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.
In the oxidative stress hypothesis of aging, the aging process is the result of cumulative damage by reactive oxygen species. Humans and chimpanzees are remarkably similar; but humans live twice as long as chimpanzees and therefore are believed to age at a slower rate. The purpose of this study was to compare biomarkers for cardiovascular disease, oxidative stress, and aging between male chimpanzees and humans. Compared with men, male chimpanzees were at increased risk for cardiovascular disease because of their significantly higher levels of fibrinogen, IGF1, insulin, lipoprotein a, and large high-density lipoproteins. Chimpanzees showed increased oxidative stress, measured as significantly higher levels of 5-hydroxymethyl-2-deoxyuridine and 8-iso-prostaglandin F, a higher peroxidizability index, and higher levels of the prooxidants ceruloplasmin and copper. In addition, chimpanzees had decreased levels of antioxidants, including α- and β-carotene, β-cryptoxanthin, lycopene, and tocopherols, as well as decreased levels of the cardiovascular protection factors albumin and bilirubin. As predicted by the oxidative stress hypothesis of aging, male chimpanzees exhibit higher levels of oxidative stress and a much higher risk for cardiovascular disease, particularly cardiomyopathy, compared with men of equivalent age. Given these results, we hypothesize that the longer lifespan of humans is at least in part the result of greater antioxidant capacity and lower risk of cardiovascular disease associated with lower oxidative stress.Abbreviations: 5OHmU, 5-hydroxymethyl-2-deoxyuridine; 8isoPGF, 8-iso-prostaglandin F; HDL, high-density lipoprotein; IGF1, insulin-like growth factor 1; LDL, low-density lipoprotein; ROS, reactive oxygen speciesAging is characterized as a progressive reduction in the capacity to withstand the stresses of everyday life and a corresponding increase in risk of mortality. According to the oxidative stress hypothesis of aging, much of the aging process can be accounted for as the result of cumulative damage produced by reactive oxygen species (ROS).6,21,28,41,97 Endogenous oxygen radicals (that is, ROS) are generated as a byproduct of normal metabolic reactions in the body and subsequently can cause extensive damage to proteins, lipids, and DNA.6,41 Various prooxidant elements, in particular free transition metals, can catalyze these destructive reactions.6 The damage caused by ROS can be counteracted by antioxidant defense systems, but the imbalance between production of ROS and antioxidant defenses, over time, leads to oxidative stress and may contribute to the rate of aging.28,97Oxidative stress has been linked to several age-related diseases including neurodegenerative diseases, ophthalmologic diseases, cancer, and cardiovascular disease.21,28,97 Of these, cardiovascular disease remains the leading cause of adult death in the United States and Europe.71 In terms of cardiovascular disease, oxidative stress has been linked to atherosclerosis, hypertension, cardiomyopathy, and chronic heart failure in humans.55,78,84 Increases in oxidant catalysts (prooxidants)—such as copper, iron, and cadmium—have been associated with hypertension, coronary artery disease, atherosclerosis, and sudden cardiac death.98,102,106 Finally, both endogenous and exogenous antioxidants have been linked to decreased risk of cardiovascular disease, although the mechanisms behind this relationship are unclear.11,52,53 However, the oxidative stress hypothesis of aging aims to explain not only the mechanism of aging and age-related diseases (such as cardiovascular disease) in humans but also the differences between aging rates and the manifestations of age-related diseases across species.The differences in antioxidant and ROS levels between animals and humans offer promise for increasing our understanding of human aging. Additional evidence supporting the oxidative stress hypothesis of aging has come from comparative studies linking differences in aging rates across taxa with both antioxidant and ROS levels.4,17-21,58,71,86,105 In mammals, maximum lifespan potential is positively correlated with both serum and tissue antioxidant levels.17,18,21,71,105 Research has consistently demonstrated that the rate of oxidative damage varies across species and is negatively correlated with maximum lifespan potential.4,19,20,58,71,86 However, few studies involved detailed comparisons of hypothesized biochemical indicators of aging and oxidative stress between humans and animals.6 This type of interspecies comparison has great potential for directly testing the oxidative stress hypothesis of aging.Much evolutionary and genetic evidence supports remarkable similarity between humans and chimpanzees.95,100 Despite this similarity, humans have a lifespan of almost twice that of chimpanzees.3,16,47 Most comparative primate aging research has focused on the use of a macaque model,62,81,88 and several biochemical markers of age-related diseases have been identified in both humans and macaque monkeys.9,22,28,81,93,97 Several other species of monkeys have also been used in research addressing oxidative stress, antioxidant defenses, and maximum lifespan potential.18,21,58,105 However, no study to date has examined biochemical indicators of oxidative stress and aging in chimpanzees and humans as a test of the oxidative stress hypothesis for aging. The purpose of this study is to compare biochemical markers for cardiovascular disease, oxidative stress, and aging directly between male chimpanzees and humans. Given the oxidative stress hypothesis for aging and the known role of oxidative stress in cardiovascular disease, we predict that chimpanzees will show higher levels of cardiovascular risk and oxidative stress than humans.  相似文献   

