Vicente M, et al. Dopamine-modified alpha-synuclein blocks chaperone-mediated autophagy. J Clin Invest. 2008;118:777–788. doi: 10.1172/JCI32806. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Yang Q, et al. Regulation of neuronal survival factor MEF2D by chaperone-mediated autophagy. Science. 2009;323:124–127. doi: 10.1126/science.1166088. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Banerjee R, et al. Mitochondrial dysfunction in the limelight of Parkinson’s disease pathogenesis. Biochim Biophys Acta. 2009;1792:651–663. doi: 10.1016/j.bbadis.2008.11.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Narendra D, et al. Parkin is recruited selectively to impaired mitochondria and promotes their autophagy. J Cell Biol. 2008;183:795–803. doi: 10.1083/jcb.200809125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Lee JY, et al. Disease-causing mutations in parkin impair mitochondrial ubiquitination, aggregation, and HDAC6-dependent mitophagy. J Cell Biol. 2010;189:671–679. doi: 10.1083/jcb.201001039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Chu CT. A pivotal role for PINK1 and autophagy in mitochondrial quality control: implications for Parkinson disease. Hum Mol Genet. 2010;19:R28–37. doi: 10.1093/hmg/ddq143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Dagda RK, et al. Loss of PINK1 function promotes mitophagy through effects on oxidative stress and mitochondrial fission. J Biol Chem. 2009;284:13843–13855. doi: 10.1074/jbc.M808515200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Vives-Bauza C, et al. PINK1-dependent recruitment of Parkin to mitochondria in mitophagy. Proc Natl Acad Sci U S A. 2010;107:378–383. doi: 10.1073/pnas.0911187107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Sha D, et al. Phosphorylation of parkin by Parkinson disease-linked kinase PINK1 activates parkin E3 ligase function and NF-kappaB signaling. Hum Mol Genet. 2010;19:352–363. doi: 10.1093/hmg/ddp501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Orr HT, Zoghbi HY. Trinucleotide repeat disorders. Annu Rev Neurosci. 2007;30:575–621. doi: 10.1146/annurev.neuro.29.051605.113042. [DOI] [PubMed] [Google Scholar]
  • 51.Lansbury PT, Lashuel HA. A century-old debate on protein aggregation and neurodegeneration enters the clinic. Nature. 2006;443:774–779. doi: 10.1038/nature05290. [DOI] [PubMed] [Google Scholar]
  • 52.Rubinsztein DC, et al. Potential therapeutic applications of autophagy. Nat Rev Drug Discov. 2007;6:304–312. doi: 10.1038/nrd2272. [DOI] [PubMed] [Google Scholar]
  • 53.Tellez-Nagel I, et al. Studies on brain biopsies of patients with Huntington’s chorea. J Neuropathol Exp Neurol. 1974;33:308–332. doi: 10.1097/00005072-197404000-00008. [DOI] [PubMed] [Google Scholar]
  • 54.Sawa A, et al. Huntingtin is cleaved by caspases in the cytoplasm and translocated to the nucleus via perinuclear sites in Huntington’s disease patient lymphoblasts. Neurobiol Dis. 2005;20:267–274. doi: 10.1016/j.nbd.2005.02.013. [DOI] [PubMed] [Google Scholar]
  • 55.Kegel KB, et al. Huntingtin expression stimulates endosomal-lysosomal activity, endosome tubulation, and autophagy. J Neurosci. 2000;20:7268–7278. doi: 10.1523/JNEUROSCI.20-19-07268.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Kim M, et al. Mutant huntingtin expression in clonal striatal cells: dissociation of inclusion formation and neuronal survival by caspase inhibition. J Neurosci. 1999;19:964–973. doi: 10.1523/JNEUROSCI.19-03-00964.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Schmelzle T, Hall MN. TOR, a central controller of cell growth. Cell. 2000;103:253–262. doi: 10.1016/s0092-8674(00)00117-3. [DOI] [PubMed] [Google Scholar]
  • 58.Ravikumar B, et al. Inhibition of mTOR induces autophagy and reduces toxicity of polyglutamine expansions in fly and mouse models of Huntington disease. Nat Genet. 2004;36:585–595. doi: 10.1038/ng1362. [DOI] [PubMed] [Google Scholar]
