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 62 毫秒
1.
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   

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

3.
Nogo-A is a well-known myelin-enriched inhibitory protein for axonal growth and regeneration in the central nervous system (CNS). Besides oligodendrocytes, our previous data revealed that Nogo-A is also expressed in subpopulations of neurons including retinal ganglion cells, in which it can have a positive role in the neuronal growth response after injury, through an unclear mechanism. In the present study, we analyzed the opposite roles of glial versus neuronal Nogo-A in the injured visual system. To this aim, we created oligodendrocyte (Cnp-Cre+/−xRtn4/Nogo-Aflox/flox) and neuron-specific (Thy1-Cretg+xRtn4flox/flox) conditional Nogo-A knock-out (KO) mouse lines. Following complete intraorbital optic nerve crush, both spontaneous and inflammation-mediated axonal outgrowth was increased in the optic nerves of the glia-specific Nogo-A KO mice. In contrast, neuron-specific deletion of Nogo-A in a KO mouse line or after acute gene recombination in retinal ganglion cells mediated by adeno-associated virus serotype 2.Cre virus injection in Rtn4flox/flox animals decreased axon sprouting in the injured optic nerve. These results therefore show that selective ablation of Nogo-A in oligodendrocytes and myelin in the optic nerve is more effective at enhancing regrowth of injured axons than what has previously been observed in conventional, complete Nogo-A KO mice. Our data also suggest that neuronal Nogo-A in retinal ganglion cells could participate in enhancing axonal sprouting, possibly by cis-interaction with Nogo receptors at the cell membrane that may counteract trans-Nogo-A signaling. We propose that inactivating Nogo-A in glia while preserving neuronal Nogo-A expression may be a successful strategy to promote axonal regeneration in the CNS.In the adult mammalian central nervous system (CNS), axons have a very limited capacity to regenerate after traumatic injury. This lack of axonal regeneration is thought to be mainly due to the presence of growth-inhibiting molecules in the injured CNS environment1, 2 and due to the low intrinsic growth capacity of mature neurons.3Nogo-A is a well-studied inhibitory protein for axonal growth, plasticity and regeneration after CNS injury.4, 5 Nogo-A is predominantly expressed in oligodendrocytes in the adult CNS, where it is thought to stabilize the neuronal circuits in healthy conditions and to inhibit neurite growth and plasticity after lesion.2 Neutralizing Nogo-A by function-blocking antibodies or genetic knockout (KO) has been shown to improve axonal sprouting and regeneration in the injured spinal cord and brain.6, 7, 8, 9, 10, 11In addition to oligodendrocytes and myelin, Nogo-A is expressed in growing and immature neurons, as well as in some adult neurons.12, 13 Neurons express Nogo-A receptors such as the Nogo-66 receptor 1 (NgR1)14 and the Nogo-A-Δ20-specific sphingosine 1-phosphate receptor 2 (S1PR2).15 They can co-express them along with Nogo-A,13 an observation that raises the possibility of cis-interactions between the ligand and its receptors within or at the cell surface of the same cell. This mechanism has previously been described for axonal guidance molecules such as Ephrins and Semaphorins, and could have a major role in the neuronal response to extracellular growth inhibitors during development.16, 17In the adult CNS, the expression of neuronal Nogo-A remains elevated mainly in plastic regions such as in the hippocampus, olfactory bulb or neocortex, and in the dorsal root ganglia.12 Nogo-A and NgR1 were shown to regulate synaptic plasticity, for example, long-term potentiation in the hippocampus and in the sensory-motor cortex,18, 19, 20, 21, 22 whereas the effects of neuronal Nogo-A after injury are not yet well understood. During development, neuronal Nogo-A influences neuronal migration,23, 24 survival,25, 26 cell spreading and neurite growth.27, 28 In injured adult retinal ganglion cells (RGCs), silencing neuronal Nogo-A resulted in a marked reduction of regenerative sprouting and decreased expression of growth-associated molecules.29 Furthermore, in the optic nerve, axonal regeneration was not improved in conventional Nogo-A KO animals, in which both glial and neuronal Nogo-A were deleted.29 The present study therefore aimed to investigate whether glial and neuronal Nogo-A differently influence axonal growth in vivo using cell type-specific Nogo-A KO mouse lines and adeno-associated virus (AAV)-mediated recombination of the Nogo-A gene in neurons. The results show that significantly more axons grew through the lesion site in the oligodendrocyte-specific Nogo-A KO mice. In contrast, neuron-specific ablation of Nogo-A in RGCs reduced the number of regenerating axons after optic nerve crush injury (ONC). In summary, we show that inactivating Nogo-A specifically in oligodendrocytes appears to be the most successful strategy to promote axonal regeneration in the adult optic nerve.  相似文献   

