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.
The DNA mismatch repair (MMR) system is a major DNA repair system that corrects DNA replication errors. In eukaryotes, the MMR system functions via mechanisms both dependent on and independent of exonuclease 1 (EXO1), an enzyme that has multiple roles in DNA metabolism. Although the mechanism of EXO1-dependent MMR is well understood, less is known about EXO1-independent MMR. Here, we provide genetic and biochemical evidence that the DNA2 nuclease/helicase has a role in EXO1-independent MMR. Biochemical reactions reconstituted with purified human proteins demonstrated that the nuclease activity of DNA2 promotes an EXO1-independent MMR reaction via a mismatch excision-independent mechanism that involves DNA polymerase δ. We show that DNA polymerase ε is not able to replace DNA polymerase δ in the DNA2-promoted MMR reaction. Unlike its nuclease activity, the helicase activity of DNA2 is dispensable for the ability of the protein to enhance the MMR reaction. Further examination established that DNA2 acts in the EXO1-independent MMR reaction by increasing the strand-displacement activity of DNA polymerase δ. These data reveal a mechanism for EXO1-independent mismatch repair.

The mismatch repair (MMR) system has been conserved from bacteria to humans (1, 2). It promotes genome stability by suppressing spontaneous and DNA damage-induced mutations (1, 3, 4, 5, 6, 7, 8, 9, 10, 11). The key function of the MMR system is the correction of DNA replication errors that escape the proofreading activities of replicative DNA polymerases (1, 4, 5, 6, 7, 8, 9, 10, 12). In addition, the MMR system removes mismatches formed during strand exchange in homologous recombination, suppresses homeologous recombination, initiates apoptosis in response to irreparable DNA damage caused by several anticancer drugs, and contributes to instability of triplet repeats and alternative DNA structures (1, 4, 5, 7, 8, 9, 10, 11, 13, 14, 15, 16, 17, 18). The principal components of the eukaryotic MMR system are MutSα (MSH2-MSH6 heterodimer), MutLα (MLH1-PMS2 heterodimer in humans and Mlh1-Pms1 heterodimer in yeast), MutSβ (MSH2-MSH3 heterodimer), proliferating cell nuclear antigen (PCNA), replication factor C (RFC), exonuclease 1 (EXO1), RPA, and DNA polymerase δ (Pol δ). Loss-of-function mutations in the MSH2, MLH1, MSH6, and PMS2 genes of the human MMR system cause Lynch and Turcot syndromes, and hypermethylation of the MLH1 promoter is responsible for ∼15% of sporadic cancers in several organs (19, 20). MMR deficiency leads to cancer initiation and progression via a multistage process that involves the inactivation of tumor suppressor genes and action of oncogenes (21).MMR occurs behind the replication fork (22, 23) and is a major determinant of the replication fidelity (24). The correction of DNA replication errors by the MMR system increases the replication fidelity by ∼100 fold (25). Strand breaks in leading and lagging strands as well as ribonucleotides in leading strands serve as signals that direct the eukaryotic MMR system to remove DNA replication errors (26, 27, 28, 29, 30). MMR is more efficient on the lagging than the leading strand (31). The substrates for MMR are all six base–base mismatches and 1 to 13-nt insertion/deletion loops (25, 32, 33, 34). Eukaryotic MMR commences with recognition of the mismatch by MutSα or MutSβ (32, 34, 35, 36). MutSα is the primary mismatch-recognition factor that recognizes both base–base mismatches and small insertion/deletion loops whereas MutSβ recognizes small insertion/deletion loops (32, 34, 35, 36, 37). After recognizing the mismatch, MutSα or MutSβ cooperates with RFC-loaded PCNA to activate MutLα endonuclease (38, 39, 40, 41, 42, 43). The activated MutLα endonuclease incises the discontinuous daughter strand 5′ and 3′ to the mismatch. A 5'' strand break formed by MutLα endonuclease is utilized by EXO1 to enter the DNA and excise a discontinuous strand portion encompassing the mismatch in a 5''→3′ excision reaction stimulated by MutSα/MutSβ (38, 44, 45). The generated gap is filled in by the Pol δ holoenzyme, and the nick is ligated by a DNA ligase (44, 46, 47). DNA polymerase ε (Pol ε) can substitute for Pol δ in the EXO1-dependent MMR reaction, but its activity in this reaction is much lower than that of Pol δ (48). Although MutLα endonuclease is essential for MMR in vivo, 5′ nick-dependent MMR reactions reconstituted in the presence of EXO1 are MutLα-independent (44, 47, 49).EXO1 deficiency in humans does not seem to cause significant cancer predisposition (19). Nevertheless, it is known that Exo1-/- mice are susceptible to the development of lymphomas (50). Genetic studies in yeast and mice demonstrated that EXO1 inactivation causes only a modest defect in MMR (50, 51, 52, 53). In agreement with these genetic studies, a defined human EXO1-independent MMR reaction that depends on the strand-displacement DNA synthesis activity of Pol δ holoenzyme to remove the mismatch was reconstituted (54). Furthermore, an EXO1-independent MMR reaction that occurred in a mammalian cell extract system without the formation of a gapped excision intermediate was observed (54). Together, these findings implicated the strand-displacement activity of Pol δ holoenzyme in EXO1-independent MMR.In this study, we investigated DNA2 in the context of MMR. DNA2 is an essential multifunctional protein that has nuclease, ATPase, and 5''→3′ helicase activities (55, 56, 57). Previous research ascertained that DNA2 removes long flaps during Okazaki fragment maturation (58, 59, 60), participates in the resection step of double-strand break repair (61, 62, 63), initiates the replication checkpoint (64), and suppresses the expansions of GAA repeats (65). We have found in vivo and in vitro evidence that DNA2 promotes EXO1-independent MMR. Our data have indicated that the nuclease activity of DNA2 enhances the strand-displacement activity of Pol δ holoenzyme in an EXO1-independent MMR reaction.  相似文献   

