Skip to main content
Cellular and Molecular Life Sciences: CMLS logoLink to Cellular and Molecular Life Sciences: CMLS
. 2012 Apr 19;69(18):3053–3067. doi: 10.1007/s00018-012-0978-5

Regulation of Parkin E3 ubiquitin ligase activity

Helen Walden 1,, R Julio Martinez-Torres 1
PMCID: PMC11115052  PMID: 22527713

Abstract

Parkin is an E3 ubiquitin ligase mutated in autosomal recessive juvenile Parkinson’s disease. In addition, it is a putative tumour suppressor, and has roles outside its enzymatic activity. It is critical for mitochondrial clearance through mitophagy, and is an essential protein in most eukaryotes. As such, it is a tightly controlled protein, regulated through an array of external interactions with multiple proteins, posttranslational modifications including phosphorylation and S-nitrosylation, and self-regulation through internal associations. In this review, we highlight some of the recent studies into Parkin regulation and discuss future challenges for gaining a full molecular understanding of the regulation of Parkin E3 ligase activity.

Keywords: Neurodegeneration, Parkinson’s disease, Ubiquitination, Posttranslational modification, Mitophagy, Mitochondria, Cancer, Regulation

Introduction

Parkinson’s disease (PD) is a neurodegenerative disorder characterised by tremors, bradykinesia and rigidity, symptoms caused by the selective and premature loss of dopaminergic neurons in the substantia nigra. PD is far from rare, affecting between 1 and 3 % of those over 65 years of age. Until recently, PD was thought to be sporadic, caused by environmental factors. However, in the last 15 years, evidence has mounted that in 5–10 % of PD cases, there are genetic components in both hereditary forms such as autosomal recessive juvenile Parkinson’s disease (ARJPD) and sporadic PD [16]. There are several genetic components, including the dominant genes α-synuclein and LRRK2, and the recessive genes DJ-1 and PINK1. However, the most frequently mutated gene, causing ~50 % of ARJPD, is PARK2, which encodes for the protein Parkin. A multitude of genetic mutations in PARK2 give rise to ARJPD including missense, exon duplications, rearrangements, deletions and truncations (Fig. 1a). In addition to the established role in ARJPD, Parkin is also a putative tumour suppressor [6]. The parkin gene is located on the long (q) arm of chromosome 6 between positions 25.2 and 27 close to a fragile chromosome site, FRA6E, known to be a hotspot for deletions and other alterations in cancer cells, which spans a large genomic region and is frequently lost in a number of human cancers [710]. With such a central role in human health and disease, it is no surprise that Parkin is highly regulated.

Fig. 1.

Fig. 1

Schematic representation of Parkin exon and protein arrangement. a Exon arrangement and mutations that occur in the parkin gene, b the protein domain organisation of Parkin in relation to the exons, with amino acid numbering according to the human sequence to highlight the domains, c the spread of pathogenic point mutations in Parkin, with 22 indicating those mutations that fall outside of the defined domains

The parkin gene (PARK2) comprises 12 exons and codes for a 465 amino acid protein (Fig. 1b). It is widely expressed in a variety of tissues, predominantly in the muscle and brain [11, 12]. Parkin is a multidomain protein comprising an N-terminal ubiquitin-like domain (Ubl), a cysteine-rich RING0 domain [13], and 2 C-terminal really interesting new gene (RING) domains (RING1 and RING2) separated by a cysteine-rich, zinc-binding in between RING (IBR) domain (Fig. 1b). Pathogenic point mutations are dispersed throughout the amino acid sequence, with clusters in each of the domains (Fig. 1c); and for a summary of mutations see ([14] and http://www.molgen.ua.ac.be/PDmutDB). In common with the majority of RING-domain containing proteins, Parkin has E3 ubiquitin ligase activity [12, 15].

Ubiquitination is a posttranslational modification required for the regulation of many cellular pathways including cell cycle control, endocytosis, inflammation and DNA repair. Conjugation of ubiquitin to a target protein requires the coordinated action of a ubiquitin-activating enzyme (E1), ubiquitin-conjugating enzyme (E2) and a ubiquitin-ligase (E3). The E3 enzyme generally confers the substrate specificity by catalysing the conjugation of ubiquitin usually to the epsilon-amino group of a lysine in the substrate. Further ubiquitin molecules can then be attached to ubiquitin itself, by virtue of it having 7 lysine residues on the surface and the amino-terminus, all of which serve as acceptors for an activated ubiquitin (for reviews, see [16, 17]). A diversity of cellular fates awaits a ubiquitinated protein, depending on the type of modification, and all possible chains can be made [18]. A K48-linked ubiquitin chain can signal for destruction of the protein by the proteasome. K11-linked chains can also target for degradation. K63-linked chains play roles in DNA repair, protein trafficking and other non-proteasomal pathways [1921]. Linear ubiquitination has been reported to function as a degradation signal, and in NFκB signalling [17, 2224]. Other polyubiquitin modifications are less well functionally characterised, but there are also numerous degradation-independent functions of ubiquitin signalling through polyubiquitination, monoubiquitination and multiple monoubiquitination [25, 26].

Parkin E3 ligase activity is absolutely dependent on the RING2 domain, as any truncation of this domain or disruption of the fold through mutation of the zinc-coordinating residues required for correct folding result in inactive Parkin. However, an N-terminal truncation encompassing only the IBR-RING2 domains is active [2729]. Parkin was originally shown to have ubiquitin ligase activity through autoubiquitination assays [12, 15]. Since then, there has been an ever-growing list of putative Parkin substrates, but confirmation of many of these substrates and their functional consequences remain controversial. Part of the complexity is no doubt due to the different techniques used to determine Parkin substrates, ranging from yeast 2-hybrid assays to co-immunoprecipitations from different cell types and proteomic analyses of Parkin-null mouse striata [15, 3052]. In addition, there is often an assumption that a ‘genuine’ substrate of Parkin should accumulate in the brains of Parkin-linked PD patients and/or Parkin-null mice. Certainly, this is a good indication of a potential substrate; however, such an assumption ignores the data that suggest Parkin can also catalyse monoubiquitination [28, 34, 43, 48, 53] and that the outcome of polyubiquitination could be non-degradative [32, 48, 53]. In addition, Parkin-null mice do not exhibit loss of nigrostriatal dopaminergic neurons, so may not faithfully recapitulate the human scenario [5456]. However, the list of diverse substrates does implicate Parkin in a variety of cellular pathways, something which would seem to align with the myriad disease-associated mutations. Consequently, the identification of genuine Parkin substrates is of great importance to gain insights into Parkin function and disease. Several groups have identified and reported putative Parkin substrates that in principle interact with Parkin and are either mono- or polyubiquitinated in a Parkin-dependent manner. The substrates are listed in Table 1, along with a summary of the pathways involved, techniques used to identify the substrate, and accumulation in patient brains or Parkin-null mice if identified. Although some of the putative substrates have been reported to accumulate in brain samples, research in experiments performed in parkin null mice and/or ARJPD patients have been conflicting [30, 42, 55], and their roles in the pathogenesis of Parkinson’s disease remain to be confirmed [57].

Table 1.

Putative Parkin substrates, the pathways involved, techniques used to identify the substrate, and accumulation in patient brains or Parkin-null mice

(tag+)Substrate Method of identification Parkin Biological pathway Accumulation References
cMyc-SEPT5 (CdCrel-1) Yeast 2-hybrid, full-length Parkin as bait Endogenous and cMyc-Parkin Synaptic function Nd [15]
cMyc-Synphilin-1 Found in Lewy body-like inclusion FLAG-Parkin Unclear, but may play a role in neurodegeneration No [32, 137]
αSP22 Co-IP from frontal cortex of normal brain tissue Endogenous Unclear Yesa [33]
cMyc-SEPT4 (CDCrel-2)b Yeast 2-hybrid, full-length Parkin as bait cMyc-Parkin Synaptic function Yes [34]
FLAG-Pael-R Yeast 2-hybrid, full-length Parkin as bait Endogenous and GST-Parkin Unknown Yes [35, 138]
Cyclin E Pull downs using Parkin to IP FLAG-Parkin Cell cycle control Yesc [36]
cMyc-p38/JTV-1 Yeast 2-hybrid, Parkin 135–290 as bait HA-Parkin Protein biosynthesis Yesd [31, 42]
GFP-synaptotagmin XI Yeast 2-hybrid, full-length Parkin as bait HA-Parkin Synaptic function No [139]
α/β-tubulin Yeast 2-hybrid, full-length Parkin as bait FLAG-Parkin Cytoskeleton No [38]
GST-polyQ proteins Co-localisation with model proteins FLAG-Parkin Neurotoxicity No [39]
FLAG-dopamine transporter N/A Untagged-Parkin Neurotransmission No [40]
HA-SIM2 (single-minded 2) N/A FLAG-Parkin Transcription No [41]
cMyc-FBP1 Interaction with AIMP2 cMyc-Parkin Transcription Yes [140]
His-Eps15# Ubiquitination performed in vitro and in cells GST/FLAG-Parkin Endocytosis No [43]
GFP-RanBP2 Yeast 2-hybrid, full-length Parkin as bait cMyc-Parkin Nuclear import/SUMOylation No [44]
FLAG-IKKγ and FLAG-TRAF2 Co-transfection and co-immunoprecipitation Untagged-Parkin NF-κB signalling No [45]
PICK1 and cMyc-PICK1b GST pulldowns from mouse synaptosomes GST/FLAG-Parkin Synaptic function No [53]
DJ-1(L166P mutant) N/A Endogenous Neurotoxicity No [46]
3xFLAG-PDCD2-1 Yeast 2-hybrid, Parkin 1-238 used as bait Endogenous and cMyc-Parkin Apoptosis/cell proliferation Yes [47]
EGFP-Bcl-2b GST pulldowns from HEK293 cells GST/FLAG-Parkin Anti-apoptosis No [48]
VDAC1/p62-SQTM1 N/A FLAG-Parkin Autophagy/Mitophagy No [49]
FLAG-Mitofusins 1 and 2 N/A Endogenous Mitochondrial fusion No [50, 6063]
cMyc-Drp1 N/A GFP-Parkin Mitochondrial fission No [141]
FLAG-PARIS Yeast 2-hybrid, full-length Parkin as bait cMyc-or His-Parkin Transcription Yes [52]
MIRO N/A YFP-Parkin Mitochondrial function No [51]

