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.
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   

2.
Neuropeptides induce signal transduction across the plasma membrane by acting through cell-surface receptors. The dynorphins, endogenous ligands for opioid receptors, are an exception; they also produce non-receptor-mediated effects causing pain and neurodegeneration. To understand non-receptor mechanism(s), we examined interactions of dynorphins with plasma membrane. Using fluorescence correlation spectroscopy and patch-clamp electrophysiology, we demonstrate that dynorphins accumulate in the membrane and induce a continuum of transient increases in ionic conductance. This phenomenon is consistent with stochastic formation of giant (~2.7 nm estimated diameter) unstructured non-ion-selective membrane pores. The potency of dynorphins to porate the plasma membrane correlates with their pathogenic effects in cellular and animal models. Membrane poration by dynorphins may represent a mechanism of pathological signal transduction. Persistent neuronal excitation by this mechanism may lead to profound neuropathological alterations, including neurodegeneration and cell death.Neuropeptides are the largest and most diverse family of neurotransmitters. They are released from axon terminals and dendrites, diffuse to pre- or postsynaptic neuronal structures and activate membrane G-protein-coupled receptors. Prodynorphin (PDYN)-derived opioid peptides including dynorphin A (Dyn A), dynorphin B (Dyn B) and big dynorphin (Big Dyn) consisting of Dyn A and Dyn B are endogenous ligands for the κ-opioid receptor. Acting through this receptor, dynorphins regulate processing of pain and emotions, memory acquisition and modulate reward induced by addictive substances.1, 2, 3, 4 Furthermore, dynorphins may produce robust cellular and behavioral effects that are not mediated through opioid receptors.5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29 As evident from pharmacological, morphological, genetic and human neuropathological studies, these effects are generally pathological, including cell death, neurodegeneration, neurological dysfunctions and chronic pain. Big Dyn is the most active pathogenic peptide, which is about 10- to 100-fold more potent than Dyn A, whereas Dyn B does not produce non-opioid effects.16, 17, 22, 25 Big Dyn enhances activity of acid-sensing ion channel-1a (ASIC1a) and potentiates ASIC1a-mediated cell death in nanomolar concentrations30, 31 and, when administered intrathecally, induces characteristic nociceptive behavior at femtomolar doses.17, 22 Inhibition of endogenous Big Dyn degradation results in pathological pain, whereas prodynorphin (Pdyn) knockout mice do not maintain neuropathic pain.22, 32 Big Dyn differs from its constituents Dyn A and Dyn B in its unique pattern of non-opioid memory-enhancing, locomotor- and anxiolytic-like effects.25Pathological role of dynorphins is emphasized by the identification of PDYN missense mutations that cause profound neurodegeneration in the human brain underlying the SCA23 (spinocerebellar ataxia type 23), a very rare dominantly inherited neurodegenerative disorder.27, 33 Most PDYN mutations are located in the Big Dyn domain, demonstrating its critical role in neurodegeneration. PDYN mutations result in marked elevation in dynorphin levels and increase in its pathogenic non-opioid activity.27, 34 Dominant-negative pathogenic effects of dynorphins are not produced through opioid receptors.ASIC1a, glutamate NMDA (N-methyl-d-aspartate) and AMPA (α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid)/kainate ion channels, and melanocortin and bradykinin B2 receptors have all been implicated as non-opioid dynorphin targets.5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 30, 31, 35, 36 Multiplicity of these targets and their association with the cellular membrane suggest that their activation is a secondary event triggered by a primary interaction of dynorphins with the membrane. Dynorphins are among the most basic neuropeptides.37, 38 The basic nature is also a general property of anti-microbial peptides (AMPs) and amyloid peptides that act by inducing membrane perturbations, altering membrane curvature and causing pore formation that disrupts membrane-associated processes including ion fluxes across the membrane.39 The similarity between dynorphins and these two peptide groups in overall charge and size suggests a similar mode of their interactions with membranes.In this study, we dissect the interactions of dynorphins with the cell membrane, the primary event in their non-receptor actions. Using fluorescence imaging, correlation spectroscopy and patch-clamp techniques, we demonstrate that dynorphin peptides accumulate in the plasma membrane in live cells and cause a profound transient increase in cell membrane conductance. Membrane poration by endogenous neuropeptides may represent a novel mechanism of signal transduction in the brain. This mechanism may underlie effects of dynorphins under pathological conditions including chronic pain and tissue injury.  相似文献   

3.
4.
