首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Aberrant posttranslational modifications (PTMs) of proteins, namely phosphorylation, induce abnormalities in the biological properties of recipient proteins, underlying neurological diseases including Parkinson''s disease (PD). Genome-wide studies link genes encoding α-synuclein (α-Syn) and Tau as two of the most important in the genesis of PD. Although several kinases are known to phosphorylate α-Syn and Tau, we focused our analysis on GSK-3β because of its accepted role in phosphorylating Tau and to increasing evidence supporting a strong biophysical relationship between α-Syn and Tau in PD. Therefore, we investigated transgenic mice, which express a point mutant (S9A) of human GSK-3β. GSK-3β-S9A is capable of activation through endogenous natural signaling events, yet is unable to become inactivated through phosphorylation at serine-9. We used behavioral, biochemical, and in vitro analysis to assess the contributions of GSK-3β to both α-Syn and Tau phosphorylation. Behavioral studies revealed progressive age-dependent impairment of motor function, accompanied by loss of tyrosine hydroxylase-positive (TH+ DA-neurons) neurons and dopamine production in the oldest age group. Magnetic resonance imaging revealed deterioration of the substantia nigra in aged mice, a characteristic feature of PD patients. At the molecular level, kinase-active p-GSK-3β-Y216 was seen at all ages throughout the brain, yet elevated levels of p-α-Syn-S129 and p-Tau (S396/404) were found to increase with age exclusively in TH+ DA-neurons of the midbrain. p-GSK-3β-Y216 colocalized with p-Tau and p-α-Syn-S129. In vitro kinase assays showed that recombinant human GSK-3β directly phosphorylated α-Syn at a single site, Ser129, in addition to its known ability to phosphorylate Tau. Moreover, α-Syn and Tau together cooperated with one another to increase the magnitude or rate of phosphorylation of the other by GSK-3β. Together, these data establish a novel upstream role for GSK-3β as one of several kinases associated with PTMs of key proteins known to be causal in PD.After Alzheimer''s disease (AD), Parkinson''s disease (PD) is the second most prevalent neurodegenerative disease, characterized by selective loss of TH+ DA-neurons of substantia nigra (SN) with diminished production of dopamine (DA).1 Genome-wide studies have identified SNCA and MAPT, genes encoding α-synuclein (α-Syn) and Tau, respectively, as having strong association to the genesis of PD.2, 3, 4 Although the precise etiology of PD remains a mystery, SNCA amplifications and mutations directly link α-Syn dysfunction to disease causation,5, 6 firmly establishing a role for α-Syn in sporadic and familial PD, respectively. α-Syn can be phosphorylated at several sites,7 and the predominance of α-Syn phosphorylated at serine 129 (S129) in Lewy bodies8 suggests its phosphorylation status at S129 has an important pathological role. Various PD models have shown that phosphorylation at S219 enhanced α-syn toxicity resulting in accelerated motor abnormalities and loss of DA-neurons.9, 10Fewer studies have examined the role of Tau (or p-Tau) in PD, but interest in the field has grown since completion of several genome-wide association studies. p-Tau has been found to colocalize with α-Syn in tissue from sporadic PD and dementia with Lewy bodies.11 We12, 13 and others14,15 have also identified p-Tau in different brain regions of PD, dementia with Lewy bodies, and AD. High levels of p-Tau have also been observed in vivo in several toxin16, 17, 18 and transgenic α-Syn models of PD,19,20 suggesting that p-Tau may be an important common factor in the neurodegeneration of not only tauopathies but also of synucleinopathies, such as PD.21, 22, 23, 24 Most studies to date have focused on the formation and accumulation of Tau and p-Tau in idiopathic PD. Yet several studies have provided evidence that leucine-rich repeat kinase-2 (LRRK2), a kinase, that when mutated is involved in familial forms of PD, can directly interact with, and activate GSK-3β, resulting in increased p-TAU formation.25,26Among the kinases known to hyperphosphorylate Tau, glycogen synthase kinase-3β (GSK-3β) may be the most important given its ability to phosphorylate Tau at the majority of its serine/threonine sites that cause associated toxicities in AD.27,28 The importance of GSK-3β is illustrated in that it is embryonically lethal when knocked out in mice. Regulation of GSK-3β is tightly controlled through a series of direct and indirect measures. Direct regulation occurs through autophosphorylation at Tyr216,29,30 resulting in a kinase-active form, p-GSK-3β-Y216, whereas phosphorylation at Ser9 results in a kinase-inactive state.31 The activity of GSK-3β can also be controlled indirectly through binding to inhibitory complexes with other cytoplasmic proteins,32,33 or through Wnt-mediated sequestration into multivesicular bodies34 resulting in the physical separation of GSK-3β from its cytoplasmic targets. Control of GSK-3β in the normal state is therefore tightly regulated, with its dysregulation and ensuing aberrant phosphorylation of targets being a common occurrence in many diverse diseases. Several studies have shown that GSK-3β is an important mediator in the injury and repair processes of neurons during cross-talk between DA-neurons and reactive astrocytes.35,36 These studies showed that astrocyte-derived Wnt1 was capable of blocking GSK-3β activation, allowing the nuclear accumulation of β-catenin and subsequent gene expression of β-catenin-dependent targets essential for neuron survival and repair during chemical or metabolic insults. The importance of regulating the active/inactive states of GSK-3β in regard to neuronal stability is further supported through the analysis of conditional (Tet-inducible) transgenic mice expressing a dominant-negative GSK-3β-K85R mutant or expressing the GSK-3β-S9A mutant.37,38 In these studies, post-natal Tet-regulated expression of either GSK-3β-K85R or GSK-3β-S9A led to neurodegeneration in the cortex, striatum, and hippocampus. What separates our TG PD model from the tet-inducible GSK-3β models is the spatial patterns of transgene expression, which is influenced by the choice of promoters. The Tet-inducible GSK-3β models are expressed using a CAMKII promoter with our human(h) GSK-3β-S9A transgene being expressed under the Thy-1 promoter. CAMKII-driven expression is limited to neurons originating from the forebrain with Thy-1 promoter-driven expression restricted to neurons in all or most brain regions.39,40 Although promoter choice effecting tissue expression ultimately decides which regions show degeneration, the important message is that both inactive and hyperactive states of GSK-3β reduce neuronal viability.In our past studies in various in vitro and in vivo models of PD and in postmortem PD tissues, we have consistently observed a positive correlation between increased α-Syn and p-Tau levels with increased GSK-3β-Y216 (the kinase-active form of GSK-3β).12, 13, 16, 19, 20 In in vitro studies of MPTP-treated SH-SY5Y cells, blockade of GSK-3β with lithium, or with the highly selective non-ATP competitive inhibitor, TDZD-8, prevented the induction of p-GSK-3β-Y216, abolished p-Tau formation, and reversed the accumulation and aggregation of both p-Tau and α-Syn, averting cell death.16 Other studies using Rotenone or MPTP/MPP+ in chemical PD models, have shown similar results of decreased neuronal viability during treatments accompanied by dose- and time-dependent increases in GSK-3β activation, with decreased cytotoxicity detected when GSK-3β was inhibited or knocked-down through the use of GSK-3β-specific small molecule inhibitors or through RNAi.41,42 This suggested to us that p-GSK-3β-Y216 may have a contributory role in the pathogenesis of PD. Using a mouse model overexpressing hGSK-3β-S9A under the Thy-1 promoter together with in vitro kinase assays allowed us to discern the role GSK-3β has in the development of PD-like pathology.43 Analysis of our hGSK-3β-S9A mouse model showed here for the first time that upon aging, these mice develop the cardinal features of parkinsonism, manifested as impaired motor behavior, with associated loss of TH+ neurons, reduced DA production, and shrinkage of SN. Invitro kinase assays confirmed that hGSK-3β was capable of phosphorylating α-Syn on Serine 129 together with the known ability to phosphorylate Tau. Remarkably, both α-Syn and Tau influenced the rate and magnitude of phosphorylation of the other by GSK-3β indicating that an intimate physical relationship exist between the trio of PD related proteins. Together, these data shown indicate the importance of GSK-3β activation, in the behavioral and physiological development of PD like pathology in a new mouse model.  相似文献   

2.