4.
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   

5.
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   

6.
Excessive N-methyl-D-aspartate receptor (NMDAR) activation and the resulting activation of neuronal nitric oxide synthase (nNOS) cause neuronal injury. Homer1b/c facilitates NMDAR-PSD95-nNOS complex interactions, and Homer1a is a negative competitor of Homer1b/c. We report that Homer1a was both upregulated by and protected against NMDA-induced neuronal injury in vitro and in vivo. The neuroprotective activity of Homer1a was associated with NMDA-induced Ca2+ influx, oxidative stress and the resultant downstream signaling activation. Additionally, we found that Homer1a functionally regulated NMDAR channel properties in neurons, but did not regulate recombinant NR1/NR2B receptors in HEK293 cells. Furthermore, we found that Homer1a detached the physical links among NR2B, PSD95 and nNOS and reduced the membrane distribution of NMDAR. NMDA-induced neuronal injury was more severe in Homer1a homozygous knockout mice (KO, Homer1a−/−) when compared with NMDA-induced neuronal injury in wild-type mice (WT, Homer1a+/+). Additionally, Homer1a overexpression in the cortex of Homer1a−/− mice alleviated NMDA-induced neuronal injury. These findings suggest that Homer1a may be a key neuroprotective endogenous molecule that protects against NMDA-induced neuronal injury by disassembling NR2B-PSD95-nNOS complexes and reducing the membrane distribution of NMDARs.Glutamate (Glu) acts on glutamate receptors, such as the N-methyl-D-aspartate receptor (NMDAR), and leads to neuronal hyper-excitability and death in a dose-dependent manner.1 NMDAR activation induces Ca2+ influx and specifically activates neuronal nitric oxide synthase (nNOS) and downstream signaling pathways.2, 3, 4 Ca2+ influx is involved in glutamate-induced apoptosis caused by the activation of apoptosis-related signaling pathways, mitochondrial dysfunction and ROS induction.3, 4 Additionally, nNOS has been reported to contribute to NMDA-induced excitotoxicity.5, 6 Considering that direct NMDAR inhibition has not yet demonstrated favorable efficacy in most clinic trails and further considering the remarkable role of nNOS in NMDA-induced neuronal death,7 measures that can effectively protect neurons from NMDA-induced neuronal injury are urgently needed and represent a worthwhile research goal.Homer proteins belong to the postsynaptic density (PSD) family and consist of two major groups: the short-form Homer proteins (Homer1a and Ania3) and the long-form Homer proteins (Homer1b/c, Homer2 and Homer3).8 Homer1b/c has a conserved N-terminal Ena/VASP homology 1 domain and binds to group I metabotropic glutamate receptors (mGluRs), inositol triphosphate receptors and Shank family proteins.9, 10, 11, 12 Homer1b/c regulates surface receptor expression,13, 14 clustering,15 transient receptor potential family channels and mGluRs coupled to ion channels.10, 16, 17, 18, 19 Additionally, because of its C-terminal coiled-coil (CC) domains, Homer1b/c can self-multimerize, form multiprotein complexes and facilitate signal transduction to downstream pathways. Homer1a, which lacks the CC domain, is believed to compete with constitutive Homer1b/c and disrupt the association of multiple Homer1b/c complexes.Notably, Homer1b/c can interact with the Glu-induced Ca2+ influx pathway by binding to Shank, a NMDAR complex adaptor protein (NMDAR-PSD95-GKAP-Shank-Homer1b/c).12, 20 Furthermore, Homer1a also interacts with Shank, NMDA, nNOS and other Homer1b/c target proteins. Homer1a has a negative regulatory role by physically replacing certain target proteins, and is involved in the regulation of a variety of cellular and molecular functions in neurological diseases.21, 22, 23, 24, 25 Nevertheless, the mechanisms of action and associations between Homer1a and NMDA-induced neuronal injury have not yet been studied. Here, we aimed to investigate the possible neuroprotective effects of Homer1a and explore the mechanisms underlying Homer1a activity in NMDA-induced neuronal injury.  相似文献   

