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 867 毫秒
1.
2.
The increase of extracellular heme is a hallmark of hemolysis or extensive cell damage. Heme has prooxidant, cytotoxic, and inflammatory effects, playing a central role in the pathogenesis of malaria, sepsis, and sickle cell disease. However, the mechanisms by which heme is sensed by innate immune cells contributing to these diseases are not fully characterized. We found that heme, but not porphyrins without iron, activated LPS-primed macrophages promoting the processing of IL-1β dependent on nucleotide-binding domain and leucine rich repeat containing family, pyrin domain containing 3 (NLRP3). The activation of NLRP3 by heme required spleen tyrosine kinase, NADPH oxidase-2, mitochondrial reactive oxygen species, and K+ efflux, whereas it was independent of heme internalization, lysosomal damage, ATP release, the purinergic receptor P2X7, and cell death. Importantly, our results indicated the participation of macrophages, NLRP3 inflammasome components, and IL-1R in the lethality caused by sterile hemolysis. Thus, understanding the molecular pathways affected by heme in innate immune cells might prove useful to identify new therapeutic targets for diseases that have heme release.Hemolysis, hemorrhage, and rhabdomyolysis cause the release of large amounts of hemoproteins to the extracellular space, which, once oxidized, release the heme moiety, a potentially harmful molecule due to its prooxidant, cytotoxic, and inflammatory effects (1, 2). Scavenging proteins such as haptoglobin and hemopexin bind hemoglobin and heme, respectively, promoting their clearance from the circulation and delivery to cells involved with heme catabolism. Heme oxygenase cleaves heme and generates equimolar amounts of biliverdin, carbon monoxide (CO) and iron (2). Studies using mice deficient for haptoglobin (Hp), hemopexin (Hx), and heme oxygenase 1 (HO-1) demonstrate the importance of these proteins in controlling the deleterious effects of heme. Both Hp−/− and Hx−/− mice have increased renal damage after acute hemolysis induced by phenyhydrazine (Phz) compared with wild-type mice (3, 4). Mice lacking both proteins present splenomegaly and liver inflammation composed of several foci with leukocyte infiltration after intravascular hemolysis (5). Hx protect mice against heme-induced endothelial damage improving liver and cardiovascular function (68). Lack of heme oxygenase 1 (Hmox1−/−) causes iron overload, increased cell death, and tissue inflammation under basal conditions and upon inflammatory stimuli (915). This salutary effect of HO-1 has been attributed to its effect of reducing heme amounts as well as generating the cytoprotective molecules, biliverdin and CO.Heme induces neutrophil migration in vivo and in vitro (16, 17), inhibits neutrophil apoptosis (18), triggers cytokine and lipid mediator production by macrophages (19, 20), and increases the expression of adhesion molecules and tissue factor on endothelial cells (2123). Heme cooperates with TNF, causing hepatocyte apoptosis in a mechanism dependent on reactive oxygen species (ROS) generation (12). Whereas heme-induced TNF production depends on functional toll-like receptor 4 (TLR4), ROS generation in response to heme is TLR4 independent (19). We recently observed that heme triggers receptor-interacting protein (RIP)1/3-dependent macrophage-programmed necrosis through the induction of TNF and ROS (15). The highly unstable nature of iron is considered critical for the ability of heme to generate ROS and to cause inflammation. ROS generated by heme has been mainly attributed to the Fenton reaction. However, recent studies suggest that heme can generate ROS through multiple sources, including NADPH oxidase and mitochondria (22, 2427).Heme causes inflammation in sterile and infectious conditions, contributing to the pathogenesis of hemolytic diseases, subarachnoid hemorrhage, malaria, and sepsis (11, 13, 24, 28), but the mechanisms by which heme operates in different conditions are not completely understood. Blocking the prooxidant effects of heme protects cells from death and prevents tissue damage and lethality in models of malaria and sepsis (12, 13, 15). Importantly, two recent studies demonstrated the pathogenic role of heme-induced TLR4 activation in a mouse model of sickle cell disease (29, 30). These results highlight the great potential of understanding the molecular mechanisms of heme-induced inflammation and cell death as a way to identify new therapeutic targets.Hemolysis and heme synergize with microbial molecules for the induction of inflammatory cytokine production and inflammation in a mechanism dependent on ROS and Syk (24). Processing of pro–IL-1β is dependent on caspase-1 activity, requiring assembly of the inflammasome, a cytosolic multiprotein complex composed of a NOD-like receptor, the adaptor protein apoptosis-associated speck-like protein containing a CARD (ASC), and caspase-1 (3133). The most extensively studied inflammasome is the nucleotide-binding domain and leucine rich repeat containing family, pyrin domain containing 3 (NLRP3). NLRP3 and pro–IL-1β expression are increased in innate immune cells primed with NF-κB inducers such as TLR agonists and TNF (34, 35). NLRP3 inflammasome is activated by several structurally nonrelated stimuli, such as endogenous and microbial molecules, pore-forming toxins, and particulate matter (34, 35). The activation of NLRP3 involves K+ efflux, increase of ROS and Syk phosphorylation. Importantly, critical roles of NLRP3 have been demonstrated in a vast number of diseases (34, 36). We hypothesize that heme causes the activation of the inflammasome and secretion of IL-1β. Here we found that heme triggered the processing and secretion of IL-1β dependently on NLRP3 inflammasome in vitro and in vivo. The activation of NLRP3 by heme was dependent on Syk, ROS, and K+ efflux, but independent of lysosomal leakage, ATP release, or cell death. Finally, our results indicated the critical role of macrophages, the NLRP3 inflammasome, and IL-1R to the lethality caused by sterile hemolysis.  相似文献   

3.