Nd not determined

aGoldberg et al. [55] found slight increase in αSP22 in Parkin-null mice

bMono-ubiquitination identified

cNot found to accumulate in Parkin-null mice [42]

dKo et al. [42] report significant accumulation in ARJPD patient brains and Parkin-null mice, but these findings were not confirmed by Periquet et al. [30, 142]

A growing body of evidence suggests that one functional consequence of Parkin dysfunction is loss of the quality control clearance of depolarised and fragmented mitochondria through mitophagy [58, 59]. Consistent with this hypothesis, one of the few putative substrates that has been identified and corroborated in independent laboratories are the mitochondrial proteins Mitofusin (1 and 2) [50, 6063], which are required for mitochondrial fusion. Selective degradation of the mitofusins may prevent fusion of damaged mitochondria and therefore promote mitophagy [64]. In addition, two recent potential substrates have been identified that are also implicated in mitochondrial homeostasis [51, 52], adding to the hypothesis that Parkin dysfunction may result in aberrant mitochondrial regulation.

What follows is a review and discussion of the various means by which Parkin activity may be regulated, outside of interactions with putative substrates.

Regulation of Parkin activity through internal interactions

Historically, Parkin has been considered as a constitutively active protein due to its apparent autoubiquitination activity in vitro and in cell culture [12, 15]. However, a recent study suggests that wild-type native Parkin adopts an autoinhibited conformation mediated by an intramolecular interaction between the N-terminal Ubl-domain and putative peptides in the C-terminal RING regions [65]. In addition, the same C-terminal regions of Parkin interact with ubiquitin and this interaction is required for efficient ligase activity even in the active state. An independent study revealed that MBP-Parkin could catalyse E2-independent autoubiquitination in the presence of high concentrations of E1 [27]. Furthermore, removal of the N-terminal regions of Parkin, including the Ubl and RING0 domains, allowed Parkin to catalyse polyubiquitin chains as well as monoubiquitination, while the presence of the Ubl domain restricted the modification to monoubiquitination. Interestingly, a study of the mechanistic reactivity of the E2 UbcH7 revealed that HHARI (human homolog of Ariadne-1), a member of the RING-IBR-RING family of E3 ligases, forms a catalytic thioester-bound intermediate with ubiquitin in a manner analogous to the HECT-type E3 ligases [66]. Taken together, these studies suggest a mode of Parkin self-regulation whereby the N-terminal Ubl domain inhibits Parkin autoubiquitination, but once released, whether by pathogenic mutations or interactions with other proteins (see later), Parkin interacts covalently or non-covalently with ubiquitin in a semi-catalytic intermediate to catalyse an efficient transfer of ubiquitin [65, 66]. The key to the catalytic activity is the recognition of charged ubiquitin, rather than E2, explaining the observation that, in the presence of high concentrations of E1, an E2 can be dispensed with [27] (Fig. 2). Many early reports on Parkin mutations hypothesised a loss-of-catalytic-activity as the mechanism of Parkin-associated pathogenesis, but several biochemical studies reveal that loss of activity is not a unifying mechanism [28, 29, 65, 67]. Coupled with earlier studies that show a C-terminal fragment comprising IBR-RING2 is sufficient for Parkin activity [28, 29], it is clear that the N-terminal Ubl and RING0 domains are not required for Parkin activity per se, but are required for regulating its activity.

Fig. 2.

Fig. 2

Parkin regulation through internal interactions. Parkin exists in an autoinhibited state which can be relieved by removal of the Ubl domain, or release of the Ubl domain from the C-terminal regions of Parkin i. Non-covalent ii or covalent interaction with ubiquitin allows for productive transfer iii. Polyubiquitination occurs in the absence of the Ubl and RING0 domains, while substrate (synphilin-1 [27]) ubiquitination is only supported by the IBR-RING2 fragment iv

Regulation of Parkin through external interactions

In addition to the putative substrates identified for Parkin E3 ligase activity, there are multiple proteins reported to interact with Parkin. The picture is complicated by the fact that Parkin itself is prone to misfolding [29, 6870], and therefore under certain conditions may appear to interact with other poorly folded proteins, but these may not be specific interactions. However, in this section, we will summarise the current evidence for regulation of Parkin activity through interactions with multiple external proteins.

PINK1 (PTEN-induced kinase 1)

PINK1 is a highly conserved mitochondrial kinase, mutations in which are also linked to ARJPD [71]. Mutations cluster in the kinase domain, suggesting a loss-of-function of the kinase activity [72, 73]. Interestingly, in common with Parkin-null mice, PINK1-null mice do not show nigrostriatal degeneration [7476]. However, a genetic link between Parkin and PINK1 has been identified in which Parkin and PINK1-mutant cells and flies exhibit similar phenotypes. Parkin overexpression can rescue loss of PINK1, but not vice versa, placing Parkin downstream of PINK1 [7782]. It is reported that PINK1 can phosphorylate Parkin [83, 84], but there is no evidence that Parkin ubiquitinates PINK1. In 2008, the Youle group reported the translocation of Parkin to dysfunctional mitochondria [85]. Subsequent studies by several groups found that PINK1 is required to translocate Parkin to dysfunctional mitochondria [49, 67, 86] (reviewed in [59]). In HeLa cells exposed to the uncoupling agent carbonyl cyanide 3-chlorophenylhydrazone (CCCP), overexpressed Parkin co-localises to the mitochondria and induces mitophagy. Such localisation appears to be abolished in fibroblasts from PINK1-null mice [86]. Interestingly, the PINK1-dependent mitochondrial localisation of Parkin is required to activate ‘latent’ Parkin activity, supporting the notion that Parkin activity is inhibited and requires activation [67]. A note of caution, however, comes from a recent study conducted in vivo which finds that, while mCherry-Parkin localises to the mitochondria in transiently transfected HeLa cells, the same construct administered in mice fails to co-localise with mitochondria in dopaminergic neurons. Equally, removal of Parkin does not modify the pathway in their readout, leading to the suggestion that Parkin-mediated mitophagy is a rare event in vivo [87]. Further studies in animal models will be required to resolve these discrepancies; however, it appears that some interaction between PINK1 and Parkin, direct or indirect, is involved in mitochondrial homeostasis. Whether the regulation of Parkin by PINK1 is limited to localisation or extends to modifications and influence in choice of substrate remains to be determined.

Parkin as part of a multiprotein complex

PINK1 and Parkin have been reported to collaborate with yet another protein implicated in recessive Parkinsonism, DJ-1 [88]. As reported by Xiong et al. [88], differentially-tagged Parkin, DJ-1 and PINK1 co-elute at a size of ~200 kDa by size-exclusion chromatography (Fig. 3a). This complex then promotes the ubiquitination and degradation of unfolded proteins. Mitochondrial fragmentation caused by α-synuclein overexpression could be rescued by co-expression of each protein, but not certain mutants [89]. However, loss of DJ-1 gives rise to a different mitochondrial phenotype than seen with loss of PINK1 or Parkin, suggesting that this complex may not be required for mitochondrial homeostasis. Furthermore, a subsequent independent study also analysed complex formation using endogenous DJ-1 with overexpressed Parkin and PINK1, separated by size-exclusion chromatography [90], and found that DJ-1 did not co-elute with PINK1 and Parkin. A very recent study also fails to find a stable complex between PINK1 and Parkin [91] in either an endogenous or overexpression setting. Therefore, at this point, it remains unclear whether Parkin is part of a complex including DJ-1 and PINK1, but since all 3 proteins offer protection against mitochondrial stress, it seems likely that there is at least a convergence of pathways to maintain mitochondrial integrity.

Fig. 3.

