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.
BRCA1 is a multifunctional protein best known for its role in DNA repair and association with breast and ovarian cancers. To uncover novel biologically significant molecular functions of BRCA1, we tested a panel of 198 approved and experimental drugs to inhibit growth of MDA-MB-231 breast cancer cells depleted for BRCA1 by siRNA. 26S proteasome inhibitors bortezomib and carfilzomib emerged as a new class of selective BRCA1-targeting agents. The effect was confirmed in HeLa and U2OS cancer cell lines using two independent siRNAs, and in mouse embryonic stem (ES) cells with inducible deletion of Brca1. Bortezomib treatment did not cause any increase in nuclear foci containing phosphorylated histone H2AX, and knockdown of BRCA2 did not entail sensitivity to bortezomib, suggesting that the DNA repair function of BRCA1 may not be directly involved. We found that a toxic effect of bortezomib on BRCA1-depleted cells is mostly due to deregulated cell cycle checkpoints mediated by RB1-E2F pathway and 53BP1. Similar to BRCA1, depletion of RB1 also conferred sensitivity to bortezomib, whereas suppression of E2F1 or 53BP1 together with BRCA1 reduced induction of apoptosis after bortezomib treatment. A gene expression microarray study identified additional genes activated by bortezomib treatment only in the context of inactivation of BRCA1 including a critical involvement of the ERN1-mediated unfolded protein response. Our data indicate that BRCA1 has a novel molecular function affecting cell cycle checkpoints in a manner dependent on the 26S proteasome activity.BRCA1 is an important tumor suppressor gene whose germ-line or somatic inactivation is implicated in a significant number of breast and ovarian cancers.1 Human BRCA1 encodes an 1863 amino-acid-long protein with a RING-finger domain at the N terminus and two BRCT domains located at the C terminus.2, 3 BRCT domains mediate interaction with phosphorylated proteins such as Abraxas, BACH1, CtIP and others involved in sensing DNA damage and assembly of the BRCA1-associated genome surveillance complex at sites of DNA breaks.4 The RING domain constitutively interacts with the BRCA1-associated RING domain protein (BARD1), forming a heterodimer having an E3 ubiquitin ligase activity.5 Ubiquitination of target proteins, including cell cycle or DNA repair-regulating proteins (e.g. CtIP (RBBP8), nucleophosmin (NPM1, B23), claspin (CLSPN) and others), occurs either at Lys48 residue of the ubiquitin leading to the 26S proteasome-mediated degradation of target proteins or at Lys6 or Lys63 having a trafficking and signaling role.6 A serine cluster coiled-coil domain spanning amino acids 1280–1524 contains multiple phosphorylation sites for ATM and ATR kinases activated by DNA damage.7 The same region also binds PALB2 protein linking BRCA1 to another major breast cancer predisposition gene BRCA2.8The most prominent function of BRCA1 is associated with its role in repair of DNA damage, particularly of double-stranded DNA breaks (DSBs), one of the most severe types of DNA lesions.9 BRCA1 is recruited to sites of DNA damage via a series of phosphorylation and ubiquitination events, where it serves as a binding scaffold for other DNA repair proteins,10, 11 ubiquitinates claspin, cyclin B and CDC25C, triggering cell cycle arrest to allow time for repair,12 and facilitates BRCA2-mediated loading of RAD51 recombinase to enable the homologous recombination (HR) mechanism of DNA repair.9 In addition, BRCA1 may contribute to maintaining genome integrity by stabilizing the heterochromatin structure via ubiquitination of histone H2A.13 BRCA1 is also required for centrosome-dependent and -independent mitotic spindle formation, providing another route, by which loss of BRCA1 could