Pathogen recognition by nucleotide-binding oligomerization domain-like receptor (NLR) results in the formation of a macromolecular protein complex (inflammasome) that drives protective inflammatory responses in the host. It is thought that the number of inflammasome complexes forming in a cell is determined by the number of NLRs being activated, with each NLR initiating its own inflammasome assembly independent of one another; however, we show here that the important foodborne pathogen Salmonella enterica serovar Typhimurium (S. Typhimurium) simultaneously activates at least two NLRs, whereas only a single inflammasome complex is formed in a macrophage. Both nucleotide-binding domain and leucine-rich repeat caspase recruitment domain 4 and nucleotide-binding domain and leucine-rich repeat pyrin domain 3 are simultaneously present in the same inflammasome, where both NLRs are required to drive IL-1β processing within the Salmonella-infected cell and to regulate the bacterial burden in mice. Superresolution imaging of Salmonella-infected macrophages revealed a macromolecular complex with an outer ring of apoptosis-associated speck-like protein containing a caspase activation and recruitment domain and an inner ring of NLRs, with active caspase effectors containing the pro–IL-1β substrate localized internal to the ring structure. Our data reveal the spatial localization of different components of the inflammasome and how different members of the NLR family cooperate to drive robust IL-1β processing during Salmonella infection.Inflammasomes are cytosolic multimeric protein complexes formed in the host cell in response to the detection of pathogen-associated molecular patterns (PAMPs) or danger-associated molecular patterns (DAMPs). Formation of the inflammasome in response to PAMPs is critical for host defense because it facilitates processing of the proinflammatory cytokines pro–IL-1β and pro–IL-18 into their mature forms (1). The inflammasome also initiates host cell death in the form of pyroptosis, releasing macrophage-resident microbes to be killed by other immune mechanisms (2). The current paradigm is that there are individual, receptor-specific inflammasomes consisting of one nucleotide-binding oligomerization domain-like receptor (NLR; leucine-rich repeat–containing) or PYHIN [pyrin domain and hematopoietic expression, interferon-inducible nature, and nuclear localization (HIN) domain-containing] receptor, the adaptor protein apoptosis-associated speck-like protein containing a caspase activation and recruitment domain (CARD; ASC), and caspase-1 (3). How the protein constituents of the inflammasome are spatially orientated is unclear.Nucleotide-binding domain and leucine-rich repeat caspase recruitment domain 4 (NLRC4) and nucleotide-binding domain and leucine-rich repeat pyrin domain 3 (NLRP3) are the best-characterized inflammasomes, especially with respect to their responses to pathogenic bacteria. The NLRC4 inflammasome is activated primarily by bacteria, including Aeromonas veronii (4), Escherichia coli (5), Listeria monocytogenes (6, 7), Pseudomonas aeruginosa (5), Salmonella enterica serovar Typhimurium (S. Typhimurium) (5, 810), and Yersinia species (11). In mouse macrophages, the NLRC4 inflammasome responds to flagellin and type III secretion system-associated needle or rod proteins (5, 8, 9) after their detection by NLR family, apoptosis inhibitory protein (NAIP) 5 or NAIP6 and NAIP1 or NAIP2, respectively (1215). Phosphorylation of NLRC4 at a single, evolutionarily conserved residue, Ser 533, by PKCδ kinase is required for NLRC4 inflammasome assembly (16). The NLRP3 inflammasome is activated by a large repertoire of DAMPs, including ATP, nigericin, maitotoxin, uric acid crystals, silica, aluminum hydroxide, and muramyl dipeptide (1720). NLRP3 is also activated by bacterial PAMPs from many species, including Aeromonas species (4, 21), L. monocytogenes (6, 7, 22), Neisseria gonorrhoeae (23), S. Typhimurium (10), Streptococcus pneumoniae (24), and Yersinia species (11). The mechanisms by which NLRC4 and NLRP3 inflammasomes contribute to host defense against bacterial pathogens are emerging; however, little is known about the dynamics governing inflammasome assembly in infections caused by bacteria that activate multiple NLRs, such as S. Typhimurium (10), A. veronii (4), and Yersinia (11).NLRP3 does not have a CARD and requires ASC to interact with the CARD of procaspase-1. This interaction requires a charged interface around Asp27 of the procaspase-1 CARD (25). Whether ASC is also required for the assembly of the NLRC4 inflammasome is less clear. NLRC4 contains a CARD that can interact directly with the CARD of procaspase-1 (26); however, ASC is required for some of the responses driven by NLRC4 (27). Macrophages infected with S. Typhimurium or other pathogens exhibit formation of a distinct cytoplasmic ASC focus or speck, which can be visualized under the microscope and is indicative of inflammasome activation (10, 28, 29). Our laboratory and others have shown that only one ASC speck is formed per cell irrespective of the stimulus used (2932). However, many bacteria activate two or more NLRs, and it is unclear whether a singular inflammasome is formed at a time or if multiple inflammasomes are formed independent of each other, with each inflammasome containing one member of the NLR family.In this study, we describe the endogenous molecular constituents of the Salmonella-induced inflammasome and their spatial orientation. In cross-section, ASC forms a large external ring with the NLRs and caspases located internally. Critically, NLRC4, NLRP3, caspase-1, and caspase-8 coexist in the same ASC speck to coordinate pro–IL-1β processing. All ASC specks observed contained both NLRC4 and NLRP3. These results suggest that Salmonella infection induces a single inflammasome protein complex containing different NLRs and recruiting multiple caspases to coordinate a multifaceted inflammatory response to infection.  相似文献   

4.
Inflammasomes are caspase-1–activating multiprotein complexes. The mouse nucleotide-binding domain and leucine rich repeat pyrin containing 1b (NLRP1b) inflammasome was identified as the sensor of Bacillus anthracis lethal toxin (LT) in mouse macrophages from sensitive strains such as BALB/c. Upon exposure to LT, the NLRP1b inflammasome activates caspase-1 to produce mature IL-1β and induce pyroptosis. Both processes are believed to depend on autoproteolysed caspase-1. In contrast to human NLRP1, mouse NLRP1b lacks an N-terminal pyrin domain (PYD), indicating that the assembly of the NLRP1b inflammasome does not require the adaptor apoptosis-associated speck-like protein containing a CARD (ASC). LT-induced NLRP1b inflammasome activation was shown to be impaired upon inhibition of potassium efflux, which is known to play a major role in NLRP3 inflammasome formation and ASC dimerization. We investigated whether NLRP3 and/or ASC were required for caspase-1 activation upon LT stimulation in the BALB/c background. The NLRP1b inflammasome activation was assessed in both macrophages and dendritic cells lacking either ASC or NLRP3. Upon LT treatment, the absence of NLRP3 did not alter the NLRP1b inflammasome activity. Surprisingly, the absence of ASC resulted in IL-1β cleavage and pyroptosis, despite the absence of caspase-1 autoprocessing activity. By reconstituting caspase-1/caspase-11−/− cells with a noncleavable or catalytically inactive mutant version of caspase-1, we directly demonstrated that noncleavable caspase-1 is fully active in response to the NLRP1b activator LT, whereas it is nonfunctional in response to the NLRP3 activator nigericin. Taken together, these results establish variable requirements for caspase-1 cleavage depending on the pathogen and the responding NLR.Anthrax is a zoonotic disease caused by the Gram-positive bacterium Bacillus anthracis. B. anthracis provokes a shock-like syndrome that can prove fatal to the host (1) and has recently gained notoriety as a potential bioterrorism agent. Anthrax pathogenicity relies on its ability to secrete three virulence proteins, which combine with each other to form two toxins. The protective antigen (PA) combines with the edema factor (EF) to form the edema toxin (2, 3). EF is an adenylate cyclase that causes edema of the infected tissue. The binary combination of PA with lethal factor (LF) gives rise to the most virulent factor, called lethal toxin (LT), responsible for the systemic symptoms and death of the infected animal. To escape the host immune response, LT impairs the host innate immunity by killing macrophages (46). The PA protein interacts with LF and binds to cell surface receptors, enabling endocytosis of the LT complex. In the acidic compartment, PA forms pores allowing the delivery of LF to the cytosol. LF is a zinc metalloprotease that was shown to cleave the N-terminal region of many MAP kinase kinases and to induce apoptosis of macrophages. LT also triggers pyroptosis through the formation of a caspase-1–activating platform, named “inflammasome” (68).Inflammasomes are multiprotein complexes of the innate immune response that control caspase-1 activity and pro–IL-1β and pro–IL-18 maturation. Most inflammasomes are composed of specific cytosolic pathogen recognition receptors (PRRs), as well as the apoptosis-associated speck-like protein containing a caspase activation and recruitment domain (CARD) (ASC) adaptor protein that enables the recruitment and activation of the caspase-1 protease. Once caspase-1 is oligomerized within an inflammasome platform, the enzyme undergoes autoproteolysis to form heterodimers of active caspase-1 (912). In the mouse, at least five distinct inflammasomes have been described, distinguished by the PRR that induces the complex formation. The PRRs capable of participating in inflammasome platform formation are either members of the nod-like receptor (NLR) family (e.g., NLRP1, NLRP3, or NLRC4) or of the PYrin and HIN (PYHIN) family (e.g., AIM2) (13, 14). ASC is composed of a pyrin domain (PYD) and a caspase activation and recruitment domain (CARD). ASC interacts with a PYD-containing PRR via its PYD domain and recruits the CARD domain of caspase-1 via its CARD domain. Thus, ASC is essential to the formation of the inflammasome by receptors such as NLRP3 or AIM2 (1518). However, its presence is dispensable for NLRC4, which contains a CARD in place of a PYD, allowing direct interaction with the CARD domain of caspase-1 (19, 20).Past studies have determined that certain mouse strains are more sensitive than others to LT cytotoxicity, and genetic studies identified NLRP1b as the factor conferring mouse strain susceptibility to anthrax LT (21). The mouse genome contains three different NLRP1 isoforms (a, b, and c) and a functional NLRP1b was found to be expressed by the mouse strains sensitive to LT (e.g., BALB/c or 129 background). Expression of NLRP1b was shown to mediate IL-1β release and caspase-1–mediated cell death in response to LT (7, 21, 22). Mouse NLRP1b differs structurally from human NLRP1 in that it lacks the N-terminal PYD (23). The absence of the PYD suggests that NLRP1b can directly engage caspase-1 without a requirement for ASC. However, studies dissecting the mechanism of NLRC4 inflammasome activation demonstrated that ASC is required for the amplification of caspase-1 autoprocessing and IL-1β secretion but not for pyroptosis (19, 20). Cell lysis mediated by LT was shown to be dependent on sodium and potassium fluxes (24), and high extracellular potassium inhibited IL-1β secretion upon LT treatment, suggesting a role for the NLRP3 inflammasome in LT sensing (22, 25). Therefore, we investigated whether NLRP3 and/or ASC were required for caspase-1 activation in response to LT. The NLRP3, ASC, and caspase-1 mouse knockout strains were backcrossed into the BALB/c background and the response of macrophages and dendritic cells (DCs) to LT intoxication was studied. Our data reveal that (i) in response to LT, ASC is dispensable for caspase-1 activation, but uncleavable caspase-1 is fully active; and (ii) upon activation of the NLRP3 inflammasome, uncleavable caspase-1 is inactive.  相似文献   

5.