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

8.
Well-balanced mitochondrial fission and fusion processes are essential for nervous system development. Loss of function of the main mitochondrial fission mediator, dynamin-related protein 1 (Drp1), is lethal early during embryonic development or around birth, but the role of mitochondrial fission in adult neurons remains unclear. Here we show that inducible Drp1 ablation in neurons of the adult mouse forebrain results in progressive, neuronal subtype-specific alterations of mitochondrial morphology in the hippocampus that are marginally responsive to antioxidant treatment. Furthermore, DRP1 loss affects synaptic transmission and memory function. Although these changes culminate in hippocampal atrophy, they are not sufficient to cause neuronal cell death within 10 weeks of genetic Drp1 ablation. Collectively, our in vivo observations clarify the role of mitochondrial fission in neurons, demonstrating that Drp1 ablation in adult forebrain neurons compromises critical neuronal functions without causing overt neurodegeneration.In addition to their crucial importance in energy conversion, mitochondria serve many other housekeeping functions, including calcium buffering, amino-acid and steroid biosynthesis as well as fatty acids beta-oxidation and regulation of cell death. During the past decade, it has become increasingly clear that processes regulating mitochondrial morphology and ultrastructure are influenced by specific cellular requirements upon which mitochondria, in a precisely regulated manner, undergo fusion and division events.1 Maintaining this balance is especially important for highly energy-consuming, polarized cells such as neurons, where single organellar units sprouting from the mitochondrial network are transported along the cytoskeleton into dendrites and spines to meet local energy requirements.2 In addition, elaborate quality-control mechanisms also rely on mitochondrial dynamics: whereas defective organelles are sequestered by fission, enabling their removal from the mitochondrial network,3, 4 fusion supports qualitative homogeneity of the syncytium through complementation.5Mitochondrial fusion and fission are mediated by large GTPases of the dynamin superfamily.6 The outer mitochondrial membrane mitofusins 1 (MFN1) and 2 (MFN2) tether mitochondrial membranes by homodimer or heterodimer formation,7 thereby initiating fusion of the organelles, a process that also involves the inner mitochondrial membrane-associated GTPase Optic Atrophy 1.8 In addition, MFN2 also mediates contacts between mitochondria and endoplasmic reticulum.9 The only known mammalian mitochondrial fission protein, Dynamin-Related Protein 1 (Drp1), translocates upon dephosphorylation by calcineurin10 to fission sites where it binds to mitochondrial fission factor.11 Drp1 translocation is preceded by ER membranes wrapping around mitochondria to constrict the organelles,12 thereby facilitating the formation of multimeric Drp1 complexes that, upon GTP hydrolysis, further tighten to complete the process of mitochondrial fission.13Genetic evidence in mice and humans indicates that mitochondrial dynamics are crucially important in neurons: in humans, a sporadic dominant-negative DRP1 mutation caused a lethal syndromic defect with abnormal brain development;14 similarly, constitutive Drp1 knockout in the mouse brain leads to lethal neurodevelopmental defects.15, 16 Although the crucial role of Drp1 during brain development is undisputed, studies on Drp1 function in postmitotic (adult) neurons are scarce; likewise, Drp1 ablation studies in primary cultures have so far failed to yield a conclusive picture. In vitro, Drp1 ablation is reported to lead to a super-elongated neuroprotective17, 18, 19, 20, 21, 22, 23, 24 or an aggregated mitochondrial phenotype associated with neurodegeneration.15, 16, 25, 26, 27 These discrepancies are probably due to different experimental conditions: neuronal health is indeed influenced by the onset and duration of Drp1 inhibition, which varies considerably among the cited reports,28 and different types of neuronal cultures studied display different sensitivity to Drp1 inhibition. In vivo, Drp1 ablation in Purkinje cells results in oxidative stress and neurodegeneration,29 demonstrating that Drp1 is essential for postmitotic neurons'' health. In contrast, transient pharmacological Drp1 inhibition is neuroprotective in several mouse ischemia models, indicating that temporarily blocking mitochondrial fission holds therapeutic potential.30, 31, 32To elucidate the consequences of blocked mitochondrial fission in the central nervous system in vivo, we bypassed the critical role of Drp1 during brain development by generating Drp1flx/flx mice15 expressing tamoxifen-inducible Cre recombinase under the control of the CaMKIIα promoter.33 Upon induced Drp1 deletion in postmitotic adult mouse forebrain neurons, mice develop progressive, neuronal subtype-specific alterations in mitochondrial shape and distribution in the absence of overt neurodegeneration. In addition, respiratory capacity, ATP content, synaptic reserve pool vesicle recruitment as well as spatial working memory are impaired, demonstrating that severely dysregulated mitochondrial dynamics can compromise critical neuronal functions in vivo without causing neuronal cell death.  相似文献   