promote chromosome instability and tumor formation.14, 15Such a critical role of BRCA1 in DNA repair is exploited therapeutically. DNA-damaging agents, particularly DNA-crosslinking agents such as platinum-containing drugs, or ionizing radiation lead to the accumulation of DNA breaks requiring HR for repair and, therefore, are particularly toxic to BRCA1-deficient tumor cells.16 Pharmacological inhibitors of poly-(ADP-ribose) polymerases (PARPs) selectively kill BRCA1-deficient cells owing to defective HR, functioning as a back-up repair mechanism in the absence of the PARP-mediated repair of single-stranded DNA breaks.17 However, multiple mechanisms allow BRCA1-deficient cells to develop resistance to these drugs including elevated expression of the efflux transporters pumping the drugs out of the cell, secondary mutations restoring a functional BRCA1 protein and loss of 53BP1 protein, which counteracts BRCA1 and HR by blocking resection of DNA ends around the breaks (see Lord and Ashworth18 for the latest review). Therefore, additional efforts to identify small-molecule agents especially targeting BRCA1 functions unrelated to its DNA repair function are warranted.Here we performed a high-throughput chemical screen of BRCA1-depleted MDA-MB-231 cells using a collection of 198 US Food and Drug Administration (FDA)-approved and experimental drugs. We found that 26S proteasome inhibitors were more toxic to BRCA1 knockdown than control cells. Response of BRCA1-deficient cells to bortezomib involved deregulation of the RB1-mediated cell cycle checkpoint, activation of a noncanonical ERN1-mediated unfolded protein response and 53BP1-related G2/M cell cycle arrest. Our results reveal novel aspects of BRCA1 function unrelated to DNA repair.  相似文献   

2.
3.
NFBD1 functions in cell cycle checkpoint activation and DNA repair following ionizing radiation (IR). In this study, we defined the NFBD1 as a tractable molecular target to radiosensitize nasopharyngeal carcinoma (NPC) cells. Silencing NFBD1 using lentivirus-mediated shRNA-sensitized NPC cells to radiation in a dose-dependent manner, increasing apoptotic cell death, decreasing clonogenic survival and delaying DNA damage repair. Furthermore, downregulation of NFBD1 inhibited the amplification of the IR-induced DNA damage signal, and failed to accumulate and retain DNA damage-response proteins at the DNA damage sites, which leaded to defective checkpoint activation following DNA damage. We also implicated the involvement of NFBD1 in IR-induced Rad51 and DNA-dependent protein kinase catalytic subunit foci formation. Xenografts models in nude mice showed that silencing NFBD1 significantly enhanced the antitumor activity of IR, leading to tumor growth inhibition of the combination therapy. Our studies suggested that a combination of gene therapy and radiation therapy may be an effective strategy for human NPC treatment.Nasopharyngeal carcinoma (NPC) is a non-lymphomatous, squamous cell carcinoma that occurs in the epithelial lining of the nasopharynx, which is a prevalent tumor in people of southern Chinese ancestry in southern China and Southeast Asia, and the incidence is still increasing.1 Although radiotherapy is routinely used to treat patients with NPC, local recurrences and distant metastasis often occur in 30–40% of NPC patients at advanced staged.2 Thus, new therapeutic strategies are required to improve the poor prognosis of NPC.Among the various types of DNA damage, DNA double-strand breaks (DSBs) are the most serious and require elaborated networks of proteins to signal and repair the damage.3 It has recently been shown that the histone H2A variant H2AX specifically controls the recruitment of DNA repair proteins to the sites of DNA damage.4 H2AX is phosphorylated extensively on a conserved serine residue