  • 59.Atwal RS, Truant R. A stress sensitive ER membrane-association domain in Huntingtin protein defines a potential role for Huntingtin in the regulation of autophagy. Autophagy. 2008;4:91–93. doi: 10.4161/auto.5201. [DOI] [PubMed] [Google Scholar]
  • 60.Zheng S, et al. Deletion of the huntingtin polyglutamine stretch enhances neuronal autophagy and longevity in mice. PLoS Genet. 2010;6:e1000838. doi: 10.1371/journal.pgen.1000838. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Martinez-Vicente M, et al. Cargo recognition failure is responsible for inefficient autophagy in Huntington’s disease. Nat Neurosci. 2010;13:567–576. doi: 10.1038/nn.2528. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Shibata M, et al. Regulation of intracellular accumulation of mutant Huntingtin by Beclin 1. J Biol Chem. 2006;281:14474–14485. doi: 10.1074/jbc.M600364200. [DOI] [PubMed] [Google Scholar]
  • 63.Metzger S, et al. Age at onset in Huntington’s disease is modified by the autophagy pathway: implication of the V471A polymorphism in Atg7. Hum Genet. 2010 Aug 10; doi: 10.1007/s00439-010-0873-9. [Epub ahead of print] [DOI] [PubMed] [Google Scholar]
  • 64.Bauer PO, et al. Harnessing chaperone-mediated autophagy for the selective degradation of mutant huntingtin protein. Nat Biotechnol. 2010;28:256–263. doi: 10.1038/nbt.1608. [DOI] [PubMed] [Google Scholar]
  • 65.Li L, et al. Altered macroautophagy in the spinal cord of SOD1 mutant mice. Autophagy. 2008;4:290–293. doi: 10.4161/auto.5524. [DOI] [PubMed] [Google Scholar]
  • 66.Hetz C, et al. XBP-1 deficiency in the nervous system protects against amyotrophic lateral sclerosis by increasing autophagy. Genes Dev. 2009;23:2294–2306. doi: 10.1101/gad.1830709. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Morimoto N, et al. Increased autophagy in transgenic mice with a G93A mutant SOD1 gene. Brain Res. 2007;1167:112–117. doi: 10.1016/j.brainres.2007.06.045. [DOI] [PubMed] [Google Scholar]
  • 68.Okamoto K, et al. Pathology of protein synthesis and degradation systems in ALS. Neuropathology. 2010;30:189–193. doi: 10.1111/j.1440-1789.2009.01088.x. [DOI] [PubMed] [Google Scholar]
  • 69.Caccamo A, et al. Rapamycin rescues TDP-43 mislocalization and the associated low molecular mass neurofilament instability. J Biol Chem. 2009;284:27416–27424. doi: 10.1074/jbc.M109.031278. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Parkinson N, et al. ALS phenotypes with mutations in CHMP2B (charged multivesicular body protein 2B) Neurology. 2006;67:1074–1077. doi: 10.1212/01.wnl.0000231510.89311.8b. [DOI] [PubMed] [Google Scholar]
  • 71.Pasquali L, et al. Autophagy, lithium, and amyotrophic lateral sclerosis. Muscle Nerve. 2009;40:173–194. doi: 10.1002/mus.21423. [DOI] [PubMed] [Google Scholar]
  • 72.Crippa V, et al. The small heat shock protein B8 (HspB8) promotes autophagic removal of misfolded proteins involved in amyotrophic lateral sclerosis (ALS) Hum Mol Genet. 2010;19:3440–3456. doi: 10.1093/hmg/ddq257. [DOI] [PubMed] [Google Scholar]
  • 73.Siciliano G, et al. Clinical trials for neuroprotection in ALS. CNS Neurol Disord Drug Targets. 2010;9:305–313. doi: 10.2174/187152710791292648. [DOI] [PubMed] [Google Scholar]
  • 74.Spencer B, et al. Beclin 1 gene transfer activates autophagy and ameliorates the neurodegenerative pathology in alpha-synuclein models of Parkinson’s and Lewy body diseases. J Neurosci. 2009;29:13578–13588. doi: 10.1523/JNEUROSCI.4390-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Crews L, et al. Selective molecular alterations in the autophagy pathway in patients with Lewy body disease and in models of alpha-synucleinopathy. PLoS One. 2010;5:e9313. doi: 10.1371/journal.pone.0009313. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Terni B, et al. Mutant ubiquitin and p62 immunoreactivity in cases of combined multiple system atrophy and Alzheimer’s disease. Acta Neuropathol. 2007;113:403–416. doi: 10.1007/s00401-006-0192-3. [DOI] [PubMed] [Google Scholar]
