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 15 毫秒
1.
Dysferlin deficiency compromises the repair of injured muscle, but the underlying cellular mechanism remains elusive. To study this phenomenon, we have developed mouse and human myoblast models for dysferlinopathy. These dysferlinopathic myoblasts undergo normal differentiation but have a deficit in their ability to repair focal injury to their cell membrane. Imaging cells undergoing repair showed that dysferlin-deficit decreased the number of lysosomes present at the cell membrane, resulting in a delay and reduction in injury-triggered lysosomal exocytosis. We find repair of injured cells does not involve formation of intracellular membrane patch through lysosome–lysosome fusion; instead, individual lysosomes fuse with the injured cell membrane, releasing acid sphingomyelinase (ASM). ASM secretion was reduced in injured dysferlinopathic cells, and acute treatment with sphingomyelinase restored the repair ability of dysferlinopathic myoblasts and myofibers. Our results provide the mechanism for dysferlin-mediated repair of skeletal muscle sarcolemma and identify ASM as a potential therapy for dysferlinopathy.Dysferlinopathy is a progressive muscle wasting disease, which is classified as limb-girdle muscular dystrophy type 2B (LGMD2B) or Miyoshi muscular dystrophy 1, based on its muscle involvement.1, 2 Dysferlin deficit leads to altered vesicle formation and trafficking,3, 4 poor repair of injured cell membranes,5, 6 and increased muscle inflammation.7, 8 Dysferlin contains C2 domains that are found in Ca2+-dependent membrane fusion proteins such as synaptotagmins.9 Thus, dysferlin is thought to regulate muscle function by regulating vesicle trafficking and fusion.10, 11, 12, 13 Dysferlin deficiency has also been implicated in conflicting reports regarding the fusion ability of dysferlinopathic myoblasts.4, 14, 15, 16 With such diverse roles for dysferlin, the mechanism through which dysferlin deficiency results in muscle pathology is unresolved. As skeletal muscle-specific re-expression of dysferlin rescues all dysferlinopathic pathologies,17, 18 myofiber repair has been suggested to be the unifying deficit underlying muscle pathology in dysferlinopathy.19 Repair of injured cell membranes requires subcellular compartments, which in mammalian cells include lysosomes,11 enlargeosomes,20 caveolae,21 dysferlin-containing vesicles,5 and mitochondria.22Cells from muscular dystrophy patients that have normal dysferlin expression exhibit normal lysosome and enlargeosome exocytosis.23 However, dysferlinopathic muscle cells exhibit enlarged LAMP2-positive lysosomes, reduced fusion of early endosomes, altered expression of proteins regulating late endosome/lysosome fusion, and reduced injury-triggered cell-surface levels of LAMP1.4, 11, 12 In non-muscle cells, lack of dysferlin reduces lysosomal exocytosis.24 These findings implicate lysosomes in dysferlin-mediated muscle cell membrane repair. In one model for lysosome-mediated cell membrane repair, Ca2+ triggers vesicle–vesicle fusion near the site of injury, forming ‘membrane patch'', which fuses to repair the wounded cell membrane.25, 26, 27, 28 In another model, lysosome exocytosis following cell membrane injury by pore-forming toxins leads to secretion of the lysosomal enzyme acid sphingomyelinase (ASM), which causes endocytosis of pores in the damaged cell membranes.21, 29, 30 Both these models have been suggested to be involved in the repair of injured muscle cells.21, 28To examine the muscle cell pathology in dysferlinopathy, we have developed dysferlinopathic mouse and human models. Use of these models shows that a lack of dysferlin does not alter myogenic differentiation but causes poor repair of even undifferentiated muscle cells. We show that dysferlin is required for tethering lysosomes to the cell membrane. Fewer lysosomes at the cell membrane in dysferlinopathic cells results in slow and reduced lysosome exocytosis following injury. This reduction in exocytosis reduces injury-triggered ASM secretion, which is responsible for the poor repair of dysferlinopathic muscle cells. Extracellular sphingomyelinase (SM) fully rescues the repair deficit in dysferlinopathic cells and mouse myofibers, offering a potential drug-based therapy for dysferlinopathy.  相似文献   

2.
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   

3.
4.
5.
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   

6.
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   

7.