Inflammasomes are multiprotein platforms that activate caspase-1, which leads to the processing and secretion of the proinflammatory cytokines IL-1β and IL-18. Previous studies demonstrated that bacterial RNAs activate the nucleotide-binding domain, leucine-rich-repeat-containing family, pyrin domain-containing 3 (NLRP3) inflammasome in both human and murine macrophages. Interestingly, only mRNA, but neither tRNA nor rRNAs, derived from bacteria could activate the murine Nlrp3 inflammasome. Here, we report that all three types of bacterially derived RNA (mRNA, tRNA, and rRNAs) were capable of activating the NLRP3 inflammasome in human macrophages. Bacterial RNA’s 5′-end triphosphate moieties, secondary structure, and double-stranded structure were dispensable; small fragments of bacterial RNA were sufficient to activate the inflammasome. In addition, we also found that 20-guanosine ssRNA can activate the NLRP3 inflammasome in human macrophages but not in murine macrophages. Therefore, human and murine macrophages may have evolved to recognize bacterial cytosolic RNA differently during bacterial infections.The innate immune system is the first line of defense against microbial infections. Germ-line–encoded pattern-recognition receptors (PRRs) of the innate immune system recognize the presence of invariant evolutionarily conserved microbial components called “pathogen-associated molecular patterns” (13). In response to microbial infections, PRRs rapidly initiate signal-transduction pathways to induce type 1 IFN production, proinflammatory cytokine production, and inflammasome activation. The inflammasome is a cytosolic large caspase-1–containing multiprotein complex that enables autocatalytic activation of caspase-1. Once caspase-1 is activated, it starts to cleave prointerleukin-1β (pro–IL-1β) and prointerleukin-18 (pro–IL-18) proteolytically into bioactive IL-1β and IL-18 (47). The mature forms of IL-1β and IL-18 play roles in a variety of infectious and inflammatory processes.Cytosolic microbial nucleic acids are important activators of the innate immune system against both bacterial and viral infections, which induce type 1-IFN and proinflammatory cytokine responses as well as inflammasome activation. The role of microbial nucleic acids in inflammasome activation has been studied mostly in murine bone marrow-derived dendritic cells (BMDCs) or bone marrow-derived macrophages (BMDMs). AIM2 has been identified as a specific cytosolic dsDNA sensor that directly binds ASC (apoptosis-associated speck-like protein containing a carboxyl-terminal CARD-like domain) and forms inflammasome complexes in human and murine cells (811).Viral dsRNA was found to activate the nucleotide-binding domain, leucine-rich-repeat-containing family, pyrin domain-containing 3 (NLRP3) inflammasome in human and murine cells (1215). Several groups have reported that cytosolic bacterial RNA activate the Nlrp3 inflammasome in murine macrophages (13, 16, 17). Our group also has reported that human THP-1–derived macrophages recognize cytosolic bacterial RNA and induce NLRP3 inflammasome activation (12). Bacterial RNA is composed of mRNA, tRNA, and three different sizes of rRNA (23s, 16s, and 5s). Sander et al. (18) reported that, of the different types of Escherichia coli RNA, only E. coli mRNA induced the secretion of IL-1β by murine BMDMs, but E. coli tRNA and E. coli rRNAs did not.We aimed to study (i) whether a variety of cytosolic bacterial RNAs could activate the inflammasome in human myeloid cells and (ii) what types of bacterial RNA activate the inflammasome in human and murine myeloid cells. Here, we demonstrate that a broad spectrum of cytosolic bacterial RNAs strongly induce the cleavage of caspase-1 and the secretion of IL-1β and IL-18 in human macrophages. Human macrophages can sense mRNA, tRNA, rRNAs, and small synthetic ssRNA through NLRP3, but murine macrophages can sense only the mRNA component. Bacterial RNA’s 5′-end triphosphate moieties, secondary structure, and double-stranded structure were dispensable, but small fragments of bacterial RNA were sufficient to activate the inflammasome. These findings suggest that upon bacterial infections the human and murine NLRP3 inflammasomes sense cytosolic bacterial RNAs differently.  相似文献   

6.
7.
Physiologically, α-synuclein chaperones soluble NSF attachment protein receptor (SNARE) complex assembly and may also perform other functions; pathologically, in contrast, α-synuclein misfolds into neurotoxic aggregates that mediate neurodegeneration and propagate between neurons. In neurons, α-synuclein exists in an equilibrium between cytosolic and membrane-bound states. Cytosolic α-synuclein appears to be natively unfolded, whereas membrane-bound α-synuclein adopts an α-helical conformation. Although the majority of studies showed that cytosolic α-synuclein is monomeric, it is unknown whether membrane-bound α-synuclein is also monomeric, and whether chaperoning of SNARE complex assembly by α-synuclein involves its cytosolic or membrane-bound state. Here, we show using chemical cross-linking and fluorescence resonance energy transfer (FRET) that α-synuclein multimerizes into large homomeric complexes upon membrane binding. The FRET experiments indicated that the multimers of membrane-bound α-synuclein exhibit defined intermolecular contacts, suggesting an ordered array. Moreover, we demonstrate that α-synuclein promotes SNARE complex assembly at the presynaptic plasma membrane in its multimeric membrane-bound state, but not in its monomeric cytosolic state. Our data delineate a folding pathway for α-synuclein that ranges from a monomeric, natively unfolded form in cytosol to a physiologically functional, multimeric form upon membrane binding, and show that only the latter but not the former acts as a SNARE complex chaperone at the presynaptic terminal, and may protect against neurodegeneration.α-Synuclein is an abundant presynaptic protein that physiologically acts to promote soluble NSF attachment protein receptor (SNARE) complex assembly in vitro and in vivo (13). Point mutations in α-synuclein (A30P, E46K, H50Q, G51D, and A53T) as well as α-synuclein gene duplications and triplications produce early-onset Parkinson''s disease (PD) (410). Moreover, α-synuclein is a major component of intracellular protein aggregates called Lewy bodies, which are pathological hallmarks of neurodegenerative disorders such as PD, Lewy body dementia, and multiple system atrophy (1114). Strikingly, neurotoxic α-synuclein aggregates propagate between neurons during neurodegeneration, suggesting that such α-synuclein aggregates are not only intrinsically neurotoxic but also nucleate additional fibrillization (1518).α-Synuclein is highly concentrated in presynaptic terminals where α-synuclein exists in an equilibrium between a soluble and a membrane-bound state, and is associated with synaptic vesicles (1922). The labile association of α-synuclein with membranes (23, 24) suggests that binding of α-synuclein to synaptic vesicles, and its dissociation from these vesicles, may regulate its physiological function. Membrane-bound α-synuclein assumes an α-helical conformation (2532), whereas cytosolic α-synuclein is natively unfolded and monomeric (refs. 25, 26, 31, and 32; however, see refs. 33 and 34 and Discussion for a divergent view). Membrane binding by α-synuclein is likely physiologically important because in in vitro experiments, α-synuclein remodels membranes (35, 36), influences lipid packing (37, 38), and induces vesicle clustering (39). Moreover, membranes were found to be important for the neuropathological effects of α-synuclein (4044).However, the relation of membrane binding to the in vivo function of α-synuclein remains unexplored, and it is unknown whether α-synuclein binds to membranes as a monomer or oligomer. Thus, in the present study we have investigated the nature of the membrane-bound state of α-synuclein and its relation to its physiological function in SNARE complex assembly. We found that soluble monomeric α-synuclein assembles into higher-order multimers upon membrane binding and that membrane binding of α-synuclein is required for its physiological activity in promoting SNARE complex assembly at the synapse.  相似文献   

8.