Metacaspases (MCAs) are cysteine peptidases expressed in plants, fungi and protozoa, with a caspase-like histidine–cysteine catalytic dyad, but differing from caspases, for example, in their substrate specificity. The role of MCAs is subject to debate: roles in cell cycle control, in cell death or even in cell survival have been suggested. In this study, using a Leishmania major MCA-deficient strain, we showed that L. major MCA (LmjMCA) not only had a role similar to caspases in cell death but also in autophagy and this through different domains. Upon cell death induction by miltefosine or H2O2, LmjMCA is processed, releasing the catalytic domain, which activated substrates via its catalytic dyad His/Cys and a proline-rich C-terminal domain. The C-terminal domain interacted with proteins, notably proteins involved in stress regulation, such as the MAP kinase LmaMPK7 or programmed cell death like the calpain-like cysteine peptidase. We also showed a new role of LmjMCA in autophagy, acting on or upstream of ATG8, involving Lmjmca gene overexpression and interaction of the C-terminal domain of LmjMCA with itself and other proteins. These results allowed us to propose two models, showing the role of LmjMCA in the cell death and also in the autophagy pathway, implicating different protein domains.Apoptosis is, in most cases, associated with and depends on the activation of cys-dependent peptidases, named caspases.1, 2 Once activated, initiator caspases induce a proteolytic cascade via the activation of effector caspases that ultimately cleave numerous substrates, thereby causing the typical morphological features of apoptosis.3, 4 Despite their essential role in apoptosis, caspases are also involved in non-apoptotic events, including inflammation, cell proliferation, cell differentiation5 and the cell survival process autophagy, a major catabolic process in eukaryotic cells that allows cells to survive nutrient starvation due to engulfment of a portion of the cytoplasm by a specific membrane, delivery to lysosomes or vacuoles and digestion by hydrolytic enzymes.6, 7, 8, 9, 10 Plants, fungi and protozoa are devoid of caspases but express metacaspases (MCAs).11MCAs are cysteine peptidases of the clan CD, family 14, with a caspase-like histidine–cysteine catalytic dyad.12, 13 However, besides their distant similarity to caspases,14 MCAs prefer arginine/lysine in the P1 position, whereas caspases prefer aspartic residues.15, 16 The role of MCAs in cell death is still enigmatic. For example, in the yeast Saccharomyces cerevisiae, YCA1 has a role in cell death,17, 18 whereas, although only partly dependent on its conserved catalytic cysteine, it also facilitates the removal of unfolded proteins, prolonging cellular life span.19 Similarly, some metacaspases have roles, outside of death, in stress acclimation pathways, as in Aspergillus fumigatus20 or in the unicellular planctonic organisms diatoms.21, 22 In Arabidopsis thaliana, AtMC1 is a positive regulator of cell death and a survival factor for aging plants,23 whereas AtMC2 negatively regulates cell death.24 Trypanosoma brucei TbMCA2, TbMCA3 and TbMCA5 and Leishmania major MCA are involved in cell cycle regulation.25, 26Leishmania are parasitic protozoa responsible for the neglected tropical disease leishmaniasis, transmitted to humans by the bite of the sand fly. In the insect, parasites proliferate as free-living flagellated forms called procyclic promastigotes within the midgut before differentiating into virulent metacyclic promastigotes and migrating to the proboscis.27, 28 In the mammalian host, promastigotes are taken up by macrophages and transform into amastigotes. Under a variety of stress stimuli, apoptosis-like morphological and biochemical features have been described in Leishmania, among which are cell shrinkage, chromatin condensation, DNA fragmentation or mitochondrial depolarization.29, 30, 31, 32, 33, 34, 35, 36, 37, 38 Despite the evidence of morphological and biochemical markers of cell death in dying Leishmania, very little is known about the cell death pathway and the implicated executioner proteins. Indeed, essential proteins involved in mammalian apoptosis, death receptors, small pro- and anti-apoptotic molecules and caspases, are apparently not encoded in the genome of Leishmania39 and the role of Leishmania MCA in cell death is still controversial, certain authors suggesting a role as a negative regulator of intracellular amastigote proliferation, instead of having a caspase-like role in the execution of cell death.40LmjMCA contains different domains: an N-terminal domain with a Mitochondrion Localization Signal (MLS),41 a caspase-like catalytic domain and a C-terminal proline-rich domain.41 On the basis of this domain structure, LmjMCA can be classified among the type I metacaspases,16 a subclass more generally defined in higher plants and characterized by the presence of an N-terminal prodomain and a short linker between the large and small subunits, as initiator caspases in metazoans.11 Upon induction of cell death by heat shock, H2O2 or drugs like miltefosine or curcumin, LmjMCA is processed and the catalytic domain is released,41 liberating the C-terminal domain. It was therefore interesting to investigate the functional roles of the different domains.In this report, we studied the role of L. major MCA (LmjMCA), using an MCA-deficient strain and overexpressing independently the catalytic and the C-terminal domains. The results confirmed that MCA was not essential to L. major survival. In contrast, LmjMCA processing, releasing its catalytic and C-terminal domains, induced cell death in L. major, whereas the overexpression of Lmjmca gene triggered autophagy after interaction of the C-terminal domain with itself and with other proteins, acting on or upstream of the autophagic protein ATG8.  相似文献   

5.