2.
A 26-y-old male sooty mangabey (Cercocebus atys) was found at necropsy to have a moderate degree of cerebral amyloid β (Aβ) angiopathy in superficial and parenchymal blood vessels of the brain. Senile (Aβ) plaques were absent, as were neurofibrillary tangles and other signs of neurodegeneration. Affected blood vessels were arterial, capillary, and, less frequently, venous in nature. Histologically, the Aβ40 isoform was more prevalent than was Aβ42. As in humans but unlike in squirrel monkeys, the density of lesions in this mangabey increased along a rostral-to-caudal gradient. Therefore mangabeys appear to conform to the general tendency of nonhuman primates by developing cerebral Aβ angiopathy in the absence of other indices of Alzheimer-type neuropathology.Abbreviations: Aβ, amyloid β, CAA, cerebral amyloid angiopathy, GFAP, glial fibrillary acidic protein, Iba 1, microglia-expressed calcium-binding proteinOne of the most common microvasculopathies in the aging human brain is cerebral amyloid angiopathy (CAA), a disorder in which various aggregation-prone proteins accumulate in the walls of parenchymal and meningeal blood vessels.4,9 Most often, the amyloidogenic protein is amyloid β (Aβ), a cleavage product of the Aβ precursor protein and the essential component of senile plaques in Alzheimer disease.13,43 In the brain vasculature, the basal lamina is a primary site of Aβ deposition.25,35 Severely affected arterioles show a loss of smooth muscle cells in the tunica media, a weakening of the vascular wall and a propensity to rupture.3,34 CAA thus increases the risk of intracerebral bleeding and may be responsible for as much as 20% of nontraumatic hemorrhagic stroke in elderly humans.15,18,35 CAA is present to various degrees in virtually all cases of Alzheimer disease,15,16,21 but it also occurs independently.24 As is the case for other proteopathies, advancing age is a significant risk factor for CAA.8,19In humans, CAA most often affects the arteries and arterioles of the brain, particularly those in the leptomeninges and cortex.2,25 CAA is less frequent in veins and capillaries,25 but capillary CAA can be prominent in some cases.26,33 The occipital lobe is affected most often1,32,37 but all cortical regions are vulnerable. CAA is variable in occurrence in the cerebellum and uncommon in deep telencephalic gray structures, white matter, and the brainstem,36 except in severely affected cases.32Although its specific role in the pathogenesis of Alzheimer disease remains uncertain, there is now strong evidence that dementia is exacerbated by CAA.14 Furthermore, CAA is independently linked to cognitive decline both in rare familial cases20 and in older humans with idiopathic CAA.2,20 Despite the prevalence of cerebrovascular amyloidosis in elderly humans, surprisingly little is known about its effect on the brain, in part because of a paucity of natural animal models that closely mimic the human disorder.17,38Nonhuman primates offer a unique opportunity to view CAA from a comparative perspective, given that they normally generate human-sequence Aβ and develop severe cerebral Aβ amyloidosis in old age, generally in the absence of other changes that characterize Alzheimer disease.12 Nonhuman primates have the additional advantage that, compared with humans, their relatively small brains enable exhaustive regional analysis of microscopic lesions, something that, for practical reasons, is seldom undertaken in the human brain. Here we present the first investigation of age-associated brain changes in sooty mangabeys, focusing in particular on Aβ deposition and related abnormalities. One of the 2 aged mangabeys analyzed had Aβ deposition in the brain which was almost exclusively in the form of CAA. Remarkably, the vessel types affected and the regional distribution of CAA more closely resembled the pattern seen in humans than that in other nonhuman primates, particularly squirrel monkeys.6 Differences and similarities in CAA among primate species could provide fresh insights into the development of cerebral amyloidosis and related disorders in older humans.  相似文献   

3.
Human induced pluripotent stem cell (iPSC)-derived neurons have been proposed to be a highly valuable cellular model for studying the pathomechanisms of Alzheimer''s disease (AD). Studies employing patient-specific human iPSCs as models of familial and sporadic forms of AD described elevated levels of AD-related amyloid-β (Aβ). However, none of the present AD iPSC studies could recapitulate the synaptotoxic actions of Aβ, which are crucial early events in a cascade that eventually leads to vast brain degeneration. Here we established highly reproducible, human iPSC-derived cortical cultures as a cellular model to study the synaptotoxic effects of Aβ. We developed a highly efficient immunopurification procedure yielding immature neurons that express markers of deep layer cortical pyramidal neurons and GABAergic interneurons. Upon long-term cultivation, purified cells differentiated into mature neurons exhibiting the generation of action potentials and excitatory glutamatergic and inhibitory GABAergic synapses. Most interestingly, these iPSC-derived human neurons were strongly susceptible to the synaptotoxic actions of Aβ. Application of Aβ for 8 days led to a reduction in the overall FM4–64 and vGlut1 staining of vesicles in neurites, indicating a loss of vesicle clusters. A selective analysis of presynaptic vesicle clusters on dendrites did not reveal a significant change, thus suggesting that Aβ impaired axonal vesicle clusters. In addition, electrophysiological patch-clamp recordings of AMPA receptor-mediated miniature EPSCs revealed an Aβ-induced reduction in amplitudes, indicating an impairment of postsynaptic AMPA receptors. A loss of postsynaptic AMPA receptor clusters was confirmed by immunocytochemical stainings for GluA1. Incubation with Aβ for 8 days did not result in a significant loss of neurites or cell death. In summary, we describe a highly reproducible cellular AD model based on human iPSC-derived cortical neurons that enables the mechanistic analysis of Aβ-induced synaptic pathomechanisms and the development of novel therapeutic approaches.In Alzheimer''s disease (AD), synapse damage and synapse loss are thought to underlie cognitive deficits.1 Oligomers of the amyloid-β (Aβ) peptide appear to induce synaptic failure as an early event in the etiology of AD.2, 3, 4 However, despite its well-established synapse-impairing effects in rodent models,5, 6, 7 the synaptotoxic actions of Aβ most relevant for the human disease have not been identified in a human model system. Several studies have investigated the synaptotoxic effects of Aβ in cultured rodent neurons and in transgenic mouse models revealing a multitude of potential mechanisms affecting synapses. Postsynaptic Aβ actions result in the loss of functional (α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA)-type) glutamate receptors,8, 9, 10 involve long-term depression-like mechanisms,9, 11, 12 and lead to the degradation of the entire postsynapse (dendritic spines).9, 11, 13 In addition, several distinct presynaptic Aβ actions on the synaptic vesicle cycle have been described.10, 14 Furthermore, Aβ-induced impairments of axonal transport regulation and Aβ-induced axon degeneration have been found in rodent neurons.15, 16, 17 This puzzling diversity of Aβ-induced synapse-related defects raises the question whether all of them are involved in the early pathomechanisms of human AD.In addition to well-established animal systems, the modelling of human neurological disease pathologies by human induced pluripotent stem cell (hiPSC) technology18 has been proposed as an innovative approach.19, 20, 21 The in vitro differentiation of hiPSCs to excitable neurons has been reported using a variety of protocols.22, 23, 24 However, quantitative analysis of both functional glutamatergic and GABAergic synapses has been difficult to achieve.19, 25, 26 In addition to studying the functional properties of iPSC-derived human neurons from healthy individuals, the in vitro differentiation of patient-derived iPSCs has been used to model complex neurodevelopmental and neurodegenerative diseases.19, 27, 28 Recently, iPSCs derived from AD patients have been reported to exhibit increased secretion of Aβ upon in vitro neuronal differentiation; however, neither a loss of synapses nor an impairment of synapse function was detected.21, 29, 30, 31, 32, 33 Here we describe a hiPSC-based, carefully optimized in vitro differentiation protocol, including a novel immunopanning step, which enabled us to study the deleterious effects of application of Aβ on human cortical neurons and on human synapses.  相似文献   