2.
Multivesicular bodies (MVBs) are endocytic compartments that enclose intraluminal vesicles (ILVs) formed by inward budding from the limiting membrane of endosomes. In T lymphocytes, these ILV contain Fas ligand (FasL) and are secreted as ''lethal exosomes'' following activation-induced fusion of the MVB with the plasma membrane. Diacylglycerol (DAG) and diacylglycerol kinase α (DGKα) regulate MVB maturation and polarized traffic, as well as subsequent secretion of pro-apoptotic exosomes, but the molecular basis underlying these phenomena remains unclear. Here we identify protein kinase D (PKD) family members as DAG effectors involved in MVB genesis and secretion. We show that the inducible secretion of exosomes is enhanced when a constitutively active PKD1 mutant is expressed in T lymphocytes, whereas exosome secretion is impaired in PKD2-deficient mouse T lymphoblasts and in PKD1/3-null B cells. Analysis of PKD2-deficient T lymphoblasts showed the presence of large, immature MVB-like vesicles and demonstrated defects in cytotoxic activity and in activation-induced cell death. Using pharmacological and genetic tools, we show that DGKα regulates PKD1/2 subcellular localization and activation. Our studies demonstrate that PKD1/2 is a key regulator of MVB maturation and exosome secretion, and constitutes a mediator of the DGKα effect on MVB secretory traffic.Exosomes are nanovesicles that form as intraluminal vesicles (ILVs) inside multivesicular bodies (MVBs) and are then secreted by numerous cell types.1 ILVs are generated by inward budding of late endosome limiting membrane in a precisely regulated maturation process.2, 3 Two main pathways are involved in MVB maturation.4, 5 In addition to the ESCRT (endosomal complex required for traffic) proteins,6 there is increasing evidence that lipids such as lyso-bisphosphatidic acid (LBPA),7 ceramides8 and diacylglycerol (DAG)9 contribute to this membrane invagination process.Exosomes participate in many biological processes related to T-cell receptor (TCR)-triggered immune responses, including T lymphocyte-mediated cytotoxicity and activation-induced cell death (AICD), antigen presentation and intercellular miRNA exchange.10, 11, 12, 13, 14, 15 The discovery of exosome involvement in these responses increased interest in the regulation of exosome biogenesis and secretory traffic, with special attention to the contribution of lipids such as ceramide and DAG, as well as DAG-binding proteins.14, 16, 17, 18, 19, 20, 21 These studies suggest that positive and negative DAG regulators may control secretory traffic. By transforming DAG into phosphatidic acid (PA), diacylglycerol kinase α (DGKα) is essential for the negative control of DAG function in T lymphocytes.22 DGKα translocates transiently to the T-cell membrane after human muscarinic type 1 receptor (HM1R) triggering or to the immune synapse (IS) after TCR stimulation; at these subcellular locations, DGKα acts as a negative modulator of phospholipase C (PLC)-generated DAG.23, 24The secretory vesicle pathway involves several DAG-controlled checkpoints at which DGKα may act; these include vesicle formation and fission at the trans-Golgi network (TGN), MVB maturation, as well as their transport, docking and fusion to the plasma membrane.9, 16, 17, 18, 19, 20 The molecular components that regulate some of these trafficking processes include protein kinase D (PKD) family members.21 PKD1 activity, for instance, regulates fission of transport vesicles from TGN via direct interaction with the pre-existing DAG pool at this site.19 The cytosolic serine/threonine kinases PKD1, PKD2 and PKD3(ref. 21) are expressed in a wide range of cells, with PKD2 the most abundant isotype in T lymphocytes.25, 26 PKD have two DAG-binding domains (C1a and C1b) at the N terminus,21 which mediate PKD recruitment to cell membranes. Protein kinase C (PKC) phosphorylation at the PKD activation loop further promotes PKD autophosphorylation and activation.27Based on our previous studies showing DGKα regulation of DAG in MVB formation and exosome secretion,9, 14, 28 and the identification of PKD1/2 association to MVB,14 we hypothesized that DGKα control of DAG mediates these events, at least in part, through PKD. Here we explored whether, in addition to its role in vesicle fission from TGN,19 PKD regulates other steps in the DAG-controlled secretory traffic pathway. Using PKD-deficient cell models, we analyzed the role of PKD1/2 in MVB formation and function, and demonstrate their implication in exosome secretory traffic.  相似文献   

3.
Necroptosis is a form of regulated necrotic cell death mediated by receptor-interacting serine/threonine-protein kinase 1 (RIPK1) and RIPK3. Necroptotic cell death contributes to the pathophysiology of several disorders involving tissue damage, including myocardial infarction, stroke and ischemia-reperfusion injury. However, no inhibitors of necroptosis are currently in clinical use. Here we performed a phenotypic screen for small-molecule inhibitors of tumor necrosis factor-alpha (TNF)-induced necroptosis in Fas-associated protein with death domain (FADD)-deficient Jurkat cells using a representative panel of Food and Drug Administration (FDA)-approved drugs. We identified two anti-cancer agents, ponatinib and pazopanib, as submicromolar inhibitors of necroptosis. Both compounds inhibited necroptotic cell death induced by various cell death receptor ligands in human cells, while not protecting from apoptosis. Ponatinib and pazopanib abrogated phosphorylation of mixed lineage kinase domain-like protein (MLKL) upon TNF-α-induced necroptosis, indicating that both agents target a component upstream of MLKL. An unbiased chemical proteomic approach determined the cellular target spectrum of ponatinib, revealing key members of the necroptosis signaling pathway. We validated RIPK1, RIPK3 and transforming growth factor-β-activated kinase 1 (TAK1) as novel, direct targets of ponatinib by using competitive binding, cellular thermal shift and recombinant kinase assays. Ponatinib inhibited both RIPK1 and RIPK3, while pazopanib preferentially targeted RIPK1. The identification of the FDA-approved drugs ponatinib and pazopanib as cellular inhibitors of necroptosis highlights them as potentially interesting for the treatment of pathologies caused or aggravated by necroptotic cell death.Programmed cell death has a crucial role in a variety of biological processes ranging from normal tissue development to diverse pathological conditions.1, 2 Necroptosis is a form of regulated cell death that has been shown to occur during pathogen infection or sterile injury-induced inflammation in conditions where apoptosis signaling is compromised.3, 4, 5, 6 Given that many viruses have developed strategies to circumvent apoptotic cell death, necroptosis constitutes an important, pro-inflammatory back-up mechanism that limits viral spread in vivo.7, 8, 9 In contrast, in the context of sterile inflammation, necroptotic cell death contributes to disease pathology, outlining potential benefits of therapeutic intervention.10 Necroptosis can be initiated by death receptors of the tumor necrosis factor (TNF) superfamily,11 Toll-like receptor 3 (TLR3),12 TLR4,13 DNA-dependent activator of IFN-regulatory factors14 or interferon receptors.15 Downstream signaling is subsequently conveyed via RIPK116 or TIR-domain-containing adapter-inducing interferon-β,8, 17 and converges on RIPK3-mediated13, 18, 19, 20 activation of MLKL.21 Phosphorylated MLKL triggers membrane rupture,22, 23, 24, 25, 26 releasing pro-inflammatory cellular contents to the extracellular space.27 Studies using the RIPK1 inhibitor necrostatin-1 (Nec-1) 28 or RIPK3-deficient mice have established a role for necroptosis in the pathophysiology of pancreatitis,19 artherosclerosis,29 retinal cell death,30 ischemic organ damage and ischemia-reperfusion injury in both the kidney31 and the heart.32 Moreover, allografts from RIPK3-deficient mice are better protected from rejection, suggesting necroptosis inhibition as a therapeutic option to improve transplant outcome.33 Besides Nec-1, several tool compounds inhibiting different pathway members have been described,12, 16, 21, 34, 35 however, no inhibitors of necroptosis are available for clinical use so far.2, 10 In this study we screened a library of FDA approved drugs for the precise purpose of identifying already existing and generally safe chemical agents that could be used as necroptosis inhibitors. We identified the two structurally distinct kinase inhibitors pazopanib and ponatinib as potent blockers of necroptosis targeting the key enzymes RIPK1/3.  相似文献   