at its carboxyl terminus in chromatin regions bearing DSBs, which is mediated by members of the phos-phoinositide-3-kinase-related protein kinase (PIKK) family.5, 6 Of these PIKKs, ataxia telangiectasia mutated (ATM) and DNA-dependent protein kinase catalytic subunit (DNA-PKcs) phosphorylate H2AX in response to DSBs in a partially redundant manner.7, 8 NFBD1 (Nuclear Factor with BRCT Domain Protein 1), also known as MDC1 (mediator of DNA damage checkpoint protein 1), is a recently identified nuclear protein that regulates many aspects of the DNA damage-response pathway, such as intra-S phase checkpoint, G2/M checkpoint, spindle assembly checkpoint and foci formation of NBS/MRE/Rad50 (MRN complex), 53BP1 and BRCA1.9, 10, 11, 12, 13 Human NFBD1 comprises 2089 amino acid residues and has a predicted molecular weight of ∼220 kDa. Motifs found in the protein include an FHA (Forkhead Associated) domain, two BRCT (BRCA1 carboxy terminal) domains and around 20 in terminal repeats of ∼41 amino acid residues each.14 Following DNA damage, NFBD1 serves as a bridging molecule and directly interacts with ATM and phospho-H2AX (γ-H2AX) through its FHA and BRCT domains, respectively, which leads to the expansion of γ-H2AX region surrounding DNA strand breaks and provides docking sites for many DNA damage and repair proteins including the MRN complex, 53BP1, BRCA1, RNF8, RNF4 and so on, ensuring genomics stability.11, 15, 16, 17, 18 In mammalian cells, DSBs are mainly repaired by two mechanisms, homologous recombination (HR) or non-homologous end-joining (NHEJ).19, 20, 21 For NHEJ repair, it is estimated that following exposure to ionizing radiation (IR), 80–90% of the DSBs in G1 are rejoined with fast kinetics in a manner dependent upon the NHEJ core components, Ku, DNA-PKcs, XRCC4 and DNA ligase IV. In contrast, HR predominates in late S- and G2-phase cells, when the sister chromatid is available to act as the template, representing those normally repaired with slow kinetics, require Rad51, Rad52, Rad54, XRCC2, XRCC3, the Rad51 paralogs and the breast cancer susceptibility genes BRCA1 and BRCA2.22, 23, 24, 25, 26Since NFBD1 contains protein–protein interaction domains, and participate in the DNA damage-response (DDR) pathway. However, the mechanism by which NFBD1 regulates so many aspects of the DNA damage-response pathway in NPC cells is not fully understood. In addition, the physiological function of NFBD1 in NPC cells has been not investigated. With these goals in mind, we generated NFBD1-knockdown NPC cells and studied the physiological function of NFBD1 in DDR.  相似文献   

4.
Autophagy is a major nutrient recycling mechanism in plants. However, its functional connection with programmed cell death (PCD) is a topic of active debate and remains not well understood. Our previous studies established the plant metacaspase AtMC1 as a positive regulator of pathogen-triggered PCD. Here, we explored the linkage between plant autophagy and AtMC1 function in the context of pathogen-triggered PCD and aging. We observed that autophagy acts as a positive regulator of pathogen-triggered PCD in a parallel pathway to AtMC1. In addition, we unveiled an additional, pro-survival homeostatic function of AtMC1 in aging plants that acts in parallel to a similar pro-survival function of autophagy. This novel pro-survival role of AtMC1 may be functionally related to its prodomain-mediated aggregate localization and potential clearance, in agreement with recent findings using the single budding yeast metacaspase YCA1. We propose a unifying model whereby autophagy and AtMC1 are part of parallel pathways, both positively regulating HR cell death in young plants, when these functions are not masked by the cumulative stresses of aging, and negatively regulating senescence in older plants.An