3.
INFLIXIMAB IS A CHIMERIC ANTI-TUMOUR NECROSIS FACTOR-α antibody that is efficacious in treating Crohn''s disease. However, its immunomodulatory properties increase susceptibility to opportunistic infections. We present a case of cutaneous Nocardia infection in a patient who was taking infliximab for Crohn''s disease. The case illustrates the challenges in the diagnosis and management of this disease and serves as a reminder of the complications associated with the use of immunomodulatory agents.A 45-year-old HIV-negative man with fistulous Crohn''s disease had a history of inadequate disease control despite ongoing prednisone therapy. He had previously taken budesonide, mesalamine, ciprofloxacin and metronidazole in attempts to induce remission of his inflammatory bowel disease. The patient was born in Canada and, other than a 1-week holiday to Mexico 10 years before presentation, had travelled only locally. He denied a family history of tuberculosis, and he had never worked in a health care facility. Infliximab was introduced, and the patient received 3 infusions of 5 mg/kg at baseline and 2 and 6 weeks later. After he received his third infusion, prednisone was tapered to 40 mg at a rate of 5 mg weekly. One month after the third infusion, in February 2000, he reported multiple erythematous papulopustular lesions on his right leg (Fig. 1). There was no associated lymphadenopathy, cough, shortness of breath, fever or constitutional symptoms. He denied a history of insect bites, but in November 1999 he had received cuts to his right leg from a metal blade when at work. He had not immersed the leg in a whirlpool or swimming pool around the time of the leg injury. The patient continued to receive further infliximab infusions (at weeks 12 and 18 after baseline), and the lesions were treated with cloxacillin for suspected Staphylococcus aureus infection. Since improvement was minimal, a skin biopsy was performed in July 2000. Granulomatous inflammation was present, and acid-fast bacilli were visualized (Fig. 2). Cultures sent for mycobacteriology and mycology were incubated at 35°C for 8 weeks, but the results from the mycobacteriology culture proved negative. A polymerase chain reaction assay for Mycobacterium tuberculosus was also negative. A clinical diagnosis of M. marinum infection was made, and the patient''s antibiotic regimen was changed to minocycline. A tuberculin skin test was not performed; since the patient was immunocompromised, a negative result would not have excluded the disease. A chest radiograph appeared normal.Open in a separate windowFig. 2: Acid-fast bacilli visualized in skin biopsy.Open in a separate windowFig. 1: Multiple erythematous papulopustular lesions on the patient''s leg 1 month after the third infusion of infliximab.The patient failed to respond to the minocycline therapy, and he was referred for infectious disease consultation in October 2000. The infliximab infusions were discontinued, and 2 more skin biopsies were performed, with acid-fast bacilli visualized in both specimens. The patient was given trimethoprim–sulfamethoxazole, and his lesions began to heal slowly but progressively.Acid-fast bacilli were recovered from the second set of biopsies, and specific instructions were given to incubate the cultures at 30°C and 35°C to ensure that M. marinum, if present, would be detected. Again, the cultures failed to recover organisms. The laboratory, using polymerase chain reaction restriction analysis of the 439-base pair segment of the gene encoding a 65-kDa heat shock protein,1 identified the presence of Nocardia species. Nevertheless, Nocardia organisms still could not be recovered in culture, and therefore final speciation could not be performed.The patient resumed taking prednisone, and the dosage was increased in order to ameliorate the symptoms of his Crohn''s disease. The trimethoprim–sulfamethoxazole therapy was continued until late 2003, and the dosage was reduced over the subsequent months. Complete healing of the lesions was eventually achieved 4 years after therapy was initiated.  相似文献   

4.
Begomovirus-DNA-β disease complexes induce different symptom phenotypes in their hosts. To investigate the genetic determinants of the phenotypic differences, Nicotiana spp. and tomato plants were inoculated with infectious clones of Tobacco curly shoot virus (TbCSV)/TbCSV DNA-β (TbCSB) and Tomato yellow leaf curl China virus (TYLCCNV)/TYLCCNV DNA-β (TYLCCNB) pseudorecombinants and showed that TYLCCNB induced characteristic vein-thickening and enation symptoms, while TbCSB only slightly exacerbated the leaf-curling symptoms, regardless of the helper virus being used. The roles of DNA-β-encoded βC1 and a 430-nucleotide fragment containing the A-rich region and the putative βC1 promoter region of the βC1 gene (referred to as AP) in symptom development were further investigated by constructing hybrid satellites in which the βC1 coding region or AP was exchanged between the two satellite molecules. A TYLCCNB hybrid with TbCSB βC1 lost the ability to elicit the vein-thickening and enation phenotypes. TbCSB hybrids containing the TYLCCNB βC1 or AP fragment failed to induce the characteristic vein thickening and enations. A TYLCCNB hybrid having the TbCSB AP fragment produced the enations, but the number of enations was less and their sizes were reduced. Differently from the phloem-specific pattern of the TYLCCNB promoter, a full-length fragment upstream of the TbCSB βC1 gene confers a constitutive β-glucuronidase expression pattern in transgenic tobacco plants. The above results indicate that the DNA-β-encoded βC1 protein is the symptom determinant, but the promoter of the βC1 gene has influence on symptom production.Geminiviruses are small plant viruses with circular single-stranded DNA (ssDNA) genomes that are encapsidated in unique twinned (geminate) particles. Members of the genus Begomovirus are transmitted by whiteflies (Bemisia tabaci) and infect dicotyledonous plants (42). Begomoviruses have either one or two circular ssDNA genomic components (DNA-A and DNA-B). The DNA-A component is capable of autonomous replication and encapsidation, whereas the DNA-B component encodes two proteins (BC1 and BV1) involved in movement (14). Recently, some monopartite begomoviruses have been found in association with a novel satellite DNA molecule, referred to as DNA-β and now known as a betasatellite (2, 5, 20, 22, 38, 45). DNA-β is approximately half the size of the viral genomic DNA, and apart from a nonanucleotide sequence (TAATATTAC), it has little sequence identity with viral genomic DNA. DNA-β depends on the helper virus for replication and encapsidation and, in turn, is required for the induction of bona fide disease symptoms. DNA-β bears a βC1 open reading frame (ORF) on the complementary-sense strand, which is conserved among distinct betasatellites in terms of position and size. Mutational analyses and constitutive expression have shown that βC1 is a strong pathogenicity/symptom determinant (7, 34, 39).Begomovirus-DNA-β disease complexes are associated with a wide range of plant species and induce different sets of symptom phenotypes in their natural hosts (25). However, the contributions of the helper virus and the satellite molecule to symptom development are not clear. Tomato yellow leaf curl China virus (TYLCCNV) and Tobacco curly shoot virus (TbCSV) are monopartite begomoviruses associated with DNA-β, but they differ in the symptom phenotypes induced in Nicotiana spp. and Solanum lycopersicum (7, 22). In the present work, we report that the symptom differences between TYLCCNV/TYLCCNV DNA-β (TYLCCNB) and TbCSV/TbCSV DNA-β (TbCSB) are determined by DNA-β and the DNA-β-encoded βC1 protein is the symptom determinant, but the promoter of the βC1 gene has influence on symptom production.  相似文献   

5.