4.
Tumor necrosis factor α (TNFα) triggers necroptotic cell death through an intracellular signaling complex containing receptor-interacting protein kinase (RIPK) 1 and RIPK3, called the necrosome. RIPK1 phosphorylates RIPK3, which phosphorylates the pseudokinase mixed lineage kinase-domain-like (MLKL)—driving its oligomerization and membrane-disrupting necroptotic activity. Here, we show that TNF receptor-associated factor 2 (TRAF2)—previously implicated in apoptosis suppression—also inhibits necroptotic signaling by TNFα. TRAF2 disruption in mouse fibroblasts augmented TNFα–driven necrosome formation and RIPK3-MLKL association, promoting necroptosis. TRAF2 constitutively associated with MLKL, whereas TNFα reversed this via cylindromatosis-dependent TRAF2 deubiquitination. Ectopic interaction of TRAF2 and MLKL required the C-terminal portion but not the N-terminal, RING, or CIM region of TRAF2. Induced TRAF2 knockout (KO) in adult mice caused rapid lethality, in conjunction with increased hepatic necrosome assembly. By contrast, TRAF2 KO on a RIPK3 KO background caused delayed mortality, in concert with elevated intestinal caspase-8 protein and activity. Combined injection of TNFR1-Fc, Fas-Fc and DR5-Fc decoys prevented death upon TRAF2 KO. However, Fas-Fc and DR5-Fc were ineffective, whereas TNFR1-Fc and interferon α receptor (IFNAR1)-Fc were partially protective against lethality upon combined TRAF2 and RIPK3 KO. These results identify TRAF2 as an important biological suppressor of necroptosis in vitro and in vivo.Apoptotic cell death is mediated by caspases and has distinct morphological features, including membrane blebbing, cell shrinkage and nuclear fragmentation.1, 2, 3, 4 In contrast, necroptotic cell death is caspase-independent and is characterized by loss of membrane integrity, cell swelling and implosion.1, 2, 5 Nevertheless, necroptosis is a highly regulated process, requiring activation of RIPK1 and RIPK3, which form the core necrosome complex.1, 2, 5 Necrosome assembly can be induced via specific death receptors or toll-like receptors, among other modules.6, 7, 8, 9 The activated necrosome engages MLKL by RIPK3-mediated phosphorylation.6, 10, 11 MLKL then oligomerizes and binds to membrane phospholipids, forming pores that cause necroptotic cell death.10, 12, 13, 14, 15 Unchecked necroptosis disrupts embryonic development in mice and contributes to several human diseases.7, 8, 16, 17, 18, 19, 20, 21, 22The apoptotic mediators FADD, caspase-8 and cFLIP suppress necroptosis.19, 20, 21, 23, 24 Elimination of any of these genes in mice causes embryonic lethality, subverted by additional deletion of RIPK3 or MLKL.19, 20, 21, 25 Necroptosis is also regulated at the level of RIPK1. Whereas TNFα engagement of TNFR1 leads to K63-linked ubiquitination of RIPK1 by cellular inhibitor of apoptosis proteins (cIAPs) to promote nuclear factor (NF)-κB activation,26 necroptosis requires suppression or reversal of this modification to allow RIPK1 autophosphorylation and consequent RIPK3 activation.2, 23, 27, 28 CYLD promotes necroptotic signaling by deubiquitinating RIPK1, augmenting its interaction with RIPK3.29 Conversely, caspase-8-mediated CYLD cleavage inhibits necroptosis.24TRAF2 recruits cIAPs to the TNFα-TNFR1 signaling complex, facilitating NF-κB activation.30, 31, 32, 33 TRAF2 also supports K48-linked ubiquitination and proteasomal degradation of death-receptor-activated caspase-8, curbing apoptosis.34 TRAF2 KO mice display embryonic lethality; some survive through birth but have severe developmental and immune deficiencies and die prematurely.35, 36 Conditional TRAF2 KO leads to rapid intestinal inflammation and mortality.37 Furthermore, hepatic TRAF2 depletion augments apoptosis activation via Fas/CD95.34 TRAF2 attenuates necroptosis induction in vitro by the death ligands Apo2L/TRAIL and Fas/CD95L.38 However, it remains unclear whether TRAF2 regulates TNFα-induced necroptosis—and if so—how. Our present findings reveal that TRAF2 inhibits TNFα necroptotic signaling. Furthermore, our results establish TRAF2 as a biologically important necroptosis suppressor in vitro and in vivo and provide initial insight into the mechanisms underlying this function.  相似文献   