emerging theme in cell death research is that cellular processes thought to be regulated by linear signaling pathways are, in fact, complex. Autophagy, initially considered merely a nutrient recycling mechanism necessary for cellular homeostasis, was recently shown to regulate cell death, mechanistically interacting with components that control apoptosis. Deficient autophagy can result in apoptosis1, 2, 3 and autophagy hyper-activation can also lead to programmed cell death (PCD).4 In addition, the pro-survival function of autophagy is mediated by apoptosis inhibition and apoptosis mediates autophagy, although this cross-regulation is not fully understood.5In plants, autophagy can also have both pro-survival and pro-death functions. Autophagy-deficient plants exhibit accelerated senescence,6, 7, 8 starvation-induced chlorosis,6, 7, 9 hypersensitivity to oxidative stress10 and endoplasmic reticulum stress.11 Further, autophagy-deficient plants cannot limit the spread of cell death after infection with tissue-destructive microbial infections.12, 13 The plant phytohormone salicylic acid (SA) mediates most of these phenotypes.8 Autophagy has an essential, pro-survival role in situations where there is an increasing load of damaged proteins and organelles that need to be eliminated, that is, during aging or stress. Autophagy has an opposing, pro-death role during developmentally regulated cell death14, 15 or during the pathogen-triggered hypersensitive response PCD (hereafter, HR) that occurs locally at the site of attempted pathogen attack.16, 17 The dual pro-death/pro-survival functions of plant autophagy remain a topic of active debate.Also under scrutiny are possible novel functions of caspases and caspase-like proteins as central regulators of pro-survival processes. Caspases were originally defined as executioners of PCD in animals, but increasing evidence indicates that several caspases have non-apoptotic regulatory roles in cellular differentiation, motility and in the mammalian immune system.18, 19, 20Yeast, protozoa and plants do not have canonical caspases, despite the occurrence of morphologically heterogeneous PCDs.21 More than a decade ago, distant caspase homologs termed metacaspases were identified in these organisms using structural homology searches.22 Metacaspases were classified into type I or type II metacaspases based on the presence or absence of an N-terminal prodomain, reminiscent of the classification in animals into initiator/inflammatory or executioner caspases, respectively. Despite the architectural analogy between caspases and metacaspases, differences in their structure, function, activation and mode of action exist.23, 24, 25Metacaspases mediate PCD in yeast,26, 27, 28, 29, 30, 31 leishmania,32, 33 trypanosoma34 and plants.24 We demonstrated that two type I metacaspases, AtMC1 and AtMC2, antagonistically regulate HR in Arabidopsis thaliana.35 Our work showed that AtMC1 is a positive regulator of HR and that this function is mediated by its catalytic activity and negatively regulated by the AtMC1 N-terminal prodomain. AtMC2 antagonizes AtMC1-mediated HR.Besides AtMC2, new examples of metacaspases with a pro-life/non-PCD role are emerging. Protozoan metacaspases are involved in cell cycle dynamics34, 36, 37, 38 and cell proliferation.39 The yeast metacaspase Yca1 alters cell cycle dynamics40 and interestingly, is required for clearance of insoluble protein aggregates, thus contributing to yeast fitness.41Here, we explore the linkage between plant autophagy and AtMC1 function in the context of pathogen-triggered HR and aging. Our data support a model wherein autophagy and AtMC1 are part of parallel pathways, both positively regulating HR cell death in young plants and negatively regulating senescence in older plants.  相似文献   

5.