6.
The role of calcium-mediated signaling has been extensively studied in plant responses to abiotic stress signals. Calcineurin B-like proteins (CBLs) and CBL-interacting protein kinases (CIPKs) constitute a complex signaling network acting in diverse plant stress responses. Osmotic stress imposed by soil salinity and drought is a major abiotic stress that impedes plant growth and development and involves calcium-signaling processes. In this study, we report the functional analysis of CIPK21, an Arabidopsis (Arabidopsis thaliana) CBL-interacting protein kinase, ubiquitously expressed in plant tissues and up-regulated under multiple abiotic stress conditions. The growth of a loss-of-function mutant of CIPK21, cipk21, was hypersensitive to high salt and osmotic stress conditions. The calcium sensors CBL2 and CBL3 were found to physically interact with CIPK21 and target this kinase to the tonoplast. Moreover, preferential localization of CIPK21 to the tonoplast was detected under salt stress condition when coexpressed with CBL2 or CBL3. These findings suggest that CIPK21 mediates responses to salt stress condition in Arabidopsis, at least in part, by regulating ion and water homeostasis across the vacuolar membranes.Drought and salinity cause osmotic stress in plants and severely affect crop productivity throughout the world. Plants respond to osmotic stress by changing a number of cellular processes (Xiong et al., 1999; Xiong and Zhu, 2002; Bartels and Sunkar, 2005; Boudsocq and Lauriére, 2005). Some of these changes include activation of stress-responsive genes, regulation of membrane transport at both plasma membrane (PM) and vacuolar membrane (tonoplast) to maintain water and ionic homeostasis, and metabolic changes to produce compatible osmolytes such as Pro (Stewart and Lee, 1974; Krasensky and Jonak, 2012). It has been well established that a specific calcium (Ca2+) signature is generated in response to a particular environmental stimulus (Trewavas and Malhó, 1998; Scrase-Field and Knight, 2003; Luan, 2009; Kudla et al., 2010). The Ca2+ changes are primarily perceived by several Ca2+ sensors such as calmodulin (Reddy, 2001; Luan et al., 2002), Ca2+-dependent protein kinases (Harper and Harmon, 2005), calcineurin B-like proteins (CBLs; Luan et al., 2002; Batistič and Kudla, 2004; Pandey, 2008; Luan, 2009; Sanyal et al., 2015), and other Ca2+-binding proteins (Reddy, 2001; Shao et al., 2008) to initiate various cellular responses.Plant CBL-type Ca2+ sensors interact with and activate CBL-interacting protein kinases (CIPKs) that phosphorylate downstream components to transduce Ca2+ signals (Liu et al., 2000; Luan et al., 2002; Batistič and Kudla, 2004; Luan, 2009). In several plant species, multiple members have been identified in the CBL and CIPK family (Luan et al., 2002; Kolukisaoglu et al., 2004; Pandey, 2008; Batistič and Kudla, 2009; Weinl and Kudla, 2009; Pandey et al., 2014). Involvement of specific CBL-CIPK pair to decode a particular type of signal entails the alternative and selective complex formation leading to stimulus-response coupling (D’Angelo et al., 2006; Batistič et al., 2010).Several CBL and CIPK family members have been implicated in plant responses to drought, salinity, and osmotic stress based on genetic analysis of Arabidopsis (Arabidopsis thaliana) mutants (Zhu, 2002; Cheong et al., 2003, 2007; Kim et al., 2003; Pandey et al., 2004, 2008; D’Angelo et al., 2006; Qin et al., 2008; Tripathi et al., 2009; Held et al., 2011; Tang et al., 2012; Drerup et al., 2013; Eckert et al., 2014). A few CIPKs have also been functionally characterized by gain-of-function approach in crop plants such as rice (Oryza sativa), pea (Pisum sativum), and maize (Zea mays) and were found to be involved in osmotic stress responses (Mahajan et al., 2006; Xiang et al., 2007; Yang et al., 2008; Tripathi et al., 2009; Zhao et al., 2009; Cuéllar et al., 2010).In this report, we examined the role of the Arabidopsis CIPK21 gene in osmotic stress response by reverse genetic analysis. The loss-of-function mutant plants became hypersensitive to salt and mannitol stress conditions, suggesting that CIPK21 is involved in the regulation of osmotic stress response in Arabidopsis. These findings are further supported by an enhanced tonoplast targeting of the cytoplasmic CIPK21 through interaction with the vacuolar Ca2+ sensors CBL2 and CBL3 under salt stress condition.  相似文献   

7.