5.
Metabolic syndrome is a condition that typically includes central obesity, insulin resistance, glucose intolerance, dyslipidemia, and hypertension. Disruption of the hypothalamic–pituitary–adrenal axis, a regulator of corticosterone secretion, occurs in some cases of metabolic syndrome and obesity, and Cushing hypercortisolemia is associated with obesity and metabolic disorders. We therefore assessed anatomic and clinical pathology in C57BL/6NCrl mice to evaluate the effects of chronic corticosterone in the drinking water at doses of 25, 50, and 100 μg/mL for 25 d. Treated mice developed obesity, glucose intolerance, electrolyte aberrations, and dyslipidemia that were dose-dependent and most severe in the 100-μg/mL treatment group. To evaluate return to normal function, additional C57BL/6NCrl mice received corticosterone-free water for 2 wk after the 25-d treatment period. According to results of gross examination, mice appeared to recover within days of exogenous corticosterone withdrawal; however, adrenal gland vacuolation and protein, lipid, and electrolyte abnormalities persisted. Together, these findings support chronic corticosterone exposure through the drinking water as a potentially useful, noninvasive method to induce some features of metabolic syndrome.Obesity and associated metabolic dysfunctions are an increasing public health concern in modern Western society. In humans, obesity and metabolic syndrome heighten the risk of developing debilitating and costly illness including diabetes, cardiovascular disease, stroke, and some forms of cancer.2,20 Mounting evidence indicates that stress and associated hormones such as cortisol (corticosterone in rodents) contribute to the development of metabolic syndrome. Furthermore, regional glucocorticoid metabolism in adipocytes is proposed to be involved in the pathogenesis of metabolic syndrome.6,16,17,27,56 Cushing syndrome, iatrogenic hypercortisolemia, and metabolic syndrome share clinical and physiologic similarities, including central obesity, insulin resistance, glucose intolerance, dyslipidemia, and hypertension.1,2,31,35,41,46 How glucocorticoids contribute to the development of these problems remains unclear.Numerous clinical and experimental studies have linked stress, diet, and lifestyle choices to changes in risk factors associated with the development of metabolic disorders.1,3,7,10,21,33,36,42,55 How corticosterone influences this risk remains unclear. Although corticosterone has beneficial short-term effects, long-term corticosterone exposure can result in damage to the physiologic systems it protects acutely.27 Disruption of this physiologic signal occurs in numerous disparate disorders, ranging from depression to Cushing syndrome.16,22,36,54 Therefore, understanding the effects of chronic high corticosterone on metabolism and physiology is of key importance.To clarify how chronic treatment with corticosterone alters the physiology of an organism, we treated adrenally intact adult male mice with corticosterone in drinking water for 4 wk. Furthermore, we examined the return of physiology 2 wk after withdrawal of chronic corticosterone administration. We used this approach as a rapid (3- to 4-wk), noninvasive method of altering plasma corticosterone levels that enabled us to retain some integrity in the diurnal rhythm present in normal animals.We previously characterized the gross metabolic consequences of exogenous noninvasive corticosterone delivery in the drinking water.20,28 In those studies, we found that high doses of corticosterone (100 μg/mL) resulted in rapid and dramatic hyperphagia; weight gain; increased adiposity; elevated plasma corticosterone, leptin, insulin, and triglyceride levels; and decreased homecage locomotion.20 Moreover, several studies have shown that a lower dose of corticosterone (25 μg/mL) resulted in an intermediate phenotype in some of these measures but had no effect on others.12,14,20,23,28,38,42,47 As such, the high corticosterone dose results in a phenotype that satisfies most of the criteria for metabolic syndrome as defined by the National Heart, Lung, and Blood Institute and the American Heart Association.15 However, little information is available on the resulting histologic, hematologic, and serum chemical profiles associated with this treatment. We sought to more fully characterize this model to support selection of the model that most accurately reflects the human disease conditions under study. In-depth characterization of the model also provides more precise measurements of response to therapies intended to ameliorate the effects of the treatment.The current study provides a detailed examination of the physiologic effect of 3 dosages of corticosterone—low (25 μg/mL), intermediate (50 μg/mL), and high (100 μg/mL) doses—in drinking water. The goal was to extend the previous findings that established this regimen as a model of metabolic syndrome by exploring the detailed physiologic changes associated with this model and to assess whether and how treated mice recover after withdrawal of the corticosterone treatment. We propose that the physiologic changes observed in the mice treated with high-dose corticosterone approximate changes observed in human patients with metabolic syndrome and that these mice potentially serve as a model for hypercortisolemia and associated obesity. In addition, we hypothesized that 2 wk of recovery from corticosterone treatment would not completely resolve cellular and clinical pathologies characterized during treatment, given the numerous changes in physiology.  相似文献   