The small diameter of the carotid artery is not compatible with the evaluation of clinically available endovascular devices in the carotid balloon-injury (BI) model. We developed an endovascular BI model in the rat descending aorta, whose size is compatible with available endovascular instruments. We also tested the hypothesis that neointima formation is enhanced in the aorta of obese Zucker rats (OZR) compared with lean Zucker rats (LZR). Left external carotid arteriotomies and BI of the thoracic and abdominal aorta were performed by using a balloon catheter. Aortograms and aortic pathology were examined at 2, 4, and 10 wk after BI. At 10 wk after BI, the abdominal aorta in OZR had narrowed 8.3% ± 1.1% relative to baseline compared with an expansion of 2.4% ± 2.2% in LZR. Simultaneously, the thoracic aorta had expanded 9.5% ± 4.3% in LZR compared with stenosis of 2.8% ± 1.6% in OZR. Calculation of the intimal:medial thickness ratio revealed significantly increased neointimal formation in the OZR descending aorta compared with that in LNR. In conclusion, this minimally invasive BI model involving the rat descending aorta is compatible with available endovascular instruments. The descending aorta of OZR demonstrates enhanced neointimal formation and constrictive vascular remodeling after BI.Abbreviations: BI, balloon injury; LZR, lean Zucker rats; OZR, obese Zucker ratsMany endovascular interventions, either cardiovascular or neuro­vascular, involve balloon-angioplasty or stenting of stenosed or vasospastic arteries. Balloon injury (BI) to the arterial wall is well known to induce restenosis, with a typical decrease in lumen diameter as a result of intimal hyperplasia and vessel wall remodeling.4,5,6,15,17,18 Restenosis occurs more frequently in patients with additional cardiovascular risk factors including diabetes, dyslipidemia, obesity, and hypertension.4,8,23 Zucker rats frequently are used as animal models of obesity, dyslipidemia, and type 2 diabetes.16,20,24The rat common carotid artery BI model is widely applied to study molecular mechanisms and the role of smooth muscle cells in arterial disease and healing.17 Data on injury in other arteries and models are not extensive.5 From a clinical perspective, the rat carotid artery is too small to test available endovascular devices. Other sites that have been used in rat models include the infrarenal aorta and the common iliac artery.4,5,11,13,15,25 The aortic model requires an open laparotomy, whereas the common iliac artery is another small-caliber vessel. The goal of the present study was to create a minimally invasive BI model in the rat descending aorta that would be compatible with many endovascular instruments and implantable devices, such as stents, that are used currently in clinical practice.15  相似文献   

8.
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.
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   

11.
12.
The purpose of this study was to conduct a comprehensive evaluation of the vascular supply to the femoral head, including the vessels that give rise to the terminal perfusing branches. Using a casting agent, we highlighted the anatomy of the external iliac and ischiatic arteries with their associated branches after anatomic dissection of 24 hips from 12 Leghorn chickens. We confirmed published findings regarding perfusion of the femoral head and identified 3 previously undescribed arterial branches to this structure. The first branch (the acetabular branch of the femoralis artery) was supplied by the femoralis artery and directly perfused the acetabulum and femoral head. The second branch (the lateral retinacular artery) was a tributary of the femoralis artery that directly supplied the femoral head. Finally, we found that the middle femoral nutrient artery supplies a previously undescribed ascending intraosseous branch (the ascending branch of the middle femoral nutrient artery) that perfuses the femoral head. Precise understanding of the major vascular branches to the femoral head would allow for complete or selective ligation of its blood supply and enable the creation of a reproducible bipedal model of femoral head osteonecrosis.Like humans, chickens are bipedal animals that rely on the hip joint to absorb the majority of the body''s weight. This anatomy, in concert with their high activity level, makes chickens an attractive model for the study of osteonecrosis of the femoral head in humans. The vast majority of animal research on osteonecrosis of the femoral head has been performed on quadrupedal animals,3,4,10,19,25,26,28,29,31,36,37,41,51,52 thus limiting its application to bipedal species because most quadruped models fail to progress to end-stage mechanical collapse similar to that in humans.6Avascular necrosis is the death of bone that occurs from ischemia due to disruption of the vascular supply to bone through direct or indirect mechanisms.38 Avascular necrosis should be differentiated from the broader term of osteonecrosis, which refers to bone death in general.32 Causes of femoral head osteonecrosis include direct and indirect disruption of vascular supply (traumatic injury, intravascular coagulation, extrinsic compression) as well as changes in cellular differentiation and cellular apoptosis.4,7,12,15,17,18,24,30-32,38,49,50 Accordingly, causes of osteonecrosis are both traumatic and nontraumatic.16,31,32The arterial anatomy in the chicken hindlimb has been outlined by several authors.20,22,27,35,42,44,45 Briefly, the external iliac and ischiatic artery arise from the abdominal aorta to provide blood supply to the chicken hind limb. The external iliac artery has 2 main branches—the femoralis and femoral circumflex arteries—that distribute blood to the chicken hindlimb. The ischiatic artery provides 3 main branches: the trochanteric artery, superior femoral nutrient artery, and middle femoral nutrient artery. Although the terminal vascular supply to the femoral head of Leghorn and Broiler chickens has been described,46,47 the origin of these terminal arteries with reference to the ischiatic and femoralis arteries and their respective branches has not been addressed. The current study will describe the blood vessels that feed these terminal branches to the chicken femoral head.  相似文献   

13.