Q Xia  Q Hu  H Wang  H Yang  F Gao  H Ren  D Chen  C Fu  L Zheng  X Zhen  Z Ying  G Wang 《Cell death & disease》2015,6(3):e1702
Neuroinflammation is a striking hallmark of amyotrophic lateral sclerosis (ALS) and other neurodegenerative disorders. Previous studies have shown the contribution of glial cells such as astrocytes in TDP-43-linked ALS. However, the role of microglia in TDP-43-mediated motor neuron degeneration remains poorly understood. In this study, we show that depletion of TDP-43 in microglia, but not in astrocytes, strikingly upregulates cyclooxygenase-2 (COX-2) expression and prostaglandin E2 (PGE2) production through the activation of MAPK/ERK signaling and initiates neurotoxicity. Moreover, we find that administration of celecoxib, a specific COX-2 inhibitor, greatly diminishes the neurotoxicity triggered by TDP-43-depleted microglia. Taken together, our results reveal a previously unrecognized non-cell-autonomous mechanism in TDP-43-mediated neurodegeneration, identifying COX-2-PGE2 as the molecular events of microglia- but not astrocyte-initiated neurotoxicity and identifying celecoxib as a novel potential therapy for TDP-43-linked ALS and possibly other types of ALS.Amyotrophic lateral sclerosis (ALS) is an adult-onset neurodegenerative disease characterized by the degeneration of motor neurons in the brain and spinal cord.1 Most cases of ALS are sporadic, but 10% are familial. Familial ALS cases are associated with mutations in genes such as Cu/Zn superoxide dismutase 1 (SOD1), TAR DNA-binding protein 43 (TARDBP) and, most recently discovered, C9orf72. Currently, most available information obtained from ALS research is based on the study of SOD1, but new studies focusing on TARDBP and C9orf72 have come to the forefront of ALS research.1, 2 The discovery of the central role of the protein TDP-43, encoded by TARDBP, in ALS was a breakthrough in ALS research.3, 4, 5 Although pathogenic mutations of TDP-43 are genetically rare, abnormal TDP-43 function is thought to be associated with the majority of ALS cases.1 TDP-43 was identified as a key component of the ubiquitin-positive inclusions in most ALS patients and also in other neurodegenerative diseases such as frontotemporal lobar degeneration,6, 7 Alzheimer''s disease (AD)8, 9 and Parkinson''s disease (PD).10, 11 TDP-43 is a multifunctional RNA binding protein, and loss-of-function of TDP-43 has been increasingly recognized as a key contributor in TDP-43-mediated pathogenesis.5, 12, 13, 14Neuroinflammation, a striking and common hallmark involved in many neurodegenerative diseases, including ALS, is characterized by extensive activation of glial cells including microglia, astrocytes and oligodendrocytes.15, 16 Although numerous studies have focused on the intrinsic properties of motor neurons in ALS, a large amount of evidence showed that glial cells, such as astrocytes and microglia, could have critical roles in SOD1-mediated motor neuron degeneration and ALS progression,17, 18, 19, 20, 21, 22 indicating the importance of non-cell-autonomous toxicity in SOD1-mediated ALS pathogenesis.Very interestingly, a vital insight of neuroinflammation research in ALS was generated by the evidence that both the mRNA and protein levels of the pro-inflammatory enzyme cyclooxygenase-2 (COX-2) are upregulated in both transgenic mouse models and in human postmortem brain and spinal cord.23, 24, 25, 26, 27, 28, 29 The role of COX-2 neurotoxicity in ALS and other neurodegenerative disorders has been well explored.30, 31, 32 One of the key downstream products of COX-2, prostaglandin E2 (PGE2), can directly mediate COX-2 neurotoxicity both in vitro and in vivo.33, 34, 35, 36, 37 The levels of COX-2 expression and PGE2 production are controlled by multiple cell signaling pathways, including the mitogen-activated protein kinase (MAPK)/ERK pathway,38, 39, 40 and they have been found to be increased in neurodegenerative diseases including AD, PD and ALS.25, 28, 32, 41, 42, 43, 44, 45, 46 Importantly, COX-2 inhibitors such as celecoxib exhibited significant neuroprotective effects and prolonged survival or delayed disease onset in a SOD1-ALS transgenic mouse model through the downregulation of PGE2 release.28Most recent studies have tried to elucidate the role of glial cells in neurotoxicity using TDP-43-ALS models, which are considered to be helpful for better understanding the disease mechanisms.47, 48, 49, 50, 51 Although the contribution of glial cells to TDP-43-mediated motor neuron degeneration is now well supported, this model does not fully suggest