9.
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   

10.
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   

11.
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   

12.
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   

13.
Acute glaucoma is a leading cause of irreversible blindness in East Asia. The mechanisms underlying retinal neuronal injury induced by a sudden rise in intraocular pressure (IOP) remain obscure. Here we demonstrate that the activation of CXCL10/CXCR3 axis, which mediates the recruitment and activation of inflammatory cells, has a critical role in a mouse model of acute glaucoma. The mRNA and protein expression levels of CXCL10 and CXCR3 were significantly increased after IOP-induced retinal ischemia. Blockade of the CXCR3 pathway by deleting CXCR3 gene significantly attenuated ischemic injury-induced upregulation of inflammatory molecules (interleukin-1β and E-selectin), inhibited the recruitment of microglia/monocyte to the superficial retina, reduced peroxynitrite formation, and prevented the loss of neurons within the ganglion cell layer. In contrast, intravitreal delivery of CXCL10 increased leukocyte recruitment and retinal cell apoptosis. Inhibition of endoplasmic reticulum (ER) stress with chemical chaperones partially blocked ischemic injury-induced CXCL10 upregulation, whereas induction of ER stress with tunicamycin enhanced CXCL10 expression in retina and primary retinal ganglion cells. Interestingly, deleting CXCR3 attenuated ER stress-induced retinal cell death. In conclusion, these results indicate that ER stress-medicated activation of CXCL10/CXCR3 pathway has an important role in retinal inflammation and neuronal injury after high IOP-induced ischemia.Acute glaucoma is the major form of glaucoma in East Asia where it is a leading cause of irreversible blindness.1 In Western countries, it is less common, but it still has much higher rate to induce vision impairment and blindness than open-angle glaucoma.2 Acute glaucoma is a medical emergent condition when intraocular pressure (IOP) is suddenly increased because of blocked drainage canals.1, 2 Prompt treatment is needed to avoid irreversible glaucomatous optic nerve damage.1 Nevertheless, in a substantial portion of patients, acute glaucoma continues progressing to blindness in spite of intensive medical treatment.3 A rapid rise in IOP that exceeds retinal perfusion pressure is known to cause retinal ischemia and induce retinal neuronal cell death.2, 4 However, the mechanisms by which elevated IOP induces retinal neuronal injury in acute glaucoma are largely unknown.Inflammation is the body''s defense system against pathogens,5 whereas excessive or uncontrolled inflammation induces tissue injury and results in diseases. In the central nervous system (CNS), inflammation has been recognized as a key player in many neurodegenerative diseases, such as Alzheimer''s disease, Parkinson''s disease, and Huntington''s disease.6 Inflammation is involved in the development of glaucoma given that the levels of inflammatory cytokines (e.g., TNF-α, interleukin-6, CCL2) and adhesion molecules (e.g., p-selectin) are increased in retina during glaucoma.7, 8 Moreover, in an animal model of acute glaucoma, inflammatory pathways including toll-like receptors and inflammasome are activated and contribute to retinal neuronal injury.4, 9 Nonetheless, the key mediators that control inflammatory cell recruitment and activation in this process remain to be elucidated.Chemokines are a family of pro-inflammatory peptides (8-15 kD) that are produced locally in tissues and mediate leukocyte and microglia directional migration and activation during inflammatory reactions by binding to specific chemokine receptors in the membrane of their respective target cells. CXCR3 is the common receptor for three chemokines belonging to the CXC subclass, namely CXCL9, CXCL10, and CXCL11,10 whereas it binds to CXCL4 with low affinity.11, 12, 13 CXCR3 is critical to mediate the recruitment of activated T cells and microglia/macrophage.10, 14, 15, 16, 17 CXCL10 expression is increased in the central nervous system in neurodegenerative diseases including Alzheimer''s disease and multiple sclerosis.18, 19, 20, 21 The activation of CXCL10/CXCR3 axis has also been shown to promote microglia recruitment and induce neuronal cell death in several models of neurodegeneration.22, 23 The role of CXCL10/CXCR3 pathway in glaucomatous optic neuropathy is unknown.In this study, we demonstrate that the CXCL10/CXCR3 axis is activated in a mouse model of acute glaucoma and the activation of this pathway is essential for retinal inflammation and neuronal injury. Moreover, the upregulation of CXCL10 and CXCR3 expression in retina is at least partially mediated by endoplasmic reticulum (ER) stress after ischemia.  相似文献   