Heterotrimeric G proteins, consisting of Gα, Gβ, and Gγ subunits, are a conserved signal transduction mechanism in eukaryotes. However, G protein subunit numbers in diploid plant genomes are greatly reduced as compared with animals and do not correlate with the diversity of functions and phenotypes in which heterotrimeric G proteins have been implicated. In addition to GPA1, the sole canonical Arabidopsis (Arabidopsis thaliana) Gα subunit, Arabidopsis has three related proteins: the extra-large GTP-binding proteins XLG1, XLG2, and XLG3. We demonstrate that the XLGs can bind Gβγ dimers (AGB1 plus a Gγ subunit: AGG1, AGG2, or AGG3) with differing specificity in yeast (Saccharomyces cerevisiae) three-hybrid assays. Our in silico structural analysis shows that XLG3 aligns closely to the crystal structure of GPA1, and XLG3 also competes with GPA1 for Gβγ binding in yeast. We observed interaction of the XLGs with all three Gβγ dimers at the plasma membrane in planta by bimolecular fluorescence complementation. Bioinformatic and localization studies identified and confirmed nuclear localization signals in XLG2 and XLG3 and a nuclear export signal in XLG3, which may facilitate intracellular shuttling. We found that tunicamycin, salt, and glucose hypersensitivity and increased stomatal density are agb1-specific phenotypes that are not observed in gpa1 mutants but are recapitulated in xlg mutants. Thus, XLG-Gβγ heterotrimers provide additional signaling modalities for tuning plant G protein responses and increase the repertoire of G protein heterotrimer combinations from three to 12. The potential for signal partitioning and competition between the XLGs and GPA1 is a new paradigm for plant-specific cell signaling.The classical heterotrimeric G protein consists of a GDP/GTP-binding Gα subunit with GTPase activity bound to an obligate dimer formed by Gβ and Gγ subunits. In the signaling paradigm largely elucidated from mammalian systems, the plasma membrane-associated heterotrimer contains Gα in its GDP-bound form. Upon receiving a molecular signal, typically transduced by a transmembrane protein (e.g. a G protein-coupled receptor), Gα exchanges GDP for GTP and dissociates from the Gβγ dimer. Both Gα and Gβγ interact with intracellular effectors to initiate downstream signaling cascades. The intrinsic GTPase activity of Gα restores Gα to the GDP-bound form, which binds Gβγ, thereby reconstituting the heterotrimer (McCudden et al., 2005; Oldham and Hamm, 2008).Signal transduction through a heterotrimeric G protein complex is an evolutionarily conserved eukaryotic mechanism common to metazoa and plants, although there are distinct differences in the functional intricacies between the evolutionary branches (Jones et al., 2011a, 2011b; Bradford et al., 2013). The numbers of each subunit encoded within genomes, and therefore the potential for combinatorial complexity within the heterotrimer, is one of the most striking differences between plants and animals. For example, the human genome encodes 23 Gα (encoded by 16 genes), five Gβ, and 12 Gγ subunits (Hurowitz et al., 2000; McCudden et al., 2005; Birnbaumer, 2007). The Arabidopsis (Arabidopsis thaliana) genome, however, only encodes one canonical Gα (GPA1; Ma et al., 1990), one Gβ (AGB1; Weiss et al., 1994), and three Gγ (AGG1, AGG2, and AGG3) subunits (Mason and Botella, 2000, 2001; Chakravorty et al., 2011), while the rice (Oryza sativa) genome encodes one Gα (Ishikawa et al., 1995), one Gβ (Ishikawa et al., 1996), and either four or five Gγ subunits (Kato et al., 2004; Chakravorty et al., 2011; Botella, 2012). As expected, genomes of polyploid plants have more copies due to genome duplication, with the soybean (Glycine max) genome encoding four Gα, four Gβ (Bisht et al., 2011), and 10 Gγ subunits (Choudhury et al., 2011). However, Arabidopsis heterotrimeric G proteins have been implicated in a surprisingly large number of phenotypes, which is seemingly contradictory given the relative scarcity of subunits. Arabidopsis G proteins have been implicated in cell division (Ullah et al., 2001; Chen et al., 2006) and morphological development in various tissues, including hypocotyls (Ullah et al., 2001, 2003), roots (Ullah et al., 2003; Chen et al., 2006; Li et al., 2012), leaves (Lease et al., 2001; Ullah et al., 2001), inflorescences (Ullah et al., 2003), and flowers and siliques (Lease et al., 2001), as well as in pathogen responses (Llorente et al., 2005; Trusov et al., 2006; Cheng et al., 2015), regulation of stomatal movement (Wang et al., 2001; Coursol et al., 2003; Fan et al., 2008) and development (Zhang et al., 2008; Nilson and Assmann, 2010), cell wall composition (Delgado-Cerezo et al., 2012), responses to various light stimuli (Warpeha et al., 2007; Botto et al., 2009), responses to multiple abiotic stimuli (Huang et al., 2006; Pandey et al., 2006; Trusov et al., 2007; Zhang et al., 2008; Colaneri et al., 2014), responses to various hormones during germination (Ullah et al., 2002), and postgermination development (Ullah et al., 2002; Pandey et al., 2006; Trusov et al., 2007). Since the Gγ subunit appeared to be the only subunit that provides diversity in heterotrimer composition in Arabidopsis, it was proposed that all functional specificity in heterotrimeric G protein signaling was provided by the Gγ subunit (Trusov et al., 2007; Chakravorty et al., 2011; Thung et al., 2012, 2013). This allowed for only three heterotrimer combinations to account for the wide range of G protein-associated phenotypes.In addition to the above typical G protein subunits, the plant kingdom contains a conserved protein family of extra-large GTP-binding proteins (XLGs). XLGs differ from typical Gα subunits in that they possess a long N-terminal extension of unknown function, but they are similar in that they all have a typical C-terminal Gα-like region, with five semiconserved G-box (G1–G5) motifs. The XLGs also possess the two sequence features that differentiate heterotrimeric G protein Gα subunits from monomeric G proteins: a helical region between the G1 and G2 motifs and an Asp/Glu-rich loop between the G3 and G4 motifs (Lee and Assmann, 1999; Ding et al., 2008; Heo et al., 2012). The Arabidopsis XLG family comprises XLG1, XLG2, and XLG3, and all three have demonstrated GTP-binding and GTPase activities, although they differ from GPA1 in exhibiting a much slower rate of GTP hydrolysis, with a Ca2+ cofactor requirement instead of an Mg2+ requirement, as for canonical Gα proteins (Heo et al., 2012). All three Arabidopsis XLGs were observed to be nuclear localized (Ding et al., 2008). Although much less is known about XLGs than canonical Gα subunits, XLG2 positively regulates resistance to the bacterial pathogen Pseudomonas syringae and was immunoprecipitated with AGB1 from tissue infected with P. syringae (Zhu et al., 2009). xlg3 mutants, like agb1 mutants, are impaired in root-waving and root-skewing responses (Pandey et al., 2008). During the preparation of this report, Maruta et al. (2015) further investigated XLG2, particularly focusing on the link between XLG2 and Gβγ in pathogen responses. Based on symptom progression in xlg mutants, they found that XLG2 is a positive regulator of resistance to both bacterial and fungal pathogens, with a minor contribution from XLG3 in resistance to Fusarium oxysporum. XLG2 and XLG3 are also positive regulators of reactive oxygen species (ROS) production in response to pathogen-associated molecular pattern elicitors. The resistance and pathogen-associated molecular pattern-induced ROS phenotypes of the agg1 agg2 and xlg2 xlg3 double mutants were not additive in an agg1 agg2 xlg2 xlg3 quadruple mutant, indicating that these two XLGs and the two Gγ subunits function in the same, rather than parallel, pathways. Unfortunately, the close proximity of XLG2 and AGB1 on chromosome 4 precluded the generation of an agb1 xlg2 double mutant; therefore, direct genetic evidence of XLG2 and AGB1 interaction is still lacking, but physical interactions between XLG2 and the Gβγ dimers were shown by yeast (Saccharomyces cerevisiae) three-hybrid and bimolecular fluorescence complementation (BiFC) assays (Maruta et al., 2015). Localization of all three XLGs was also reexamined, indicating that XLGs are capable of localizing to the plasma membrane in addition to the nucleus (Maruta et al., 2015).Interestingly, several other plant G protein-related phenotypes, in addition to pathogen resistance, have been observed only in Gβ and Gγ mutants, with opposite phenotypes observed in Gα (gpa1) mutants. Traditionally, the observation of opposite phenotypes in Gα versus Gβγ mutants in plants and other organisms has mechanistically been attributed to signaling mediated by free Gβγ, which increases in abundance in the absence of Gα. However, an intriguing alternative is that XLG proteins fulfill a Gα-like role in forming heterotrimeric complexes with Gβγ and function in non-GPA1-based G protein signaling processes. If XLGs function like Gα subunits, the corresponding increase in subunit diversity could potentially account for the diversity of G protein phenotypes. In light of this possibility, we assessed the heterotrimerization potential of all possible XLG and Gβγ dimer combinations, XLG localization and its regulation by Gβγ, and the effect of xlg mutation on selected known phenotypes associated with heterotrimeric G proteins. Our results provide compelling evidence for the formation of XLG-Gβγ heterotrimers and reveal that plant G protein signaling is substantially more complex than previously thought.  相似文献   

6.