To grant faithful chromosome segregation, the spindle assembly checkpoint (SAC) delays mitosis exit until mitotic spindle assembly. An exceedingly prolonged mitosis, however, promotes cell death and by this means antimicrotubule cancer drugs (AMCDs), that impair spindle assembly, are believed to kill cancer cells. Despite malformed spindles, cancer cells can, however, slip through SAC, exit mitosis prematurely and resist killing. We show here that the Fcp1 phosphatase and Wee1, the cyclin B-dependent kinase (cdk) 1 inhibitory kinase, play a role for this slippage/resistance mechanism. During AMCD-induced prolonged mitosis, Fcp1-dependent Wee1 reactivation lowered cdk1 activity, weakening SAC-dependent mitotic arrest and leading to mitosis exit and survival. Conversely, genetic or chemical Wee1 inhibition strengthened the SAC, further extended mitosis, reduced antiapoptotic protein Mcl-1 to a minimum and potentiated killing in several, AMCD-treated cancer cell lines and primary human adult lymphoblastic leukemia cells. Thus, the Fcp1-Wee1-Cdk1 (FWC) axis affects SAC robustness and AMCDs sensitivity.The spindle assembly checkpoint (SAC) delays mitosis exit to coordinate anaphase onset with spindle assembly. To this end, SAC inhibits the ubiquitin ligase Anaphase-Promoting Complex/Cyclosome (APC/C) to prevent degradation of the anaphase inhibitor securin and cyclin B, the major mitotic cyclin B-dependent kinase 1 (cdk1) activator, until spindle assembly.1 However, by yet poorly understood mechanisms, exceedingly prolonging mitosis translates into cell death induction.2, 3, 4, 5, 6, 7 Although mechanistic details are still missing on how activation of cell death pathways is linked to mitosis duration, prolongation of mitosis appears crucial for the ability of antimicrotubule cancer drugs (AMCDs) to kill cancer cells.2, 3, 4, 5, 6, 7 These drugs, targeting microtubules, impede mitotic spindle assembly and delay mitosis exit by chronically activating the SAC. Use of these drugs is limited, however, by toxicity and resistance. A major mechanism for resistance is believed to reside in the ability of cancer cells to slip through the SAC and exit mitosis prematurely despite malformed spindles, thus resisting killing by limiting mitosis duration.2, 3, 4, 5, 6, 7 Under the AMCD treatment, cells either die in mitosis or exit mitosis, slipping through the SAC, without or abnormally dividing.2, 3, 4 Cells that exit mitosis either die at later stages or survive and stop dividing or proliferate, giving rise to resistance.2, 3, 4 Apart from a role for p53, what dictates cell fate is still unknown; however, it appears that the longer mitosis is protracted, the higher the chances for cell death pathway activation are.2, 3, 4, 5, 6, 7Although SAC is not required per se for killing,6 preventing SAC adaptation should improve the efficacy of AMCD by increasing mitosis duration.2, 3, 4, 5, 6, 7 Therefore, further understanding of the mechanisms by which cells override SAC may help to improve the current AMCD therapy. Several kinases are known to activate and sustain SAC, and cdk1 itself appears to be of primary relevance.1, 8, 9 By studying mitosis exit and SAC resolution, we recently reported a role for the Fcp1 phosphatase to bring about cdk1 inactivation.10, 11 Among Fcp1 targets, we identified cyclin degradation pathway components, such as Cdc20, an APC/C co-activator, USP44, a deubiquitinating enzyme, and Wee1.10, 11 Wee1 is a crucial kinase that controls the G2 phase by performing inhibitory phosphorylation of cdk1 at tyr-15 (Y15-cdk1). Wee1 is also in a feedback relationship with cdk1 itself that, in turn, can phosphorylate and inhibit Wee1 in an autoamplification loop to promote the G2-to-M phase transition.12 At mitosis exit, Fcp1 dephosphorylated Wee1 at threonine 239, a cdk1-dependent inhibitory phosphorylation, to dampen down the cdk1 autoamplification loop, and Cdc20 and USP44, to promote APC/C-dependent cyclin B degradation.10, 11, 12 In this study we analysed the Fcp1 relevance in SAC adaptation and AMCD sensitivity.  相似文献   

6.
7.
8.