Neuritin 1 (Nrn1) is an extracellular glycophosphatidylinositol-linked protein that stimulates axonal plasticity, dendritic arborization and synapse maturation in the central nervous system (CNS). The purpose of this study was to evaluate the neuroprotective and axogenic properties of Nrn1 on axotomized retinal ganglion cells (RGCs) in vitro and on the in vivo optic nerve crush (ONC) mouse model. Axotomized cultured RGCs treated with recombinant hNRN1 significantly increased survival of RGCs by 21% (n=6–7, P<0.01) and neurite outgrowth in RGCs by 141% compared to controls (n=15, P<0.05). RGC transduction with AAV2-CAG–hNRN1 prior to ONC promoted RGC survival (450%, n=3–7, P<0.05) and significantly preserved RGC function by 70% until 28 days post crush (dpc) (n=6, P<0.05) compared with the control AAV2-CAG–green fluorescent protein transduction group. Significantly elevated levels of RGC marker, RNA binding protein with multiple splicing (Rbpms; 73%, n=5–8, P<0.001) and growth cone marker, growth-associated protein 43 (Gap43; 36%, n=3, P<0.01) were observed 28 dpc in the retinas of the treatment group compared with the control group. Significant increase in Gap43 (100%, n=5–6, P<0.05) expression was observed within the optic nerves of the AAV2–hNRN1 group compared to controls. In conclusion, Nrn1 exhibited neuroprotective, regenerative effects and preserved RGC function on axotomized RGCs in vitro and after axonal injury in vivo. Nrn1 is a potential therapeutic target for CNS neurodegenerative diseases.Central nervous system (CNS) trauma and neurodegenerative disorders trigger a cascade of intrinsic and extrinsic cellular events resulting in regenerative failure and subsequent damage to neurons.1, 2, 3, 4, 5 The intrinsic factors include deregulation in growth-promoting factors, apoptotic factors, intracellular signaling molecules and trophic factors.6 Similarly, the extrinsic factors correlate to growth inhibition due to inhibitory cues3, 7, 8, 9, 10, 11, 12, 13 that include myelin and myelin associated inhibitors, glial scarring,5, 14 slow clearance of axonal debris,7 incorrect development of neuronal projections6 and CNS inflammation.15, 16 Progressive degeneration of mature retinal ganglion cells (RGCs) has been associated with loss of trophic support,8, 9 detrimental inflammatory processes/immune regulation10, 11 and apoptotic effectors.9, 12, 13, 15, 17After injury, mammalian RGC axons show only a short-lived sprouting response but no long-distance regeneration through the optic nerve (ON).16 Glial responses around the affected area are initiated by injured CNS axons.18 Axons undergoing Wallerian degeneration are surrounded by astrocytes that upregulate glial fibrillary acidic protein (Gfap) expression and these reactive astrocytes contribute to trauma-induced neurodegeneration.19 Glial scarring inhibits axonal transport after ON crush (ONC)5, 14 decreasing transport of proteins involved in neuroprotection and synaptic plasticity. Regenerative failure is a critical endpoint of these destructive triggers culminating in neuronal apoptosis3, 20, 21 and inhibition of functional recovery. Intrinsic factors affecting axonal regeneration after CNS injury are crucial for recovery and thus, dysregulation of genes involved in axonal plasticity and outgrowth can prove detrimental to the neuronal recovery.22, 23, 24Current neuroprotection approaches include promoting survival of RGCs by intraocular injections of recombinant factors like ciliary neurotrophic factor (CNTF) and peripheral nerve (PN) transplantations in vitro25 and in vivo after injury.26 Studies performed with glial cell-line-derived neurotrophic factor and neurturin protect RGCs from axotomy-induced apoptosis.27 Further, in the ON injury model, RGC survival was promoted after deletion of CCAAT/enhancer binding protein homologous protein28 and enhanced regeneration observed with co-deletion of kruppel-like factor 4 (Klf4) and suppressor of cytokine signaling 3 (Socs3).29 Intraocular administration of neurotrophin-4 (NT-4) and brain-derived neurotrophic factor (BDNF) after ON transection has also exerted neuroprotective