6.
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   

7.
Glial cell line-derived neurotrophic factor (GDNF) is a potent survival and regeneration-promoting factor for dopaminergic neurons in cell and animal models of Parkinson disease (PD). GDNF is currently tested in clinical trials on PD patients with so far inconclusive results. The receptor tyrosine kinase Ret is the canonical GDNF receptor, but several alternative GDNF receptors have been proposed, raising the question of which signaling receptor mediates here the beneficial GDNF effects. To address this question we overexpressed GDNF in the striatum of mice deficient for Ret in dopaminergic neurons and subsequently challenged these mice with 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP). Strikingly, in this established PD mouse model, the absence of Ret completely abolished GDNF''s neuroprotective and regenerative effect on the midbrain dopaminergic system. This establishes Ret signaling as absolutely required for GDNF''s effects to prevent and compensate dopaminergic system degeneration and suggests Ret activation as the primary target of GDNF therapy in PD.Glial cell line-derived neurotrophic factor (GDNF) is the founding member of the four ligands in the GDNF family, which belong to the transforming growth factor-β superfamily.1 GDNF was characterized as a potent survival factor for many neurons in culture such as dopaminergic, motor, sympathetic, parasympathetic, sensory and enteric neurons.1, 2 In addition, in dopaminergic neuron cultures GDNF stimulates neuronal differentiation, neurite outgrowth, synapse formation and dopamine release.1, 2As degeneration of midbrain dopaminergic neurons in the substantia nigra pars compacta (SNpc) represents a major hallmark of Parkinson disease (PD), the most common neurodegenerative movement disorder, GDNF has raised considerable interest as a therapeutic molecule for the treatment of PD.3, 4, 5 PD affects >2% of individuals over the age of 60 years, but no curative treatment is available to date, mainly due to a lack of understanding disease etiology.6, 7, 8 Preclinical studies in the established 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) and 6-hydroxydopamine (6-OHDA) rodent and primate models of PD demonstrated a substantial neuroprotection and regeneration effect by striatal provided GDNF or its close relative neurturin.3, 4, 9 However, clinical phase II trials on PD patients using GDNF or neurturin did so far not convincingly recapitulate their beneficial effects on the dopaminergic system in humans most likely due to technical problems and the selection of advanced PD patients.10, 11, 12, 13GDNF signaling is highly complex as this neurotrophic factor can bind to a variety of receptors, thus being able to induce pleiotropic effects. GDNF efficiently binds to the GPI-linked GDNF family receptor α1 (GFRα1).1, 2 It has been shown that the GDNF/GFRα1 complex can activate not only the canonical GDNF receptor Ret, a receptor tyrosine kinase which signals through the sarcoma protein (Src)/rat sarcoma (Ras)/mitogen-activated protein kinase (MAPK), phosphatidylinositol-4,5-bisphosphate 3-kinase (PI3K)/Akt, NF-κB (nuclear factor ''kappa-light-chain-enhancer'' of activated B cells), JNK (c-Jun N-terminal kinases) and PLCγ (phospholipase γ) pathway, but also with other signaling inducing receptors.1, 2, 4, 5, 13 So far, at least four alternative GDNF receptors have been described which are all expressed in midbrain dopaminergic neurons, NCAM,14, 15 the integrins αV and βI,14, 16 syndecan 317 and N-cadherin.18 Interestingly, Ret is not essential during pre- and postnatal development of the mouse dopaminergic system,19, 20, 21, 22, 23 but specifically required for the maintenance of SNpc dopaminergic neurons and their striatal innervation in aged mice.23, 24, 25 In contrast, GDNF seems most likely under physiological conditions to be dispensable during development and maintenance of midbrain dopaminergic neurons in mice, although conflicting results exist.26, 27, 28 Thus, Ret might be activated by a GDNF-independent mechanism to stimulate SNpc dopaminergic neuron survival. In addition, the in vivo function of the alternative GDNF receptors in the dopaminergic system under physiological and pathophysiological conditions, like PD, and their dependence on GDNF has not yet been addressed in detail. This raised the important question which GDNF receptor might be required to mediate GDNF''s reported neuroprotective and regenerative effect in the dopaminergic system in PD animal models and potentially in PD patients.5, 29Previously, we showed in dopaminergic neuron-specific Ret knockout mice that Ret receptor loss does not result in a higher vulnerability of midbrain dopaminergic neurons against MPTP but to less resprouting of left over dopaminergic neuron axons in the striatum after MPTP intoxication.30 In adult mice endogenous GDNF levels are rather low.26, 31 Therefore, we could not rule out in that study the possibility, that higher levels of GDNF—as also used in the clinical GDNF trials in PD patients—might have neuroprotective and regenerating effects even in the absence of the Ret receptor. Here we addressed now this question by viral overexpression of GDNF in MPTP-treated mice lacking expression of Ret again specifically in dopaminergic neurons.23, 30 We found that in the absence of Ret in dopaminergic neurons even a substantial overexpression of GDNF in the striatum does not have a neuroprotective and regenerative effect. Thus, despite the expression of alternative GDNF receptors on midbrain dopaminergic neurons, the presence of the canonical GDNF receptor Ret seems to be mandatory for mediating GDNF''s beneficial survival and axonal resprouting effect in these neurons.  相似文献   