4.
Evidence indicates that nitrosative stress and mitochondrial dysfunction participate in the pathogenesis of Alzheimer''s disease (AD). Amyloid beta (Aβ) and peroxynitrite induce mitochondrial fragmentation and neuronal cell death by abnormal activation of dynamin-related protein 1 (DRP1), a large GTPase that regulates mitochondrial fission. The exact mechanisms of mitochondrial fragmentation and DRP1 overactivation in AD remain unknown; however, DRP1 serine 616 (S616) phosphorylation is likely involved. Although it is clear that nitrosative stress caused by peroxynitrite has a role in AD, effective antioxidant therapies are lacking. Cerium oxide nanoparticles, or nanoceria, switch between their Ce3+ and Ce4+ states and are able to scavenge superoxide anions, hydrogen peroxide and peroxynitrite. Therefore, nanoceria might protect against neurodegeneration. Here we report that nanoceria are internalized by neurons and accumulate at the mitochondrial outer membrane and plasma membrane. Furthermore, nanoceria reduce levels of reactive nitrogen species and protein tyrosine nitration in neurons exposed to peroxynitrite. Importantly, nanoceria reduce endogenous peroxynitrite and Aβ-induced mitochondrial fragmentation, DRP1 S616 hyperphosphorylation and neuronal cell death.Nitric oxide (NO) is a neurotransmitter and neuromodulator required for learning and memory.1 NO is generated by NO synthases, a group of enzymes that produce NO from L-arginine. In addition to its normal role in physiology, NO is implicated in pathophysiology. When overproduced, NO combines with superoxide anions (O2·), byproducts of aerobic metabolism and mitochondrial oxidative phosphorylation, to form peroxynitrite anions (ONOO) that are highly reactive and neurotoxic. Accumulation of these reactive oxygen species (ROS) and reactive nitrogen species (RNS), known as oxidative and nitrosative stress, respectively, is a common feature of aging, neurodegeneration and Alzheimer''s disease (AD).1Nitrosative stress caused by peroxynitrite has a critical role in the etiology and pathogenesis of AD.2, 3, 4, 5, 6, 7 Peroxynitrite is implicated in the formation of the two hallmarks of AD, Aβ aggregates and neurofibrillary tangles containing hyperphosphorylated Tau protein.1, 4, 7 In addition, peroxynitrite promotes the nitrotyrosination of presenilin 1, the catalytic subunit of the γ-secretase complex, which shifts production of Aβ to amyloid beta (Aβ)42 and increases the Aβ42/Aβ40 ratio, ultimately resulting in an increased propensity for aggregation and neurotoxicity.5 Furthermore, nitration of Aβ tyrosine 10 enhances its aggregation.6 Peroxynitrite can also modify enzymes, such as triosephosphate isomerase,4 and activate kinases, including Jun amino-terminal kinase and p38 mitogen-activated protein kinase, which enhance neuronal cell death.8, 9 Moreover, peroxynitrite can trigger the release of free metals such as Zn2+ from intracellular stores with consequent inhibition of mitochondrial function and enhancement of neuronal cell death.10, 11, 12 Finally, peroxynitrite can irreversibly inhibit complexes I and IV of the mitochondrial respiratory chain.11, 13Because mitochondria have a critical role in neurons as energy producers to fuel vital processes such as synaptic transmission and axonal transport,14 and mitochondrial dysfunction is a well-documented and early event in AD,15 it is important to consider how peroxynitrite and nitrosative stress affect mitochondria. Although the ultimate cause of mitochondrial dysfunction in AD remains unclear, an imbalance in mitochondrial fission and fusion is one possibility.1, 14, 16, 17, 18 Notably, peroxynitrite, N-methyl D-aspartate (NMDA) receptor activation and Aβ can induce mitochondrial fragmentation by activating mitochondrial fission and/or inhibiting fusion.16 Mitochondrial fission and fusion is regulated by large GTPases of the dynamin family, including dynamin-related protein 1 (DRP1) that is required for mitochondrial division,19 and inhibition of mitochondrial division by overexpression of the GTPase-defective DRP1K38A mutant provides protection against peroxynitrite-, NMDA- and Aβ-induced mitochondrial fragmentation and neuronal cell death.16The exact mechanism of peroxynitrite-induced mitochondrial fragmentation remains unclear. A recent report suggested that S-nitrosylation of DRP1 at cysteine 644 increases DRP1 activity and is the cause of peroxynitrite-induced mitochondrial fragmentation in AD;20 however, the work remains controversial, suggesting that alternative pathways might be involved.21 For example, peroxynitrite also causes rapid DRP1 S616 phosphorylation that promotes its translocation to mitochondria and organelle division.21, 22 In mitotic cells, DRP1 S616 phosphorylation is mediated by Cdk1/cyclinB1 and synchronizes mitochondrial division with cell division.23 Interestingly, DRP1 is S616 hyperphosphorylated in AD brains, suggesting that this event might contribute to mitochondrial fragmentation in the disease.21, 22 A recent report indicates that Cdk5/p35 is responsible for DRP1 S616 phosphorylation,24 and notably aberrant Cdk5/p35/p25 signaling is associated with AD pathogenesis.25 Thus, we explored here the possible role of DRP1 S616 hyperphosphorylation in Aβ- and peroxynitrite-mediated mitochondrial fragmentation.Under normal conditions, accumulated mitochondrial superoxide anions and hydrogen peroxide (H2O2) can be neutralized by superoxide dismutase (SOD) and catalase. Nitrosative stress in aging and AD might be explained by a loss of antioxidant enzymes. Previous studies suggest that expression of SOD subtypes is decreased in the human AD brain.26, 27 Furthermore, SOD1 deletion in a mouse model of AD increased the burden of amyloid plaques.26 By contrast, overexpression of SOD2 in a mouse model of AD decreased the Aβ42/Aβ40 ratio and alleviated memory deficits.28, 29 There is currently a lack of antioxidants that can effectively quench superoxide anions, H2O2 or peroxynitrite and provide lasting effects. Cerium is a rare earth element and cerium oxide (CeO2) nanoparticles, or nanoceria, shuttle between their 3+ or 4+ states. Oxidation of Ce4+ to Ce3+ causes oxygen vacancies and defects on the surface of the crystalline lattice structure of the nanoparticles, generating a cage for redox reactions to occur.30 Accordingly, nanoceria mimic the catalytic activities of antioxidant enzymes, such as SOD31, 32 and catalase,33 and are able to neutralize peroxynitrite.34 Because of these antioxidant properties, we hypothesized that nanoceria could detoxify peroxynitrite and protect against Aβ-induced DRP1 S616 hyperphosphorylation, mitochondrial fragmentation and neuronal cell death.  相似文献   