14.
15.
Synaptic dysfunction and degeneration is an early pathological feature of aging and age-related diseases, including Alzheimer''s disease (AD). Aging is associated with increased generation and deposition of advanced glycation endproducts (AGEs), resulting from nonenzymatic glycation (or oxidation) proteins and lipids. AGE formation is accelerated in diabetes and AD-affected brain, contributing to cellular perturbation. The extent of AGEs'' involvement, if at all, in alterations in synaptic structure and function is currently unknown. Here we analyze the contribution of neuronal receptor of AGEs (RAGE) signaling to AGE-mediated synaptic injury using novel transgenic neuronal RAGE knockout mice specifically targeted to the forebrain and transgenic mice expressing neuronal dominant-negative RAGE (DN-RAGE). Addition of AGEs to brain slices impaired hippocampal long-term potentiation (LTP). Similarly, treatment of hippocampal neurons with AGEs significantly decreases synaptic density. Such detrimental effects are largely reversed by genetic RAGE depletion. Notably, brain slices from mice with neuronal RAGE deficiency or DN-RAGE are resistant to AGE-induced LTP deficit. Further, RAGE deficiency or DN-RAGE blocks AGE-induced activation of p38 signaling. Taken together, these data show that neuronal RAGE functions as a signal transducer for AGE-induced synaptic dysfunction, thereby providing new insights into a mechanism by which the AGEs–RAGE-dependent signaling cascade contributes to synaptic injury via the p38 MAP kinase signal transduction pathway. Thus, RAGE blockade may be a target for development of interventions aimed at preventing the progression of cognitive decline in aging and age-related neurodegenerative diseases.Advanced glycation endproducts (AGEs) are members of a heterogeneous class of molecules, which modify cellular function by distinct mechanisms, including ligation and activation of signal transduction receptors. The products of non-enzymatic glycation (or oxidation) of proteins and lipids, AGEs contribute to the normal aging process and when accelerated have a causative role in the vasculature complications of diabetes mellitus and several neurodegenerative diseases, including Alzheimer''s (AD), Parkinson''s, and Huntington''s diseases.1, 2, 3, 4, 5 In diabetic patients, the concentration of circulating AGEs (serum AGE level) has been reported at 7.2–22 mU/ml (equivalent to 30–88 μg/ml AGE-BSA), which is significantly higher than that of non-diabetic patients (3 mU/ml, equivalent to 12 μg/ml AGE-BSA).6, 7, 8 The brain AGE level was also increased to 5-6 μM (equivalent to 325–390 μg/ml AGE-BSA) in the diabetic animal model.9 Excess AGE accumulation is detrimental to neurons and is believed to be a key to the pathogenesis of cognitive decline in normal aging and specific chronic diseases of aging. For example, in a recent clinical study, peripheral AGE levels were associated with cognitive decline in older adults with and without diabetes.10 