an astrocyte-based non-cell autonomous mechanism. For example, recent studies have shown that TDP-43-mutant astrocytes do not affect the survival of motor neurons,50, 51 indicating a previously unrecognized non-cell autonomous TDP-43 proteinopathy that associates with cell types other than astrocytes.Given that the role of glial cell types other than astrocytes in TDP-43-mediated neuroinflammation is still not fully understood, we aim to compare the contribution of microglia and astrocytes to neurotoxicity in a TDP-43 loss-of-function model. Here, we show that TDP-43 has a dominant role in promoting COX-2-PGE2 production through the MAPK/ERK pathway in primary cultured microglia, but not in primary cultured astrocytes. Our study suggests that overproduction of PGE2 in microglia is a novel molecular mechanism underlying neurotoxicity in TDP-43-linked ALS. Moreover, our data identify celecoxib as a new potential effective treatment of TDP-43-linked ALS and possibly other types of ALS.  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.
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   

10.
11.
Poly(ADP-ribose) polymerase-1 (PARP1) plays a regulatory role in apoptosis, necrosis and other cellular processes after injury. Status epilepticus (SE) induces neuronal and astroglial death that show regional-specific patterns in the rat hippocampus and piriform cortex (PC). Thus, we investigated whether PARP1 regulates the differential neuronal/glial responses to pilocarpine (PILO)-induced SE in the distinct brain regions. In the present study, both CA1 and CA3 neurons showed PARP1 hyperactivation-dependent neuronal death pathway, whereas PC neurons exhibited PARP1 degradation-mediated neurodegeneration following SE. PARP1 degradation was also observed in astrocytes within the molecular layer of the dentate gyrus. PARP1 induction was detected in CA1–3-reactive astrocytes, as well as in reactive microglia within the PC. Although PARP1 inhibitors attenuated CA1–3 neuronal death and reactive gliosis in the CA1 region, they deteriorated the astroglial death in the molecular layer of the dentate gyrus and in the stratum lucidum of the CA3 region. Ex vivo study showed the similar regional and cellular patterns of PARP1 activation/degradation. Taken together, our findings suggest that the cellular-specific PARP1 activation/degradation may distinctly involve regional-specific neuronal damage, astroglial death and reactive gliosis in response to SE independently of hemodynamics.Poly(ADP-ribose) polymerase-1 (PARP1) repairs single-stranded DNA (ssDNA) breaks following various injuries. As PARP1 utilizes NAD+ to form poly(ADP-ribose) polymers (PAR) during this process, extensive PARP1 activation results in energy failure, promoting necrotic cell death because of NAD+ depletion.1, 2, 3, 4, 5, 6 Furthermore, PARP1 is a useful hallmark of apoptosis because full-length PARP1 is cleaved by the apoptotic proteases, caspase-3 and -7, into p85 and p25 fragments during apoptosis.7, 8 In contrast, the degradation of full-length PARP1 protein without cleavage into apoptotic fragments is mediated by caspase-independent ubiquitylation that plays a regulatory role in apoptosis, necrosis and other PARP1-regulated cellular processes.9, 10, 11, 12 Therefore, it is likely that the distinct profiles of PARP1 (activation, cleavage or degradation) may involve the differential cellular responses following harmful stimuli.Status epilepticus (SE) is a medical emergency with significant mortality.13 SE is a continuous seizure activity involving severe and prolonged hypoxia that induces sustained neuronal damage, astroglial death and reactive astrogliosis.14, 15, 16, 17, 18, 19, 20, 21, 22, 23 In particular, astroglial responses show regional-specific patterns following SE. Briefly, astroglial death was observed in the molecular layer of the dentate gyrus and the piriform cortex (PC) before or after neuronal death. In contrast, reactive astrogliosis was detected in other regions of the hippocampus and cortex.19, 20, 21, 22, 23, 24, 25 Based on the properties of PARP1 responses to stimuli, it is likely that PARP1 may be one of the potential molecules to involve neuronal damage and regional-specific astroglial responses to SE. In order to address this hypothesis, we first investigated the characteristics of PARP1 responses to SE in the rat hippocampus and PC. We then examined whether PARP1 regulates the neuronal/glial responses to SE, and finally whether hemodynamics involves PARP1 responses to SE using ex vivo model.  相似文献   

12.