Fig. 3

Regulation of Parkin activity through complex formation. a Parkin, PINK1 and DJ-1 form a complex required for mitochondrial integrity. b Parkin may be part of an SCF-like complex. In a typical complex (left), Skp1 is the adaptor for the F-box protein, which then recruits the substrate. Cullin-1 associates with the RING protein, Rbx1, which in turn recruits the E2. In the SCF-like complex, Parkin adopts the roles of both Skp1 and Rbx1 (right). c Parkin forms a complex with Hsp70 and CHIP

Parkin has also been proposed to be a component of a Skp1-Cullin-Fbox (SCF)-like complex [36] (Fig. 3b). Along with Cullin-1 and the F-box protein, hSel10, Parkin is reported to ubiquitinate and mediate the proteasomal degradation of Cyclin E. The SCF or Cullin-RING ligases typically provide a Cullin-1 scaffold with a RING protein for efficient ubiquitin transfer, Skp1, to act as an adaptor for the F-box, [92, 93], while the F-box protein recruits the substrate. In the scenario proposed by Staropoli et al. [36], Parkin would provide the RING component and act as the adaptor to recruit the F-box, as Parkin does not co-immunoprecipitate either Skp1 or the RING protein, Rbx1 (Fig. 3b).

A third potential Parkin complex is the interplay between Parkin, the chaperone Hsp70, and the U-box protein CHIP (carboxy terminus of Hsc70 interacting protein) [94]. Both Hsp70 and Parkin bind the putative Parkin substrate, unfolded PaelR. Hsp70 apparently inhibits Parkin E3 ligase activity towards PaelR, allowing it to refold. However, CHIP inhibits Hsp70 activity [95, 96] and displaces Hsp70 from PaelR, thus allowing the ubiquitination of PaelR [94] (Fig. 3c). This intriguing complex provides an elegant model for competing refolding and degradation. It is worth noting that a stable endogenous Parkin complex, obtained from brains, elutes at 110 kDa, which is too small for the complexes described above [97]. It remains to be determined which of these complexes, if any, predominate in regulating Parkin activity towards different substrates.

Negative regulators

14-3-3η is a member of the 14-3-3 chaperone like protein family. Myc-14-3-3η was found to bind FLAG-Parkin in a manner that inhibited Parkin ubiquitination activity both towards itself and towards synphilin-1 [98]. This inhibitory effect is relieved by overexpression of α-synuclein, and is specific to the η isoform of 14-3-3. Sato et al. [98] speculate that the inhibition of Parkin autoubiquitination allows for levels of Parkin to remain constant, and this is relieved in the presence of increasing concentrations of α-synuclein, allowing activation of Parkin.

BAG5 (bcl-2-associated athanogene 5) is a member of the BAG-domain-containing family that are thought to confer resistance to cell death, and has been shown to interact with the chaperone Hsp70, and Parkin [99]. This interaction has been found to enhance the degeneration of dopaminergic neurons through inhibition of both the ubiquitin ligase activity of Parkin and the chaperone activity of Hsp70, leading to the sequestration of Parkin into protein aggregates and blocking the neuroprotective function of Parkin when the proteasome system is overwhelmed [99]. Recently, BAG5 has also been described as a negative regulator of CHIP [100], which is thought to function together with Parkin (see above) [94], raising the intriguing possibility of BAG5 being a major negative regulator of the refolding-or-degradation pathway administered by the CHIP/Hsp70/Parkin complex. It is possible that 14-3-3η and BAG5 both secure wild-type Parkin in an autoinhibited conformation, or that they block additional protein binding sites for E2s or putative substrates (Fig. 4a). Further studies will be required to determine the molecular details of such interactions and the mechanism(s) of the apparent negative regulation of Parkin by external binding partners.

Fig. 4.

Fig. 4

Regulation of Parkin through external interactions. Negative regulation of Parkin through interaction with 14-3-3η and BAG5 (left). The inactivation could be achieved by stabilising an autoinhibited conformation of Parkin, or by obscuring substrate and/or E2-binding sites. Positive regulation of Parkin through interaction with SUMO-1, Eps15 and Endophilin-1A (right). The activation could be achieved through release of the autoinhibition, or through stabilising an E2- or substrate-bound form

Positive regulators

The small ubiquitin-like protein SUMO-1 is reported to associate non-covalently with Parkin [101]. Parkin is found predominantly in the cytosol [102], but is also present in the nucleus [31, 69, 103, 104]. The association with SUMO-1 enhances the nuclear translocation of Parkin and increases its autoubiquitination [101]. Interestingly, although an increase in ubiquitinated Parkin was observed, no appreciable difference in Parkin protein levels was detected upon overexpression of SUMO-1 [101], suggesting that the increased autoubiquitination activity does not necessarily result in proteasome-dependent degradation of Parkin. Parkin is implicated in gene transcription as a repressor of p53 [104], but this activity is independent of its ligase function, suggesting a functional requirement for subcellular localisation of Parkin outside substrate ubiquitination.

In our recent study describing the autoinhibition of Parkin [65], we observed that the addition of Ubl-domain-interacting proteins, Eps15 [43] and Endophilin-1A [105], released the autoinhibition allowing Parkin self-ubiquitination to occur. This raises the possibility of effectors of Parkin function and also suggests a potential for substrate-mediated activation [106]. However, it is as yet unclear whether such activation of Parkin includes a greater activity towards substrates rather than autoubiquitination, or if effector-led activation would result in proteasomal degradation of autoubiquitinated Parkin. Chew et al. [27] found that only the C-terminal IBR-RING2 of Parkin is capable of ubiquitinating synphilin-1, even though other Parkin constructs, including full-length MBP-tagged Parkin, are competent for self-ubiquitination. This suggests that there are differences between self and substrate ubiquitination. It is possible that positive regulators of Parkin may simply relieve an autoinhibited conformation, or enhance E2 or substrate binding (Fig. 4b). The possibility of allosteric regulation of Parkin by effectors is an exciting area for further exploration.

Components of the ubiquitination/deubiquitination machinery

Parkin is reported to act with multiple ubiquitin-conjugating E2 enzymes, including UbcH7, UbcH8 and the K63-linkage specific dimer, UbcH13/Uev1a. It is not yet clear what parameters influence the pairing up of Parkin with a particular E2, and deconvoluting this remains an important question as E2s often influence the choice of ubiquitin modification [107, 108].

Another important mechanism for regulating reversible ubiquitination is deubiquitination, mediated by the deubiquitinatinases (DUBs) [109]. Very recently, ataxin-3 has been identified as a binding partner of Parkin [110]. Ataxin-3 is a polyglutamine repeat protein involved in the neurodegenerative disorder, Machado-Joseph disease [111]. It contains 3 ubiquitin-interacting motifs (UIMs), a polyglutamine region which is expanded in the disease state, and an N-terminal Josephin deubiquitinase domain. Durcan et al. [110] report bimodal binding between Parkin and ataxin-3, with one interaction between the Ubl domain of Parkin and the UIMs of ataxin-3, and a second interaction between the IBR-RING2 of Parkin and the Josephin domain of ataxin-3. Polyubiquitination (but not mono-) of Parkin enhances this interaction, and ataxin-3 deubiquitinates Parkin. Interestingly, the polyglutamine expansion mutant of ataxin-3 promotes the removal of Parkin via the autophagy pathway and reduces Parkin levels in vivo [110, 112]. More recently, the same group have described an intriguing regulation of Parkin autoubiquitination that is coupled to the E2 Ubc7 [113]. Within a complex comprising ataxin-3, Parkin and Ubc7, the charged ubiquitin thioester-bound to the E2 is conjugated onto ataxin-3 rather than Parkin (Fig. 5). Furthermore, the authors find that, in contrast to their earlier report, ataxin-3 is unable to deubiquitinate chains that have already been assembled on Parkin, and is only able to deubiquitinate Parkin when Parkin is actively autoubiquitinating. These studies suggest an intricate interplay between E2, E3 and DUB enzymes that will require further studies to resolve.

Fig. 5.

Fig. 5

Regulation of Parkin by the deubiquitinase ataxin-3. Parkin autoubiquitination is inhibited when Ataxin-3 interacts with Parkin via the E2. Subsequent charged ubiquitin molecules are transferred to Ataxin-3 but not to Parkin

There are multiple other interactors of Parkin that as yet are not identified as having regulatory effects on Parkin activity, including Klokin1 and Ambra1, both implicated in mitochondrial homeostasis [114, 115]. Intriguingly, the proliferating cell nuclear antigen (PCNA) is also reported to interact with Parkin and therefore suggests a role for Parkin in promoting DNA repair [116]. These proteins may or may not turn out to have regulatory activity regarding choice of substrate, subcellular localisation, or mediating additional Parkin–protein interactions.

Regulation of Parkin through posttranslational modifications

The regulation of Parkin function via posttranslational modifications has been recently and extensively reviewed [117, 118]. Below, we summarise the current understanding of the different posttranslational modifications and their effects specifically on Parkin E3 ligase activity.

Phosphorylation

Phosphorylation of Parkin has been observed by casein kinase-1, protein kinase A, protein kinase C, cyclin-dependent kinase 5 (cdk5), c-Abl, and, as discussed above, potentially by PINK1 [83, 84, 119123]. Interestingly, with the exception of the proposed phosphorylation by PINK1, modification of Parkin by phosphorylation appears to inactivate Parkin in each of the systems studied. The interaction between Parkin and Endophilin-1A (described above in “Positive regulators”) appears to be mediated by phosphorylation, in that full-length Parkin only efficiently interacts with Endophilin-1A in conditions that favour phosphorylation [105]. Indeed, it is tempting to speculate that phosphorylation of Parkin may enhance its interactions with substrates and/or E2s, although formal evidence for such activities is still required.