effects on axotomized RGCs. In addition, PNs transplanted adjacent to ONs, ex vivo PN grafts with lenti-viral transduced Schwann cells, and stimulation of inflammatory processes have strong pro-regenerative effects on injured RGCs.26, 30, 31, 32, 33In addition, using adeno-associated-virus (AAV) therapy, AAV mediated expression of CNTF in bcl2 overexpressing transgenic mice increases cell viability and axonal regeneration,34 whereas BDNF promotes survival of RGCs.35 Likewise, experiments with AAV–BDNF, –CNTF and –growth-associated protein 43 (GAP43) have shown that AAV–CNTF was the most crucial for promoting both long-term survival and regeneration.36 The positive effects of CNTF are observed mainly through simultaneous deletion of both PTEN and SOCS337 and the concurrent activation of mTOR and STAT3 pathways.38 Although CNTF shows robust increase and sustained axon regeneration in injured ONs of rodents, it causes axonal misguidance and aberrant growth.39 Furthermore, it has been shown that CNTF acts as a chemoattractant. CNTF administration onto autologous PN grafts transplanted within transected ON increased regeneration, but these effects were significantly reduced after removal of macrophages from this site.40 In addition, the effects of CNTF using PN grafts at ON transection sites are further subject to debate, as previously it has been shown that Ad-CNTF injections preserved RGC axons but did not induce regeneration of axotomized RGCs.41 Thus, other studies have addressed RGC survivability and axonal regeneration with CNTF and other growth factors,35, 36 but most trophic factors affect neuronal survival and regeneration differentially.Previous studies targeting neuronal apoptosis by overexpressing intrinsic growth factors, inhibiting apoptosis and enhancing regeneration in CNS trauma models have established that a multifactorial approach is required for successful and long-lasting therapeutic outcomes.6, 36 Current gaps still exist for a key gene that could effectively target neuroprotection, enhance neuron regeneration and sustain neuronal function.One key gene implicated in neuronal plasticity is Neuritin 1 (Nrn1), also known as candidate plasticity gene 15. It has multiple functions and was first identified and characterized when screening for candidate plasticity genes in the rat hippocampal dentate gyrus activated by kainate.42, 43, 44 Nrn1 is highly conserved across species45 and translates to an extracellular, glycophosphatidylinositol-linked protein (GPI-linked protein), which can be secreted as a soluble form. Nrn1 stimulates axonal plasticity, dendritic arborization and synapse maturation in the CNS.46 During early embryonic development, Nrn1 promotes the survival of neural progenitors and differentiated neurons,47 while later in development it promotes axonal and dendritic growth and stabilization, allowing maturation and formation of synapses.43, 46, 48 In the adult brain, Nrn1 has been correlated with activity-dependent functional plasticity45, 49 and is expressed in post mitotic neurons.Nrn1 may be a crucial gene for neuroprotection and regeneration because growth factors such as nerve growth factor (NGF), BDNF and NT-3 as well as neuronal activity can potentiate the expression of Nrn1.44, 50 In addition, we reported that Nrn1 mRNA expression appears to be biphasic after ON axonal trauma, indicating a transient attempt by RGCs at neuroprotection/neuroregeneration in response to ONC injury.51 The dynamic regulation of Nrn1 coupled with neurotrophic effects may promote axonal regeneration in the CNS. To overcome CNS trauma, a new therapy geared towards neuroprotection and effective axonal regeneration is required to enhance a future multifactorial approach. The purpose of this study is to evaluate the therapeutic effects of Nrn1 in mouse RGC cultures as well as in the mouse ONC model. We have identified a distinct neuroprotective and regenerative strategy that prevents neurodegeneration after ON injury. AAV2–hNRN1 expression vectors partially rescued RGCs from apoptosis, maintained RGC function, and initiated regeneration of injured axons.  相似文献   

14.