8.
A 5.5-y-old intact male cynomolgus macaque (Macaca fasicularis) presented with inappetence and weight loss 57 d after heterotopic heart and thymus transplantation while receiving an immunosuppressant regimen consisting of tacrolimus, mycophenolate mofetil, and methylprednisolone to prevent graft rejection. A serum chemistry panel, a glycated hemoglobin test, and urinalysis performed at presentation revealed elevated blood glucose and glycated hemoglobin (HbA1c) levels (727 mg/dL and 10.1%, respectively), glucosuria, and ketonuria. Diabetes mellitus was diagnosed, and insulin therapy was initiated immediately. The macaque was weaned off the immunosuppressive therapy as his clinical condition improved and stabilized. Approximately 74 d after discontinuation of the immunosuppressants, the blood glucose normalized, and the insulin therapy was stopped. The animal''s blood glucose and HbA1c values have remained within normal limits since this time. We suspect that our macaque experienced new-onset diabetes mellitus after transplantation, a condition that is commonly observed in human transplant patients but not well described in NHP. To our knowledge, this report represents the first documented case of new-onset diabetes mellitus after transplantation in a cynomolgus macaque.Abbreviations: NODAT, new-onset diabetes mellitus after transplantationNew-onset diabetes mellitus after transplantation (NODAT, formerly known as posttransplantation diabetes mellitus) is an important consequence of solid-organ transplantation in humans.7-10,15,17,19,21,25-28,31,33,34,37,38,42 A variety of risk factors have been identified including increased age, sex (male prevalence), elevated pretransplant fasting plasma glucose levels, and immunosuppressive therapy.7-10,15,17,19,21,25-28,31,33,34,37,38,42 The relationship between calcineurin inhibitors, such as tacrolimus and cyclosporin, and the development of NODAT is widely recognized in human medicine.7-10,15,17,19,21,25-28,31,33,34,37,38,42 Cynomolgus macaques (Macaca fasicularis) are a commonly used NHP model in organ transplantation research. Cases of natural and induced diabetes of cynomolgus monkeys have been described in the literature;14,43,45 however, NODAT in a macaque model of solid-organ transplantation has not been reported previously to our knowledge.  相似文献   

9.
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   

10.
In the central nervous system (CNS), hyperglycemia leads to neuronal damage and cognitive decline. Recent research has focused on revealing alterations in the brain in hyperglycemia and finding therapeutic solutions for alleviating the hyperglycemia-induced cognitive dysfunction. Adiponectin is a protein hormone with a major regulatory role in diabetes and obesity; however, its role in the CNS has not been studied yet. Although the presence of adiponectin receptors has been reported in the CNS, adiponectin receptor-mediated signaling in the CNS has not been investigated. In the present study, we investigated adiponectin receptor (AdipoR)-mediated signaling in vivo using a high-fat diet and in vitro using neural stem cells (NSCs). We showed that AdipoR1 protects cell damage and synaptic dysfunction in the mouse brain in hyperglycemia. At high glucose concentrations in vitro, AdipoR1 regulated the survival of NSCs through the p53/p21 pathway and the proliferation- and differentiation-related factors of NSCs via tailless (TLX). Hence, we suggest that further investigations are necessary to understand the cerebral AdipoR1-mediated signaling in hyperglycemic conditions, because the modulation of AdipoR1 might alleviate hyperglycemia-induced neuropathogenesis.Adiponectin secreted by the adipose tissue1, 2 exists in either a full-length or globular form.3, 4, 5, 6 Adiponectin can cross the blood–brain barrier, and various forms of adiponectin are found in the cerebrospinal fluid.7, 8, 9, 10, 11 Adiponectin exerts its effect by binding to the adiponectin receptor 1 (AdipoR1) and adiponectin receptor 2 (AdipoR2)12, 13 that have different affinities for the various circulating adiponectins.12, 14, 15, 16, 17 Several studies reported that both receptor subtypes are expressed in the central nervous system (CNS).7, 12, 18 As adiponectin modulates insulin sensitivity and inflammation,19 its deficiency induces insulin resistance and glucose intolerance in animals fed a high-fat diet (HFD).19, 20, 21 In addition, adiponectin can ameliorate the glucose homeostasis and increase insulin sensitivity.22, 23, 24 Adiponectin, which is the most well-known adipokine, acts mainly as an anti-inflammatory regulator,25, 26 and is associated with the onset of neurological disorders.27 In addition, a recent study reported that adiponectin promotes the proliferation of hippocampal neural stem cells (NSCs).28 Considering that adiponectin acts by binding to the adiponectin receptors, investigation of the adiponectin receptor-mediated signaling in the brain is crucial to understand the cerebral effects of adiponectin and the underlying cellular mechanisms.The prevalence of type II diabetes mellitus (DM2) and Alzheimer''s disease increases with aging.29 According to a cross-sectional study, in people with DM2, the risk of dementia is 2.5 times higher than that in the normal population.30, 31 A study performed between 1980 and 2002 suggested that an elevated blood glucose level is associated with a greater risk for dementia in elderly patients with DM2.32 In addition, according to a 9-year-long longitudinal cohort study, the risk of developing Alzheimer''s disease was 65% higher in people with diabetes than in control subjects.33 A community-based cohort study also reported that higher plasma glucose concentrations are associated with an increased risk for dementia, because the higher glucose level has detrimental effects on the brain.31 High blood glucose level causes mitochondria-dependent apoptosis,34, 35, 36 and aggravates diverse neurological functions.37, 38 Inflammation and oxidative stress, which are commonly observed in people with diabetes, inhibit neurogenesis.39, 40, 41 Similarly, neurogenesis is decreased in mice and rats with genetically induced type I diabetes.42, 43 In addition, diabetic rodents have a decreased proliferation rate of neural progenitors.43, 44 Furthermore, several studies suggested that an HFD leads to neuroinflammation, the impairment of synaptic plasticity, and cognitive decline.45, 46Here, we investigated whether AdipoR1-mediated signaling is associated with cell death in the brain of mice on a HFD, and whether high glucose level modifies the proliferation and differentiation capacity of NSCs in vitro. Our study provides novel findings about the role of AdipoR1-mediated signaling in hyperglycemia-induced neuropathogenesis.  相似文献   