S-nitrosylation

S-nitrosylation is a reversible posttranslational modification involving the covalent attachment of a nitric oxide (NO) group to the thiol of a cysteine residue to form an S-nitrosothiol species. NO is a free radical signalling molecule generated by various nitric oxide synthases, and there is increasing evidence that nitrosative stress is a key player in protein misfolding and neurodegeneration [124128]. Two studies report the S-nitrosylation of Parkin, although they differ in the location of the modified cysteine(s) [129, 130]. However, Parkin is highly cysteine-rich, and co-ordinates 8 Zinc atoms, all of which are required for correct Parkin folding [13]. Therefore, it seems plausible that S-nitrosylation of any of the zinc-coordinating cysteines would impact on Parkin function. Chung et al. [129] report a decrease in Parkin activity, while Yao et al. [130] report an increase. However, the modulation of function seen by Yao et al. is biphasic, with an initial increase in self-ubiquitination, followed 6 h later by a decrease. Thus, the two reports seem to converge on an eventual loss of function of Parkin.

Oxidation

As with S-nitrosylation, oxidative stress is also implicated in neurodegeneration. A recent study identifies the oxidation of cysteines within Parkin (spread throughout the cysteine-rich RING0, RING1, IBR and RING2 domains) via comprehensive mass spectrometry [131]. Also in common with S-nitrosylation, the authors report a biphasic initial increase in Parkin activity, followed by a loss of Parkin itself. Coupled with an earlier report that Parkin is inactivated following oxidative stress [132], it seems likely that oxidation of Parkin leads to eventual loss of Parkin function, most likely through changes in solubility, aggregation and degradation.

Ubiquitination

Perhaps the most perplexing of all the modifications of Parkin is that it is still not clear whether Parkin autoubiquitination is something that occurs in the normal situation in cells. Certainly, it is clear that, when Parkin is heavily ubiquitinated in cells, it is degraded in a proteasome-dependent manner (e.g. [65, 131]), thus effectively inactivating Parkin. However, it is also plausible that ubiquitination of Parkin is a regulatory mechanism required for efficient ubiquitination of substrates [133].

Neddylation

A recent report suggests that Parkin can be conjugated with the ubiquitin-like protein NEDD8 [134]. Through a series of truncations and lysine-to-arginine point mutations, the authors attempt to locate a specific lysine residue that is neddylated. As their results reveal that all the mutants are neddylated, they conclude that the sites for neddylation occur throughout the sequence. However, it is interesting to note that the authors do not state whether they mutated the lysine residue in the Myc tag N-terminal to each Parkin species. Certainly, in the case of ubiquitination, Parkin has been shown to ubiquitinate N-terminal tags that are fused, in vitro and in cells [65, 135], and therefore it is possible that a similar mechanism is occurring in the case of the observed neddylation. The authors also report that neddylation enhances the interaction of Parkin with UbcH8 (but not UbcH7) and with the putative substrate, the p38 subunit of aminoacyl transferase, and that the outcome of these interactions is an increased ubiquitin ligase activity [134]. Indeed, a subsequent paper also reports increased Parkin E3 ligase activity upon neddylation, coupled with a stabilisation of PINK1 [136]. It will be interesting to see if Parkin activity, in common with the cullin-RING ligases [92, 93], is regulated by neddylation.

Perspective

Studies reporting effects on Parkin function can often be conflicting. This is likely due to the wide array of different techniques, cell types, organisms, expression systems, levels of purity of recombinant proteins, N- or C-terminal tags, temporal aspects of experiments and so on. However, one apparently unifying thread is that Parkin-related Parkinson’s disease is due to a loss of Parkin function. One common mechanism for a loss of Parkin function appears to be degradation of Parkin itself by the proteasome. By this means, pathogenic Parkin mutations that are ‘activating’ ultimately result in loss of Parkin; the cases of S-nitrosylation and oxidation which appear to be both activating and inactivating are likely due to an initial increase in Parkin autoubiquitination, followed by a subsequent destruction of Parkin by the proteasome, which is therefore inactivated. However, in the case of regulatory binding partners, the opposite appears to be the case, whereby binding to SUMO-1, Eps15 or Endophilin-1A enhances Parkin activity in vitro, but not degradation in cells. A summary of Parkin functions and regulations is shown in Fig. 6.

Fig. 6.

Fig. 6

Summary of regulation of Parkin function. Parkin is activated by effectors to carry out its normal functions in ubiquitination of proteasomal substrates, and to promote mitophagy. When Parkin is compromised by pathogenic mutations, or oxidation/nitrosylation, Parkin is inactivated through proteasomal degradation

One mechanistic question that has yet to be addressed is whether, after a certain threshold of self-ubiquitination has been achieved, Parkin retains activity towards itself, substrates or both. Indeed, separating the effects of Parkin autoubiquitination and substrate ubiquitination remains a challenge, exacerbated by a lack of consensus on which criteria denote a genuine substrate. Certainly, one breakthrough would be the identification of a ‘model’ Parkin substrate that could be ubiquitinated by wild-type Parkin in vitro. Such a tool would allow for a detailed mechanistic understanding of how Parkin transfers ubiquitin to a substrate, and how that transfer is regulated. In the cellular environment, it will be important to determine what the native state of Parkin is, what conformational changes it undergoes and whether substrates are direct or indirect. Is it possible that Parkin alone directly recognises such a diverse array of substrates? Or is it likely to have indirect substrates mediated through complexes such as the SCF/CRLs? Do effectors act as substrate-binding adaptors? Or does Parkin perform a generic protein quality control role on one level, with a more specific subset of substrates on another? In addition, how Parkin interacts with different E2s and under what conditions each E2 is selected is a challenge for understanding the multiple modifications apparently catalysed by Parkin. Finally, a molecular understanding of how Parkin is recruited into various complexes and locations will be critical to our knowledge of Parkin dysfunction. Understanding how Parkin is regulated in terms of activity, protein levels, substrate recognition and subcellular localisation will be key to restoring compromised Parkin activity.

Acknowledgment

This work was supported by Cancer Research UK. H.W. is a European Molecular Biology Organisation (EMBO) Young Investigator.