Cell-based therapies are increasingly recognized as a potential strategy to treat retinal neurodegenerative disease. Their administration, however, is normally indirect and complex, often with an inability to assess in real time their effects on cell death and their migration/integration into the host retina. In the present study, using a partial optic nerve transection (pONT) rat model, we describe a new method of Schwann cell (SC) delivery (direct application to injured optic nerve sheath, SC/DONS), which was compared with intravitreal SC delivery (SC/IVT). Both SC/DONS and SC/IVT were able to be assessed in vivo using imaging to visualize retinal ganglion cell (RGC) apoptosis and SC retinal integration. RGC death in the pONT model was best fitted to the one-phase exponential decay model. Although both SC/DONS and SC/IVT altered the temporal course of RGC degeneration in pONT, SC/DONS resulted in delayed but long-lasting effects on RGC protection, compared with SC/IVT treatment. In addition, their effects on primary and secondary degeneration, and axonal regeneration, were also investigated, by histology, whole retinal counting, and modelling of RGC loss. SC/DONS was found to significantly reduce RGC apoptosis in vivo and significantly increase RGC survival by targeting secondary rather than primary degeneration. Both SC/DONS and SC/IVT were found to promote RGC axonal regrowth after optic nerve injury, with evidence of GAP-43 expression in RGC somas and axons. SC/DONS may have the potential in the treatment of optic neuropathies, such as glaucoma. We show that SC transplantation can be monitored in real time and that the protective effects of SCs are associated with targeting secondary degeneration, with implications for translating cell-based therapies to the clinic.In the central (CNS) and peripheral (PNS) nervous systems, injury from initial lesions can lead to widespread damage to neurons beyond the primary injury site; a phenomenon known as ‘secondary degeneration''. Studies in spinal cord injury have revealed secondary rather than primary degeneration to be the major contributor to neuronal death and functional impairment, and it is increasingly recognized as a therapeutic target.1,2 Secondary degeneration also occurs in optic neuropathies, including glaucoma, ischaemic optic neuropathy, and Leber''s hereditary optic neuropathy.3, 4, 5 Retinal neuronal loss in these conditions is reported to occur long after the initial insult,6 implying that secondary mechanisms may have an important role in optic neuropathic damage and that targeting of secondary neuronal loss may represent a novel therapeutic strategy.Partial optic nerve transection (pONT) represents a reliable and reproducible model for studying secondary degeneration, in which a primary lesion is only made to dorsal axons and leaves those in ventral optic nerve (ON) intact but vulnerable to secondary degeneration.4,7 Secondary degeneration is thought to be initiated by a cascade of reactive metabolic events, including glutamate excitotoxicity, Ca2+ overload, excess free radical formation, oxidative stress, mitochondrial dysfunction, and increased proteoglycan expression, leading to cell death.7, 8, 9, 10, 11, 12, 13, 14 Activated astrocytes are reported to be a major contributor to spreading and acceleration of secondary degeneration.8,9As in most CNS pathways, the mature ON possesses only a limited ability to repair itself after injury, resulting in permanent vision loss due to the death of retinal ganglion cells (RGCs), the retinal output neurons that transmit visual information to the brain.15 Compared with the CNS, the PNS has a remarkable ability to regrow after injury, a process in which Schwann cells (SCs) are thought to have a key role.16,17SCs are the principal glia of the PNS and support normal neuronal function.18,19 Upon axonal injury, SCs are reported to shed their myelin sheaths and de-differentiate into progenitor stem cells, which are capable of replacing damaged tissue and providing a permissive environment for neuronal survival and axonal regrowth.18,19 SCs are believed to achieve this through releasing neurotrophic factors and producing cell adhesion molecules and extracellular matrix components.20 The neuroprotective and regenerative mechanisms between SCs and neurons are thought to operate on a local basis via adhesion molecules, allowing contact-mediated signalling between cells,16,17,20,21 and extracellular free ligands, facilitating specific binding to the receptors in the target neurons.16,17,20 However, a novel regulatory mechanism has emerged, representing a more efficient and advanced communication machinery, that is, vesicular transfer between SCs and axons.16 We have recently demonstrated that the highly efficient response of SCs to PN injury is triggered by Ephrin-B/EphB2 signalling in fibroblasts, which guide SC sorting and migration during nerve repair.21Due to the regenerative ability of SCs in PNS repair, transplantation of SCs to the injured ON has been previously attempted.22, 23, 24, 25, 26, 27, 28 To date, however, the protective effects of SCs on retinal neurons have been only assessed after either intravitreal administration or suturing artificial SC grafts onto transected ON, using postmortem histological observations, with incomplete delineation of the mechanisms involved.22, 23, 24, 25, 26, 27, 28Here we use a pONT model to investigate a new method of SC delivery (direct application to injured ON sheath, SC/DONS), using in vivo imaging and histological techniques, and compare its effects on RGC apoptosis and loss to intravitreal SC delivery (SC/IVT). Furthermore, we analyse whether these actions target primary or secondary degeneration, to determine their potential in the treatment of optic neuropathy.  相似文献   

15.