11.
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   

12.
13.
Overgrowth of white adipose tissue (WAT) in obesity occurs as a result of adipocyte hypertrophy and hyperplasia. Expansion and renewal of adipocytes relies on proliferation and differentiation of white adipocyte progenitors (WAP); however, the requirement of WAP for obesity development has not been proven. Here, we investigate whether depletion of WAP can be used to prevent WAT expansion. We test this approach by using a hunter-killer peptide designed to induce apoptosis selectively in WAP. We show that targeted WAP cytoablation results in a long-term WAT growth suppression despite increased caloric intake in a mouse diet-induced obesity model. Our data indicate that WAP depletion results in a compensatory population of adipose tissue with beige adipocytes. Consistent with reported thermogenic capacity of beige adipose tissue, WAP-depleted mice display increased energy expenditure. We conclude that targeting of white adipocyte progenitors could be developed as a strategy to sustained modulation of WAT metabolic activity.Obesity, a medical condition predisposing to diabetes, cardiovascular diseases, cancer, and complicating other life-threatening diseases, is becoming an increasingly important social problem.1, 2, 3 Development of pharmacological approaches to reduction of body fat has remained a daunting task.4 Approved obesity treatments typically produce only moderate and temporary effects.2,5 White adipocytes are the differentiated cells of white adipose tissue (WAT) that store triglycerides in lipid droplets.6,7 In contrast, adipocytes of brown adipose tissue (BAT) dissipate excess energy through adaptive thermogenesis. Under certain conditions, white adipocytes can become partially replaced with brown-like ‘beige'' (‘brite'') adipocytes that simulate the thermogenic function of BAT adipocytes.7,8 Obesity develops in the context of positive energy balance as a result of hypertrophy and hyperplasia of white adipocytes.9Expansion and renewal of the white adipocyte pool in WAT continues in adulthood.10,11 This process is believed to rely on proliferation and self-renewal of mesenchymal precursor cells12 that we term white adipocyte progenitors (WAPs). WAPs reside within the population of adipose stromal cells (ASCs)13 and are functionally similar to bone marrow mesenchymal stem cells (MSCs).14, 15, 16 ASCs can be isolated from the stromal/vascular fraction (SVF) of WAT based on negativity for hematopoietic (CD45) and endothelial (CD31) markers.17,18 ASCs support vascularization as mural/adventitial cells secreting angiogenic factors5,19 and, unlike bone marrow MSCs, express CD34.19,20 WAPs have been identified within the ASC population based on expression of mesenchymal markers, such as platelet-derived growth factor receptor-β (PDGFRβ, aka CD140b) and pericyte markers.17,18 Recently, a distinct ASC progenitor population capable of differentiating into both white and brown adipocytes has been identified in WAT based on PDGFRα (CD140a) expression and lack of PDGFRβ expression.21,22 The physiological relevance of the two precursor populations residing in WAT has not been explored.We have previously established an approach to isolate peptide ligands binding to receptors selectively expressed on the surface of cell populations of interest.23, 24, 25, 26, 27 Such cell-targeted peptides can be used for targeted delivery of experimental therapeutic agents in vivo. A number of ‘hunter-killer'' peptides28 composed of a cell-homing domain binding to a surface marker and of KLAKLAK2 (sequence KLAKLAKKLAKLAK), a moiety inducing apoptosis upon receptor-mediated internalization, has been described by our group.26,29 Such bimodal peptides have been used for depletion of malignant cells and organ-specific endothelial cells in preclinical animal models.26,30,31 Recently, we isolated a cyclic peptide WAT7 (amino acid sequence CSWKYWFGEC) based on its specific binding to ASCs.20 We identified Δ-decorin (ΔDCN), a proteolytic cleavage fragment of decorin, as the WAT7 receptor specifically expressed on the surface of CD34+PDGFRβ+CD31-CD45- WAPs and absent on MSCs in other organs.20Here, we investigated whether WAPs are required for obesity development in adulthood. By designing a new hunter-killer peptide that directs KLAKLAK2 to WAPs through WAT7/ΔDCN interaction, we depleted WAP in the mouse diet-induced obesity model. We demonstrate that WAP depletion suppresses WAT growth. We show that, in response to WAP deficiency, WAT becomes populated with beige adipocytes. Consistent with the reported thermogenic function of beige adipocytes,32,33 the observed WAT remodeling is associated with increased energy expenditure. We identify a population of PDGFRα-positive, PDGFRβ-negative ASCs reported recently22 as a population surviving WAP depletion and responsible for WAT browning.  相似文献   