References

  • 1.Gasser T. Update on the genetics of Parkinson’s disease. Mov Disord. 2007;22(Suppl 17):S343–S350. doi: 10.1002/mds.21676. [DOI] [PubMed] [Google Scholar]
  • 2.Hardy J. Genetic analysis of pathways to Parkinson disease. Neuron. 2010;68:201–206. doi: 10.1016/j.neuron.2010.10.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Bonifati V. Autosomal recessive Parkinsonism. Parkinsonism Relat Disord. 2012;18(Suppl 1):S4–S6. doi: 10.1016/S1353-8020(11)70004-9. [DOI] [PubMed] [Google Scholar]
  • 4.Dawson TM, Dawson VL. Rare genetic mutations shed light on the pathogenesis of Parkinson disease. J Clin Invest. 2003;111:145–151. doi: 10.1172/JCI17575. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Dawson TM, Dawson VL. The role of parkin in familial and sporadic Parkinson’s disease. Mov Disord. 2010;25(Suppl 1):S32–S39. doi: 10.1002/mds.22798. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Devine MJ, Plun-Favreau H, Wood NW. Parkinson’s disease and cancer: two wars, one front. Nat Rev Cancer. 2011;11:812–823. doi: 10.1038/nrc3150. [DOI] [PubMed] [Google Scholar]
  • 7.Parsons DW, Jones S, Zhang X, Lin JC, Leary RJ, et al. An integrated genomic analysis of human glioblastoma multiforme. Science. 2008;321:1807–1812. doi: 10.1126/science.1164382. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Cesari R, Martin ES, Calin GA, Pentimalli F, Bichi R, et al. Parkin, a gene implicated in autosomal recessive juvenile Parkinsonism, is a candidate tumor suppressor gene on chromosome 6q25–q27. Proc Natl Acad Sci USA. 2003;100:5956–5961. doi: 10.1073/pnas.0931262100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Veeriah S, Taylor BS, Meng S, Fang F, Yilmaz E, et al. Somatic mutations of the Parkinson’s disease-associated gene PARK2 in glioblastoma and other human malignancies. Nat Genet. 2010;42:77–82. doi: 10.1038/ng.491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Weir BA, Woo MS, Getz G, Perner S, Ding L, et al. Characterizing the cancer genome in lung adenocarcinoma. Nature. 2007;450:893–898. doi: 10.1038/nature06358. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Kitada T, Asakawa S, Hattori N, Matsumine H, Yamamura Y, 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]
  • 12.Shimura H, Hattori N, Kubo S, Mizuno Y, Asakawa S, et al. Familial Parkinson disease gene product, parkin, is a ubiquitin-protein ligase. Nat Genet. 2000;25:302–305. doi: 10.1038/77060. [DOI] [PubMed] [Google Scholar]
  • 13.Hristova VA, Beasley SA, Rylett RJ, Shaw GS. Identification of a novel Zn2+-binding domain in the autosomal recessive juvenile Parkinson-related E3 ligase parkin. J Biol Chem. 2009;284:14978–14986. doi: 10.1074/jbc.M808700200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Nuytemans K, Theuns J, Cruts M, Van Broeckhoven C. Genetic etiology of Parkinson disease associated with mutations in the SNCA, PARK2, PINK1, PARK7, and LRRK2 genes: a mutation update. Hum Mutat. 2010;31:763–780. doi: 10.1002/humu.21277. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Zhang Y, Gao J, Chung KK, Huang H, Dawson VL, et al. Parkin functions as an E2-dependent ubiquitin- protein ligase and promotes the degradation of the synaptic vesicle-associated protein, CDCrel-1. Proc Natl Acad Sci USA. 2000;97:13354–13359. doi: 10.1073/pnas.240347797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Komander D. The emerging complexity of protein ubiquitination. Biochem Soc Trans. 2009;37:937–953. doi: 10.1042/BST0370937. [DOI] [PubMed] [Google Scholar]
  • 17.Emmerich CH, Schmukle AC, Walczak H. The emerging role of linear ubiquitination in cell signaling. Sci Signal. 2011;4:re5. doi: 10.1126/scisignal.2002187. [DOI] [PubMed] [Google Scholar]
  • 18.Xu P, Duong DM, Seyfried NT, Cheng D, Xie Y, et al. Quantitative proteomics reveals the function of unconventional ubiquitin chains in proteasomal degradation. Cell. 2009;137:133–145. doi: 10.1016/j.cell.2009.01.041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Hoeller D, Hecker CM, Dikic I. Ubiquitin and ubiquitin-like proteins in cancer pathogenesis. Nat Rev Cancer. 2006;6:776–788. doi: 10.1038/nrc1994. [DOI] [PubMed] [Google Scholar]
  • 20.Chen ZJ, Sun LJ. Nonproteolytic functions of ubiquitin in cell signaling. Mol Cell. 2009;33:275–286. doi: 10.1016/j.molcel.2009.01.014. [DOI] [PubMed] [Google Scholar]
  • 21.Spence J, Sadis S, Haas AL, Finley D. A ubiquitin mutant with specific defects in DNA repair and multiubiquitination. Mol Cell Biol. 1995;15:1265–1273. doi: 10.1128/mcb.15.3.1265. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Gerlach B, Cordier SM, Schmukle AC, Emmerich CH, Rieser E, et al. Linear ubiquitination prevents inflammation and regulates immune signalling. Nature. 2011;471:591–596. doi: 10.1038/nature09816. [DOI] [PubMed] [Google Scholar]
  • 23.Tokunaga F, Nakagawa T, Nakahara M, Saeki Y, Taniguchi M, et al. SHARPIN is a component of the NF-kappaB-activating linear ubiquitin chain assembly complex. Nature. 2011;471:633–636. doi: 10.1038/nature09815. [DOI] [PubMed] [Google Scholar]
  • 24.Ikeda F, Deribe YL, Skanland SS, Stieglitz B, Grabbe C, et al. SHARPIN forms a linear ubiquitin ligase complex regulating NF-kappaB activity and apoptosis. Nature. 2011;471:637–641. doi: 10.1038/nature09814. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Ulrich HD, Walden H. Ubiquitin signalling in DNA replication and repair. Nat Rev Mol Cell Biol. 2010;11:479–489. doi: 10.1038/nrm2921. [DOI] [PubMed] [Google Scholar]
  • 26.Winget JM, Mayor T. The diversity of ubiquitin recognition: hot spots and varied specificity. Mol Cell. 2010;38:627–635. doi: 10.1016/j.molcel.2010.05.003. [DOI] [PubMed] [Google Scholar]
  • 27.Chew KC, Matsuda N, Saisho K, Lim GG, Chai C, et al. Parkin mediates apparent E2-independent monoubiquitination in vitro and contains an intrinsic activity that catalyzes polyubiquitination. PLoS ONE. 2011;6:e19720. doi: 10.1371/journal.pone.0019720. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Matsuda N, Kitami T, Suzuki T, Mizuno Y, Hattori N, et al. Diverse effects of pathogenic mutations of Parkin that catalyze multiple monoubiquitylation in vitro. J Biol Chem. 2006;281:3204–3209. doi: 10.1074/jbc.M510393200. [DOI] [PubMed] [Google Scholar]
  • 29.Hampe C, Ardila-Osorio H, Fournier M, Brice A, Corti O. Biochemical analysis of Parkinson’s disease-causing variants of Parkin, an E3 ubiquitin-protein ligase with monoubiquitylation capacity. Hum Mol Genet. 2006;15:2059–2075. doi: 10.1093/hmg/ddl131. [DOI] [PubMed] [Google Scholar]
  • 30.Periquet M, Corti O, Jacquier S, Brice A. Proteomic analysis of parkin knockout mice: alterations in energy metabolism, protein handling and synaptic function. J Neurochem. 2005;95:1259–1276. doi: 10.1111/j.1471-4159.2005.03442.x. [DOI] [PubMed] [Google Scholar]
  • 31.Corti O, Hampe C, Koutnikova H, Darios F, Jacquier S, et al. The p38 subunit of the aminoacyl-tRNA synthetase complex is a Parkin substrate: linking protein biosynthesis and neurodegeneration. Hum Mol Genet. 2003;12:1427–1437. doi: 10.1093/hmg/ddg159. [DOI] [PubMed] [Google Scholar]
  • 32.Chung KK, Zhang Y, Lim KL, Tanaka Y, Huang H, et al. Parkin ubiquitinates the alpha-synuclein-interacting protein, synphilin-1: implications for Lewy-body formation in Parkinson disease. Nat Med. 2001;7:1144–1150. doi: 10.1038/nm1001-1144. [DOI] [PubMed] [Google Scholar]
  • 33.Shimura H, Schlossmacher MG, Hattori N, Frosch MP, Trockenbacher A, et al. Ubiquitination of a new form of alpha-synuclein by parkin from human brain: implications for Parkinson’s disease. Science. 2001;293:263–269. doi: 10.1126/science.1060627. [DOI] [PubMed] [Google Scholar]
  • 34.Choi P, Snyder H, Petrucelli L, Theisler C, Chong M, et al. SEPT5_v2 is a parkin-binding protein. Brain Res Mol Brain Res. 2003;117:179–189. doi: 10.1016/S0169-328X(03)00318-8. [DOI] [PubMed] [Google Scholar]
  • 35.Imai Y, Soda M, Inoue H, Hattori N, Mizuno Y, et al. An unfolded putative transmembrane polypeptide, which can lead to endoplasmic reticulum stress, is a substrate of Parkin. Cell. 2001;105:891–902. doi: 10.1016/S0092-8674(01)00407-X. [DOI] [PubMed] [Google Scholar]
  • 36.Staropoli JF, McDermott C, Martinat C, Schulman B, Demireva E, et al. Parkin is a component of an SCF-like ubiquitin ligase complex and protects postmitotic neurons from kainate excitotoxicity. Neuron. 2003;37:735–749. doi: 10.1016/S0896-6273(03)00084-9. [DOI] [PubMed] [Google Scholar]