Secondary hepatic amyloidosis in nonhuman primates carries a grave prognosis once animals become clinically ill. The purpose of this study was to establish serologic parameters that potentially could be used to identify rhesus macaques undergoing subclinical development of secondary hepatic amyloidosis. A retrospective analysis was completed by using serum biochemical profiles from 26 histologically diagnosed amyloidotic macaques evaluated at 2 stages of disease, clinical and subclinical (3 to 32 mo prior to clinical signs of disease). Standard serum biochemistry values for cases were compared with institutional age- and gender-specific references ranges by construction of 95% confidence intervals for the difference between means. In addition, 19 histologically diagnosed amyloidotic macaques and 19 age-matched controls were assayed for changes in various parameters by using routinely banked, frozen (–80 °C) sera available from clinical and subclinical time points. Clinically amyloidotic animals displayed increased levels of alkaline phosphatase, aspartate aminotransferase, lactate dehydrogenase, gamma glutamyltranspeptidase, and macrophage colony-stimulating factor and significantly decreased quantities of albumin and total cholesterol. Subclinical amyloidotic animals displayed increased levels of alkaline phosphatase, aspartate aminotransferase, lactate dehydrogenase, and serum amyloid A and decreased concentrations of albumin and total cholesterol. The serologic parameters studied indicate a temporal relationship of these factors not previously described, show a clear pattern of disease progression, and could be useful in subclinical disease detection.Abbreviations: mCSF, macrophage colony stimulating factor; SAA, serum amyloid AAmyloid is an eosinophilic substance made of insoluble fibrillar protein.32 When deposited extracellularly, amyloid causes displacement of tissue form and disruption of organ function.32 Persistent accretion of amyloid can result in organ failure and ultimately animal death.22 Clinical signs of disease depend on the tissues affected and the degree of involvement.32 Amyloidosis has been well documented in humans, other mammals, birds, and reptiles.38 In humans, amyloidosis plays a key role in many diseases, including Alzheimer disease, type II diabetes, rheumatoid arthritis, and Down syndrome.15,20,35,38Amyloidosis generally is classified into 3 categories: primary, secondary, and hereditary. Primary amyloidosis consists of the immunoglobulin- and myeloma-associated types. Secondary (reactive) amyloidosis is associated with chronic inflammation.24 Common causes of secondary amyloidosis in humans include rheumatoid arthritis, idiopathic colitis, infectious diseases, such as tuberculosis and leprosy, and malignant tumors, such as mesothelioma and Hodgkins disease.28 Hereditary amyloid syndromes are rare and include Mediterranean fever, Muckle–Wells syndrome, and familial amyloid cardiomyopathy.32,38Secondary amyloidosis is the most common form of amyloidosis in animals.38 Amyloidosis occurs in many species of nonhuman primates including the common marmoset (Callithrix jacchus),23 squirrel monkey (Saimiri sciureus),34 rhesus macaque (Macaca mulatta),9,10 pigtailed macaque (Macaca nemestrina),18,27 crab-eating macaque (Macaca fascicularis),27 barbary ape (Macaca sylvanus),6 baboon (Papio spp.),17 mandrill (Papio sphinx), and chimpanzee (Pan troglodytes).16,39 Although a definitive cause of secondary amyloidosis has not been identified in nonhuman primates, this condition has been associated with chronic inflammation due to rheumatoid arthritis,6 viral infection,18 parasitism,1 respiratory disease,27,30 trauma,30 and bacterial enterocolitis.27,30,31 Shigella spp. have received particular attention as a common etiology linking enterocolitis with amyloidosis.4,7,38Previous research on amyloidosis in nonhuman primates has yielded clinical and serologic profiles in end-stage amyloidotic animals, but little is known about the serologic status in the subclinical stages of disease. Amyloid can accumulate for as long as 3 y before severe organ disruption occurs14 and clinical signs of amyloidosis become evident.16 With appropriate analysis, detection of amyloidosis could occur much earlier than typically now achieved, thus allowing for targeted preventative therapy to potentially halt the progression of this insidious disease.  相似文献   

13.