Many neuropathological and experimental studies suggest that the degeneration of dopaminergic terminals and axons precedes the demise of dopaminergic neurons in the substantia nigra, which finally results in the clinical symptoms of Parkinson disease (PD). The mechanisms underlying this early axonal degeneration are, however, still poorly understood. Here, we examined the effects of overexpression of human wildtype alpha-synuclein (αSyn-WT), a protein associated with PD, and its mutant variants αSyn-A30P and -A53T on neurite morphology and functional parameters in rat primary midbrain neurons (PMN). Moreover, axonal degeneration after overexpression of αSyn-WT and -A30P was analyzed by live imaging in the rat optic nerve in vivo. We found that overexpression of αSyn-WT and of its mutants A30P and A53T impaired neurite outgrowth of PMN and affected neurite branching assessed by Sholl analysis in a variant-dependent manner. Surprisingly, the number of primary neurites per neuron was increased in neurons transfected with αSyn. Axonal vesicle transport was examined by live imaging of PMN co-transfected with EGFP-labeled synaptophysin. Overexpression of all αSyn variants significantly decreased the number of motile vesicles and decelerated vesicle transport compared with control. Macroautophagic flux in PMN was enhanced by αSyn-WT and -A53T but not by αSyn-A30P. Correspondingly, colocalization of αSyn and the autophagy marker LC3 was reduced for αSyn-A30P compared with the other αSyn variants. The number of mitochondria colocalizing with LC3 as a marker for mitophagy did not differ among the groups. In the rat optic nerve, both αSyn-WT and -A30P accelerated kinetics of acute axonal degeneration following crush lesion as analyzed by in vivo live imaging. We conclude that αSyn overexpression impairs neurite outgrowth and augments axonal degeneration, whereas axonal vesicle transport and autophagy are severely altered.Growing evidence suggests that Parkinson''s disease (PD) pathology starts at the presynaptic terminals and the distal axons and is then propagated back to the soma in a ''dying back'' pattern.1, 2 Accordingly, at the time of clinical onset, there is only a 30% loss of total substantia nigra pars compacta neurons but a far more severe loss of striatal dopaminergic markers (70–80%), suggesting that axonal terminals of the nigrostriatal pathway are affected earlier.1 It is thus essential to understand the pathomechanisms specifically affecting the axon in PD in order to interfere with early disease progression.Neurodegeneration in PD is accompanied by the appearance of intraneuronal protein aggregates, denoted Lewy bodies (LBs).3 Interestingly, also LB pathology is initially found in the distal axons before becoming evident in the neuronal somata, and dystrophic neurites, so called ''Lewy neurites'', outnumber LBs in the early stages of PD.2, 4, 5 A main component of LBs is the protein alpha-synuclein (αSyn) that is not only widely used as a histopathological marker for PD but is also believed to have a major role in PD pathogenesis.6, 7 The importance of αSyn is further underlined by the discovery of αSyn point mutations (e.g. Ala53Thr (A53T), Ala30Pro (A30P)) and multiplications of the αSyn gene, all of which cause autosomal dominant forms of PD.8, 9, 10 However, neither the physiological functions nor the pathogenetic mechanisms of αSyn are well understood.7The biological effects of αSyn expression strongly depend on the model system. Wild-type (WT) human αSyn does not lead to major clinical or histological abnormalities when expressed in transgenic mice,11, 12 but its overexpression mediated by adeno-associated viral vectors (AAV) results in severe neurodegeneration, suggesting a dose-dependent toxic effect.13, 14 Different human αSyn-A30P and -A53T transgenic mouse lines develop severe motor impairments, partly resembling symptoms of human PD, accompanied by a degeneration of the nigrostriatal neuronal system and LB-like pathology.11, 12, 15 In line with the pathological findings in human PD, the axonal compartment is affected early and most prominently in these animal models.Different putative pathomechanisms of αSyn toxicity have been explored. For example, the cytoskeleton is an important molecular target of αSyn. Multimeric forms of αSyn were shown to impair the polymerization of tubulin and microtubule formation.16, 17 Overexpression of αSyn increased actin instability and induced actin bundling in cultured hippocampal neurons.18 There are, however, divergent data on the resulting effects of αSyn overexpression on neurite outgrowth and integrity in different model systems.19, 20, 21, 22Moreover, a dysregulation of autophagy has been implicated in PD pathology. Aberrant αSyn is normally degraded by autophagy and only to a negligible degree by the proteasome.23 Several studies have shown that the inhibition of autophagy results in an accumulation and increased toxicity of αSyn, whereas the activation of autophagy has therapeutic effects in PD models.23, 24, 25, 26 However, the direct effects of αSyn and its mutants on autophagy seem to rely strongly on the model system and the published data are highly controversial.24, 26, 27, 28, 29, 30, 31, 32Given the central role of axonal degeneration in PD, it is likely that disturbances of axonal transport are involved.33 In support of this proposition, the motor protein kinesin was shown to be decreased early and stage-dependently in PD patients, preceding the loss of substantia nigra neurons.34 αSyn itself is actively transported along the axons, mainly by the slow component of axonal transport, but the role of αSyn in axonal vesicle transport is unclear.35Here, we present a comprehensive analysis of the effects of αSyn on neurite morphology and examine important pathomechanisms.  相似文献   

7.
8.