5.
The reduced clearance of amyloid-β peptide (Aβ) from the brain partly accounts for the neurotoxic accumulation of Aβ in Alzheimer''s disease (AD). Recently, it has been suggested that P-glycoprotein (P-gp), which is an efflux transporter expressed on the luminal membrane of the brain capillary endothelium, is capable of transporting Aβ out of the brain. Although evidence has shown that restoring P-gp reduces brain Aβ in a mouse model of AD, the molecular mechanisms underlying the decrease in P-gp expression in AD is largely unknown. We found that Aβ1–42 reduced P-gp expression in the murine brain endothelial cell line bEnd.3, which was consistent with our in vivo data that P-gp expression was significantly reduced, especially near amyloid plaques in the brains of five familial AD mutations (5XFAD) mice that are used as an animal model for AD. A neutralizing antibody against the receptor for advanced glycation end products (RAGE) and an inhibitor of nuclear factor-kappa B (NF-κB) signaling prevented the decrease in Aβ1–42-induced P-gp expression, suggesting that Aβ reduced P-gp expression through NF-κB signaling by interacting with RAGE. In addition, we observed that the P-gp reduction by Aβ was rescued in bEnd.3 cells receiving inductive signals or factors from astrocytes making contacts with endothelial cells (ECs). These results support that alterations of astrocyte–EC contacts were closely associated with P-gp expression. This suggestion was further supported by the observation of a loss of astrocyte polarity in the brains of 5XFAD mice. Taken together, we found that P-gp downregulation by Aβ was mediated through RAGE–NF-κB signaling pathway in ECs and that the contact between astrocytes and ECs was an important factor in the regulation of P-gp expression.Alzheimer''s disease (AD) is a neurodegenerative disorder that is characterized by a progressive loss of cognitive function leading to dementia. The major pathological hallmark of AD is the deposition of neurotoxic amyloid-β peptide (Aβ) within the brain.1 The amyloid hypothesis proposes that the accumulation of Aβ is caused by an imbalance between Aβ production and clearance.2 Although genetic alterations increase the production of Aβ in rare familial AD, reduced Aβ clearance from the brain likely accounts for sporadic AD, which is much more common.3 The mechanisms that are involved in clearing Aβ from the brain include enzymatic degradation, perivascular drainage, and the most significant, active transport across the blood–brain barrier (BBB).4The BBB regulates molecular exchanges at the interface between the blood and the brain.5 It plays a critical role in maintaining the brain microenvironment.6 The BBB, which is formed by cerebral endothelial cells (ECs) and which, interacts with astrocytes, neurons, pericytes, and the extracellular matrix, is organized into a neurovascular unit.7, 8 Although the relationship between BBB breakdown and AD pathology is unclear,9 it has been proposed that the BBB loses its Aβ clearing capability, thus increasing amyloid deposition in the outer capillary membrane and resulting in the distortion of the neurovascular unit with neuronal loss.10Recently, it has been suggested that P-glycoprotein (P-gp), which is an ATP-driven efflux transporter that is highly expressed in the luminal membrane of the brain capillary endothelium, is also involved in the clearance of Aβ from the brain.11 P-gp, which is able to transport various kinds of substrates, has been shown to play an important role in clearing toxic substances in the brain and protecting it from harmful molecules in the circulation.12 Along with other BBB properties, P-gp expression is induced when ECs are in contact with astrocytes in vitro and in vivo.13, 14 ECs respond to inductive signals or factors from astrocytes that encircle the capillary endothelium.13Several lines of evidence have shown that P-gp plays an important role in Aβ clearance. It has been shown in vitro that P-gp mediates the transport of Aβ and that blocking P-gp function reduces the clearance of Aβ.15, 16 In addition, cerebral Aβ deposition in elderly non-demented individuals has been demonstrated to be inversely correlated with brain capillary P-gp expression.17 Furthermore, in P-gp knockout mice, Aβ deposition is increased by the reduced efflux of Aβ,18 while it has been shown that restoring P-gp at the BBB reduces brain Aβ in a mouse model of AD.19 However, the molecular mechanisms underlying the decrease in P-gp expression that is observed in AD have not been identified. We found that Aβ decreased P-gp expression by increasing nuclear factor-kappa B (NF-κB) through an interaction with the receptor for advanced glycation end products (RAGE). Moreover, we observed that the P-gp reduction by Aβ was rescued by inductive signals or factors from astrocytes that made contact with ECs in bEnd.3 cells. These results suggested that alterations in astrocyte–EC contact in AD likely decrease P-gp expression by Aβ. Together, we identified a mechanism by which the Aβ–RAGE interaction mediated the downregulation of P-gp in the BBB by increasing NF-κB signaling in AD and that astrocyte–EC contact played a critical role in maintaining P-gp expression.  相似文献   

6.
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   

7.
Bak and Bax mediate apoptotic cell death by oligomerizing and forming a pore in the mitochondrial outer membrane. Both proteins anchor to the outer membrane via a C-terminal transmembrane domain, although its topology within the apoptotic pore is not known. Cysteine-scanning mutagenesis and hydrophilic labeling confirmed that in healthy mitochondria the Bak α9 segment traverses the outer membrane, with 11 central residues shielded from labeling. After pore formation those residues remained shielded, indicating that α9 does not line a pore. Bak (and Bax) activation allowed linkage of α9 to neighboring α9 segments, identifying an α9:α9 interface in Bak (and Bax) oligomers. Although the linkage pattern along α9 indicated a preferred packing surface, there was no evidence of a dimerization motif. Rather, the interface was invoked in part by Bak conformation change and in part by BH3:groove dimerization. The α9:α9 interaction may constitute a secondary interface in Bak oligomers, as it could link BH3:groove dimers to high-order oligomers. Moreover, as high-order oligomers were generated when α9:α9 linkage in the membrane was combined with α6:α6 linkage on the membrane surface, the α6-α9 region in oligomerized Bak is flexible. These findings provide the first view of Bak carboxy terminus (C terminus) membrane topology within the apoptotic pore.Mitochondrial permeabilization during apoptosis is regulated by the Bcl-2 family of proteins.1, 2, 3 Although the Bcl-2 homology 3 (BH3)-only members such as Bid and Bim trigger apoptosis by binding to other family members, the prosurvival members block apoptosis by sequestering their pro-apoptotic relatives. Two remaining members, Bak and Bax, form the apoptotic pore within the mitochondrial outer membrane (MOM).Bak and Bax are globular proteins comprising nine α-helices.4, 5 They are activated by BH3-only proteins binding to the α2–α5 surface groove,6, 7, 8, 9, 10, 11, 12 or for Bax, to the α1/α6 ‘rear pocket''.13 Binding triggers dissociation of the latch domain (α6–α8) from the core domain (α2–α5), together with exposure of N-terminal epitopes and the BH3 domain.6, 7, 14, 15, 16 The exposed BH3 domain then binds to the hydrophobic groove in another Bak or Bax molecule to generate symmetric homodimers.6, 7, 14, 17, 18 In addition to dimerizing, parts of activated Bak and Bax associate with the lipid bilayer.19 In Bax, the α5 and α6 helices may insert into the MOM,20 although recent studies indicate that they lie in-plane on the membrane surface, with the hydrophobic α5 sandwiched between the membrane and a BH3:groove dimer interface.7, 21, 22, 23 The dimers can be linked via cysteine residues placed in α6,18, 24, 25 and more recently via cysteine residues in either α3 or α5,6, 21 allowing detection of the higher-order oligomers associated with pore formation.26, 27 However, whether these interactions are required for high-order oligomers and pore formation remains unclear.Like most Bcl-2 members, Bak and Bax are targeted to the MOM via a hydrophobic C-terminal region. The C terminus targets Bak to the MOM in healthy cells,28 whereas the Bax C terminus is either exposed29 or sequestered within the hydrophobic groove until apoptotic signals trigger Bax translocation.5, 30, 31 The hydrophobic stretch is important, as substituting polar or charged residues decreased targeting of Bak and Bax.10, 32 Mitochondrial targeting is also controlled by basic residues at the far C termini,32, 33, 34 and by interaction with VDAC235, 36 via the Bak and Bax C termini.37, 38 Retrotranslocation of Bak and Bax was also altered by swapping the C termini.39The membrane topology of the Bak and Bax C termini before and after apoptosis has not been examined directly, due in part to difficulty in reconstituting oligomers of full-length Bak in artificial membranes. Nor is it known whether the C termini contribute to pore formation by promoting oligomerization or disturbing the membrane. To address these questions synthetic peptides based on the Bak and Bax C termini have been studied in model membranes. The peptides adopt a predominantly α-helical secondary structure,40, 41, 42, 43 with orientation affected by lipid composition.42, 44, 45 The peptides could also permeabilize lipid vesicles,41, 43, 46, 47 suggesting that the C termini in full-length Bak and Bax may contribute to pore formation.Here we examined the membrane topology of the C termini within full-length Bak and Bax in the MOM, both before and after apoptotic pore formation. After pore formation the α9 helices of Bak (and of Bax) became juxtaposed but did not line the surface of a pore. The α9:α9 interaction occurred after Bak activation and conformation change, but was promoted by formation of BH3:groove dimers. Combining linkage at more than one interface indicated that the Bak α9:α9 interface can link BH3:groove dimers to high-order oligomers, and moreover, that the α6–α9 region is flexible in oligomerized Bak.  相似文献   