Diabetes complications affect the brain, increasing risk for depression, dementia, and AD. In fact, patients with type 2 diabetes are at twofold to threefold increased relative risk for AD11, 12, 13, 14, 15, 16, 17, 18 and accelerated cognitive dysfunction.Long-lived proteins such as β-amyloid peptide (Aβ) and hyperphosphorylated tau protein that accumulate in AD brain are highly susceptible to AGE modification.19, 20, 21, 22 AGE-modified Aβ or tau protein results in increased oxidative stress and chronic inflammation, accelerating AD pathology and neuronal perturbation.19, 20, 22, 23, 24, 25 Moreover, Aβ or tau glycation results in increased aggregation and subsequent formation of senile plaques or neurofibrillary tangles, the major pathological feature of AD,19, 22 suggesting that AGE modification is an important risk factor for neurodegenerative diseases.26 Although increased accumulation of AGEs in brain, as seen in aging, diabetes, or neurodegenerative diseases, speeds up oxidative damage to neurons contributing to synaptic dysfunction and cognitive decline, its underlying mechanisms are not well understood.Receptor for advanced glycation endproduct (RAGE) was first identified as a cell surface receptor of the immunoglobulin superfamily for AGEs.27, 28 Increased expression of RAGE occurs in neuronal and non-neuronal cells in the peripheral and central nervous system in aging, diabetes, and AD-affected individuals, where RAGE ligands are upregulated.29, 30 Although it has been shown that AGEs–RAGE interaction contributes to cellular perturbation relevant to the pathogenesis of the cardiovascular disease and the diabetes vascular complications,31, 32, 33 little is known about the role of AGEs and its interaction with RAGE on synaptic dysfunction. To understand the mechanisms involved in AGE-mediated synaptic damage, the following questions need to be addressed: (1) ‘Do AGEs alter synaptic structure and function? If so, are these changes dependent on RAGE signaling?'' (2) ‘Does RAGE blockage by genetic depletion protect from AGE-induced synaptic dysfunction and loss?'' and (3) ‘What is the impact of neuronal RAGE in AGE-induced aberrant synaptic function?''. Thus it is important to evaluate the impact of AGEs–RAGE interaction on synaptic dysfunction and to explore the mechanism underlying AGE–RAGE-dependent signal transduction and its contribution to synaptic damage.Here we investigate neuronal RAGE signaling in AGE-induced synaptic injury using our novel conditional RAGE knockout mice targeted to cortical neurons as well as transgenic mice that overexpress signal transduction-deficient mutants of RAGE in neurons. Given that neuronal and non-neuronal cells in the brain may contribute to AGE-induced sustained neuronal and synaptic stress and dysfunction, we assessed the impact of global RAGE deletion in this setting and further delineated the mechanism by which RAGE-dependent activation of p38 MAP kinase potentiates AGE-insulted synaptic injury.  相似文献   