In the oxidative stress hypothesis of aging, the aging process is the result of cumulative damage by reactive oxygen species. Humans and chimpanzees are remarkably similar; but humans live twice as long as chimpanzees and therefore are believed to age at a slower rate. The purpose of this study was to compare biomarkers for cardiovascular disease, oxidative stress, and aging between male chimpanzees and humans. Compared with men, male chimpanzees were at increased risk for cardiovascular disease because of their significantly higher levels of fibrinogen, IGF1, insulin, lipoprotein a, and large high-density lipoproteins. Chimpanzees showed increased oxidative stress, measured as significantly higher levels of 5-hydroxymethyl-2-deoxyuridine and 8-iso-prostaglandin F, a higher peroxidizability index, and higher levels of the prooxidants ceruloplasmin and copper. In addition, chimpanzees had decreased levels of antioxidants, including α- and β-carotene, β-cryptoxanthin, lycopene, and tocopherols, as well as decreased levels of the cardiovascular protection factors albumin and bilirubin. As predicted by the oxidative stress hypothesis of aging, male chimpanzees exhibit higher levels of oxidative stress and a much higher risk for cardiovascular disease, particularly cardiomyopathy, compared with men of equivalent age. Given these results, we hypothesize that the longer lifespan of humans is at least in part the result of greater antioxidant capacity and lower risk of cardiovascular disease associated with lower oxidative stress.Abbreviations: 5OHmU, 5-hydroxymethyl-2-deoxyuridine; 8isoPGF, 8-iso-prostaglandin F; HDL, high-density lipoprotein; IGF1, insulin-like growth factor 1; LDL, low-density lipoprotein; ROS, reactive oxygen speciesAging is characterized as a progressive reduction in the capacity to withstand the stresses of everyday life and a corresponding increase in risk of mortality. According to the oxidative stress hypothesis of aging, much of the aging process can be accounted for as the result of cumulative damage produced by reactive oxygen species (ROS).6,21,28,41,97 Endogenous oxygen radicals (that is, ROS) are generated as a byproduct of normal metabolic reactions in the body and subsequently can cause extensive damage to proteins, lipids, and DNA.6,41 Various prooxidant elements, in particular free transition metals, can catalyze these destructive reactions.6 The damage caused by ROS can be counteracted by antioxidant defense systems, but the imbalance between production of ROS and antioxidant defenses, over time, leads to oxidative stress and may contribute to the rate of aging.28,97Oxidative stress has been linked to several age-related diseases including neurodegenerative diseases, ophthalmologic diseases, cancer, and cardiovascular disease.21,28,97 Of these, cardiovascular disease remains the leading cause of adult death in the United States and Europe.71 In terms of cardiovascular disease, oxidative stress has been linked to atherosclerosis, hypertension, cardiomyopathy, and chronic heart failure in humans.55,78,84 Increases in oxidant catalysts (prooxidants)—such as copper, iron, and cadmium—have been associated with hypertension, coronary artery disease, atherosclerosis, and sudden cardiac death.98,102,106 Finally, both endogenous and exogenous antioxidants have been linked to decreased risk of cardiovascular disease, although the mechanisms behind this relationship are unclear.11,52,53 However, the oxidative stress hypothesis of aging aims to explain not only the mechanism of aging and age-related diseases (such as cardiovascular disease) in humans but also the differences between aging rates and the manifestations of age-related diseases across species.The differences in antioxidant and ROS levels between animals and humans offer promise for increasing our understanding of human aging. Additional evidence supporting the oxidative stress hypothesis of aging has come from comparative studies linking differences in aging rates across taxa with both antioxidant and ROS levels.4,17-21,58,71,86,105 In mammals, maximum lifespan potential is positively correlated with both serum and tissue antioxidant levels.17,18,21,71,105 Research has consistently demonstrated that the rate of oxidative damage varies across species and is negatively correlated with maximum lifespan potential.4,19,20,58,71,86 However, few studies involved detailed comparisons of hypothesized biochemical indicators of aging and oxidative stress between humans and animals.6 This type of interspecies comparison has great potential for directly testing the oxidative stress hypothesis of aging.Much evolutionary and genetic evidence supports remarkable similarity between humans and chimpanzees.95,100 Despite this similarity, humans have a lifespan of almost twice that of chimpanzees.3,16,47 Most comparative primate aging research has focused on the use of a macaque model,62,81,88 and several biochemical markers of age-related diseases have been identified in both humans and macaque monkeys.9,22,28,81,93,97 Several other species of monkeys have also been used in research addressing oxidative stress, antioxidant defenses, and maximum lifespan potential.18,21,58,105 However, no study to date has examined biochemical indicators of oxidative stress and aging in chimpanzees and humans as a test of the oxidative stress hypothesis for aging. The purpose of this study is to compare biochemical markers for cardiovascular disease, oxidative stress, and aging directly between male chimpanzees and humans. Given the oxidative stress hypothesis for aging and the known role of oxidative stress in cardiovascular disease, we predict that chimpanzees will show higher levels of cardiovascular risk and oxidative stress than humans.  相似文献   

16.
17.
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   

18.