8.
Given the complex heterogeneity of pathological changes occurring in Alzheimer''s disease (AD), any therapeutic effort absolutely requires a multi-targeted approach, because attempts addressing only a single event may result ineffective. Palmitoylethanolamide (PEA), a naturally occurring lipid amide between palmitic acid and ethanolamine, seems to be a compound able to fulfill the criteria of a multi-factorial therapeutic approach. Here, we describe the anti-inflammatory and neuroprotective activities of systemic administration of PEA in adult male rats given intrahippocampal injection of beta amyloid 1–42 (Aβ 1–42). Moreover, to investigate the molecular mechanisms responsible for the effects induced by PEA, we co-administered PEA with the GW6471, an antagonist of peroxisome proliferator-activated receptor-α (PPAR-α). We found that Aβ 1–42 infusion results in severe changes of biochemical markers related to reactive gliosis, amyloidogenesis, and tau protein hyperphosphorylation. Interestingly, PEA was able to restore the Aβ 1–42-induced alterations through PPAR-α involvement. In addition, results from the Morris water maze task highlighted a mild cognitive deficit during the reversal learning phase of the behavioral study. Similarly to the biochemical data, also mnestic deficits were reduced by PEA treatment. These data disclose novel findings about the therapeutic potential of PEA, and suggest novel strategies that hopefully could have the potential not just to alleviate the symptoms but also to modify disease progression.Alzheimer''s disease (AD) is the most common late-onset and progressive neurodegenerative disorder.1 It is characterized by progressive dementia including episodic memory impairments and involvement of other cognitive domains and skills.2 Although senile plaques (SP) and neurofibrillary tangles (NFTs) are considered pathognomic,3 the concept that these are the only significant pathological changes occurring in AD brain is misleading. In fact, both in vitro and in vivo findings have demonstrated that beta amyloid (Aβ) fragments promote a marked neuroinflammatory response, accounting for the synthesis of different cytokines and pro-inflammatory mediators.4 A chronic inflammation goes beyond physiological control and eventually detrimental effects override the beneficial effects. In fact, after their release, pro-inflammatory signaling molecules may act autocrinally to self-perpetuate reactive gliosis and paracrinally to kill neighboring neurons, thus amplifying the neuropathological damage. Once started, the inappropriate and prolonged inflammatory process may contribute independently to neural dysfunction, cell death, and disease progression.5, 6 Besides the sustained production of pathogenic substances, astrocytes fail to provide their neuro-supportive functions, making neurons more vulnerable to toxic molecules.7 Therefore, targeting neuroinflammation might be an effective therapeutic strategy in AD.8, 9, 10, 11At present, no therapies in clinical use are able to effectively impact the disease course. Therefore, new drugs able to simultaneously ameliorate the numerous pathogenic mechanism involved in AD are therapeutically promising.Palmitoylethanolamide (PEA) has attracted much attention for its proven anti-inflammatory and neuroprotective properties reported in many neuropathological conditions other than AD. PEA, a naturally occurring amide of ethanolamine and palmitic acid, is a lipid messenger that mimics several endocannabinoid-driven actions, even though it does not bind to cannabinoid receptors.12, 13, 14, 15, 16, 17, 18, 19, 20 PEA is abundant in the central nervous system (CNS) and it is conspicuously produced by glial cells.14, 15 Many of its beneficial properties have been considered to be dependent on the activation of the peroxisome proliferator-activated receptor-alpha (PPAR-α).12, 21 Recent reports from our group showed the ability of PEA to attenuate in vitro the Aβ-induced upregulation of a wide range of inflammatory mediators by interacting at the PPAR-α nuclear site.22, 23 Both PPAR-α and PEA are clearly detected in the CNS and their expression may largely change during pathological conditions.24 It has been observed that Aβ significantly blunts PPAR-α in primary rat astrocytes, suggesting the possibility that the downregulation of this receptor may represent one of the molecular mechanisms by which Aβ induces astrocyte activation and possibly exerts toxicity. In addition, PEA is able to reverse the downregulation of PPAR-α induced by Aβ, thus mitigating the overexpression of pro-inflammatory molecules and signals.25Therefore, we explored the anti-inflammatory and neuroprotective effects of PEA in an in vivo model of AD. The present study also incorporates pharmacological experiments to test the hypothesis that PPAR-α was involved in the effects induced by PEA administration. To this purpose, we co-administered PEA with the GW6471 (N-((2S)-2-(((1Z)-1-methyl-3-oxo-3-(4-(trifluoromethyl)phenyl)prop-1-enyl)amino)-3-(4-(2-(5-methyl-2-phenyl-1,3-oxazol-4-yl)ethoxy)phenyl)propyl)propanamide), an antagonist of PPAR-α receptor. These experiments are much needed because PEA is currently clinically used for a condition other than AD. Moreover, its safety and tolerability in humans further demonstrate the translational value of this work.  相似文献   