Being activated by depolarizing voltages and increases in cytoplasmic Ca2+, voltage- and calcium-activated potassium (BK) channels and their modulatory β-subunits are able to dampen or stop excitatory stimuli in a wide range of cellular types, including both neuronal and nonneuronal tissues. Minimal alterations in BK channel function may contribute to the pathophysiology of several diseases, including hypertension, asthma, cancer, epilepsy, and diabetes. Several gating processes, allosterically coupled to each other, control BK channel activity and are potential targets for regulation by auxiliary β-subunits that are expressed together with the α (BK)-subunit in almost every tissue type where they are found. By measuring gating currents in BK channels coexpressed with chimeras between β1 and β3 or β2 auxiliary subunits, we were able to identify that the cytoplasmic regions of β1 are responsible for the modulation of the voltage sensors. In addition, we narrowed down the structural determinants to the N terminus of β1, which contains two lysine residues (i.e., K3 and K4), which upon substitution virtually abolished the effects of β1 on charge movement. The mechanism by which K3 and K4 stabilize the voltage sensor is not electrostatic but specific, and the α (BK)-residues involved remain to be identified. This is the first report, to our knowledge, where the regulatory effects of the β1-subunit have been clearly assigned to a particular segment, with two pivotal amino acids being responsible for this modulation.High-conductance voltage- and calcium-activated potassium (BK) channels are homotetrameric proteins of α-subunits encoded by the slo1 gene (1). These channels are expressed in virtually all mammalian tissues, where they detect and integrate membrane voltage and calcium concentration changes dampening the responsiveness of cells when confronted with excitatory stimuli. They are abundant in the CNS and nonneuronal tissues, such as smooth muscle or hair cells. This wide distribution is associated with an outstandingly large functional diversity, in which BK channel activity appears optimally adapted to the particular physiological demands of each cell type (2). On the other hand, small alterations in BK channel function may contribute to the pathophysiology of hypertension, asthma, cancer, epilepsy, diabetes, and other conditions in humans (38). Alternative splicing, posttranslational modifications, and regulation by auxiliary proteins have been proposed to contribute to this functional diversity (1, 2, 916).The BK channel α-subunit is formed by a single polypeptide of about 1,200 amino acids that contains all of the key structural elements for ion permeation, gating, and modulation by ions and other proteins. Tetramers of α-subunits form functional BK channels. Each subunit has seven hydrophobic transmembrane segments (S0–S6), where the voltage-sensor domain (VSD) and pore domain (PD) reside (2). The N terminus faces the extracellular side of the membrane, whereas the C terminus is intracellular. The latter contains four hydrophobic α-helices (S7–S10) and the main Ca2+ binding sites (2). VSDs formed by segments S1–S4 harbor a series of charged residues across the membrane that contributes to voltage sensing (2). Upon membrane depolarization, each VSD undergoes a rearrangement (17) that prompts the opening of a highly K+-selective pore formed by the four PDs that come together at the symmetry center of the tetramer.Although BK channel expression is ubiquitous, in most physiological scenarios their functioning is provided by their coassembly with auxiliary proteins, such as β-subunits. This coassembly brings channel activity into the proper cell/tissue context (11, 13). Four different β-subunits have been cloned (β1–β4) (1824), all of which have been observed to modify BK channel function. Albeit to a different extent, all β-subunits modify the Ca2+ sensitivity, voltage dependence, and gating properties of BK channels, hence modifying plasma membrane excitability balance. Regarding auxiliary β-subunits, β1- and β2-subunits increase apparent Ca2+ sensitivity and decelerate macroscopic current kinetics (14, 20, 21, 2530); β2 and β3 induce fast inactivation as well as an instantaneous outward rectification (20, 21, 24, 31, 32); and β4 slows down activation and deactivation kinetics (12, 23) and modifies Ca2+ sensitivity (12, 33, 34).It should be kept in mind that β-subunits are potential targets for different molecules that modulate channel function, such as alcohol (35), estrogens (15), hormones (36), and fatty acids (37, 38). Additionally, scorpion toxin affinity in BK channels would tend to increase when β1 is coexpressed with the α-subunit (22).To identify the molecular elements that give β1 the ability to modulate the voltage sensor of BK channels, we constructed chimeric proteins of β1/β2- and β1/β3-subunits by swapping their N and C termini, the transmembrane (TM) segments, and the extracellular loops and recorded their gating currents. Two lysine residues that are unique to the N terminus of β1 were identified to be sufficient for BK voltage-sensor modulation.  相似文献   

9.
10.