  • 37.Hyun DH, Lee M, Halliwell B, Jenner P. Proteasomal inhibition causes the formation of protein aggregates containing a wide range of proteins, including nitrated proteins. J Neurochem. 2003;86:363–373. doi: 10.1046/j.1471-4159.2003.01841.x. [DOI] [PubMed] [Google Scholar]
  • 38.Ren Y, Zhao J, Feng J. Parkin binds to alpha/beta tubulin and increases their ubiquitination and degradation. J Neurosci. 2003;23:3316–3324. doi: 10.1523/JNEUROSCI.23-08-03316.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Tsai YC, Fishman PS, Thakor NV, Oyler GA. Parkin facilitates the elimination of expanded polyglutamine proteins and leads to preservation of proteasome function. J Biol Chem. 2003;278:22044–22055. doi: 10.1074/jbc.M212235200. [DOI] [PubMed] [Google Scholar]
  • 40.Jiang H, Jiang Q, Feng J. Parkin increases dopamine uptake by enhancing the cell surface expression of dopamine transporter. J Biol Chem. 2004;279:54380–54386. doi: 10.1074/jbc.M409282200. [DOI] [PubMed] [Google Scholar]
  • 41.Okui M, Yamaki A, Takayanagi A, Kudoh J, Shimizu N, et al. Transcription factor single-minded 2 (SIM2) is ubiquitinated by the RING-IBR-RING-type E3 ubiquitin ligases. Exp Cell Res. 2005;309:220–228. doi: 10.1016/j.yexcr.2005.05.018. [DOI] [PubMed] [Google Scholar]
  • 42.Ko HS, von Coelln R, Sriram SR, Kim SW, Chung KK, et al. Accumulation of the authentic parkin substrate aminoacyl-tRNA synthetase cofactor, p38/JTV-1, leads to catecholaminergic cell death. J Neurosci. 2005;25:7968–7978. doi: 10.1523/JNEUROSCI.2172-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Fallon L, Belanger CM, Corera AT, Kontogiannea M, Regan-Klapisz E, et al. A regulated interaction with the UIM protein Eps15 implicates parkin in EGF receptor trafficking and PI(3)K-Akt signalling. Nat Cell Biol. 2006;8:834–842. doi: 10.1038/ncb1441. [DOI] [PubMed] [Google Scholar]
  • 44.Um JW, Min DS, Rhim H, Kim J, Paik SR, et al. Parkin ubiquitinates and promotes the degradation of RanBP2. J Biol Chem. 2006;281:3595–3603. doi: 10.1074/jbc.M504994200. [DOI] [PubMed] [Google Scholar]
  • 45.Henn IH, Bouman L, Schlehe JS, Schlierf A, Schramm JE, et al. Parkin mediates neuroprotection through activation of IkappaB kinase/nuclear factor-kappaB signaling. J Neurosci. 2007;27:1868–1878. doi: 10.1523/JNEUROSCI.5537-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Olzmann JA, Li L, Chudaev MV, Chen J, Perez FA, et al. Parkin-mediated K63-linked polyubiquitination targets misfolded DJ-1 to aggresomes via binding to HDAC6. J Cell Biol. 2007;178:1025–1038. doi: 10.1083/jcb.200611128. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Fukae J, Sato S, Shiba K, Sato K, Mori H, et al. Programmed cell death-2 isoform1 is ubiquitinated by parkin and increased in the substantia nigra of patients with autosomal recessive Parkinson’s disease. FEBS Lett. 2009;583:521–525. doi: 10.1016/j.febslet.2008.12.055. [DOI] [PubMed] [Google Scholar]
  • 48.Chen D, Gao F, Li B, Wang H, Xu Y, et al. Parkin mono-ubiquitinates BCL-2 and regulates autophagy. J Biol Chem. 2010;285:38214–38223. doi: 10.1074/jbc.M110.101469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Geisler S, Holmstrom KM, Skujat D, Fiesel FC, Rothfuss OC, 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]
  • 50.Ziviani E, Tao RN, Whitworth AJ. Drosophila parkin requires PINK1 for mitochondrial translocation and ubiquitinates mitofusin. Proc Natl Acad Sci USA. 2010;107:5018–5023. doi: 10.1073/pnas.0913485107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Wang X, Winter D, Ashrafi G, Schlehe J, Wong YL, et al. PINK1 and Parkin target Miro for phosphorylation and degradation to arrest mitochondrial motility. Cell. 2011;147:893–906. doi: 10.1016/j.cell.2011.10.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Shin JH, Ko HS, Kang H, Lee Y, Lee YI, et al. PARIS (ZNF746) repression of PGC-1alpha contributes to neurodegeneration in Parkinson’s disease. Cell. 2011;144:689–702. doi: 10.1016/j.cell.2011.02.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Joch M, Ase AR, Chen CX, MacDonald PA, Kontogiannea M, et al. Parkin-mediated monoubiquitination of the PDZ protein PICK1 regulates the activity of acid-sensing ion channels. Mol Biol Cell. 2007;18:3105–3118. doi: 10.1091/mbc.E05-11-1027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Itier JM, Ibanez P, Mena MA, Abbas N, Cohen-Salmon C, et al. Parkin gene inactivation alters behaviour and dopamine neurotransmission in the mouse. Hum Mol Genet. 2003;12:2277–2291. doi: 10.1093/hmg/ddg239. [DOI] [PubMed] [Google Scholar]
  • 55.Goldberg MS, Fleming SM, Palacino JJ, Cepeda C, Lam HA, et al. Parkin-deficient mice exhibit nigrostriatal deficits but not loss of dopaminergic neurons. J Biol Chem. 2003;278:43628–43635. doi: 10.1074/jbc.M308947200. [DOI] [PubMed] [Google Scholar]
  • 56.Von Coelln R, Thomas B, Savitt JM, Lim KL, Sasaki M, et al. Loss of locus coeruleus neurons and reduced startle in parkin null mice. Proc Natl Acad Sci USA. 2004;101:10744–10749. doi: 10.1073/pnas.0401297101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Moore DJ, Zhang L, Troncoso J, Lee MK, Hattori N, et al. Association of DJ-1 and parkin mediated by pathogenic DJ-1 mutations and oxidative stress. Hum Mol Genet. 2005;14:71–84. doi: 10.1093/hmg/ddi007. [DOI] [PubMed] [Google Scholar]
  • 58.Youle RJ, Narendra DP. Mechanisms of mitophagy. Nat Rev Mol Cell Biol. 2011;12:9–14. doi: 10.1038/nrm3028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Pilsl A, Winklhofer KF. Parkin, PINK1 and mitochondrial integrity: emerging concepts of mitochondrial dysfunction in Parkinson’s disease. Acta Neuropathol. 2012;123:173–188. doi: 10.1007/s00401-011-0902-3. [DOI] [PubMed] [Google Scholar]
  • 60.Gegg ME, Cooper JM, Chau KY, Rojo M, Schapira AH, et al. Mitofusin 1 and mitofusin 2 are ubiquitinated in a PINK1/parkin-dependent manner upon induction of mitophagy. Hum Mol Genet. 2010;19:4861–4870. doi: 10.1093/hmg/ddq419. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Rakovic A, Grunewald A, Kottwitz J, Bruggemann N, Pramstaller PP, et al. Mutations in PINK1 and Parkin impair ubiquitination of Mitofusins in human fibroblasts. PLoS ONE. 2011;6:e16746. doi: 10.1371/journal.pone.0016746. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Poole AC, Thomas RE, Yu S, Vincow ES, Pallanck L. The mitochondrial fusion-promoting factor mitofusin is a substrate of the PINK1/parkin pathway. PLoS ONE. 2010;5:e10054. doi: 10.1371/journal.pone.0010054. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Glauser L, Sonnay S, Stafa K, Moore DJ. Parkin promotes the ubiquitination and degradation of the mitochondrial fusion factor mitofusin 1. J Neurochem. 2011;118:636–645. doi: 10.1111/j.1471-4159.2011.07318.x. [DOI] [PubMed] [Google Scholar]
  • 64.Tanaka A, Cleland MM, Xu S, Narendra DP, Suen DF, et al. Proteasome and p97 mediate mitophagy and degradation of mitofusins induced by Parkin. J Cell Biol. 2010;191:1367–1380. doi: 10.1083/jcb.201007013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Chaugule VK, Burchell L, Barber KR, Sidhu A, Leslie SJ, et al. Autoregulation of Parkin activity through its ubiquitin-like domain. EMBO J. 2011;30:2853–2867. doi: 10.1038/emboj.2011.204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Wenzel DM, Lissounov A, Brzovic PS, Klevit RE. UBCH7 reactivity profile reveals parkin and HHARI to be RING/HECT hybrids. Nature. 2011;474:105–108. doi: 10.1038/nature09966. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Matsuda N, Sato S, Shiba K, Okatsu K, Saisho K, 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]
  • 68.Henn IH, Gostner JM, Lackner P, Tatzelt J, Winklhofer KF. Pathogenic mutations inactivate parkin by distinct mechanisms. J Neurochem. 2005;92:114–122. doi: 10.1111/j.1471-4159.2004.02854.x. [DOI] [PubMed] [Google Scholar]
  • 69.Cookson MR, Lockhart PJ, McLendon C, O’Farrell C, Schlossmacher M, et al. RING finger 1 mutations in Parkin produce altered localization of the protein. Hum Mol Genet. 2003;12:2957–2965. doi: 10.1093/hmg/ddg328. [DOI] [PubMed] [Google Scholar]
  • 70.Wang C, Ko HS, Thomas B, Tsang F, Chew KC, et al. Stress-induced alterations in parkin solubility promote parkin aggregation and compromise parkin’s protective function. Hum Mol Genet. 2005;14:3885–3897. doi: 10.1093/hmg/ddi413. [DOI] [PubMed] [Google Scholar]
  • 71.Valente EM, Abou-Sleiman PM, Caputo V, Muqit MM, Harvey K, 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]