Structural Bases for Function in Cytochromes c. An Interpretation of Comparative X-ray and Biochemical Data (Salemme, F. R., Kraut, J., and Kamen, M. D. (1973) J. Biol. Chem. 248, 7701–7716) Martin David Kamen (1913–2002) was born in Toronto, Canada, but grew up in Chicago. He enrolled at the University of Chicago in 1930, intending to study English. However, the Great Depression took a toll on his family''s finances, and his father suggested he switch his major to chemistry in order to make a living after graduating. By his junior year, Kamen was hooked on chemistry. He graduated in 1933 with honors in physical chemistry, working with William Draper Harkins to determine ammonia gas emission spectra excited by electrodeless discharge. He remained in Harkins'' lab for graduate school, earning his doctorate in physical chemistry in 1936 for an article on neutron scattering (1) that was accepted as his dissertation.Open in a separate windowMartin D. KamenBecause economic conditions were still bleak, Kamen followed the suggestion of one of his mentors, David Gans, and applied for a research post with Nobel laureate Ernest O. Lawrence, who had developed the cyclotron at the radiation laboratory in Berkeley, California. Kamen used his savings to move to Berkeley and worked at the laboratory without pay for 6 months before Lawrence offered him a staff position as a chemist. In addition to troubleshooting the cyclotrons and preparing samples of radioisotopes, Kamen performed numerous photosynthetic studies with Samuel Ruben, using carbon-11. Because carbon-11 had a half-life of only 21 min, Lawrence assigned Kamen and Ruben the task of finding carbon-14. The pair succeeded by bombarding graphite in the cyclotron, producing carbon-14, which had a 5730-year half-life (2).Kamen and Ruben planned to use their discovery to create a company that would construct and operate several cyclotrons dedicated to carbon-14 production and expand on the laboratory''s radioisotope program. However, the war intervened, and all non-war-related research at Berkeley was halted. Kamen was assigned to head a program studying the separation of uranium isotopes for the Manhattan Project. But, unexpectedly in 1944, he was declared a security risk and dismissed from the lab. A few years later, Kamen was called before the House Un-American Activities Committee, being wrongly linked to an espionage ring working for the USSR. Subsequently, the State Department refused to issue Kamen a passport, and the Chicago Tribune named him as a suspected spy. During the next decade, he fought recurring rumors and accusations that he had leaked atomic bomb secrets. Eventually, he won a libel suit against the Chicago Tribune, and the State Department reinstated his passport.In 1945, Kamen moved to the Mallinckrodt Institute of Radiology at the Washington University School of Medicine where he supervised cyclotron production of radioisotopes for medical research. His own research interests gradually shifted away from nuclear physics and radiochemistry to biochemistry, and he began several collaborations involving the use of radioisotopic tracers in biological and biomedical research.Kamen also initiated a series of experiments using carbon-14 to study photosynthesis in bacteria. This resulted in a number of important discoveries, including hydrogen photoevolution (3) and nitrogen fixation (4) in Rhodospirillum rubrum. While working with the bacteria, Kamen and Leo Vernon discovered that R. rubrum contained a c-type cytochrome (5), which they later named “cytochrome c2.”Twenty years after it was discovered, the structure of cytochrome c2 was solved (6). By comparing this structure with the recently solved structure of eukaryotic mitochondrial cytochrome c (7), Kamen and his colleagues were able to deduce information about the structural, functional, and evolutionary relationships in the cytochromes c. This is the subject of the Journal of Biological Chemistry (JBC) Classic reprinted here.Despite the fact that both eukaryotic cytochrome c and cytochrome c2 serve analogous functions in their respective physiological electron transport chains, i.e. they both transport electrons to the terminal and most oxidizing electron carrier of each system, Kamen was able to find several differences between the molecules. For example, he noted that cytochrome c2 has a more positive electrochemical potential and does not exhibit the large oxidation state-dependent conformational change characteristic of mitochondrial cytochrome c. Open in a separate windowKamen continued to study other bacterial cytochromes, showing that at least 12 subgroups of the cytochromes c exist. This resulted in new perspectives on potential variations in structure and function of the heme group in relation to protein.In 1957 Kamen moved to Brandeis University to help establish the graduate department of biochemistry, and in 1961 he joined the University of California, San Diego chemistry department to help found their new campus. He remained there until 1975, when he became director of the Chemical-Biological Development Laboratory at the University of Southern California. Kamen continued to teach into his eighties, being one of six faculty members of the Oregon Institute of Science and Medicine.Kamen received numerous awards and honors for his contributions to science, including the American Chemical Society''s Award for Applications of Nuclear Chemistry (1963), the American Society of Plant Physiologists'' Charles F. Kettering Research Award (1968), the American Society of Biological Chemists'' Merck Award (1982), the John Scott Medal of the City of Philadelphia (1988), the World Cultural Council''s Einstein Award (1990), and the U.S. Department of Energy''s Enrico Fermi Award (1996). He was a member of the National Academy of Sciences, the American Academy of Arts and Sciences, and the American Philosophical Society. 1  相似文献   

9.
In animals, heterotrimeric G proteins, comprising Gα, Gβ, and Gγ subunits, are molecular switches whose function tightly depends on Gα and Gβγ interaction. Intriguingly, in Arabidopsis (Arabidopsis thaliana), multiple defense responses involve Gβγ, but not Gα. We report here that the Gβγ dimer directly partners with extra-large G proteins (XLGs) to mediate plant immunity. Arabidopsis mutants deficient in XLGs, Gβ, and Gγ are similarly compromised in several pathogen defense responses, including disease development and production of reactive oxygen species. Genetic analysis of double, triple, and quadruple mutants confirmed that XLGs and Gβγ functionally interact in the same defense signaling pathways. In addition, mutations in XLG2 suppressed the seedling lethal and cell death phenotypes of BRASSINOSTEROID INSENSITIVE1-associated receptor kinase1-interacting receptor-like kinase1 mutants in an identical way as reported for Arabidopsis Gβ-deficient mutants. Yeast (Saccharomyces cerevisiae) three-hybrid and bimolecular fluorescent complementation assays revealed that XLG2 physically interacts with all three possible Gβγ dimers at the plasma membrane. Phylogenetic analysis indicated a close relationship between XLGs and plant Gα subunits, placing the divergence point at the dawn of land plant evolution. Based on these findings, we conclude that XLGs form functional complexes with Gβγ dimers, although the mechanism of action of these complexes, including activation/deactivation, must be radically different form the one used by the canonical Gα subunit and are not likely to share the same receptors. Accordingly, XLGs expand the repertoire of heterotrimeric G proteins in plants and reveal a higher level of diversity in heterotrimeric G protein signaling.Heterotrimeric GTP-binding proteins (G proteins), classically consisting of Gα, Gβ, and Gγ subunits, are essential signal transduction elements in most eukaryotes. In animals and fungi, ligand perception by G protein-coupled receptors leads to replacement of GDP with GTP in Gα, triggering activation of the heterotrimer (Li et al., 2007; Oldham and Hamm, 2008). Upon activation, GTP-bound Gα and Gβγ are released and interact with downstream effectors, thereby transmitting signals to multiple intracellular signaling cascades. Signaling terminates when the intrinsic GTPase activity of Gα hydrolyzes GTP to GDP and the inactive heterotrimer reforms at the receptor. The large diversity of mammalian Gα subunits confers specificity to the multiple signaling pathways mediated by G proteins (Wettschureck and Offermanns, 2005). Five distinct classes of Gα have been described in animals (Gαi, Gαq, Gαs, Gα12 and Gαv), with orthologs found in evolutionarily primitive organisms such as sponges (Oka et al., 2009). Humans possess four classes of Gα involving 23 functional isoforms encoded by 16 genes (McCudden et al., 2005), while only a single prototypical Gα is usually found per plant genome (Urano et al., 2013). Multiple copies of Gα are present in some species with recently duplicated genomes, such as soybean (Glycine max) with four Gα genes (Blanc and Wolfe, 2004; Bisht et al., 2011). In the model plant Arabidopsis (Arabidopsis thaliana), a prototypical Gα subunit (GPA1) is involved in a number of important processes, including cell proliferation (Ullah et al., 2001), inhibition of inward K+ channels and activation of anion channels in guard cells by mediating the abscisic acid pathway (Wang et al., 2001; Coursol et al., 2003), blue light responses (Warpeha et al., 2006, 2007), and germination and postgermination development (Chen et al., 2006; Pandey et al., 2006).It is well established that heterotrimeric G proteins play a fundamental role in plant innate immunity. In Arabidopsis, two different Gβγ dimers (Gβγ1 and Gβγ2) are generally considered to be the predominant elements in G protein defense signaling against a variety of fungal pathogens (Llorente et al., 2005; Trusov et al., 2006, 2007, 2009; Delgado-Cerezo et al., 2012; Torres et al., 2013). By contrast, these studies attributed a small or no role to Gα, because mutants deficient in Gα displayed only slightly increased resistance against the fungal pathogens (Llorente et al., 2005; Trusov et al., 2006; Torres et al., 2013). The Gβγ-mediated signaling also contributes to defense against a model bacterial pathogen Pseudomonas syringae, by participating in programmed cell death (PCD) and inducing reactive oxygen species (ROS) production in response to at least three pathogen-associated molecular patterns (PAMPs; Ishikawa, 2009; Liu et al., 2013; Torres et al., 2013). Gα is not involved in PCD or PAMP-triggered ROS production (Liu et al., 2013; Torres et al., 2013). Nonetheless, Arabidopsis Gα plays a positive role in defense against P. syringae, probably by mediating stomatal function and hence physically restricting bacterial entry to the leaf interior (Zhang et al., 2008; Zeng and He, 2010; Lee et al., 2013). Given the small contribution from Gα, the involvement of heterotrimeric G proteins in Arabidopsis resistance could be explained in two ways: either the Gβγ dimer acts independently from Gα, raising a question of how is it activated upon a pathogen attack, or Gα is replaced by another protein for heterotrimer formation.The Arabidopsis genome contains