Glioblastoma (GBM), the most prevalent type of primary intrinsic brain cancer in adults, remains universally fatal despite maximal therapy, including radiotherapy and chemotherapy. Cytotoxic therapy generates double-stranded DNA breaks (DSBs), most commonly repaired by homologous recombination (HR). We hypothesized that cancer cells coopt meiotic repair machinery as DSBs are generated during meiosis and repaired by molecular complexes distinct from genotoxic responses in somatic tissues. Indeed, we found that gliomas express meiotic repair genes and their expression informed poor prognosis. We interrogated the function of disrupted meiotic cDNA1 (DMC1), a homolog of RAD51, the primary recombinase used in mitotic cells to search and recombine with the homologous DNA template. DMC1, whose only known function is as an HR recombinase, was expressed by GBM cells and induced by radiation. Although targeting DMC1 in non-neoplastic cells minimally altered cell growth, DMC1 depletion in GBM cells decreased proliferation, induced activation of CHK1 and expression of p21CIP1/WAF1, and increased RPA foci, suggesting increased replication stress. Combining loss of DMC1 with ionizing radiation inhibited activation of DNA damage responses and increased radiosensitivity. Furthermore, loss of DMC1 reduced tumor growth and prolonged survival in vivo. Our results suggest that cancers coopt meiotic genes to augment survival under genotoxic stress, offering molecular targets with high therapeutic indices.Glioblastomas (GBMs) rank among the deadliest of all human cancers, with only modest improvement in patient survival over recent decades. More than 12 000 GBM patients are diagnosed annually in the United States.1, 2 Despite aggressive treatment consisting of maximal safe surgical resection, concurrent radiotherapy and chemotherapy, and adjuvant chemotherapy, median survival remains dismal at 12–15 months.3, 4 Although numerous molecular targets have been identified in GBM, no molecularly targeted therapy has demonstrated a survival benefit. Radiotherapy remains the cornerstone of post-surgical GBM therapy with modest additional benefit offered by concurrent administration of the oral methylator, temozolomide. However, radioresistance and tumor recurrence is universal in GBM.4, 5, 6 Radiation also damages non-neoplastic brain tissue, resulting in cognitive impairment and decreased quality-of-life.7 Focal high-dose radiation reduces toxicity to non-neoplastic tissue, but tumor invasion into normal brain regions limits the survival benefit of highly focused radiotherapy techniques, like gamma knife and proton beam, establishing a need for improved combinatorial treatments, such as radiosensitizers.8, 9 To date, no radiosensitizer has successfully increased survival with acceptable toxicity in a clinical trial. Based on this background, we sought novel molecular targets that mediate responses to genotoxic stress and have limited function in normal cells.During mitosis, cells inspect the integrity of their DNA and repair replication errors through cell-state and error-specific mechanisms.10 Unrepaired or large regions of DNA damage overwhelm replication mechanisms to induce cell death.10, 11 DNA double-strand breaks (DSBs) are detrimental as they cause large-scale chromosomal rearrangements.10 The homologous recombination (HR) pathway is primarily used to repair DSBs during S- and G2-phases, providing access to both sister and homologous chromosomes as repair templates.7, 12 RADiation sensitive 51 (RAD51) is a key recombinase important in HR and replication fork maintenance, functioning in both mitotic and meiotic cells.7, 12, 13, 14, 15 Phosphorylated RAD51 replaces replication protein A (RPA) upon DNA loading.16 Recombination mediated by RAD51 with the intact DNA template strand results in a relatively error-free repair.12In contrast to mitosis, germ cells undergoing meiosis actively generate genetic diversity through induction of programmed DSBs, which are repaired through HR.17, 18, 19 In meiotic HR, RAD51 functions in conjunction with the meiosis-specific recombinase, disrupted meiotic cDNA1 (DMC1). RAD51 and DMC1 are loaded onto DNA by a meiosis-specific accessory protein complex, homologous-pairing protein 2 (HOP2)–meiotic nuclear divisions 1 (MND1), to promote homologous strand invasion and dissociation-loop (D-loop) formation.20, 21 D-loops formed using the DMC1–RAD51 complex are more resistant to dissociation as opposed to D-loops formed by RAD51 alone, increasing the likelihood of DNA crossover events.20 In addition, DMC1-directed crossovers preferentially utilize the homologous chromosome further increasing genetic variation.22GBM cells commonly harbor genetic lesions that promote unrestrained proliferation