  • 77.Geisler S, et al. PINK1/Parkin-mediated mitophagy is dependent on VDAC1 and p62/SQSTM1. Nat Cell Biol. 2010;12:119–131. doi: 10.1038/ncb2012. [DOI] [PubMed] [Google Scholar]
  • 78.Gal J, et al. Sequestosome 1/p62 links familial ALS mutant SOD1 to LC3 via an ubiquitin-independent mechanism. J Neurochem. 2009;111:1062–1073. doi: 10.1111/j.1471-4159.2009.06388.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Munch C, et al. Point mutations of the p150 subunit of dynactin (DCTN1) gene in ALS. Neurology. 2004;63:724–726. doi: 10.1212/01.wnl.0000134608.83927.b1. [DOI] [PubMed] [Google Scholar]
  • 80.Ravikumar B, et al. Dynein mutations impair autophagic clearance of aggregate-prone proteins. Nat Genet. 2005;37:771–776. doi: 10.1038/ng1591. [DOI] [PubMed] [Google Scholar]
  • 81.Kitada T, et al. Mutations in the parkin gene cause autosomal recessive juvenile parkinsonism. Nature. 1998;392:605–608. doi: 10.1038/33416. [DOI] [PubMed] [Google Scholar]
  • 82.Valente EM, et al. Hereditary early-onset Parkinson’s disease caused by mutations in PINK1. Science. 2004;304:1158–1160. doi: 10.1126/science.1096284. [DOI] [PubMed] [Google Scholar]
  • 83.Matsuda N, et al. PINK1 stabilized by mitochondrial depolarization recruits Parkin to damaged mitochondria and activates latent Parkin for mitophagy. J Cell Biol. 2010;189:211–221. doi: 10.1083/jcb.200910140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Kawajiri S, et al. PINK1 is recruited to mitochondria with parkin and associates with LC3 in mitophagy. FEBS Lett. 2010;584:1073–1079. doi: 10.1016/j.febslet.2010.02.016. [DOI] [PubMed] [Google Scholar]
  • 85.Berger Z, et al. Rapamycin alleviates toxicity of different aggregate-prone proteins. Hum Mol Genet. 2006;15:433–442. doi: 10.1093/hmg/ddi458. [DOI] [PubMed] [Google Scholar]
  • 86.Sarkar S, et al. Small molecules enhance autophagy and reduce toxicity in Huntington’s disease models. Nat Chem Biol. 2007;3:331–338. doi: 10.1038/nchembio883. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Ravikumar B, et al. Raised intracellular glucose concentrations reduce aggregation and cell death caused by mutant huntingtin exon 1 by decreasing mTOR phosphorylation and inducing autophagy. Hum Mol Genet. 2003;12:985–994. doi: 10.1093/hmg/ddg109. [DOI] [PubMed] [Google Scholar]
  • 88.Fornai F, et al. Autophagy and amyotrophic lateral sclerosis: The multiple roles of lithium. Autophagy. 2008;4:527–530. doi: 10.4161/auto.5923. [DOI] [PubMed] [Google Scholar]
  • 89.Sarkar S, et al. Lithium induces autophagy by inhibiting inositol monophosphatase. J Cell Biol. 2005;170:1101–1111. doi: 10.1083/jcb.200504035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Sarkar S, Rubinsztein DC. Inositol and IP3 levels regulate autophagy: biology and therapeutic speculations. Autophagy. 2006;2:132–134. doi: 10.4161/auto.2387. [DOI] [PubMed] [Google Scholar]
  • 91.Sarkar S, et al. Trehalose, a novel mTOR-independent autophagy enhancer, accelerates the clearance of mutant huntingtin and alpha-synuclein. J Biol Chem. 2007;282:5641–5652. doi: 10.1074/jbc.M609532200. [DOI] [PubMed] [Google Scholar]
  • 92.Williams A, et al. Novel targets for Huntington’s disease in an mTOR-independent autophagy pathway. Nat Chem Biol. 2008;4:295–305. doi: 10.1038/nchembio.79. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Rodriguez-Navarro JA, et al. Trehalose ameliorates dopaminergic and tau pathology in parkin deleted/tau overexpressing mice through autophagy activation. Neurobiol Dis. 2010;39:423–438. doi: 10.1016/j.nbd.2010.05.014. [DOI] [PubMed] [Google Scholar]

RESOURCES