Metabolic syndrome (MetS), a compilation of associated risk factors, increases the risk of type 2 diabetes and coronary artery disease (CAD, atherosclerosis), which can progress to the point of artery occlusion. Stents are the primary interventional treatment for occlusive CAD, and patients with MetS and hyperinsulinemia have increased restenosis. Because of its thrifty genotype, the Ossabaw pig is a model of MetS. We tested the hypothesis that, when fed high-fat diet, Ossabaw swine develop more features of MetS, greater native CAD, and greater stent-induced CAD than do Yucatan swine. Animals of each breed were divided randomly into 2 groups and fed 2 different calorie-matched diets for 40 wk: control diet (C) and high-fat, high-cholesterol atherogenic diet (H). A bare metal stent was placed in the circumflex artery, and pigs were allowed to recover for 3 wk. Characteristics of MetS, macrovascular and microvascular CAD, in-stent stenosis, and Ca2+ signaling in coronary smooth muscle cells were evaluated. MetS characteristics including, obesity, glucose intolerance, hyperinsulinemia, and elevated arterial pressure were elevated in Ossabaw swine compared to Yucatan swine. Ossabaw swine with MetS had more extensive and diffuse native CAD and in-stent stenosis and impaired coronary blood flow regulation compared with Yucatan. In-stent atherosclerotic lesions in Ossabaw coronary arteries were less fibrous and more cellular. Coronary smooth muscle cells from Ossabaw had impaired Ca2+ efflux and intracellular sequestration versus cells from Yucatan swine. Therefore, Ossabaw swine are a superior model of MetS, subsequent CAD, and cellular Ca2+ signaling defects, whereas Yucatan swine are leaner and relatively resistant to MetS and CAD.Abbreviations: CAD, coronary artery disease; CSM, coronary smooth muscle; IVGTT, intravenous glucose tolerance test; MetS, metabolic syndrome; SERCA, sarco–endoplasmic reticulum Ca2+ ATPase; ET1, endothelin 1; SOCE, store-operated Ca2+ entryAtherosclerotic coronary artery disease (CAD) is increased at least 2-fold in patients with metabolic syndrome (MetS)27 and is accompanied by marked microvascular dysfunction that further impairs coronary blood flow.10 MetS generally is diagnosed by the presence of 3 or more of the following conditions: obesity, insulin resistance, glucose intolerance, dyslipidemia, and hypertension.17,28 There is strong support for the role of the hyperinsulinemia component of MetS in increased restenosis after percutaneous coronary interventions.74,75,84,85 Further, our group has shown that severe coronary microvascular dysfunction occurs in MetS.5 Because MetS (so-called ‘prediabetes’) affects as much as 27% of the United States population, is increasing dramatically in prevalence,94 and can progress to type 2 diabetes, there is great need for basic research using animal models that accurately mimic MetS and the accompanying CAD. Clearly, there is need for study of MetS-induced CAD and in-stent stenosis and the underlying cellular and molecular mechanisms.Mice, rats, and swine are known to recapitulate MetS;3,12,36,60,71,72 however, none of these models fully reproduce the combined symptoms of MetS and CAD. Further, transgenic mouse models are simply not adequate for coronary vascular interventions using stents identical to those used in humans,18,23,38,55,57,79,83,86 a step that is essential for translation to the clinic. Yucatan and domestic swine are commonly used large animal models for study of cardiovascular disease due to their ability to mimic the neointimal formation and thrombosis observed in humans.86 For example, several laboratories have produced severe CAD in swine,8,24,51,61,62,68,91 but through toxin-induced pancreatic β-cell ablation and feeding of an atherogenic diet, rather than as a natural development subsequent to MetS or diabetes. Currently, there is a paucity of large animal models that reproduce MetS and CAD.3Research on the obesity-prone Ossabaw miniature swine59 clearly indicates that these animals develop MetS and cardiovascular disease when fed a high-calorie atherogenic diet,4,5,9,16,19,42,50,52,83,92 Female Ossabaw swine on this type of diet nearly doubled their percentage body fat in only 9 wk, showed insulin resistance, impaired glucose tolerance, dyslipidemia (profound increase in the ratio of low-density to high-density lipoprotein cholesterol, hypertriglyceridemia), hypertension, and early coronary atherosclerosis.16 These data contrast with those from male Yucatan miniature pigs, which did not develop MetS even after 20 wk on a comparable excess calorie atherogenic diet.8,68,95 Yucatan swine do not develop MetS through diet manipulation, unlike Ossabaw swine, which consistently recapitulate all MetS characteristics. However, important differences in study design have not allowed direct comparison between Yucatan and Ossabaw swine.Cytosolic Ca2+ signaling is involved in ‘phenotypic modulation’ of coronary smooth muscle (CSM), as characterized by proliferation and migration in several in vitro cell culture models33,35,89,90 and in vivo rodent models of the peripheral circulation (for example, reference 51). The Yucatan swine model of diabetic dyslipidemia shows altered Ca2+ extrusion,96 Ca2+ sequestration by the sarcoplasmic reticulum,32,34,98 and Ca2+ influx through voltage-gated Ca2+ channels.98 Currently, Ca2+ signaling has not been compared directly between MetS Ossabaw and Yucatan swine CSM. Therefore, the purpose of the present study was to test the hypothesis that compared with Yucatan swine on calorie-matched standard chow (for example, Yucatan maintenance diet8,95) and atherogenic diets, Ossabaw swine have a greater propensity to MetS and CAD with impaired coronary microvascular dysfunction and Ca2+ handling in CSM.  相似文献   

19.