at least three genes encoding Gα-like proteins that have been classified as extra-large G proteins (XLGs; Lee and Assmann, 1999; Ding et al., 2008). XLGs comprise two structurally distinct regions. The C-terminal region is similar to the canonical Gα, containing the conserved helical and GTPase domains, while the N-terminal region is a stretch of approximately 400 amino acids including a putative nuclear localization signal (Ding et al., 2008). GTP binding and hydrolysis were confirmed for all three XLG proteins, although their enzymatic activities are very slow and require Ca2+ as a cofactor, whereas canonical Gα utilizes Mg2+ (Heo et al., 2012). Several other features differentiate XLGs from Gα subunits. Comparative analysis of XLG1 and Gα at the DNA level showed that the genes are organized in seven and 13 exons, respectively, without common splicing sites (Lee and Assmann, 1999). XLGs have been reported to localize to the nucleus (Ding et al., 2008). Analysis of knockout mutants revealed a nuclear function for XLG2, as it physically interacts with the Related To Vernalization1 (RTV1) protein, enhancing the DNA binding activity of RTV1 to floral integrator gene promoters and resulting in flowering initiation (Heo et al., 2012). Therefore, it appears that XLGs may act independently of G protein signaling. On the other hand, functional similarities between XLGs and the Arabidopsis Gβ subunit (AGB1) were also discovered. For instance, XLG3- and Gβ-deficient mutants were similarly impaired in root gravitropic responses (Pandey et al., 2008). Knockout of all three XLG genes caused increased root length, similarly to the Gβ-deficient mutant (Ding et al., 2008). Furthermore, as observed in Gβ-deficient mutants, xlg2 mutants displayed increased susceptibility to P. syringae, indicating a role in plant defense (Zhu et al., 2009). Nevertheless, a genetic analysis of the possible functional interaction between XLGs and Gβ has not been established.In this report, we performed in-depth genetic analyses to test the functional interaction between the three XLGs and Gβγ dimers during defense-related responses in Arabidopsis. We also examined physical interaction between XLG2 and the Gβγ dimers using yeast (Saccharomyces cerevisiae) three-hybrid (Y3H) and bimolecular fluorescent complementation (BiFC) assays. Our findings indicate that XLGs function as direct partners of Gβγ dimers in plant defense signaling. To estimate relatedness of XLGs and Gα proteins, we carried out a phylogenetic analysis. Based on our findings, we conclude that plant XLG proteins most probably originated from a canonical Gα subunit and retained prototypical interaction with Gβγ dimers. They function together with Gβγ in a number of processes including plant defense, although they most probably evolved activation/deactivation mechanisms very different from those of a prototypical Gα.  相似文献   

10.
A novel clathrin adaptor-like complex, adaptor protein (AP)-3, has recently been described in yeast and in animals. To gain insight into the role of yeast AP-3, a genetic strategy was devised to isolate gene products that are required in the absence of the AP-3 μ chain encoded by APM3. One gene identified by this synthetic lethal screen was VPS45. The Vps pathway defines the route that several proteins, including carboxypeptidase Y, take from the late Golgi to the vacuole. However, vacuolar alkaline phosphatase (ALP) is transported via an alternate, intracellular route. This suggested that the apm3-Δ vps45 synthetic phenotype could be caused by a block in both the alternate and the Vps pathways. Here we demonstrate that loss of function of the AP-3 complex results in slowed processing and missorting of ALP. ALP is no longer localized to the vacuole membrane by immunofluorescence, but is found in small punctate structures throughout the cell. This pattern is distinct from the Golgi marker Kex2p, which is unaffected in AP-3 mutants. We also show that in the apm3-Δ mutant some ALP is delivered to the vacuole by diversion into the Vps pathway. Class E vps mutants accumulate an exaggerated prevacuolar compartment containing membrane proteins on their way to the vacuole or destined for recycling to the Golgi. Surprisingly, in AP-3 class E vps double mutants these proteins reappear on the vacuole. We suggest that some AP-3–dependent cargo proteins that regulate late steps in Golgi to vacuole transport are diverted into the Vps pathway allowing completion of transfer to the vacuole in the class E vps mutant.The formation of vesicles for transport between membrane-bound organelles requires assembly of coat proteins that are recruited from the cytosol. These proteins direct the sequestration and concentration of cargo as well as invagination of the membrane. One of the best studied classes of coats involved in vesicle budding is comprised of clathrin and its adaptor proteins (APs)1, AP-1 and AP-2 (Schmid, 1997). In clathrin-mediated vesicle transport the AP complexes play the dual role of cargo selection and recruitment of clathrin to the membrane. These adaptors are heterotetramers containing two large chains (adaptins, α or γ and β), one medium chain (μ), and one small chain (σ). AP-1 (γ, β1, μ1, and σ1) functions in sorting at the TGN, whereas AP-2 (α, β2, μ2, and σ2) is involved in receptor capture at the PM during endocytosis.Although there is a great deal of evidence supporting the involvement of adaptors in clathrin-mediated vesicle budding, recent studies in animal cells have led to the discovery of a novel adaptor-like complex, AP-3, that seems to function independently of clathrin (Newman et al., 1995; Simpson et al., 1996). AP-3 has identical subunit architecture to AP-1 and AP-2, with two adaptin-like subunits (δ and β3), a medium chain (μ3), and a small chain (σ3) (Simpson et al., 1996, 1997; Dell''Angelica et al., 1997a , b ). AP-3 antibodies label a perinuclear region, perhaps the TGN, and punctate structures extending to the cell periphery, which may be endosomal compartments (Simpson et al., 1996, 1997; Dell''Angelica et al., 1997a ). However, the mammalian AP-3 complex does not colocalize with clathrin or AP-1 and AP-2 adaptors in cells and it does not copurify with brain clathrin-coated vesicles (Newman et al., 1995; Simpson et al., 1996, 1997; Dell''Angelica et al., 1997b ). Clues to the function of AP-3 have come from the discovery that the garnet gene of Drosophila encodes a protein closely related to δ adaptin (Ooi et al., 1997; Simpson et al., 1997). Mutations in garnet cause decreased pigmentation of the eyes and other tissues and a reduced number of pigment granules, which may be lysosome-like organelles (Ooi et al., 1997; Simpson et al., 1997). Thus, AP-3 is proposed to function in clathrin-independent transport between the TGN, endosomes and/or lysosomes, although its exact sorting function is still not known.Over the last several years, yeast homologues of the mammalian adaptor subunits have been identified, allowing for the examination of specific functions of these proteins in a genetically tractable organism. Genes encoding subunits sufficient for at least three complete AP complexes have been identified by sequence homology (Phan et al., 1994; Rad et al., 1995; Stepp et al., 1995) or by function (Panek et al., 1997). APL1-APL6 encode large chain/ adaptin-related subunits, APM1-APM4 encode μ-like chains, and APS1-APS3 are genes for σ-related proteins. Apl2p (β), Apl4p (γ), Apm1p (μ1), and Aps1p (σ1) are thought to be subunits of an AP-1–like complex that functions with clathrin at the late Golgi/TGN (Phan et al., 1994; Rad et al., 1995; Stepp et al., 1995; Payne, G., personal communication). Mutations in the yeast AP-1 genes enhance the growth and the α-factor processing defects of a temperature sensitive (ts) allele of the clathrin heavy chain gene (Phan et al., 1994; Rad et al., 1995; Stepp et al., 1995; Payne, G., personal communication). The latter phenotype is a hallmark of clathrin-deficient yeast, in which late Golgi/ TGN proteins, such as the α-factor processing enzymes Kex2p and dipeptidyl amino peptidase-A (DPAP)-A, are not retained in the late Golgi but escape to the cell surface (Seeger and Payne, 1992b ). To date, no yeast adaptor subunit has been shown to be important for endocytosis, although Apl3p, Apm4p, and Aps2p are most homologous to mammalian AP-2 α, μ2 and σ2, respectively.Recently, a yeast adaptor related to AP-3 of animal cells was described (Panek et al., 1997). It is comprised of Apl5p, Apl6p, Apm3p, and Aps3p, which show preferential homology to mammalian δ, β3, μ3, and σ3, respectively. Mutations in each of these subunits were isolated by their ability to suppress the lethality resulting from loss of function of PM casein kinase 1 encoded by a gene pair, YCK1 and YCK2. Yck activity was found to be required for constitutive endocytosis of the a-factor receptor (Ste3p), and AP-3 subunit mutations partially rescued this internalization defect (Panek et al., 1997). However, the AP complex itself is not necessary for endocytosis, nor is it required for sorting of carboxypeptidase Y (CPY) or retention of late Golgi proteins. Furthermore, unlike disruption of the yeast AP-1 complex, loss of AP-3 function causes no synthetic phenotype in combination with chc1 mutations, suggesting it may function independently of clathrin. Although these data indicated that Apl5p, Apl6p, Apm3p, and Aps3p comprise an AP-3-like adaptor, its precise sorting role was still not known.In this report we describe a genetic approach to determine the function of the yeast AP-3 complex. A colony sectoring screen was performed to identify genes that are essential in the absence of Apm3p, the yeast AP-3 μ chain. Such synthetic lethal screens can be used to identify functional homologues, genes whose proteins function in intersecting or parallel pathways, and genes whose proteins physically interact (Bender and Pringle, 1991). We have cloned the gene for the apm three synthetic lethal mutant, mts1-1, and found it encodes Vps45p, a protein involved in vacuolar protein sorting (Vps; Cowles et al., 1994; Piper et al., 1994). The Vps pathway is defined by >40 complementation groups whose proteins are required for the transport of a number of soluble and membrane-bound proteins, including CPY, protease A (PrA), and carboxypeptidase S (CPS) from the late Golgi/TGN to the vacuole (Stack et al., 1995; Cowles et al., 1997). This pathway is also essential for proper assembly of the vacuolar ATPase (Raymond et al., 1992). However, the type II vacuolar membrane protein alkaline phosphatase (ALP) follows an alternate intracellular pathway to the vacuole (Raymond et al., 1992; Nothwehr et al., 1995; Cowles et al., 1997; Piper et al., 1997). Few vps mutants prevent localization of ALP to the vacuolar membrane and its arrival at the vacuole is not dependent upon transport through the cell surface. The requirement for Apm3p in the absence of Vps45p suggested the possibility that at least one of these routes to the vacuole must be functional for survival and led us to examine ALP sorting in the AP-3 mutants. We show here that yeast AP-3 is essential for the transport of ALP via the alternative pathway to the vacuole.  相似文献   

11.