The role of estrogen receptor (ER) α as a target in treatment of breast cancer is clear, but those of ERβ1 and ERβ2 in the breast remain unclear. We have examined expression of all three receptors in surgically excised breast samples from two archives: (i): 187 invasive ductal breast cancer from a Japanese study; and (ii) 20 lobular and 24 ductal cancers from the Imperial College. Samples contained normal areas, areas of hyperplasia, and in situ and invasive cancer. In the normal areas, ERα was expressed in not more than 10% of epithelium, whereas approximately 80% of epithelial cells expressed ERβ. We found that whereas ductal cancer is a highly proliferative, ERα-positive, ERβ-negative disease, lobular cancer expresses both ERα and ERβ but with very few Ki67-positive cells. ERβ2 was expressed in 32% of the ductal cancers, of which 83% were postmenopausal. In all ERβ2-positive cancers the interductal space was filled with dense collagen, and cell nuclei expressed hypoxia-inducible factor 1α. ERβ2 expression was not confined to malignant cells but was strong in stromal, immune, and endothelial cells. In most of the high-grade invasive ductal cancers neither ERα nor ERβ was expressed, but in the high-grade lobular cancer ERβ was lost and ERα and Ki67 expression were abundant. The data show a clear difference in ER expression between lobular and ductal breast cancer and suggest (i) that tamoxifen may be more effective in late than in early lobular cancer and (ii) a potential role for ERβ agonists in preventing in situ ductal cancers from becoming invasive.Despite decades of research, the etiology of breast cancer remains unclear. It is currently thought that most breast cancers occur in the normal terminal duct lobular unit and progress in a stepwise fashion over time (1). Ductal carcinoma in situ (DCIS) means the cancer has not spread beyond the duct into any normal surrounding breast tissue and is thought by some to be the direct precursor of invasive ductal carcinoma (IDC).Estrogens play an important role in normal breast development as well as breast cancer progression (2). Most of the effects of estrogen are mediated through its two receptors: estrogen receptor α (ERα) and β (ERβ) (3). ERα is expressed in 50–80% of breast tumors, and its presence is the main indicator for antihormonal therapy (4). ERβ was first discovered in 1996, and its role in breast cancer is still being explored (57).The first step in understanding the role of ERβ in breast cancer was to define the expression pattern of ERβ in the normal human breast and in various stages of cancer. Since its discovery, several laboratories have reported ERβ expression in clinical samples (828). Most of these studies investigated the expression of ERβ in invasive breast cancer samples (1215, 17, 19, 2123). Some studies have reported ERβ expression in invasive breast cancer and normal breast tissue (11, 18, 2628), but few have compared the expression of ERβ in the normal tissue, DCIS, and IDC within the same sample. Usually tumor samples are taken from one patient and normal tissue from another patient (810). Samples taken from different patients have intrinsic limitation (i.e., they cannot account for variations between different patients). In addition, because tumors are heterogeneous, core biopsies do not fully reflect the histological and biological diversity of breast tumors (29).The roles of ERβ1 and its splice variant ERβ2 in breast cancer are still unclear. As reviewed by Murphy and Leygue (30), some studies show a loss of ERβ1 as ductal cancer progresses, but others do not. Some studies show ERβ2 as a marker of bad prognosis (31), and others not (19). Some of these differences may be due to differences in antibody use and differences in tissue fixation and handling.When ERα and ERβ are coexpressed in breast cancer it is unclear whether tamoxifen treatment will be successful. This is because tamoxifen acts as an agonist of ERβ at activator protein 1 (AP-1) sites (32) and thus should oppose the antiproliferative effects of the tamoxifen–ERα complex. Yan et al. (33) have found that expression of ERβ predicts tamoxifen benefit in patients with ERα-negative early breast cancer, whereas Esslimani-Sahla et al (23) have found that low ERβ level is an independent marker, better than ERα level, to predict tamoxifen resistance. Although apparently saying different things, these two results actually agree with each other: in ERα-negative breast cancer, estrogen is not driving proliferation, so tamoxifen via ERβ may interfere with another growth signaling pathway. In ERα-positive cancers whose proliferation is driven by E2, tamoxifen with ERβ would oppose the antiproliferative effects of the ERα–tamoxifen complex.Investigation of the expression pattern of ERβ in normal tissue, DCIS, and IDC is important to understand the function of this receptor in the progression of breast cancer. We have a set of samples obtained from surgical excision of breast tumors from women before pharmacological intervention. The cohorts include lobular cancer, which has not yet been thoroughly studied for ERβ expression. Lobular cancer is an ERα-positive form of breast cancer characterized by loss of E-cadherin and relatively low proliferation rate. It is accompanied by a resistance to anoikis (34). It accounts for 10–15% of diagnosed breast cancer, and there are still many questions about the optimal therapeutic approach to this cancer. We have explored the changes in expression of the two ERs using identical protocols and reagents in different developmental stages of breast cancer within each patient.  相似文献   

11.
12.
Blue light activation of the photoreceptor CRYPTOCHROME (CRY) evokes rapid depolarization and increased action potential firing in a subset of circadian and arousal neurons in Drosophila melanogaster. Here we show that acute arousal behavioral responses to blue light significantly differ in mutants lacking CRY, as well as mutants with disrupted opsin-based phototransduction. Light-activated CRY couples to membrane depolarization via a well conserved redox sensor of the voltage-gated potassium (K+) channel β-subunit (Kvβ) Hyperkinetic (Hk). The neuronal light response is almost completely absent in hk/ mutants, but is functionally rescued by genetically targeted neuronal expression of WT Hk, but not by Hk point mutations that disable Hk redox sensor function. Multiple K+ channel α-subunits that coassemble with Hk, including Shaker, Ether-a-go-go, and Ether-a-go-go–related gene, are ion conducting channels for CRY/Hk-coupled light response. Light activation of CRY is transduced to membrane depolarization, increased firing rate, and acute behavioral responses by the Kvβ subunit redox sensor.CRYPTOCHROME (CRY) is a photoreceptor that mediates rapid membrane depolarization and increased spontaneous action potential firing rate in response to blue light in arousal and circadian neurons in Drosophila melanogaster (1, 2). CRY regulates circadian entrainment by targeting circadian clock proteins to proteasomal degradation in response to light (36). CRY is expressed in a small subset of central brain circadian, arousal, and photoreceptor neurons in D. melanogaster and other insects, including the large lateral ventral neuron (LNv; l-LNv) subset (1, 2, 7, 8). The l-LNvs are light-activated arousal neurons (1, 2, 911), whereas the small lateral ventral neurons (s-LNvs) are critical for circadian function (5, 12). Previous results suggest that light activated arousal is likely attenuated in cry-null mutants. In addition to modulating light-activated firing rate, membrane excitability in the LNv neurons helps maintain circadian rhythms (9, 13, 14), and LNv firing rate is circadian regulated (2, 16).Based on our previous work suggesting that l-LNv electrophysiological light response requires a flavin-specific redox reaction and modulation of membrane K+ channels, we investigated the molecular mechanism for CRY phototransduction to determine how light-activated CRY is coupled to rapid membrane electrical changes. Sequence and structural data suggest that the cytoplasmic Kvβs are redox sensors based on a highly conserved aldo-keto-reductase domain (AKR) (1721). Although no functional role for redox sensing by Kvβ subunits has been established yet in vivo, studies with heterologously expressed WT and mutant Kvβ subunits show that they confer modulatory sensitivity to coexpressed K+ channels in response to oxidizing and reducing chemical agents (2224). Mammals express six Kvβ genes, whereas Drosophila expresses a single Kvβ designated HYPERKINETIC (Hk) (18). We find that the light-activated redox reaction of the flavin adenine dinucleotide (FAD) chromophore in CRY has a distinct phototransduction mechanism that evokes membrane electrical responses via the Kvβ subunit Hk, which we show is a functional redox sensor in vivo.  相似文献   

13.