16.
Tissue-type plasminogen activator (tPA) is a pleiotropic serine protease of the central nervous system (CNS) with reported neurotrophic and neurotoxic functions. Produced and released under its single chain form (sc), the sc-tPA can be cleaved by plasmin or kallikrein in a two chain form, tc-tPA. Although both sc-tPA and tc-tPA display a similar fibrinolytic activity, we postulated here that these two conformations of tPA (sc-tPA and tc-tPA) could differentially control the effects of tPA on neuronal survival. Using primary cultures of mouse cortical neurons, our present study reveals that sc-tPA is the only one capable to promote N-methyl-D-aspartate receptor (NMDAR)-induced calcium influx and subsequent excitotoxicity. In contrast, both sc-tPA and tc-tPA are capable to activate epidermal growth factor receptors (EGFRs), a mechanism mediating the antiapoptotic effects of tPA. Interestingly, we revealed a tPA dependent crosstalk between EGFR and NMDAR in which a tPA-dependent activation of EGFRs leads to downregulation of NMDAR signaling and to subsequent neurotrophic effects.Tissue-type plasminogen activator (tPA) is secreted by endothelial cells and promotes fibrinolysis via the conversion of fibrin-bound plasminogen into plasmin.1 Neurons and some glial cells also secrete tPA.2, 3, 4, 5 tPA is secreted as a single-chain form (sc-tPA), which can be processed into a two-chain form (tc-tPA) by plasmin or kallikreins.6, 7 Interestingly, sc-tPA is proteolytically active even without proteolytic processing. In addition to its vascular functions, tPA displays critical roles in the brain parenchyma with roles in cell migration, neuronal plasticity and survival,8, 9, 10, 11, 12, 13, 14 acting either as an enzyme or as a cytokine-like molecule. Among its actions, tPA is well described to promote neurotoxicity, likely through promotion of N-methyl-D-aspartate receptor (NMDAR) activity.15, 16, 17 Recently, we reported that only sc-tPA can promote NMDAR signaling and neurotoxicity.18 Interestingly, data from wild-type mice,19 transgenic mice overexpressing tPA in neurons20 or in vitro21 also report neuroprotective effects of tPA.9, 10 The proposed mechanisms involved a tPA-dependent and non-proteolytic activation of either epidermal growth factor receptors (EGFRs)22 on oligodendrocytes or NMDARs.20Here we explored a link between tPA conformations (sc-tPA and tc-tPA), EGFR- and NMDAR-dependent signaling pathways. Our findings identify sc-tPA as a selective positive modulator of NMDAR signaling in neurons when present at high concentrations and both sc-tPA and tc-tPA as positive modulators of EGFR signaling, this even at low concentrations. We also reveal a crosstalk between these two families of receptors, with the tPA-dependent activation of EGFRs reducing NMDAR signaling. By these mechanisms, sc-tPA and tc-tPA control neuronal death and survival.  相似文献   