12.
13.
The role of extracellular α-synuclein (α-syn) in the initiation and the spreading of neurodegeneration in Parkinson''s disease (PD) has been studied extensively over the past 10 years. However, the nature of the α-syn toxic species and the molecular mechanisms by which they may contribute to neuronal cell loss remain controversial. In this study, we show that fully characterized recombinant monomeric, fibrillar or stabilized forms of oligomeric α-syn do not trigger significant cell death when added individually to neuroblastoma cell lines. However, a mixture of preformed fibrils (PFFs) with monomeric α-syn becomes toxic under conditions that promote their growth and amyloid formation. In hippocampal primary neurons and ex vivo hippocampal slice cultures, α-syn PFFs are capable of inducing a moderate toxicity over time that is greatly exacerbated upon promoting fibril growth by addition of monomeric α-syn. The causal relationship between α-syn aggregation and cellular toxicity was further investigated by assessing the effect of inhibiting fibrillization on α-syn-induced cell death. Remarkably, our data show that blocking fibril growth by treatment with known pharmacological inhibitor of α-syn fibrillization (Tolcapone) or replacing monomeric α-syn by monomeric β-synuclein in α-syn mixture composition prevent α-syn-induced toxicity in both neuroblastoma cell lines and hippocampal primary neurons. We demonstrate that exogenously added α-syn fibrils bind to the plasma membrane and serve as nucleation sites for the formation of α-syn fibrils and promote the accumulation and internalization of these aggregates that in turn activate both the extrinsic and intrinsic apoptotic cell death pathways in our cellular models. Our results support the hypothesis that ongoing aggregation and fibrillization of extracellular α-syn play central roles in α-syn extracellular toxicity, and suggest that inhibiting fibril growth and seeding capacity constitute a viable strategy for protecting against α-syn-induced toxicity and slowing the progression of neurodegeneration in PD and other synucleinopathies.The discovery of α-synuclein (α-syn) as the main component of Lewy bodies (LBs) and the identification of gene duplication and missense mutations in the α-syn gene in some familial forms of Parkinson''s disease (PD) have reinforced the central role of α-syn in the etiology of both sporadic and familial cases of PD.1 Nevertheless, the relationship between α-syn aggregation and neurodegeneration in PD remains elusive.2A possible role for extracellular α-syn in the pathogenicity of PD emerged from the observation that newly grafted neurons in PD patients exhibit α-syn pathology similar to that of neighboring diseased cells.3, 4 Despite the consensus that α-syn is mainly an intracellular protein, α-syn has been detected in the cerebrospinal fluid under both pathological and healthy conditions.5 In addition, in vivo rodent models and cellular studies have shown that monomers6 and aggregated forms6, 7 of α-syn are secreted into the extracellular space via several mechanisms,7, 8 including the nonclassical endoplasmic reticulum/Golgi-independent exocytosis8 or the exosomal route,9, 10 and are then internalized by neighboring cells.7 This suggests that extracellular α-syn may play a critical role in the spreading of α-syn pathology throughout the brain and contributes to PD progression.Additional evidence for a causal role of extracellular α-syn in PD come from in vivo studies and cell culture models: (1) intracranial injections of pathological forms of α-syn, isolated from LBs or old mice, as well as recombinant α-syn fibrils, were shown to nucleate further α-syn aggregation, pathology spreading and trigger neurodegeneration in vivo in wild-type (WT) or transgenic mice11, 12, 13, 14 and rhesus monkeys;15 (2) recombinant extracellular α-syn aggregates are internalized in cultured cells and seed the aggregation of endogenous α-syn;12, 16, 17, 18 and (3) extracellular α-syn activates microglia that initiates or enhances nigral neurodegeneration.19, 20, 21 Although the toxic effects of exogenous recombinant α-syn have been thoroughly investigated in different cellular models,7, 17, 18, 22, 23, 24, 25 the relative contribution of monomeric, oligomeric and fibrillar forms of α-syn to the overall toxicity remains controversial. Therefore, the identification of toxic α-syn species and the molecular mechanisms by which they contribute to neurodegeneration is required to better understand how extracellular α-syn contributes to PD pathogenesis and to develop novel strategies for the diagnosis and treatment of PD and other synucleinopathies. In our study we explored the relationship between neurotoxicity and the aggregation state or amyloid formation propensity of α-syn in various cellular models.  相似文献   

14.
15.
16.