Nod-like receptor family, pyrin domain-containing 3 (NLRP3), is involved in the early stages of the inflammatory response by sensing cellular damage or distress due to viral or bacterial infection. Activation of NLRP3 triggers its assembly into a multimolecular protein complex, termed “NLRP3 inflammasome.” This event leads to the activation of the downstream molecule caspase-1 that cleaves the precursor forms of proinflammatory cytokines, such as interleukin 1 beta (IL-1β) and IL-18, and initiates the immune response. Recent studies indicate that the reactive oxygen species produced by mitochondrial respiration is critical for the activation of the NLRP3 inflammasome by monosodium urate, alum, and ATP. However, the precise mechanism by which RNA viruses activate the NLRP3 inflammasome is not well understood. Here, we show that loss of mitochondrial membrane potential [ΔΨ(m)] dramatically reduced IL-1β secretion after infection with influenza, measles, or encephalomyocarditis virus (EMCV). Reduced IL-1β secretion was also observed following overexpression of the mitochondrial inner membrane protein, uncoupling protein-2, which induces mitochondrial proton leakage and dissipates ΔΨ(m). ΔΨ(m) was required for association between the NLRP3 and mitofusin 2, a mediator of mitochondrial fusion, after infection with influenza virus or EMCV. Importantly, the knockdown of mitofusin 2 significantly reduced the secretion of IL-1β after infection with influenza virus or EMCV. Our results provide insight into the roles of mitochondria in NLRP3 inflammasome activation.Nod-like receptor family, pyrin domain-containing 3 (NLRP3) can be activated by a wide variety of stimuli, such as endogenous danger signals from damaged cells, bacterial components, environmental irritants, and DNA and RNA viruses (1). It forms a multiprotein complex called the NLRP3 inflammasome, which includes an adaptor protein apoptosis-associated speck-like protein containing a caspase recruitment domain (ASC) and procaspase-1. The NLRP3 inflammasome-mediated cytokine release requires two signaling pathways (2). The first signal is induced by Toll-like receptors (TLRs), interleukin 1 receptor (IL-1R), or tumor necrosis factor receptor, and leads to the synthesis of inactive NLRP3, pro–IL-1β, and pro–IL-18 in the cytosol. The second signal is triggered by NLRP3 agonists, which induce the activation of caspase-1. Caspase-1 catalyzes the proteolytic processing of pro–IL-1β and pro–IL-18, and their conversion to mature forms, and stimulates their secretion across the plasma membrane (1). These inflammasome-dependent cytokines play a key role in the induction of adaptive immunity and the initiation of tissue healing after influenza virus infection (35). Migration of dendritic cells (DCs) to the draining lymph nodes and priming of CD8 T cells during influenza virus infection require IL-1R signaling in respiratory DCs (6). By contrast, chronic activation of the NLRP3 inflammasome has been linked to many inflammatory diseases (7, 8). Therefore, increasing the number of studies dedicated to the investigation of the molecular mechanisms of NLRP3 inflammasome activation will be crucial for improving our understanding of the pathogenesis of infectious and autoinflammatory diseases.Mitochondria are compartmentalized by two membrane bilayers (outer and inner membranes) and are involved in a wide variety of functions in eukaryotic cells. Within the past decade, novel functions of mitochondria have been discovered demonstrating their crucial role in innate antiviral immunity in vertebrates (9). A direct link between mitochondria and innate immunity was first highlighted with the finding that an adaptor protein, mitochondrial antiviral signaling (MAVS; also known as IPS-1, VISA, or Cardif) (1013), triggered retinoic acid-inducible gene 1 (RIG-I) and melanoma differentiation-associated protein 5-mediated type I interferon (IFN) induction. In addition to their role in antiviral immunity, mitochondria also function as a platform for the activation of the NLRP3 inflammasome by producing mitochondrial reactive oxygen species (mROS) (14, 15). In this context, NLRP3 agonists trigger the generation of mROS from damaged mitochondria, resulting in the dissociation of thioredoxin (TRX) from TRX-interacting protein, which associates with NLRP3 to facilitate inflammasome formation (16). Furthermore, cytosolic mitochondrial DNA (mtDNA) released from damaged mitochondria has been reported to activate the NLRP3 inflammasome (17) and absent in melanoma 2 inflammasome (15), recently identified as a cytoplasmic DNA sensor for the inflammasome (1821). Although mitochondria are essential for host-cell defense, the mechanism of their involvement in the activation of the NLRP3 inflammasome is still unclear. In the present study, we demonstrate that the mitofusin 2 (Mfn2) is required for the full activation of the NLRP3 inflammasomes in macrophages.  相似文献   

14.
15.
16.
Next-generation sequencing is transforming our understanding of human genetic variation but assessing the functional impact of novel variants presents challenges. We analyzed missense variants in the integrin αIIbβ3 receptor subunit genes ITGA2B and ITGB3 identified by whole-exome or -genome sequencing in the ThromboGenomics project, comprising ∼32,000 alleles from 16,108 individuals. We analyzed the results in comparison with 111 missense variants in these genes previously reported as being associated with Glanzmann thrombasthenia (GT), 20 associated with alloimmune thrombocytopenia, and 5 associated with aniso/macrothrombocytopenia. We identified 114 novel missense variants in ITGA2B (affecting ∼11% of the amino acids) and 68 novel missense variants in ITGB3 (affecting ∼9% of the amino acids). Of the variants, 96% had minor allele frequencies (MAF) < 0.1%, indicating their rarity. Based on sequence conservation, MAF, and location on a complete model of αIIbβ3, we selected three novel variants that affect amino acids previously associated with GT for expression in HEK293 cells. αIIb P176H and β3 C547G severely reduced αIIbβ3 expression, whereas αIIb P943A partially reduced αIIbβ3 expression and had no effect on fibrinogen binding. We used receiver operating characteristic curves of combined annotation-dependent depletion, Polyphen 2-HDIV, and sorting intolerant from tolerant to estimate the percentage of novel variants likely to be deleterious. At optimal cut-off values, which had 69–98% sensitivity in detecting GT mutations, between 27% and 71% of the novel αIIb or β3 missense variants were predicted to be deleterious. Our data have implications for understanding the evolutionary pressure on αIIbβ3 and highlight the challenges in predicting the clinical significance of novel missense variants.Next-generation sequencing is transforming our understanding of human genetic variation (1) and providing profound insights into the impact of both inherited and de novo variants on human health (2, 3). At the same time, the data from these studies present serious challenges in providing information to individuals who are found to have variant forms of different proteins. To highlight these challenges, in this report we describe our experience in analyzing missense variants of the platelet αIIbβ3 integrin receptor from The Human Genome Mutation Database (HGMD), the 1000 Genomes project (1000G), the United Kingdom 10K Whole Exome Sequencing project (U.K.10KWES), the United Kingdom 10K Whole Genome Sequencing project (U.K.10KWGS), and The National Heart, Lung and Blood Institute Exome Sequencing Project (ESP); the latter four sources encompass ∼32,000 alleles derived from 16,108 individuals.The αIIbβ3 receptor has a number of virtues as a model system. First, it is required for hemostasis because platelet aggregation requires cross-linking of the activated form of αIIbβ3 by macromolecular ligands (4). Thus, defects in its biogenesis, activation, or ligand binding lead to the rare bleeding diathesis Glanzmann thrombasthenia (GT), an autosomal recessive disorder (5, 6). Patients with GT come to medical attention because of their hemorrhagic symptoms, and thus have been carefully analyzed clinically and with tests of platelet function for nearly 50 y (5, 7). The biochemical and molecular abnormalities in GT have been studied for nearly 40 y (4, 6, 810). In the past 10 y, high-resolution crystallography, electron microscopy, and computational studies of the αIIbβ3 receptor have provided atomic-level information on the correlation between receptor structure and function (1121). In addition, ethnic groups with relatively high prevalence of GT have been defined that share the same genetic abnormality based on founder mutations, and the dates that some of the mutations entered the population have been estimated (2228). An on-line registry of GT abnormalities, including patient phenotypes was developed in 1997 (29) and currently contains 51 αIIb and 43 β3 missense variants linked to the disorder (sinaicentral.mssm.edu/intranet/research/glanzmann). The frequency of GT in the general population has not been established but it has a world-wide distribution, and based on data from hematologic practices, it is rare except in areas with a high rate of consanguineous mating (30).Second, alloimmune disorders, including neonatal thrombocytopenia and posttransfusion purpura, due to amino acid substitutions in either αIIb or β3, have been characterized at the molecular biological level and correlated with mechanisms of immunologic recognition (31).Third, inherited macrothrombocytopenia and anisothrombocytopenia have been associated with heterozygous missense variants or deletions in αIIb or β3. All of these appear to induce constitutive activation of the receptor and impair proplatelet formation (3238).Fourth, αIIbβ3 contributes to pathological platelet thrombus formation in human ischemic cardiovascular disease and αIIbβ3 is a validated target for antithrombotic therapy (3941).Fifth, αIIbβ3 is a member of the large integrin family of receptors, which includes 24 receptors derived from 18 α- and 8 β-subunits (41, 42). These receptors are involved in important biologic processes, including development, cell migration, homing, cell survival, and adaptive immunity (4143). More is known about the structure–function relationships of αIIbβ3 than the other members of the group, and so it serves as the paradigmatic integrin receptor (44, 45).Sixth, 3D molecular models have been built based on crystallographic and NMR data to analyze the effects of novel amino acid substitutions on receptor structure and function and the generation of alloantigens (15, 4653). The data from these models and assessments of the severity of the amino acid change in the variants have the potential to aid in predicting whether a novel variant is likely to affect receptor function and immunogenicity (5459).  相似文献   

17.