17.
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   

18.
19.
A synthetic peptide (sPIF) analogous to the mammalian embryo-derived PreImplantation Factor (PIF) enables neuroprotection in rodent models of experimental autoimmune encephalomyelitis and perinatal brain injury. The protective effects have been attributed, in part, to sPIF''s ability to inhibit the biogenesis of microRNA let-7, which is released from injured cells during central nervous system (CNS) damage and induces neuronal death. Here, we uncover another novel mechanism of sPIF-mediated neuroprotection. Using a clinically relevant rat newborn brain injury model, we demonstrate that sPIF, when subcutaneously administrated, is able to reduce cell death, reverse neuronal loss and restore proper cortical architecture. We show, both in vivo and in vitro, that sPIF activates cyclic AMP dependent protein kinase (PKA) and calcium-dependent protein kinase (PKC) signaling, leading to increased phosphorylation of major neuroprotective substrates GAP-43, BAD and CREB. Phosphorylated CREB in turn facilitates expression of Gap43, Bdnf and Bcl2 known to have important roles in regulating neuronal growth, survival and remodeling. As is the case in sPIF-mediated let-7 repression, we provide evidence that sPIF-mediated PKA/PKC activation is dependent on TLR4 expression. Thus, we propose that sPIF imparts neuroprotection via multiple mechanisms at multiple levels downstream of TLR4. Given the recent FDA fast-track approval of sPIF for clinical trials, its potential clinical application for treating other CNS diseases can be envisioned.Perinatal brain injury in the context of premature birth is a major cause of neonatal morbidity and mortality.1 Depending on the degree of prematurity, 15–20% of the affected newborns die during the postnatal period and ~25% of survivors suffer significant long-term disability including cerebral palsy, epilepsy and increased hyperactivity.2 Therapeutic approaches to counteract the disastrous cascades of neonatal brain injury have been proposed. Unfortunately, in premature infants at risk, no neuroprotective agent has proven safe and effective so far.3Secreted from developing embryos, PreImplantation Factor (PIF) can be detected in the maternal circulation during pregnancy,4, 5 and its presence has been correlated with live birth.5, 6, 7 PIF has been implicated in promoting embryo implantation through modulating maternal immune tolerance.5, 8, 9 Consistent with the immune function, a synthetic PIF analog (sPIF) of 15 amino acids (MVRIKPGSANKPSDD) that was subcutaneously administrated was able to both reverse and prevent paralysis through inhibiting neuroinflammation in a murine model of experimental autoimmune encephalomyelitis.10 The neuroprotective property of sPIF was further underscored by its ability to mitigate neuronal loss and microglial activation in a rat model of perinatal brain injury.11 The neuroprotective effects were attributed, at least in part, to sPIF''s ability to downregulate microRNA let-7 in the injured brain. Abundantly expressed in the central nervous system (CNS), let-7 released from dying cells during brain injury induces neuronal death, exacerbating CNS damage.12, 13 sPIF inhibited the biogenesis of let-7 in both neuronal and immune cells through Toll-like Receptor 4 (TLR4).11 As PIF imparts multitargeted effects,10 it is almost certain that inhibiting let-7 is not the only mechanism of PIF action.In search of additional mechanisms, we chose to focus on cyclic AMP-dependent protein kinase (PKA) and calcium-dependent protein kinase (PKC). PKA/PKC signaling is downstream of TLR4,14, 15 and TLR4 was required for sPIF-induced neuroprotective effects.11 PKA/PKC are important signaling molecules in a variety of cellular functions, including cell growth and differentiation, neuronal plasticity and cellular response to hypoxia–ischemia.16, 17, 18, 19 Mechanistically, PKA/PKC activation leads to phosphorylation of serine and threonine residues on target proteins, thereby modulating protein stability, protein–protein interactions and catalytic activity.20 In the case of brain injury, activation of the PKA/PKC signaling pathways imparts neuroprotection by increasing expression of anti-apoptotic and neurotrophic molecules while reducing pro-apoptotic molecules in neurons.21, 22, 23 Not surprisingly, PKA/PKC pathways have been recognized as potent targets for neuroprotective strategies.In the current study, we have examined and revealed a novel mechanism of PIF action. sPIF confers neuroprotection in a rat model of perinatal brain injury by modulating PKA/PKC signaling, which is recapitulated in vitro using neuronal cells. Overall, our data support clinical translation of sPIF treatment for hypoxic–ischemic brain injuries.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号