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   

14.
15.
16.
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   

17.
Reactive gliosis is an early pathological feature common to most neurodegenerative diseases, yet its regulation and impact remain poorly understood. Normally astrocytes maintain a critical homeostatic balance. After stress or injury they undergo rapid parainflammatory activation, characterized by hypertrophy, and increased polymerization of type III intermediate filaments (IFs), particularly glial fibrillary acidic protein and vimentin. However, the consequences of IF dynamics in the adult CNS remains unclear, and no pharmacologic tools have been available to target this mechanism in vivo. The mammalian retina is an accessible model to study the regulation of astrocyte stress responses, and their influence on retinal neuronal homeostasis. In particular, our work and others have implicated p38 mitogen-activated protein kinase (MAPK) signaling as a key regulator of glutamate recycling, antioxidant activity and cytokine secretion by astrocytes and related Müller glia, with potent influences on neighboring neurons. Here we report experiments with the small molecule inhibitor, withaferin A (WFA), to specifically block type III IF dynamics in vivo. WFA was administered in a model of metabolic retinal injury induced by kainic acid, and in combination with a recent model of debridement-induced astrocyte reactivity. We show that WFA specifically targets IFs and reduces astrocyte and Müller glial reactivity in vivo. Inhibition of glial IF polymerization blocked p38 MAPK-dependent secretion of TNF-α, resulting in markedly reduced neuronal apoptosis. To our knowledge this is the first study to demonstrate that pharmacologic inhibition of IF dynamics in reactive glia protects neurons in vivo.Astrocyte reactivity (reactive gliosis) is an early pathological feature common to most neurodegenerative diseases, yet its regulation and impact remains poorly understood. In the healthy central nervous system (CNS), astrocytes coordinate homeostatic vascular perfusion, free radical detoxification and neurotransmitter recycling.1, 2 Injury or stress induces a phenotypic switch, whose cardinal features are cellular hypertrophy and increased expression and polymerization of type III intermediate filaments (IFs), particularly glial fibrillary acidic protein (GFAP).3, 4, 5 The role of intermediate filaments in reactive gliosis remains unclear.3, 6, 7, 8, 9 Genetic deletion of IFs GFAP and vimentin have been shown to promote axonal outgrowth and regeneration in developing neurons and models of CNS injury,10, 11, 12 yet result in developmental defects to inner retinal function13 and increased damage in models of Alzheimer''s disease.14 Genetically, GFAP gain of function mutations associated with Alexander''s disease induce a p38 mitogen-activated protein kinase (MAPK)-dependent pathology.15 However, no pharmacologic tools have been available to specifically modulate and explore this reactive switch in the context of pathological CNS injury. Consequently, strategies to therapeutically target the reactive switch have remain challenging to explore.Withaferin A (WFA) is a small molecule withanolide that is a potent and specific inhibitor of type III intermediate filament dynamics.16, 17, 18 Its activity has been most closely studied with respect to vimentin rearrangement and phosphorylation in the context of angiogenesis, fibrosis and cancer, through downstream effects on inflammatory signaling and cell proliferation.19, 20, 21, 22, 23, 24 Interestingly, WFA has been reported to regulate vimentin-mediated activation of MAPKs in a context dependent manner, as well as NFκB.25, 26 Recently Bargagna-Mohan et al.27 reported that, in addition to vimentin, WFA also binds covalently to GFAP at cysteine 294. In these studies WFA impaired GFAP filament assembly and polymerization in cultured astrocytes, and in vivo in retinal astrocytes and related Müller glia in a model of injury-induced gliosis.27 Therefore, WFA presents a novel tool to test the pharmacologic blockade of intermediate filament remodeling during gliosis. However, the consequences of WFA disruption of IFs on neuronal damage has not been studied.We have previously used the retina as a uniquely accessible model to study the regulation of astrocyte stress responses, and their influence on retinal neuronal survival.28, 29, 30 In the human and rodent eye retinal ganglion cells (RGCs) and amacrine cells of the inner retina maintain a delicate homeostatic balance and are particularly vulnerable to excitotoxic and metabolic damage, mediated in part through non-cell autonomous interactions with neighboring glia.31, 32, 33, 34 In addition, our work and others has implicated signaling through p38 MAPKs as key regulators of glutamate recycling, antioxidant activity, and cytokine secretion in neighboring stress-activated retinal astrocytes and Müller glia.29, 35, 36, 37 Here we take advantage of a model of induced retinal astrocyte reactivity to establish whether WFA, and the selective p38 MAPK inhibitor SB203580 (SB), affect neuronal apoptosis in a mouse model of excitotoxic injury.  相似文献   

18.