Aspergillus oryzae was found to secrete two distinct β-glucosidases when it was grown in liquid culture on various substrates. The major form had a molecular mass of 130 kDa and was highly inhibited by glucose. The minor form, which was induced most effectively on quercetin (3,3′,4′,5,7-pentahydroxyflavone)-rich medium, represented no more than 18% of total β-glucosidase activity but exhibited a high tolerance to glucose inhibition. This highly glucose-tolerant β-glucosidase (designated HGT-BG) was purified to homogeneity by ammonium sulfate precipitation, gel filtration, and anion-exchange chromatography. HGT-BG is a monomeric protein with an apparent molecular mass of 43 kDa and a pI of 4.2 as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and isoelectric focusing polyacrylamide gel electrophoresis, respectively. Using p-nitrophenyl-β-d-glucoside as the substrate, we found that the enzyme was optimally active at 50°C and pH 5.0 and had a specific activity of 1,066 μmol min−1 mg of protein−1 and a Km of 0.55 mM under these conditions. The enzyme is particularly resistant to inhibition by glucose (Ki, 1.36 M) or glucono-δ-lactone (Ki, 12.5 mM), another powerful β-glucosidase inhibitor present in wine. A comparison of the enzyme activities on various glycosidic substrates indicated that HGT-BG is a broad-specificity type of fungal β-glucosidase. It exhibits exoglucanase activity and hydrolyzes (1→3)- and (1→6)-β-glucosidic linkages most effectively. This enzyme was able to release flavor compounds, such as geraniol, nerol, and linalol, from the corresponding monoterpenyl-β-d-glucosides in a grape must (pH 2.9, 90 g of glucose liter−1). Other flavor precursors (benzyl- and 2-phenylethyl-β-d-glucosides) and prunin (4′,5,7-trihydroxyflavanone-7-glucoside), which contribute to the bitterness of citrus juices, are also substrates of the enzyme. Thus, this novel β-glucosidase is of great potential interest in wine and fruit juice processing because it releases aromatic compounds from flavorless glucosidic precursors.β-Glucoside glucohydrolases, commonly called β-glucosidases, catalyze the hydrolysis of alkyl- and aryl-β-glucosides, as well as diglucosides and oligosaccharides. These enzymes are widely used in various biotechnological processes, including the production of fuel ethanol from cellulosic agricultural residues (4, 27, 48) and the synthesis of useful β-glucosides (21, 38). In the flavor industry, β-glucosidases are also key enzymes in the enzymatic release of aromatic compounds from glucosidic precursors present in fruits and fermentating products (13, 39). Indeed, many natural flavor compounds, such as monoterpenols, C-13 norisoprenoids, and shikimate-derived compounds, accumulate in fruits as flavorless precursors linked to mono- or diglycosides and require enzymatic or acidic hydrolysis for the liberation of their fragrances (41, 45). Finally, β-glucosidases can also improve the organoleptic properties of citrus fruit juices, in which the bitterness is in part due to a glucosidic compound, naringin (4′,5,7-trihydroxyflavanone-7-rhamnoglucoside), whose hydrolysis requires, in succession, an α-rhamnosidase and a β-glucosidase (33).It is now well-established that certain monoterpenols of grapes (e.g., linalol, geraniol, nerol, citronelol, α-terpineol, and linalol oxide), which are linked to diglycosides, such as 6-O-α-l-rhamnopyranosyl-, 6-O-α-l-arabinofuranosyl-, and 6-O-β-d-apiofuranosyl-β-d-glucosides, contribute significantly to the flavor of wine (15, 44). The enzymatic hydrolysis of these compounds requires a sequential reaction; first, an α-l-rhamnosidase, an α-l-arabinofuranosidase, or a β-d-apiofuranosidase cleaves the (1→6) osidic linkage, and then, the flavor compounds are liberated from the monoglucosides by the action of a β-glucosidase (18, 19). Unlike acidic hydrolysis, enzymatic hydrolysis is highly efficient and does not result in modifications of the aromatic character (16). However, grape and yeast glucosidases exhibit limited activity on monoterpenyl-glucosides during winemaking, and a large fraction of the aromatic precursors remains unprocessed (9, 16, 35). The addition of exogenous β-glucosidase during or following fermentation has been found to be the most effective way to improve the hydrolysis of the glycoconjugated aroma compounds in order to enhance wine flavor (2, 14, 39, 40). The ideal β-glucosidase should function and be stable at a low pH value (pH 2.5 to 3.8) and should be active at a high concentration of glucose (10 to 20%) and in the presence of 10 to 15% ethanol. However, most microbial β-glucosidases are very sensitive to glucose inhibition (4, 12, 47), as well as to inhibition by glucono-δ-lactone, another powerful β-glucosidase inhibitor produced by grape-attacking fungi which can be found in wine must at concentrations up to 2 g/liter (10).The need for more suitable enzymes has led us and other workers to search for novel β-glucosidases with the desired properties. Recently, we showed that an extracellular glucose-tolerant and pH-stable β-glucosidase can be produced by Aspergillus strains (17). However, the enzyme of interest represented only a minor fraction of total β-glucosidase activity, and the major form was highly sensitive to glucose inhibition. Aspergillus oryzae appeared to be the best producer of the minor form when it was grown on quercetin (3,3′,4′,5,7-pentahydroxyflavone), a phenolic flavonoid found in plant cell walls. This paper presents further data on the production and characterization of this novel highly glucose-tolerant β-glucosidase (designated HGT-BG) purified from the extracellular culture filtrate of A. oryzae grown on quercetin.  相似文献   

17.
18.
The cDNA coding for Penicillium purpurogenum α-galactosidase (αGal) was cloned and sequenced. The deduced amino acid sequence of the α-Gal cDNA showed that the mature enzyme consisted of 419 amino acid residues with a molecular mass of 46,334 Da. The derived amino acid sequence of the enzyme showed similarity to eukaryotic αGals from plants, animals, yeasts, and filamentous fungi. The highest similarity observed (57% identity) was to Trichoderma reesei AGLI. The cDNA was expressed in Saccharomyces cerevisiae under the control of the yeast GAL10 promoter. Almost all of the enzyme produced was secreted into the culture medium, and the expression level reached was approximately 0.2 g/liter. The recombinant enzyme purified to homogeneity was highly glycosylated, showed slightly higher specific activity, and exhibited properties almost identical to those of the native enzyme from P. purpurogenum in terms of the N-terminal amino acid sequence, thermoactivity, pH profile, and mode of action on galacto-oligosaccharides.α-Galactosidase (αGal) (EC 3.2.1.22) is of particular interest in view of its biotechnological applications. αGal from coffee beans demonstrates a relatively broad substrate specificity, cleaving a variety of terminal α-galactosyl residues, including blood group B antigens on the erythrocyte surface. Treatment of type B erythrocytes with coffee bean αGal results in specific removal of the terminal α-galactosyl residues, thus generating serological type O erythrocytes (8). Cyamopsis tetragonoloba (guar) αGal effectively liberates the α-galactosyl residue of galactomannan. Removal of a quantitative proportion of galactose moieties from guar gum by αGal improves the gelling properties of the polysaccharide and makes them comparable to those of locust bean gum (18). In the sugar beet industry, αGal has been used to increase the sucrose yield by eliminating raffinose, which prevents normal crystallization of beet sugar (28). Raffinose and stachyose in beans are known to cause flatulence. αGal has the potential to alleviate these symptoms, for instance, in the treatment of soybean milk (16).αGals are also known to occur widely in microorganisms, plants, and animals, and some of them have been purified and characterized (5). Dey et al. showed that αGals are classified into two groups based on their substrate specificity. One group is specific for low-Mr α-galactosides such as pNPGal (p-nitrophenyl-α-d-galactopyranoside), melibiose, and the raffinose family of oligosaccharides. The other group of αGals acts on galactomannans and also hydrolyzes low-Mr substrates to various extents (6).We have studied the substrate specificity of αGals by using galactomanno-oligosaccharides such as Gal3Man3 (63-mono-α-d-galactopyranosyl-β-1,4-mannotriose) and Gal3Man4 (63-mono-α-d-galactopyranosyl-β-1,4-mannotetraose). The structures of these galactomanno-oligosaccharides are shown in Fig. Fig.1.1. Mortierella vinacea αGal I (11) and yeast αGals (29) are specific for the Gal3Man3 having an α-galactosyl residue (designated the terminal α-galactosyl residue) attached to the O-6 position of the nonreducing end mannose of β-1,4-mannotriose. On the other hand, Aspergillus niger 5-16 αGal (12) and Penicillium purpurogenum αGal (25) show a preference for the Gal3Man4 having an α-galactosyl residue (designated the stubbed α-galactosyl residue) attached to the O-6 position of the third mannose from the reducing end of β-1,4-mannotetraose. The M. vinacea αGal II (26) acts on both substrates to almost equal extents. The difference in specificity may be ascribed to the tertiary structures of these enzymes. Open in a separate windowFIG. 1Structures of galactomanno-oligosaccharides.Genes encoding αGals have been cloned from various sources, including humans (3), plants (20, 32), yeasts (27), filamentous fungi (4, 17, 24, 26), and bacteria (1, 2, 15). αGals from eukaryotes show a considerable degree of similarity and are grouped into family 27 (10).Here we describe the cloning of P. purpurogenum αGal cDNA, its expression in Saccharomyces cerevisiae, and the purification and characterization of the recombinant enzyme.  相似文献   

19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号