Epigenetic regulation of gene expression by DNA methylation plays a central role in the maintenance of cellular homeostasis. Here we present evidence implicating the DNA methylation program in the regulation of hypoxia-inducible factor (HIF) oxygen-sensing machinery and hypoxic cell metabolism. We show that DNA methyltransferase 3a (DNMT3a) methylates and silences the HIF-2α gene (EPAS1) in differentiated cells. Epigenetic silencing of EPAS1 prevents activation of the HIF-2α gene program associated with hypoxic cell growth, thereby limiting the proliferative capacity of adult cells under low oxygen tension. Naturally occurring defects in DNMT3a, observed in primary tumors and malignant cells, cause the unscheduled activation of EPAS1 in early dysplastic foci. This enables incipient cancer cells to exploit the HIF-2α pathway in the hypoxic tumor microenvironment necessary for the formation of cellular masses larger than the oxygen diffusion limit. Reintroduction of DNMT3a in DNMT3a-defective cells restores EPAS1 epigenetic silencing, prevents hypoxic cell growth, and suppresses tumorigenesis. These data support a tumor-suppressive role for DNMT3a as an epigenetic regulator of the HIF-2α oxygen-sensing pathway and the cellular response to hypoxia.Metazoan life is dependent upon the use of molecular oxygen for an array of metabolic processes. Tissue hypoxia occurs during periods of imbalance between oxygen supply and consumption. One of the primary cellular responses to hypoxia is the activation of the hypoxia-inducible factor (HIF) program (14). HIF consists of oxygen-regulated α-subunits HIF-1α and HIF-2α and a constitutively expressed β-subunit (HIF-β). In the presence of oxygen, a series of nonheme Fe(II)- and 2-oxoglutarate–dependent dioxygenase oxygen sensors, referred to as HIF prolylhydroxylases (HIF PHDs), promote the hydroxylation of key proline residues on the HIF-α subunits (5, 6). This serves as a recognition site for the von Hippel-Lindau (VHL) tumor-suppressor protein, which mediates ubiquitination and proteasomal degradation of HIF-1α and HIF-2α (79). Hypoxia inhibits HIF PHDs, allowing HIF-1α and HIF-2α to evade VHL recognition and assemble with HIF-β to produce the active heterodimeric HIF factor. Once activated, HIF-1α and HIF-2α cooperate through common and distinct pathways to regulate hypoxic gene expression and cellular adaptation to hypoxia (10).A notable feature of the HIF response is the differential expression pattern of HIF-1α and HIF-2α in normal tissues. HIF-1α mRNA is ubiquitous and constitutively expressed in adult cells. In stark contrast, HIF-2α mRNA is detected in a few cell types of adult tissues and is typically not expressed by epithelia (11). This suggests a physiological necessity to fine-tune the HIF program depending upon the cellular settings by negatively regulating the HIF-2α gene (EPAS1) upstream of the HIF oxygen-sensing enzymes. The negative regulation of EPAS1 is often compromised in cancers, as HIF-2α mRNA is observed in the vast majority of overt tumors (1113). This is particularly evident in renal cancer. Elegant studies by the Maxwell group (13) and others (14) revealed that HIF-2α mRNA is absent in human kidney tubule epithelia but present in dysplastic foci of the nephron. In these incipient renal tumor cells, HIF-2α may function as an oncoprotein (15), collaborating with, or activating, multiple growth-promoting pathways including cancer stewards c-myc (16), ras (17), and EGFR (18, 19). Silencing of HIF-2α suppresses tumorigenesis of various genetically diverse cancers, further highlighting its central role in malignancy (16, 17, 20, 21), although this depends on the experimental context (22). Therefore, EPAS1 is silent in adult epithelia but undergoes unscheduled activation in several malignancies, driving proliferation in the hypoxic tumor microenvironment (23).A clue to the mechanisms involved in the unscheduled activation of EPAS1 during early tumorigenesis may reside in its promoter, which harbors an enrichment of cytosine and guanine bases that often serve as sites of DNA methylation and epigenetic gene silencing (2427). Cytosine methylation is catalyzed by a family of DNA methyltransferases (DNMTs) including DNMT1, DNMT3a, and DNMT3b. DNMT1 maintains the methylation pattern from the template strand to the newly synthesized strand during DNA replication (28). DNMT3a and DNMT3b are de novo methyltransferases that establish postreplicative methylation patterns (29). Alterations in DNA methylation patterns are common in tumors and likely play a central role in aberrant gene expression that characterizes the malignant phenotype (26, 30, 31). This is particularly evident for DNMT3a, as recent studies have identified mutations in DNMT3a in patients with acute myeloid leukemia (32, 33) or down-regulation of DNMT3a mRNA in a variety of solid tumors (34). It is suggested that DNMT3a is a tumor-suppressor gene and that its mutation, or mRNA down-regulation, contributes to reducing global DNMT3a methyltransferase activity (35, 36). Currently, a key challenge is to link aberrant methylation profiles commonly observed in malignant lesions, including alterations in the DNMT3a epigenetic program, to genes that directly promote the tumorigenic phenotype.Here we show that DNMT3a methylates and silences EPAS1 in normal cells. Loss of DNMT3a observed in primary tumors and malignant cells causes unscheduled EPAS1 activation. This allows emerging cancer cells to exploit the HIF-2α program that facilitates cancer cell traverse of the hypoxic barrier and formation of tumors larger than the diffusion limit of oxygen. We suggest that the DNMT3a epigenetic program is a gatekeeper of the hypoxic cancer cell phenotype.  相似文献   

18.