9.
10.
11.
12.
Heme oxygenase-1 (HO-1), an inducible enzyme up-regulated in Alzheimer''s disease, catabolises heme to biliverdin, Fe2+ and carbon monoxide (CO). CO can protect neurones from oxidative stress-induced apoptosis by inhibiting Kv2.1 channels, which mediates cellular K+ efflux as an early step in the apoptotic cascade. Since apoptosis contributes to the neuronal loss associated with amyloid β peptide (Aβ) toxicity in AD, we investigated the protective effects of HO-1 and CO against Aβ1-42 toxicity in SH-SY5Y cells, employing cells stably transfected with empty vector or expressing the cellular prion protein, PrPc, and rat primary hippocampal neurons. Aβ1-42 (containing protofibrils) caused a concentration-dependent decrease in cell viability, attributable at least in part to induction of apoptosis, with the PrPc-expressing cells showing greater susceptibility to Aβ1-42 toxicity. Pharmacological induction or genetic over-expression of HO-1 significantly ameliorated the effects of Aβ1-42. The CO-donor CORM-2 protected cells against Aβ1-42 toxicity in a concentration-dependent manner. Electrophysiological studies revealed no differences in the outward current pre- and post-Aβ1-42 treatment suggesting that K+ channel activity is unaffected in these cells. Instead, Aβ toxicity was reduced by the L-type Ca2+ channel blocker nifedipine, and by the CaMKKII inhibitor, STO-609. Aβ also activated the downstream kinase, AMP-dependent protein kinase (AMPK). CO prevented this activation of AMPK. Our findings indicate that HO-1 protects against Aβ toxicity via production of CO. Protection does not arise from inhibition of apoptosis-associated K+ efflux, but rather by inhibition of AMPK activation, which has been recently implicated in the toxic effects of Aβ. These data provide a novel, beneficial effect of CO which adds to its growing potential as a therapeutic agent.Amongst the earliest of events leading to neuronal loss in Alzheimer''s disease (AD) is the loss of functional synapses,1, 2, 3 apparent long before deposition of amyloid β peptide (Aβ)-containing plaques.4 Although other parts of the neurone (e.g. the axon or soma) appear intact, their health at this early stage of disease progression is not clear. However, neurones ultimately die in AD and there is clear evidence that numerous events indicative of apoptosis occur even at early stages of disease progression.5, 6, 7, 8 Thus, targeting of apoptotic mechanisms may be of therapeutic value in AD as well as in other neurodegenerative disorders. Furthermore, apoptosis is established as a mechanism of neuronal loss following other types of pathological stresses including ischemia associated with stroke,9 which can predispose individuals to the development of AD.10, 11, 12Apoptosis is strongly influenced by intracellular K+ levels13 which regulate caspase activation, mitochondrial membrane potential and volume, osmolarity and cell volume.13, 14 K+ loss via K+ channels is a key early stage in apoptosis,15, 16, 17, 18, 19 and K+ channel inhibitors can protect against apoptosis triggered by numerous insults including oxidative stress.20, 21 Evidence suggests a particularly important role for the voltage-gated channel Kv2.1 in this process: expression of dominant negative Kv2.1 constructs (thus lacking functional Kv2.1 channels) protects against oxidant-induced apoptosis, and over-expression of Kv2.1 increases susceptibility to apoptosis.22, 23 Pro-apoptotic agents cause a rapid increase in the surface expression of Kv2.1 channels,24 but whether or not this occurs in AD remains to be determined. Alternative pathways recently reported to promote cell death include activation of the AMP-dependent protein kinase (AMP kinase) which can act either as a Tau kinase25 or to inhibit the mTOR pathway26 and thus contribute to neurodegeneration.Heme oxygenases (HO) are enzymes widely distributed throughout the body. In the central nervous system, HO-2 is constitutively expressed in neurones and astrocytes, while HO-1 is inducible in both cell types.27, 28, 29, 30 Both HO-1 and HO-2 break down heme to liberate biliverdin, ferrous iron (Fe2+) and carbon monoxide (CO). This catalysis is of biological significance since it is crucial to iron and bile metabolism, and also generates a highly effective antioxidant in bilirubin (from biliverdin via bilirubin reductase). Numerous stimuli can induce HO-1 gene expression,31 including oxidative stress32 and Aβ peptides.33 Importantly, HO-1 is strikingly up-regulated in AD patients, a finding considered indicative of oxidative stress.27, 34, 35 Induction of HO-1 is clearly a neuroprotective response (although in some cases can exert detrimental effects27). However, there is growing evidence that CO can be neuroprotective, for example against the damage of focal ischemia.36 Our recent studies have demonstrated that CO provides protection against oxidant-induced apoptosis by selectively inhibiting Kv2.1.23, 37 In the present study, we have investigated whether HO-1, or its product CO, can provide protection against Aβ-induced toxicity in the human neuroblastoma, SH-SY5Y, and in rat primary hippocampal neurones, and whether this involves regulation of K+ channels. We show that both HO-1 and CO protect cells against the toxicity of protofibrillar Aβ1-42 but that protection does not arise from inhibition of apoptosis-associated K+ efflux, but rather by inhibition of AMPK activation.  相似文献   

13.
Brain thiamine homeostasis has an important role in energy metabolism and displays reduced activity in Alzheimer''s disease (AD). Thiamine deficiency (TD) induces regionally specific neuronal death in the animal and human brains associated with a mild chronic impairment of oxidative metabolism. These features make the TD model amenable to investigate the cellular mechanisms of neurodegeneration. Once activated by various cellular stresses, including oxidative stress, PKR acts as a pro-apoptotic kinase and negatively controls the protein translation leading to an increase of BACE1 translation. In this study, we used a mouse TD model to assess the involvement of PKR in neuronal death and the molecular mechanisms of AD. Our results showed that the TD model activates the PKR-eIF2α pathway, increases the BACE1 expression levels of Aβ in specific thalamus nuclei and induces motor deficits and neurodegeneration. These effects are reversed by PKR downregulation (using a specific inhibitor or in PKR knockout mice).Thiamine (vitamin B1) deficiency (TD) induces regionally selective neurodegeneration in the mammal''s brains, particularly in specific thalamus nuclei (submedial thalamus nuclei (SmTN) and ventral lateral nuclei (VLN)) and cerebellum.1, 2 TD-induced neuronal loss is associated with a chronic impairment of oxidative metabolism and neuroinflammation associated with glial activation.3, 4, 5 Diminished thiamine-dependent processes in humans is not only associated with Wernicke–Korsakoff syndrome but also accompany other neurodegenerative disorders, such as Alzheimer''s disease (AD).6 Experimental TD has been used to model the pathogenesis and investigate the cellular mechanisms of neurodegenerative disorders. In mice, TD-induced oxidative stress enhances Aβ accumulation by regulating β-site APP-cleaving enzyme 1 (BACE1) maturation. These effects are amplified in AD mouse model.7, 8The double-stranded RNA-dependent protein kinase (PKR) is one of the four mammalian serine–threonine kinases—the others being HRI (heme-regulated eukaryotic translation initiation factor-2α (eIF2α) kinase), GCN2 (general control nonderepressible 2) and PERK (protein kinase RNA-like endoplasmic reticulum kinase)—that catalyzes the phosphorylation of the α subunit of eIF2 in response to stress signals, leading to an inhibition of protein synthesis.9, 10 Activation of PKR is induced by various triggers such as viral infection and endoplasmic reticulum or oxidative stresses11, 12 and could be controlled by an interaction with its specific activator PACT (PKR activator), also named RAX in rodents. PKR phosphorylation is known to have a significant role in AD,13 and cerebrospinal fluid (CSF) PKR levels could be used as a potential diagnostic biomarker in AD patients.14, 15 PKR is a proapoptotic kinase16 and can exert a number of toxic effects on neurons that could contribute to the functional and pathological alterations in AD brains. PKR also contributes to neurodegeneration and to the pathological molecular mechanisms observed in AD. PKR can mediate Tau phosphorylation induced by Aβ exposure in cell cultures.17 Additionally, several investigators have demonstrated that eIF2α phosphorylation, via PKR-induced cellular stress, leads to increased BACE1 mRNA translation when, paradoxically, global protein translation is inhibited.18, 19, 20, 21 These alterations of BACE1 translational control could be explained by a stress-dependent phenomenon of translation initiation.22, 23, 24 Moreover, PKR and eIF2α are highly phosphorylated in SmTN and the cerebellum of TD mouse model. Analyses performed on cerebellar granule neurons exposed to a thiamine metabolic antagonist suggest that TD-induced neuronal death could be partially mediated by PKR activation.25To date, all studies that have explored the deleterious role of PKR activation in neurodegeneration indicate that inhibition of PKR is a promising approach to inhibit pathological mechanisms. Moreover, recent studies have shown that the genetic lack of PKR enhances learning and memory in several behavioral tasks while increasing network excitability.26 The goal of this study was to assess the role of PKR downregulation on neurodegeneration and Aβ production in a mouse model of neurodegeneration.  相似文献   