  • 72.Deas E, Wood NW, Plun-Favreau H. Mitophagy and Parkinson’s disease: The PINK1-parkin link. Biochim Biophys Acta. 2011;1813:623–633. doi: 10.1016/j.bbamcr.2010.08.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Deas E, Plun-Favreau H, Wood NW. PINK1 function in health and disease. EMBO Mol Med. 2009;1:152–165. doi: 10.1002/emmm.200900024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Zhou H, Falkenburger BH, Schulz JB, Tieu K, Xu Z, et al. Silencing of the Pink1 gene expression by conditional RNAi does not induce dopaminergic neuron death in mice. Int J Biol Sci. 2007;3:242–250. doi: 10.7150/ijbs.3.242. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Kitada T, Pisani A, Porter DR, Yamaguchi H, Tscherter A, et al. Impaired dopamine release and synaptic plasticity in the striatum of PINK1-deficient mice. Proc Natl Acad Sci USA. 2007;104:11441–11446. doi: 10.1073/pnas.0702717104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Gispert S, Ricciardi F, Kurz A, Azizov M, Hoepken HH, et al. Parkinson phenotype in aged PINK1-deficient mice is accompanied by progressive mitochondrial dysfunction in absence of neurodegeneration. PLoS ONE. 2009;4:e5777. doi: 10.1371/journal.pone.0005777. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Clark IE, Dodson MW, Jiang C, Cao JH, Huh JR, et al. Drosophila pink1 is required for mitochondrial function and interacts genetically with parkin. Nature. 2006;441:1162–1166. doi: 10.1038/nature04779. [DOI] [PubMed] [Google Scholar]
  • 78.Greene JC, Whitworth AJ, Kuo I, Andrews LA, Feany MB, et al. Mitochondrial pathology and apoptotic muscle degeneration in Drosophila parkin mutants. Proc Natl Acad Sci USA. 2003;100:4078–4083. doi: 10.1073/pnas.0737556100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Park J, Lee SB, Lee S, Kim Y, Song S, et al. Mitochondrial dysfunction in Drosophila PINK1 mutants is complemented by parkin. Nature. 2006;441:1157–1161. doi: 10.1038/nature04788. [DOI] [PubMed] [Google Scholar]
  • 80.Yang Y, Gehrke S, Imai Y, Huang Z, Ouyang Y, et al. Mitochondrial pathology and muscle and dopaminergic neuron degeneration caused by inactivation of Drosophila Pink1 is rescued by Parkin. Proc Natl Acad Sci USA. 2006;103:10793–10798. doi: 10.1073/pnas.0602493103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Lutz AK, Exner N, Fett ME, Schlehe JS, Kloos K, et al. Loss of parkin or PINK1 function increases Drp1-dependent mitochondrial fragmentation. J Biol Chem. 2009;284:22938–22951. doi: 10.1074/jbc.M109.035774. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Exner N, Treske B, Paquet D, Holmstrom K, Schiesling C, et al. Loss-of-function of human PINK1 results in mitochondrial pathology and can be rescued by parkin. J Neurosci. 2007;27:12413–12418. doi: 10.1523/JNEUROSCI.0719-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Kim Y, Park J, Kim S, Song S, Kwon SK, et al. PINK1 controls mitochondrial localization of Parkin through direct phosphorylation. Biochem Biophys Res Commun. 2008;377:975–980. doi: 10.1016/j.bbrc.2008.10.104. [DOI] [PubMed] [Google Scholar]
  • 84.Sha D, Chin LS, Li L. 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]
  • 85.Narendra D, Tanaka A, Suen DF, Youle RJ. 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]
  • 86.Narendra DP, Jin SM, Tanaka A, Suen DF, Gautier CA, et al. PINK1 is selectively stabilized on impaired mitochondria to activate Parkin. PLoS Biol. 2010;8:e1000298. doi: 10.1371/journal.pbio.1000298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Sterky FH, Lee S, Wibom R, Olson L, Larsson NG. Impaired mitochondrial transport and Parkin-independent degeneration of respiratory chain-deficient dopamine neurons in vivo. Proc Natl Acad Sci USA. 2011;108:12937–12942. doi: 10.1073/pnas.1103295108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Xiong H, Wang D, Chen L, Choo YS, Ma H, et al. Parkin, PINK1, and DJ-1 form a ubiquitin E3 ligase complex promoting unfolded protein degradation. J Clin Invest. 2009;119:650–660. doi: 10.1172/JCI37617. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Kamp F, Exner N, Lutz AK, Wender N, Hegermann J, et al. Inhibition of mitochondrial fusion by alpha-synuclein is rescued by PINK1, Parkin and DJ-1. EMBO J. 2010;29:3571–3589. doi: 10.1038/emboj.2010.223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Thomas KJ, McCoy MK, Blackinton J, Beilina A, van der Brug M, et al. DJ-1 acts in parallel to the PINK1/parkin pathway to control mitochondrial function and autophagy. Hum Mol Genet. 2011;20:40–50. doi: 10.1093/hmg/ddq430. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Lazarou M, Jin SM, Kane LA, Youle R (2012) Role of PINK1 binding to the TOM complex and alternate intracellular membranes in recruitment and activation of the E3 ligase Parkin. Dev Cell. doi:10.1016/j.bbr.2011.03.031 [DOI] [PMC free article] [PubMed]
  • 92.Petroski MD, Deshaies RJ. Function and regulation of cullin-RING ubiquitin ligases. Nat Rev Mol Cell Biol. 2005;6:9–20. doi: 10.1038/nrm1547. [DOI] [PubMed] [Google Scholar]
  • 93.Duda DM, Scott DC, Calabrese MF, Zimmerman ES, Zheng N, et al. Structural regulation of cullin-RING ubiquitin ligase complexes. Curr Opin Struct Biol. 2011;21:257–264. doi: 10.1016/j.sbi.2011.01.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Imai Y, Soda M, Hatakeyama S, Akagi T, Hashikawa T, et al. CHIP is associated with Parkin, a gene responsible for familial Parkinson’s disease, and enhances its ubiquitin ligase activity. Mol Cell. 2002;10:55–67. doi: 10.1016/S1097-2765(02)00583-X. [DOI] [PubMed] [Google Scholar]
  • 95.Ballinger CA, Connell P, Wu Y, Hu Z, Thompson LJ, et al. Identification of CHIP, a novel tetratricopeptide repeat-containing protein that interacts with heat shock proteins and negatively regulates chaperone functions. Mol Cell Biol. 1999;19:4535–4545. doi: 10.1128/mcb.19.6.4535. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Jiang J, Ballinger CA, Wu Y, Dai Q, Cyr DM, et al. CHIP is a U-box-dependent E3 ubiquitin ligase: identification of Hsc70 as a target for ubiquitylation. J Biol Chem. 2001;276:42938–42944. doi: 10.1074/jbc.M101968200. [DOI] [PubMed] [Google Scholar]
  • 97.Van Humbeeck C, Waelkens E, Corti O, Brice A, Vandenberghe W. Parkin occurs in a stable, non-covalent, approximately 110 kDa complex in brain. Eur J Neurosci. 2008;27:284–293. doi: 10.1111/j.1460-9568.2007.06000.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Sato S, Chiba T, Sakata E, Kato K, Mizuno Y, et al. 14-3-3eta is a novel regulator of parkin ubiquitin ligase. EMBO J. 2006;25:211–221. doi: 10.1038/sj.emboj.7600774. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Kalia SK, Lee S, Smith PD, Liu L, Crocker SJ, et al. BAG5 inhibits parkin and enhances dopaminergic neuron degeneration. Neuron. 2004;44:931–945. doi: 10.1016/j.neuron.2004.11.026. [DOI] [PubMed] [Google Scholar]
  • 100.Kalia LV, Kalia SK, Chau H, Lozano AM, Hyman BT, et al. Ubiquitinylation of alpha-synuclein by carboxyl terminus Hsp70-interacting protein (CHIP) is regulated by Bcl-2-associated athanogene 5 (BAG5) PLoS ONE. 2011;6:e14695. doi: 10.1371/journal.pone.0014695. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Um JW, Chung KC. Functional modulation of parkin through physical interaction with SUMO-1. J Neurosci Res. 2006;84:1543–1554. doi: 10.1002/jnr.21041. [DOI] [PubMed] [Google Scholar]
  • 102.Shimura H, Hattori N, Kubo S, Yoshikawa M, Kitada T, et al. Immunohistochemical and subcellular localization of Parkin protein: absence of protein in autosomal recessive juvenile Parkinsonism patients. Ann Neurol. 1999;45:668–672. doi: 10.1002/1531-8249(199905)45:5<668::AID-ANA19>3.0.CO;2-Z. [DOI] [PubMed] [Google Scholar]
  • 103.Stichel CC, Augustin M, Kuhn K, Zhu XR, Engels P, et al. Parkin expression in the adult mouse brain. Eur J Neurosci. 2000;12:4181–4194. [PubMed] [Google Scholar]
  • 104.da Costa CA, Sunyach C, Giaime E, West A, Corti O, et al. Transcriptional repression of p53 by parkin and impairment by mutations associated with autosomal recessive juvenile Parkinson’s disease. Nat Cell Biol. 2009;11:1370–1375. doi: 10.1038/ncb1981. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Trempe JF, Chen CX, Grenier K, Camacho EM, Kozlov G, et al. SH3 domains from a subset of BAR proteins define a Ubl-binding domain and implicate parkin in synaptic ubiquitination. Mol Cell. 2009;36:1034–1047. doi: 10.1016/j.molcel.2009.11.021. [DOI] [PubMed] [Google Scholar]