20.
Young rats treated daily with intraperitoneal 4-vinylcyclohexene diepoxide (VCD) undergo selective destruction of primordial follicles, resulting in gradual ovarian failure resembling the menopausal transition in women. To determine whether VCD has similar effects on ovaries of older rats, adult and peripubertal Sprague–Dawley rats were injected intraperitoneally daily for 30 d with vehicle or VCD at 40 or 80 mg/kg. Body weight, food intake, complete blood counts, and markers of liver injury and renal function were measured during VCD treatment. Complete gross necropsy and microscopic observations were performed on day 31, and ovarian follicles were counted. At 80 mg/kg, VCD destroyed primordial and primary follicles to a similar extent in both adult and peripubertal animals, although adult rats likely started with fewer follicles and therefore approached follicle depletion. Treatment with VCD did not affect body weight, but food intake was reduced in both adult and peripubertal rats treated with 80 mg/kg VCD. Adult rats treated with 80 mg/kg VCD had neutrophilia and increased BUN and creatinine; in addition, 4 of these rats were euthanized on days 25 or 26 due to peritonitis. VCD treatment did not increase alanine aminotransferase levels, a marker of liver injury, although the 80-mg/kg dose increased liver weights. In conclusion, VCD effectively destroys small preantral follicles in adult Sprague–Dawley rats, making them a suitable model of the menopausal transition of women. However, because adult rats were more sensitive to the irritant properties of VCD, the use of a lower dose should be considered.Abbreviations: VCD, 4-vinylcyclohexene diepoxideStudies attempting to model the human menopause have relied heavily on using animals from which the ovaries have been removed surgically (ovariectomy). This approach has important limitations because women who enter natural menopause still have ovaries, which continue to produce hormones. Therefore, studies using ovariectomized animals cannot model the hormonal changes associated with the menopausal transition and postmenopausal period. However, rodent models of the menopausal transition and menopause that more closely mimic those of women have recently been developed.32,33,36 Mice or rats treated with daily intraperitoneal injections of the chemical 4-vinylcyclohexene diepoxide (VCD) undergo selective destruction of primordial and primary follicles.25 This treatment results in a gradual onset of ovarian failure because remaining larger follicles continue to develop and then ovulate or undergo atresia until they are depleted.36 These studies also demonstrate that the length of time to ovarian failure is dependent on VCD dose and duration of treatment.33,37 Moreover, in VCD-treated mice, the resulting follicle-depleted, stroma-intact ovary retains the ability to produce androgens.36 Therefore, taken together, these characteristics indicate that VCD-treated animals could be used to model the menopausal transition of women and enable research on diseases affecting women postmenopausally.The ability of VCD to destroy preantral follicles in rats by repeated dosing has been well documented.16,23,24,37 However, to our knowledge, all of the VCD studies using rats that have been published to date have used peripubertal or young (28 to 58 d) Fisher 344 rats. Although younger animals have been useful in separating the effect of age from the effect of hormonal changes associated with VCD-induced ovarian failure,22,27,32,37 the use of older rodents may provide a more appropriate model for studying the combined effects of aging and hormonal aspects of menopause (for example, osteoporosis, cognitive decline, ovarian cancer).Both young and adult Sprague–Dawley rats have been used extensively to model menopausal effects on osteoporosis,3,4,13,38,49 brain and cognitive functioning,2,14,15,29,34 lipids and cardiovascular health,30,35,53 bladder health and incontinence,6,21,31 and breast cancer.8,18,43,44 These studies used ovariectomized Sprague–Dawley rats ranging in age from 42 to 210 d. The use of this chemically induced model of menopause would be enhanced by determining whether VCD affects Sprague–Dawley rats differently and whether VCD has deleterious effects on nonovarian tissues. Furthermore, although more than a dozen publications have reported that repeated VCD dosing does not adversely affect young rodents,19,32,33,36,56 similar data have not been reported for adult Sprague–Dawley rats. The purpose of this study was to determine whether VCD affects the ovaries of peripubertal (28 d) and adult Sprague–Dawley rats differently.  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