14.
Proinflammatory cytokines exert cytotoxic effects on β-cells, and are involved in the pathogenesis of type I and type II diabetes and in the drastic loss of β-cells following islet transplantation. Cytokines induce apoptosis and alter the function of differentiated β-cells. Although the MAP3 kinase tumor progression locus 2 (Tpl2) is known to integrate signals from inflammatory stimuli in macrophages, fibroblasts and adipocytes, its role in β-cells is unknown. We demonstrate that Tpl2 is expressed in INS-1E β-cells, mouse and human islets, is activated and upregulated by cytokines and mediates ERK1/2, JNK and p38 activation. Tpl2 inhibition protects β-cells, mouse and human islets from cytokine-induced apoptosis and preserves glucose-induced insulin secretion in mouse and human islets exposed to cytokines. Moreover, Tpl2 inhibition does not affect survival or positive effects of glucose (i.e., ERK1/2 phosphorylation and basal insulin secretion). The protection against cytokine-induced β-cell apoptosis is strengthened when Tpl2 inhibition is combined with the glucagon-like peptide-1 (GLP-1) analog exendin-4 in INS-1E cells. Furthermore, when combined with exendin-4, Tpl2 inhibition prevents cytokine-induced death and dysfunction of human islets. This study proposes that Tpl2 inhibitors, used either alone or combined with a GLP-1 analog, represent potential novel and effective therapeutic strategies to protect diabetic β-cells.It is now clear that chronic inflammation is a hallmark of type I and type II diabetes, affecting both β-cell mass and insulin secretion.1 Type I diabetes is characterized by drastic decreases in β-cell mass and insulin secretion, in part mediated by proinflammatory cytokines produced following autoimmune activation.1 Proinflammatory cytokines, particularly interleukin-1β (IL-1β), in combination with interferon-γ (IFN-γ) and/or tumor necrosis factor-α (TNF-α), promote death by apoptosis and decrease function of differentiated β-cells, leading to β-cell destruction.1 Pancreatic islet transplantation is a promising alternative therapy for some type I diabetic patients.2 However, clinical outcome is not always achieved because of significant loss of islet mass during and after transplantation.3 Up to 80% of transplanted islets can die during the post-transplantation period as a result of apoptosis because of several mechanisms, notably the instant blood-mediated inflammatory response (IBMIR) and the release of a mix of cytokines including IL-1β, TNF-α and IFN-γ.4Immune-modulatory strategies for type I diabetes therapy and improvement of islet transplantation outcomes have emerged, targeting a single specific cytokine, such as IL-1β or TNF-α.2, 5 However, these strategies may only target inflammation partially.2 Indeed, multiple cytokines, originating from surrounding immune cells and/or β-cells themselves, are more likely to be present simultaneously4, 6 and act synergistically to induce β-cell death and dysfunction.7, 8, 9 Preclinical and clinical studies demonstrated that glucagon-like peptide-1 (GLP-1) analogs, in addition to regulating glucose homeostasis in vivo, contribute to the restoration of normoglycemia after islet transplantation.10, 11, 12, 13 GLP-1 receptor (GLP-1R) analogs protect β-cell survival and function from proinflammatory cytokine attack.12, 14, 15 However, some studies have shown only modest and short-term anti-inflammatory effects of GLP-1 analogs when used alone.11, 13, 16Mitogen-activated protein kinases (MAPKs) (i.e., extracellular-regulated kinase-1/2 (ERK1/2), c-Jun N-terminal kinase (JNK) and p38 MAPK) play important roles in cytokine-induced β-cell dysfunction and death.1 Conversely, ERK1/2 are involved in the beneficial effects of glucose and GLP-1 analogs.17, 18, 19 In this context, upstream protein kinases that specifically control the activation of MAPK in response to a combination of inflammatory cytokines (IL-1β, TNF-α and IFN-γ), rather than a single cytokine, may be useful targets for therapeutic interventions against pancreatic β-cell failure.The serine/threonine kinase tumor progression locus 2 (Tpl2) (also known as COT (Cancer Osaka Thyroid) in humans) is a member of the MAP3K family (the MAP3K8) whose activation stimulates primarily the ERK1/2 pathway, but also JNK and/or p38 MAPK in some cell types, specifically in response to various inflammatory stimuli.20, 21, 22 Dysregulation of Tpl2 expression and signaling is associated with acute and chronic inflammatory diseases,20, 21, 22 and several studies highlight a critical function of Tpl2 in the control of inflammatory responses and survival in adipocytes, fibroblasts and immune and epithelial cells.21, 22, 23, 24However, there is currently nothing known about the effects of Tpl2 in β-cells. The aim of this study was to determine whether Tpl2 may be a new key inflammatory regulator in β-cells or islets. We demonstrate that Tpl2 contributes to cytokine-induced β-cell apoptosis and dysfunction, and suggest that Tpl2 inhibition, either alone or combined with a GLP-1 receptor agonist, represents a potential new therapeutic strategy for the treatment of diabetes.  相似文献   