  • 106.Liu F, Walters KJ. Policing Parkin with a UblD. EMBO J. 2011;30:2757–2758. doi: 10.1038/emboj.2011.223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Wickliffe KE, Lorenz S, Wemmer DE, Kuriyan J, Rape M. The mechanism of linkage-specific ubiquitin chain elongation by a single-subunit E2. Cell. 2011;144:769–781. doi: 10.1016/j.cell.2011.01.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Ye Y, Rape M. Building ubiquitin chains: E2 enzymes at work. Nat Rev Mol Cell Biol. 2009;10:755–764. doi: 10.1038/nrm2780. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Komander D, Clague MJ, Urbe S. Breaking the chains: structure and function of the deubiquitinases. Nat Rev Mol Cell Biol. 2009;10:550–563. doi: 10.1038/nrm2731. [DOI] [PubMed] [Google Scholar]
  • 110.Durcan TM, Kontogiannea M, Thorarinsdottir T, Fallon L, Williams AJ, et al. The Machado-Joseph disease-associated mutant form of ataxin-3 regulates parkin ubiquitination and stability. Hum Mol Genet. 2011;20:141–154. doi: 10.1093/hmg/ddq452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Costa MD, Paulson HL (2011) Toward understanding Machado-Joseph disease. Prog Neurobiol (in press) [DOI] [PMC free article] [PubMed]
  • 112.Durcan TM, Fon EA. Mutant ataxin-3 promotes the autophagic degradation of parkin. Autophagy. 2011;7:233–234. doi: 10.4161/auto.7.2.14224. [DOI] [PubMed] [Google Scholar]
  • 113.Durcan TM, Kontogiannea M, Bedard N, Wing SS, Fon EA. Ataxin-3 deubiquitination is coupled to Parkin ubiquitination via E2 ubiquitin-conjugating enzyme. J Biol Chem. 2012;287:531–541. doi: 10.1074/jbc.M111.288449. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Kuroda Y, Sako W, Goto S, Sawada T, Uchida D, et al. Parkin interacts with Klokin1 for mitochondrial import and maintenance of membrane potential. Hum Mol Genet. 2012;21:991–1003. doi: 10.1093/hmg/ddr530. [DOI] [PubMed] [Google Scholar]
  • 115.Van Humbeeck C, Cornelissen T, Hofkens H, Mandemakers W, Gevaert K, et al. Parkin interacts with Ambra1 to induce mitophagy. J Neurosci. 2011;31:10249–10261. doi: 10.1523/JNEUROSCI.1917-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Kao SY. DNA damage induces nuclear translocation of parkin. J Biomed Sci. 2009;16:67. doi: 10.1186/1423-0127-16-67. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Rankin CA, Roy A, Zhang Y, Richter M. Parkin, a top level manager in the cell’s sanitation department. Open Biochem J. 2011;5:9–26. doi: 10.2174/1874091X01105010009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Rubio de la Torre E, Gomez-Suaga P, Martinez-Salvador M, Hilfiker S. Posttranslational modifications as versatile regulators of parkin function. Curr Med Chem. 2011;18:2477–2485. doi: 10.2174/092986711795843254. [DOI] [PubMed] [Google Scholar]
  • 119.Yamamoto A, Friedlein A, Imai Y, Takahashi R, Kahle PJ, et al. Parkin phosphorylation and modulation of its E3 ubiquitin ligase activity. J Biol Chem. 2005;280:3390–3399. doi: 10.1074/jbc.M407724200. [DOI] [PubMed] [Google Scholar]
  • 120.Avraham E, Rott R, Liani E, Szargel R, Engelender S. Phosphorylation of Parkin by the cyclin-dependent kinase 5 at the linker region modulates its ubiquitin-ligase activity and aggregation. J Biol Chem. 2007;282:12842–12850. doi: 10.1074/jbc.M608243200. [DOI] [PubMed] [Google Scholar]
  • 121.Rubio de la Torre E, Luzon-Toro B, Forte-Lago I, Minguez-Castellanos A, Ferrer I, et al. Combined kinase inhibition modulates parkin inactivation. Hum Mol Genet. 2009;18:809–823. doi: 10.1093/hmg/ddn407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Ko HS, Lee Y, Shin JH, Karuppagounder SS, Gadad BS, et al. Phosphorylation by the c-Abl protein tyrosine kinase inhibits parkin’s ubiquitination and protective function. Proc Natl Acad Sci USA. 2010;107:16691–16696. doi: 10.1073/pnas.1006083107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Imam SZ, Zhou Q, Yamamoto A, Valente AJ, Ali SF, et al. Novel regulation of parkin function through c-Abl-mediated tyrosine phosphorylation: implications for Parkinson’s disease. J Neurosci. 2011;31:157–163. doi: 10.1523/JNEUROSCI.1833-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Gu Z, Nakamura T, Lipton SA. Redox reactions induced by nitrosative stress mediate protein misfolding and mitochondrial dysfunction in neurodegenerative diseases. Mol Neurobiol. 2010;41:55–72. doi: 10.1007/s12035-010-8113-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Tsang AH, Lee YI, Ko HS, Savitt JM, Pletnikova O, et al. S-nitrosylation of XIAP compromises neuronal survival in Parkinson’s disease. Proc Natl Acad Sci USA. 2009;106:4900–4905. doi: 10.1073/pnas.0810595106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Nakamura T, Wang L, Wong CC, Scott FL, Eckelman BP, et al. Transnitrosylation of XIAP regulates caspase-dependent neuronal cell death. Mol Cell. 2010;39:184–195. doi: 10.1016/j.molcel.2010.07.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Gu Z, Kaul M, Yan B, Kridel SJ, Cui J, et al. S-nitrosylation of matrix metalloproteinases: signaling pathway to neuronal cell death. Science. 2002;297:1186–1190. doi: 10.1126/science.1073634. [DOI] [PubMed] [Google Scholar]
  • 128.Cho DH, Nakamura T, Fang J, Cieplak P, Godzik A, 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]
  • 129.Chung KK, Thomas B, Li X, Pletnikova O, Troncoso JC, et al. S-Nitrosylation of parkin regulates ubiquitination and compromises parkin’s protective function. Science. 2004;304:1328–1331. doi: 10.1126/science.1093891. [DOI] [PubMed] [Google Scholar]
  • 130.Yao D, Gu Z, Nakamura T, Shi ZQ, Ma Y, et al. Nitrosative stress linked to sporadic Parkinson’s disease: S-nitrosylation of parkin regulates its E3 ubiquitin ligase activity. Proc Natl Acad Sci USA. 2004;101:10810–10814. doi: 10.1073/pnas.0404161101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Meng F, Yao D, Shi Y, Kabakoff J, Wu W, et al. Oxidation of the cysteine-rich regions of parkin perturbs its E3 ligase activity and contributes to protein aggregation. Mol Neurodegener. 2011;6:34. doi: 10.1186/1750-1326-6-34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Winklhofer KF, Henn IH, Kay-Jackson PC, Heller U, Tatzelt J. Inactivation of parkin by oxidative stress and C-terminal truncations: a protective role of molecular chaperones. J Biol Chem. 2003;278:47199–47208. doi: 10.1074/jbc.M306769200. [DOI] [PubMed] [Google Scholar]
  • 133.Nagy V, Dikic I. Ubiquitin ligase complexes: from substrate selectivity to conjugational specificity. Biol Chem. 2010;391:163–169. doi: 10.1515/bc.2010.021. [DOI] [PubMed] [Google Scholar]
  • 134.Um JW, Han KA, Im E, Oh Y, Lee K, et al. (2012) Neddylation positively regulates the ubiquitin E3 ligase activity of parkin. J Neurosci Res (in press) [DOI] [PubMed]
  • 135.Burchell L, Chaugule VK, Walden H (2012) Small, N-terminal tags activate Parkin E3 ubiquitin ligase activity by disrupting its autoinhibited conformation. PLoS ONE (in press) [DOI] [PMC free article] [PubMed]
  • 136.Choo YS, Vogler G, Wang D, Kalvakuri S, Iliuk A, et al. Regulation of parkin and PINK1 by neddylation. Hum Mol Genet. 2012 doi: 10.1093/hmg/dds070. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Lim KL, Chew KC, Tan JM, Wang C, Chung KK, et al. Parkin mediates nonclassical, proteasomal-independent ubiquitination of synphilin-1: implications for Lewy body formation. J Neurosci. 2005;25:2002–2009. doi: 10.1523/JNEUROSCI.4474-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Yang Y, Nishimura I, Imai Y, Takahashi R, Lu B. Parkin suppresses dopaminergic neuron-selective meurotoxicity induced by Pael-R in Drosophila . Neuron. 2003;37:911–924. doi: 10.1016/S0896-6273(03)00143-0. [DOI] [PubMed] [Google Scholar]
  • 139.Huynh DP, Scoles DR, Nguyen D, Pulst SM. The autosomal recessive juvenile Parkinson disease gene product, parkin, interacts with and ubiquitinates synaptotagmin XI. Hum Mol Genet. 2003;12:2587–2597. doi: 10.1093/hmg/ddg269. [DOI] [PubMed] [Google Scholar]
  • 140.Ko HS, Kim SW, Sriram SR, Dawson VL, Dawson TM. Identification of far upstream element-binding protein-1 as an authentic Parkin substrate. J Biol Chem. 2006;281:16193–16196. doi: 10.1074/jbc.C600041200. [DOI] [PubMed] [Google Scholar]
  • 141.Wang H, Song P, Du L, Tian W, Yue W, et al. Parkin ubiquitinates Drp1 for proteasome-dependent degradation: implication of dysregulated mitochondrial dynamics in Parkinson disease. J Biol Chem. 2011;286:11649–11658. doi: 10.1074/jbc.M110.144238. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Corti O, Brice A. Of Parkin and Parkinson’s: light and dark sides of a multifaceted E3 ubiquitin-protein ligase. Drug Discov Today Dis Mech. 2007;4:121–127. doi: 10.1016/j.ddmec.2007.11.002. [DOI] [Google Scholar]

Articles from Cellular and Molecular Life Sciences: CMLS are provided here courtesy of Springer

RESOURCES