Cell stress is implicated in triggering bouts of systemic inflammation in patients with autoinflammatory disorders. Blood monocytes from patients affected by NLRP3-mediated cryopyrin-associated periodic syndromes (CAPS) release greater amounts of IL-1β than monocytes from unaffected subjects. Here we show that stress lowers the threshold of activation; blood monocytes from CAPS patients maintain the high levels of secreted IL-1β (fivefold) and IL-18 (10-fold) when stimulated with 1,000-fold less LPS than that required for full IL-1β secretion in control subjects. Unexpectedly, IL-1α secretion is increased 10-fold, indicating that inflammatory episodes in CAPS may not be entirely a result of IL-1β but may also involve IL-1α. In CAPS monocytes, LPS induces the externalization of copious amounts of ATP (10-fold), which drive IL-1β, IL-18, and IL-1α release via activation of the P2X purinoceptor 7. This enhanced ATP release appears to be the link between cell stress and increased cytokine secretion in CAPS. In the later phase after LPS stimulation, CAPS monocytes undergo oxidative stress, which impairs production of the anti-inflammatory IL-1 receptor antagonist (IL-1Ra). Remarkably, IL-1Ra secretion is fully restored by treatment with antioxidants. In two patients with the same NLRP3 mutation, but different disease severity, monocytes from the mildly affected patient exhibited more efficient redox response, lower ATP secretion, and more balanced cytokine production. Thus, the robustness of the individual antioxidant response increases the tolerance to stress and reduces the negative effect of the disease. Pharmacologic block of P2X purinoceptor 7 and improved stress tolerance may represent novel treatment strategies in stress-associated inflammatory diseases.Cryopyrin-associated periodic syndromes (CAPS) are autoinflammatory diseases linked to mutations in the gene NLRP3; the disease is characterized by recurrent episodes of fever and systemic inflammation (1). The pathophysiology of CAPS is mainly caused by the dysregulated secretion of IL-1β, which has been validated by dramatic therapeutic responses to the blocking of the IL-1 receptor with anakinra or the neutralization of IL-1β with canakinumab (2, 3). The NLRP3 inflammasome is a multiprotein complex that requires activating signals to assemble and generate active caspase-1, which in turn converts the inactive IL-1β and IL-18 precursors into their mature active forms (4). Extracellular ATP is a common inflammasome-activating event (5). ATP is released during inflammation by activated platelets, dying leukocytes, and injured parenchymal cells and binds to P2X purinoceptor 7 (P2X7R) on inflammatory cells, triggering a series of intracellular processes, only partially understood, that nevertheless lead to inflammasome activation (6). Human monocytes from healthy subjects stimulated by Toll-like receptor (TLR) agonists secrete their endogenous ATP, which autocrinally activates P2X7R (7, 8), thereby partially liberating monocytes from requiring exogenous ATP. ATP secretion requires reactive oxygen species (ROS) production (9), which is induced in inflammatory cells by triggering TLR (1012).Although in some cell types, including mouse macrophages or human myelomonocytic cell lines, ROS inhibitors such as diphenyleneiodonium [DPI (13)] impair pro-IL-1β synthesis and NLRP3 priming (14, 15) in primary human monocytes, pro-IL-1β biosynthesis and intracellular accumulation are unaffected by treatment with DPI (12, 15), indicating that ROS inhibition in monocytes hinders IL-1β secretion by acting at the posttranslational level.Primary monocytes from patients affected by autoinflammatory diseases are under stress (1620). In CAPS patients, blood monocytes display levels of both ROS and antioxidants considerably higher than in healthy subjects (16); this basal stress causes an aberrant redox response to TLR stimulation, with further increase of ROS followed by collapse of intracellular antioxidant systems, resulting in net oxidative stress. A functional consequence of oxidative stress is the impaired production of cytokines downstream of IL-1, such as IL-1 receptor antagonist (IL-1Ra) and IL-6 (18).Here, we identify a series of events induced by stress in monocytes from CAPS patients that contribute to the pathogenesis of the disease. First, we provide evidence that increased ATP secretion secondary to LPS-induced ROS is significantly contributing to the increased secretion of IL-1β and IL-18 and, unexpectedly, IL-1α. We also show that stress at baseline lowers the threshold of activation, with secretion of these three cytokines induced by markedly lower concentrations of LPS, concentrations unable to activate secretion in monocytes from healthy subjects. Moreover, we demonstrate that the presence of antioxidants restores the secretion of the anti-inflammatory cytokine IL-1Ra, which is inhibited by oxidative stress. In addition, comparative studies in two patients from the same family, with the same NLRP3 mutation, suggest that the basal degree of stress and the efficiency of redox response of the individual patients influence the severity of the disease phenotype.  相似文献   

19.
20.
Advances in computational design methods have made possible extensive engineering of soluble proteins, but designed β-barrel membrane proteins await improvements in our understanding of the sequence determinants of folding and stability. A subset of the amino acid residues of membrane proteins interact with the cell membrane, and the design rules that govern this lipid-facing surface are poorly understood. We applied a residue-level depth potential for β-barrel membrane proteins to the complete redesign of the lipid-facing surface of Escherichia coli OmpA. Initial designs failed to fold correctly, but reversion of a small number of mutations indicated by backcross experiments yielded designs with substitutions to up to 60% of the surface that did support folding and membrane insertion.The β-barrel membrane proteins comprise one of the two structural classes of integral membrane proteins. They are found within the outer membranes of bacteria, mitochondria, and chloroplasts, where they perform a range of structural, transport, and catalytic functions (1). In addition to their biological interest they are increasingly relevant to biotechnology, serving as scaffolds for bacterial surface display (2, 3) and atomically precise pores for nanopore-based DNA sequencing. Although the suitability of natural β-barrel membrane proteins for biotechnology has been improved by protein engineering (310), the ability to design membrane proteins de novo would deliver tools customized to meet the demands of each application.De novo design provides a stringent test of our understanding of the determinants of protein folding and stability. Protein design software [e.g., Rosetta (11, 12)] has made tremendous strides in addressing the design problem for small water-soluble proteins (1315), and design of simplified model α-helical membrane proteins including single transmembrane helices and small bundles (1620) has also been accomplished. In contrast, a designed β-barrel membrane protein has yet to be reported, perhaps as a consequence of the unique design challenges presented by the folding pathway and architecture of these proteins. Unlike the α-helical membrane proteins, nascent β-barrel membrane proteins must transit the periplasm to the outer membrane, where folding and membrane insertion are thought to occur in concert (21, 22). An extensive network of chaperones maintains the solubility of the unfolded barrel and guides membrane insertion. The C-terminal β-strand is known to interact with the BAM chaperone complex (2325), which assists the folding of all β-barrel membrane proteins. However, despite recent progress (2630), we do not fully understand how interactions between chaperones and transiting membrane proteins are directed by sequence-encoded information.Further complicating design is the inside-out architecture of β-barrel membrane proteins. In place of a hydrophobic core is either a central water-filled pore or a solid core composed of polar side chains. The lipid bilayer becomes increasingly hydrophobic at greater depths within the membrane (31), and this environmental anisotropy is reflected in the amino acid composition of the barrel surface. Aliphatic side chains are prevalent toward the center of the membrane, and aromatic side chains are common in the lipid head group regions, where they encircle the barrel in external- and periplasmic-side girdles (32).Recently we developed Ezβ, a membrane depth-dependent, residue-level potential calculated from an ensemble of experimentally determined outer membrane protein structures (33, 34). Ezβ can be used to estimate energetics of membrane insertion to predict transmembrane protein orientation within the bilayer, and to detect oligomerization sites on β-barrel surfaces (34). Ezβ and related statistical functions (35, 36) can recapitulate properties of natural outer membrane proteins (37, 38) and predict the effects of mutations on protein stability and oligomerization (39). Similar potentials have driven computational approaches that have fully redesigned α-helical membrane protein surfaces to convert membrane proteins into water-soluble ones (4042).Here, we considered whether the complete redesign of the lipid-facing surface of an outer membrane protein using a statistical potential such as Ezβ preserves its structure and function. This approach allowed us to investigate whether membrane insertion requires only a lipid-facing surface composed of depth-appropriate hydrophobic residues, or whether folding requires sequence-specific interstrand interactions, chaperone-recruiting sequences, evolutionarily optimized aromatic girdles, folding nucleation sites, or other design features lost during the population averaging inherent in parameter fitting of statistical potentials.Previous studies have explored the sensitivity of the β-barrel fold and its chaperone recognition mechanisms to mutations. The canonical eight-stranded β-barrel membrane protein OmpA tolerates a limited number of mutations to the lipid-facing surface, provided hydrophobicity is maintained (43, 44). More radically, the eight-stranded barrel OmpX has been duplicated to form a 16-stranded barrel capable of membrane insertion (45). However, the lipid-facing residues of transmembrane β-strands are conserved across homologous β-barrel membrane proteins beyond the extent expected from hydrophobicity alone (46, 47), implying a functional role that has yet to be elucidated.To explore the sequence constraints on β-barrel membrane proteins, we extensively redesigned the lipid-facing surface of E. coli OmpA. We created a series of OmpA variants with entirely or partially redesigned lipid-facing surfaces and tested their ability to insert into the outer membrane of E. coli. Our results indicate that the surfaces of β-barrel membrane proteins are amenable to large-scale redesign, provided that energetically destabilizing substitutions are avoided.  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