but also stimulate genotoxic stress responses. Neoplastic cells do not require perfect fidelity of repair. In fact, dysfunctional repair accelerates genetic evolution of clones, but cancer cells must acquire mechanisms to bypass cell death or senescence in response to exogenous stressors.11, 23 Radiotherapy targets proliferating cancer cells by production of reactive oxygen species, leading to generation of DSBs and activation of the DNA damage response (DDR) pathway.11, 24 DSBs generated as a result of ionizing radiation (IR) are repaired through HR or non-homologous end joining (NHEJ).7, 12, 25, 26 Terminally differentiated neurons are post-mitotic and rely on NHEJ as a means to repair DNA DSBs. Therefore, inhibition of the NHEJ pathway may result in unfavorable normal neural cell toxicity.26The HR pathway is an attractive target as it is linked to increased genetic variation and loss of heterozygosity (LOH).12, 27 Multiple HR checkpoints have been proposed as potential therapeutic targets for GBM.28, 29, 30, 31 Although the prognostic value of RAD51 expression in GBM is unresolved,29, 32, 33 RAD51 is consistently elevated in GBM compared with normal brain.33 Reducing RAD51 expression radiosensitizes GBM cells,29 but may have a limited therapeutic index because of the potentially toxic effects on non-neoplastic cells. In this study, we investigated the aberrant activity of meiotic HR regulators in glioma, focusing on the meiosis-specific DMC1. Activation of meiotic repair genes in neoplastic cells selectively provides tumor cells with a repair mechanism to evade cell death caused by DNA damage, yet increase genetic diversity to drive clonal evolution.  相似文献   

9.
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   

10.
11.
12.
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   

13.
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   

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.
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.
18.
Spontaneous neoplasms in Mongolian gerbils have an incidence of 20% to 26.8%, but osteosarcomas occur at a much lower rate. Here we report a 1-y-old Mongolian gerbil with a spontaneous osteosarcoma at the level of the proximal tibia, with metastases to the pectoral muscles and lungs. Grossly, the tibial mass obliterated the tibia and adjacent muscles, and an axillary mass with a bloody, cavitary center expanded the pectoral muscles. Microscopically, the tibial mass was an infiltrative, osteoblastic mesenchymal neoplasm, and the axillary mass was an anaplastic mesenchymal neoplasm with hemorrhage. The lung contained multiple metastatic foci. Immunohistochemistry for osteonectin was strongly positive in the tibial, axillary, and pulmonary metastases. Although osteosarcoma is the most common primary malignant bone neoplasm that occurs spontaneously in all laboratory and domestic animal species and humans, it arises less frequently than does other neoplasms. The current case of spontaneous osteoblastic osteosarcoma of the proximal tibia and metastases to the pectoral muscles and lung in a Mongolian gerbil is similar in presentation, histology, and predilection site of both osteoblastic and telangiectatic osteosarcomas in humans. In addition, this case is an unusual manifestation of osteosarcoma in the appendicular skeleton of a Mongolian gerbil.Mongolian gerbils are used frequently in biologic research,1,2,4,9,10,12-14 particularly in oncogenic studies and filariasis research studying Brugia malayi.2 There have been several reports1,6,10,11,13-15 of spontaneous neoplasms, particularly in gerbils 2 y of age and older, typically occurring with the highest incidences in the skin, reproductive tract, and adrenal glands; however, neoplasms have also been reported in the thyroid, thymus, liver, kidney, pancreas, and bone.1,6,10,11,13-15 The incidence of spontaneous neoplasms occurring in the subfamily Gerbillinae ranges from 20% to 26.8%,1,6,10,11,13-15 depending on the study, age, and sex of the animals.With a lower incidence than those reported for other neoplasms, osteosarcomas in gerbils have been described in the ramus of the mandible and as an extraskeletal mass throughout the peritoneum.10,11 The usual age of onset for osteosarcomas in Mongolian gerbils is approximately 3 y (36 to 39 mo); however, no tumor type has been reported at less than 2 y of age in this species.10,11 Here we report a spontaneous osteosarcoma that occurred at the level of the proximal tibia, with metastases to the pectoral muscles and lung, in a 1-y-old Mongolian gerbil.  相似文献   

19.
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   

20.
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   

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