15.
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   

16.
17.
18.
Lysosomal dysfunction has been implicated both pathologically and genetically in neurodegenerative disorders, such as Alzheimer''s disease and Parkinson''s disease (PD). Lysosomal gene deficiencies cause lysosomal storage disorders, many of which involve neurodegeneration. Heterozygous mutations of some of these genes, such as GBA1, are associated with PD. CTSD is the gene encoding Cathepsin D (CTSD), a lysosomal protein hydrolase, and homozygous CTSD deficiency results in neuronal ceroid-lipofuscinosis, which is characterized by the early onset, progressive neurodegeneration. CTSD deficiency was also associated with deposition of α-synuclein aggregates, the hallmark of PD. However, whether partial deficiency of CTSD has a role in the late onset progressive neurodegenerative disorders, including PD, remains unknown. Here, we generated cell lines harboring heterozygous nonsense mutations in CTSD with genomic editing using the zinc finger nucleases. Heterozygous mutation in CTSD resulted in partial loss of CTSD activity, leading to reduced lysosomal activity. The CTSD mutation also resulted in increased accumulation of intracellular α-synuclein aggregates and the secretion of the aggregates. When α-synuclein was introduced in the media, internalized α-synuclein aggregates accumulated at higher levels in CTSD+/− cells than in the wild-type cells. Consistent with these results, transcellular transmission of α-synuclein aggregates was increased in CTSD+/− cells. The increased transmission of α-synuclein aggregates sustained during the successive passages of CTSD+/− cells. These results suggest that partial loss of CTSD activity is sufficient to cause a reduction in lysosomal function, which in turn leads to α-synuclein aggregation and propagation of the aggregates.Maintaining protein homeostasis (proteostasis) is crucial in not only maintenance of physiological functions of cells, but survival of cells. Proteostasis is a particularly important issue for the survival of post-mitotic cells, such as neurons, while dividing cells can dilute aged and misfolded proteins during the mitosis process.1, 2 For the clearance of protein burden, cells utilize two major protein degradation systems, ubiquitin proteasome system and lysosomal degradation, the latter degrades endosomal and autophagosomal cargos.3, 4, 5, 6 Dysregulation of ubiquitin proteasome system and lysosome has been shown to cause protein conformational diseases, including neurodegenerative disorders and metabolic disorders.7, 8 Genetic studies have suggested that impairment of lysosomal functions has important roles in the pathogenesis of neurodegenerative diseases. Mutations in ATP13A2, GBA1 and VPS35 have been associated with PD.9, 10, 11, 12 Mutations in progranulin and charged multivesicular body protein 2B (CHMP2B) have been identified as genetic causes of amyotrophic lateral sclerosis and frontotemporal dementia.13, 14, 15 Postmortem brain tissues of neurodegenerative diseases have exhibited deposition of endosomal and autophagic vesicles.16 Therefore, neurodegenerative proteinopathies might be attributed to lysosomal dysfunction.Pathological examinations of patient tissues have exhibited that protein aggregates, such as amyloid beta (Aβ), tau and α-synuclein aggregates, spread to larger brain regions as disease progresses.17 In animal models, intracerebrally injected α-synuclein aggregates could spread into larger brain regions both in α-synuclein transgenic and non-transgenic mice.18, 19, 20, 21 Inoculation of Aβ or tau aggregates into either non-transgenic or transgenic models of AD also exhibited propagation of those aggregates.22, 23, 24, 25, 26, 27, 28 Studies have suggested that cell-to-cell transmission of protein aggregates is the underlying mechanism of the pathological propagation.29, 30Mounting evidence have suggested that lysosomal function is important for the clearance of the transferred aggregates in recipient neurons during cell-to-cell aggregate transmission.31 This has been extensively studied in cell culture models for α-synuclein transmission. Previous studies showed α-synculein aggregates can be internalized and transported through the endolysosomal pathway.32 Lyososomal dysfunction led to increased accumulation of the internalized α-synuclein aggregates, suggesting that the lysosomal activity in recipient cells is critical in the clearance of the transmitted α-synuclein aggregates.32, 33Lysosomal storage diseases (LSDs) are caused by defects in the lysosomal degradation process. Mutations in genes encoding lysosomal catabolic enzymes and transporters manifest excessive deposition of the enzyme substrates in various organs.34 Though different LSDs show different symptoms, most of LSD patients exhibit neurological symptoms such as mental retardation, motor dysfunction and progressive neurodegeneration, as well as specific pathological changes in the nervous system.35, 36 In addition, some of progressive neurodegenerative disorders such as AD, PD and Huntington''s disease also show similar pathological features with LSD: accumulations of endosomal and autophagosomal vesicles and undegraded macromolecules, and inflammatory responses in brain.16Gaucher''s disease (GD) is the most common LSD, which is inherited in an autosomal recessive manner. Homozygous mutations of GBA1 gene, encoding β-glucocerebrosidase 1 (GCase 1), a lysosomal hydrolase, is responsible for GD.37 Evidence has suggested that GD is closely related to PD. Patients with type-1 GD, the most common form of GD, frequently develop parkinsonism.38 Heterozygous carriers of GBA1 mutations are at a higher risk for PD.39, 40 It has been shown that about 75% of Lewy bodies, a pathological hallmark of PD, colocalized with GCase 1 in brains of PD and DLB patients with heterozygous GBA1 mutations.41 These results suggest that lysosomal enzyme deficiency is associated with the development of PD.Cathepsin D (CTSD) is a major lysosomal endopeptidase, which is critical in the degradation of long-lived proteins.42 Genetic and clinical studies have shown that the homozygous deficiency of CTSD results in the early onset, progressive neurodegeneration, such as congenital neuronal ceroid-lipofuscinosis.43 The heterozygous missense mutations in CTSD have been known to cause the early onset motor and visual problems, brain atrophy, and progressive psychomotor symptoms.44 However, the effects of CTSD deficiency on the late onset progressive neurodegenerative disorders, including AD and PD, remain unclear. Nevertheless, it has become clear that CTSD activity is crucial in the degradation of pathogenic protein aggregates.45, 46Herein, we generated a cell line with a heterozygous nonsense mutation in CTSD and investigated the roles of the CTSD activity in lysosomal function, α-synuclein aggregation and transcellular transmission of α-synuclein aggregates.  相似文献   

19.
Transforming growth factor-β1 (TGF-β1) is an important regulator of fibrogenesis in heart disease. In many other cellular systems, TGF-β1 may also induce autophagy, but a link between its fibrogenic and autophagic effects is unknown. Thus we tested whether or not TGF-β1-induced autophagy has a regulatory function on fibrosis in human atrial myofibroblasts (hATMyofbs). Primary hATMyofbs were treated with TGF-β1 to assess for fibrogenic and autophagic responses. Using immunoblotting, immunofluorescence and transmission electron microscopic analyses, we found that TGF-β1 promoted collagen type Iα2 and fibronectin synthesis in hATMyofbs and that this was paralleled by an increase in autophagic activation in these cells. Pharmacological inhibition of autophagy by bafilomycin-A1 and 3-methyladenine decreased the fibrotic response in hATMyofb cells. ATG7 knockdown in hATMyofbs and ATG5 knockout (mouse embryonic fibroblast) fibroblasts decreased the fibrotic effect of TGF-β1 in experimental versus control cells. Furthermore, using a coronary artery ligation model of myocardial infarction in rats, we observed increases in the levels of protein markers of fibrosis, autophagy and Smad2 phosphorylation in whole scar tissue lysates. Immunohistochemistry for LC3β indicated the localization of punctate LC3β with vimentin (a mesenchymal-derived cell marker), ED-A fibronectin and phosphorylated Smad2. These results support the hypothesis that TGF-β1-induced autophagy is required for the fibrogenic response in hATMyofbs.Interstitial fibrosis is common to many cardiovascular disease etiologies including myocardial infarction (MI),1 diabetic cardiomyopathy2 and hypertension.3 Fibrosis may arise due to maladaptive cardiac remodeling following injury and is a complex process resulting from activation of signaling pathways, such as TGF-β1.4 TGF-β1 signaling has broad-ranging effects that may affect cell growth, differentiation and the production of extracellular matrix (ECM) proteins.5, 6 Elevated TGF-β1 is observed in post-MI rat heart7 and is associated with fibroblast-to-myofibroblast phenoconversion and concomitant activation of canonical Smad signaling.8 The result is a proliferation of myofibroblasts, which then leads to inappropriate deposition of fibrillar collagens, impaired cardiac function and, ultimately, heart failure.9, 10Autophagy is necessary for cellular homeostasis and is involved in organelle and protein turnover.11, 12, 13, 14 Autophagy aids in cell survival by providing primary materials, for example, amino acids and fatty acids for anabolic pathways during starvation conditions.15, 16 Alternatively, autophagy may be associated with apoptosis through autodigestive cellular processes, cellular infection with pathogens or extracellular stimuli.17, 18, 19, 20 The overall control of cardiac fibrosis is likely due to the complex functioning of an array of regulatory factors, but to date, there is little evidence linking autophagy with fibrogenesis in cardiac tissue.11, 12, 13, 14, 15, 16, 17, 18, 21, 22Recent studies have demonstrated that TGF-β1 may not only promote autophagy in mouse fibroblasts and human tubular epithelial kidney cells15, 23, 24 but can also inhibit this process in fibroblasts extracted from human patients with idiopathic pulmonary fibrosis.25 Moreover, it has recently been reported that autophagy can negatively15 and positively25, 26, 27 regulate the fibrotic process in different model cell systems. In this study, we have explored the putative link between autophagy and TGF-β1-induced fibrogenesis in human atrial myofibroblasts (hATMyofbs) and in a model of MI rat heart.  相似文献   

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