14.
Neutral sphingomyelinase (nSMase) activation in response to environmental stress or inflammatory cytokine stimuli generates the second messenger ceramide, which mediates the stress-induced apoptosis. However, the signaling pathways and activation mechanism underlying this process have yet to be elucidated. Here we show that the phosphorylation of nSMase1 (sphingomyelin phosphodiesterase 2, SMPD2) by c-Jun N-terminal kinase (JNK) signaling stimulates ceramide generation and apoptosis and provide evidence for a signaling mechanism that integrates stress- and cytokine-activated apoptosis in vertebrate cells. An nSMase1 was identified as a JNK substrate, and the phosphorylation site responsible for its effects on stress and cytokine induction was Ser-270. In zebrafish cells, the substitution of Ser-270 for alanine blocked the phosphorylation and activation of nSMase1, whereas the substitution of Ser-270 for negatively charged glutamic acid mimicked the effect of phosphorylation. The JNK inhibitor SP600125 blocked the phosphorylation and activation of nSMase1, which in turn blocked ceramide signaling and apoptosis. A variety of stress conditions, including heat shock, UV exposure, hydrogen peroxide treatment, and anti-Fas antibody stimulation, led to the phosphorylation of nSMase1, activated nSMase1, and induced ceramide generation and apoptosis in zebrafish embryonic ZE and human Jurkat T cells. In addition, the depletion of MAPK8/9 or SMPD2 by RNAi knockdown decreased ceramide generation and stress- and cytokine-induced apoptosis in Jurkat cells. Therefore the phosphorylation of nSMase1 is a pivotal step in JNK signaling, which leads to ceramide generation and apoptosis under stress conditions and in response to cytokine stimulation. nSMase1 has a common central role in ceramide signaling during the stress and cytokine responses and apoptosis.The sphingomyelin pathway is initiated by the hydrolysis of sphingomyelin to generate the second messenger ceramide.1 Sphingomyelin hydrolysis is a major pathway for stress-induced ceramide generation. Neutral sphingomyelinase (nSMase) is activated by a variety of environmental stress conditions, such as heat shock,1, 2, 3 oxidative stress (hydrogen peroxide (H2O2), oxidized lipoproteins),1 ultraviolet (UV) radiation,1 chemotherapeutic agents,4 and β-amyloid peptides.5, 6 Cytokines, including tumor necrosis factor (TNF)-α,7, 8, 9 interleukin (IL)-1β,10 Fas ligand,11 and their associated proteins, also trigger the activation of nSMase.12 Membrane-bound Mg2+-dependent nSMase is considered to be a strong candidate for mediating the effects of stress and inflammatory cytokines on ceramide.3Among the four vertebrate nSMases, nSMase1 (SMPD2) was the first to be cloned and is localized in the endoplasmic reticulum (ER) and Golgi apparatus.13 Several studies have focused on the potential signaling roles of nSMase1, and some reports have suggested that nSMase1 is important for ceramide generation in response to stress.5, 6, 14, 15 In addition, nSMase1 is responsible for heat-induced apoptosis in zebrafish embryonic cultured (ZE) cells, and a loss-of-function study showed a reduction in ceramide generation, caspase-3 activation, and apoptosis in zebrafish embryos.16 However, nSMase1-knockout mice showed no lipid storage diseases or abnormalities in sphingomyelin metabolism.17 Therefore, the molecular mechanisms by which nSMase1 is activated have yet to be elucidated.Environmental stress and inflammatory cytokines1, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27 stimulate stress-activated protein kinase (SAPK)/c-Jun N-terminal kinase (JNK) signaling, which involves the sequential activation of members of the mitogen-activated protein kinase (MAPK) family, including MAPK/ERK kinase kinase (MEKK)1/MAPK kinase (MKK)4, and/or SAPK/ERK kinase (SEK)1/MKK7, JNK, and c-jun. Both the JNK and sphingomyelin signaling pathways coordinately mediate the induction of apoptosis.1 However, possible crosstalk between the JNK and sphingomyelin signaling pathways has not yet been characterized. Previously, we used SDS-PAGE to determine that nSMase1 polypeptides migrated at higher molecular masses,16 suggesting that the sphingomyelin signaling pathway might cause the production of a chemically modified phosphorylated nSMase1, which is stimulated under stressed conditions in ZE cells.16 Here, we demonstrate that JNK signaling results in the phosphorylation of Ser-270 of nSMase1, which initiates ceramide generation and apoptosis. We also provide evidence for a signaling mechanism that integrates cytokine- and stress-activated apoptosis in vertebrate cells. We studied stress-induced ceramide generation in two cell types: ZE cells and human leukemia Jurkat T-lymphoid cells. Stress-induced apoptosis has been investigated in these systems previously.16, 28  相似文献   

15.
16.
17.
Group B Streptococcus (Streptococcus agalactiae, GBS) is a gram-positive commensal and occasional opportunistic pathogen of the human vaginal, respiratory, and intestinal tracts that can cause sepsis, pneumonia, or meningitis in human neonates, infants, and immunosuppressed persons. We report here on a spontaneous outbreak of postnatal GBS-associated disease in rats. Ten of 26 (38.5%) 21- to 24-d-old rat pups died or were euthanized due to a moribund state in a colony of rats transgenic for the human diphtheria toxin receptor on a Munich–Wistar–Frömter genetic background. Four pups had intralesional coccoid bacteria in various organs without accompanying inflammation. GBS was isolated from the liver of 2 of these pups and from skin abscesses in 3 littermates. A connection with the transgene could not be established. A treatment protocol was evaluated in the remaining breeding female rats. GBS is a potentially clinically significant spontaneous infection in various populations of research rats, with some features that resemble late-onset postnatal GBS infection in human infants.Abbreviations: GBS, Group B Streptococcus; MWF, Munich Wistar Frömter; hDTR, human diphtheria toxin receptorStreptococci are gram-positive, coccoid bacteria that typically are classified according to their hemolytic capacity. α-hemolytic streptococci produce a zone of partial hemolysis that appears greenish on blood agar, whereas β-hemolytic streptococci produce a zone of complete hemolysis, and γ-hemolytic organisms produce no hemolysis on blood agar.24 The β-hemolytic streptococci are further subdivided into Lancefield groups (A through G), according to cell-wall carbohydrate antigens.24,29,39 The group B β-hemolytic Streptococcus (GBS) have been speciated as Streptococcus agalactiae.28,39 It was first isolated as a causative agent of mastitis in cattle.29 This organism has since been recognized as a cause of severe infection in human neonates.28,39 In humans, GBS is harbored asymptomatically in the maternal genitourinary tract.24,28 Infants can be infected and present with serious systemic disease in the first week of life (early-onset GBS) or from 1 wk to 3 mo of age (late-onset GBS).39 In laboratory animals, rats have been used experimentally as models for neonatal1,6,7,20,37,38,43,44,47,50,51 or adult45 GBS infection, but to our knowledge, GBS has not been associated with spontaneous disease in rats.  相似文献   

18.
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   

19.
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   

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