Damage and loss of the postmitotic photoreceptors is a leading cause of blindness in many diseases of the eye. Although the mechanisms of photoreceptor death have been extensively studied, few studies have addressed mechanisms that help sustain these non-replicating neurons for the life of an organism. Autophagy is an intracellular pathway where cytoplasmic constituents are delivered to the lysosomal pathway for degradation. It is not only a major pathway activated in response to cellular stress, but is also important for cytoplasmic turnover and to supply the structural and energy needs of cells. We examined the importance of autophagy in photoreceptors by deleting the essential autophagy gene Atg5 specifically in rods. Loss of autophagy led to progressive degeneration of rod photoreceptors beginning at 8 weeks of age such that by 44 weeks few rods remained. Cone photoreceptor numbers were only slightly diminished following rod degeneration but their function was significantly decreased. Rod cell death was apoptotic but was not dependent on daily light exposure or accelerated by intense light. Although the light-regulated translocation of the phototransduction proteins arrestin and transducin were unaffected in rods lacking autophagy, Atg5-deficient rods accumulated transducin-α as they degenerated suggesting autophagy might regulate the level of this protein. This was confirmed when the light-induced decrease in transducin was abolished in Atg5-deficient rods and the inhibition of autophagy in retinal explants cultures prevented its degradation. These results demonstrate that basal autophagy is essential to the long-term health of rod photoreceptors and a critical process for maintaining optimal levels of the phototransduction protein transducin-α. As the lack of autophagy is associated with retinal degeneration and altered phototransduction protein degradation in the absence of harmful gene products, this process may be a viable therapeutic target where rod cell loss is the primary pathologic event.Autophagy is an intracellular pathway where cytoplasmic constituents are delivered to the lysosomes for degradation. Defective autophagy can contribute to the age-dependent accumulation of damaged proteins and organelles leading to altered cellular homeostasis and loss of function.1, 2, 3, 4, 5 The metabolic roles of autophagy can be classified into two types, basal and induced. In nutrient-rich conditions, autophagy is suppressed but still occurs at low levels (basal autophagy); however, when cells are subjected to stress (starvation, injury, hypoxia), autophagy is activated immediately (induced autophagy).6 Induced autophagy maintains the amino acid pool inside cells to adapt to starvation while constitutive autophagy has been shown to function as a cell-repair mechanism that is important for long-lived postmitotic cells.7, 8, 9, 10, 11 Defects in autophagy have been associated with neurodegenerative diseases,12, 13, 14, 15 diabetes,16, 17 lysosomal storage disease18 and the loss of vision.19 In addition to macroautophagy, microautophagy and chaperone-mediated autophagy (CMA) have been described. Although little is known about microautophagy in mammalian cells, macroautophagy (hereafter autophagy) is a major pathway for bulk degradation of cytoplasmic components. CMA is a more selective pathway for degradation of cytosolic proteins that can compensate for the loss of macroautophagy.2, 20, 21, 22Inherited retinal degenerative diseases such as retinitis pigmentosa or Leber''s congenital amaurosis are characterized by premature and progressive death of rod and cone photoreceptor cells.23 These diseases are characterized by the loss of night vision due to the death of rods followed by the loss of cones leading to diminished visual acuity and a reduction in the quality of life for patients. Disease is typically associated with the production of harmful gene products that promote pathology by inhibiting critical pathways resulting in cell death.24, 25, 26 Strategies to prevent photoreceptor death during retinal degenerative disease such as gene replacement therapies or inhibition of cell death pathways have been undertaken with some success;27, 28, 29 however, effective treatments for these blinding disorders are lacking.Another strategy that could be used in conjunction with other therapies might be to enhance survival by stimulating autophagy. Augmenting autophagy would increase the supply of nutrients to stressed cells and accelerate removal of damaged proteins thereby prolonging cell survival beyond what would be possible by only preventing cell death. Although canonical22, 30, 31, 32, 33 and noncanonical autophagic mechanisms34 have been detected in the eye, our knowledge of basic autophagy functions in this organ is still limited. In order to understand how autophagy maintains retinal homeostasis and function, we undertook studies to examine the consequences of deleting the essential autophagy gene Atg5 in rod photoreceptors.  相似文献   

19.
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   

20.
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   

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