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 36 毫秒
1.
2.
3.
4.
5.
6.
7.
8.
Human skin is constantly exposed to solar ultraviolet radiation (UVR), the most prevalent environmental carcinogen. Humans have the unique ability among mammals to respond to UVR by increasing their skin pigmentation, a protective process driven by melanin synthesis in epidermal melanocytes. The molecular mechanisms used by melanocytes to detect and respond to long-wavelength UVR (UVA) are not well understood. We recently identified a UVA phototransduction pathway in melanocytes that is mediated by G protein-coupled receptors and leads to rapid calcium mobilization. Here we report that in human epidermal melanocytes physiological doses of UVR activate a retinal-dependent current mediated by transient receptor potential A1 (TRPA1) ion channels. The TRPA1 photocurrent is UVA-specific and requires G protein and phospholipase C signaling, thus contributing to UVA-induced calcium responses to mediate downstream cellular effects and providing evidence for TRPA1 function in mammalian phototransduction. Remarkably, TRPA1 activation is required for the UVR-induced and retinal-dependent early increase in cellular melanin. Our results show that TRPA1 is essential for a unique extraocular phototransduction pathway in human melanocytes that is activated by physiological doses of UVR and results in early melanin synthesis.Skin provides a protective barrier to the external environment. Solar ultraviolet radiation (UVR), a prevalent risk factor for skin cancer, evokes a unique response in human skin that leads to increased pigmentation (1). UVR-induced pigmentation is a protective response mediated by melanin synthesis in human epidermal melanocytes (HEMs) to shield DNA from UVR-induced damage (2). While the mechanism by which short-wavelength UVR (UVB) results in increased pigmentation has been characterized (1), very little is known about phototransduction of long-wavelength UVR (UVA) in human skin.Transient receptor potential (TRP) ion channels were first characterized in Drosophila phototransduction as calcium- (Ca2+) permeable ion channels, essential for light activation of rhabdomeric photoreceptors (35). More recently, members of the TRP family have been implicated in a wide range of sensory functions, including photosensation, chemosensation, thermosensation, and nociception (6, 7). In mammalian phototransduction, TRPC channels mediate a light-sensitive current in intrinsically photoreceptive retinal ganglion cells, which are responsible for nonimage-forming visual processes in the human retina (811). TRP channels are also expressed in the epidermis and are activated by noxious compounds, acidic pH, and temperature (1214).We recently characterized a retinal-dependent, UVA-activated pathway in HEMs that is mediated by G protein activation and leads to a rapid Ca2+ release from internal stores and Ca2+-dependent early melanin production (15). Here we show that UVA phototransduction in human melanocytes activates a TRP channel, subfamily A member 1 (TRPA1), leading to a retinal-dependent current and a rapid Ca2+ influx. TRPA1 activation is G protein and phospholipase C (PLC)-dependent and causes a Ca2+ influx necessary for UVA-induced early melanogenesis. Our results demonstrate that UVR doses corresponding to seconds to minutes of sun exposure activate TRPA1 channels in melanocytes to increase melanin production, providing evidence for TRPA1 function in mammalian extraocular phototransduction and melanin synthesis.  相似文献   

9.
10.
We report on crystal structures of ternary Thermus thermophilus Argonaute (TtAgo) complexes with 5′-phosphorylated guide DNA and a series of DNA targets. These ternary complex structures of cleavage-incompatible, cleavage-compatible, and postcleavage states solved at improved resolution up to 2.2 Å have provided molecular insights into the orchestrated positioning of catalytic residues, a pair of Mg2+ cations, and the putative water nucleophile positioned for in-line attack on the cleavable phosphate for TtAgo-mediated target cleavage by a RNase H-type mechanism. In addition, these ternary complex structures have provided insights into protein and DNA conformational changes that facilitate transition between cleavage-incompatible and cleavage-compatible states, including the role of a Glu finger in generating a cleavage-competent catalytic Asp-Glu-Asp-Asp tetrad. Following cleavage, the seed segment forms a stable duplex with the complementary segment of the target strand.Argonaute (Ago) proteins, critical components of the RNA-induced silencing complex, play a key role in guide strand-mediated target RNA recognition, cleavage, and product release (reviewed in refs. 13). Ago proteins adopt a bilobal scaffold composed of an amino terminal PAZ-containing lobe (N and PAZ domains), a carboxyl-terminal PIWI-containing lobe (Mid and PIWI domains), and connecting linkers L1 and L2. Ago proteins bind guide strands whose 5′-phosphorylated and 3′-hydroxyl ends are anchored within Mid and PAZ pockets, respectively (47), with the anchored guide strand then serving as a template for pairing with the target strand (8, 9). The cleavage activity of Ago resides in the RNase H fold adopted by the PIWI domain (10, 11), whereby the enzyme’s Asp-Asp-Asp/His catalytic triad (1215) initially processes loaded double-stranded siRNAs by cleaving the passenger strand and subsequently processes guide-target RNA duplexes by cleaving the target strand (reviewed in refs. 1618). Such Mg2+ cation-mediated endonucleolytic cleavage of the target RNA strand (19, 20) resulting in 3′-OH and 5′-phosphate ends (21) requires Watson–Crick pairing of the guide and target strands spanning the seed segment (positions 2–2′ to 8–8′) and the cleavage site (10′–11′ step on the target strand) (9). Insights into target RNA recognition and cleavage have emerged from structural (9), chemical (22), and biophysical (23) experiments.Notably, bacterial and archaeal Ago proteins have recently been shown to preferentially bind 5′-phosphoryated guide DNA (14, 15) and use an activated water molecule as the nucleophile (reviewed in ref. 24) to cleave both RNA and DNA target strands (9). Structural studies have been undertaken on bacterial and archaeal Ago proteins in the free state (10, 15) and bound to a 5′-phosphorylated guide DNA strand (4) and added target RNA strand (8, 9). The structural studies of Thermus thermophilus Ago (TtAgo) ternary complexes have provided insights into the nucleation, propagation, and cleavage steps of target RNA silencing in a bacterial system (9). These studies have highlighted the conformational transitions on proceeding from Ago in the free state to the binary complex (4) to the ternary complexes (8, 9) and have emphasized the requirement for a precisely aligned Asp-Asp-Asp triad and a pair of Mg2+ cations for cleavage chemistry (9), typical of RNase H fold-mediated enzymes (24, 25). Structural studies have also been extended to binary complexes of both human (5, 6) and yeast (7) Agos bound to 5′-phosphorylated guide RNA strands.Despite these singular advances in the structural biology of RNA silencing, further progress was hampered by the modest resolution (2.8- to 3.0-Å resolution) of TtAgo ternary complexes with guide DNA (4) and added target RNAs (8, 9). This precluded identification of water molecules coordinated with the pair of Mg2+ cations, including the key water that acts as a nucleophile and targets the cleavable phosphate between positions 10′-11′ on the target strand. We have now extended our research to TtAgo ternary complexes with guide DNA and target DNA strands, which has permitted us to grow crystals of ternary complexes that diffract to higher (2.2–2.3 Å) resolution in the cleavage-incompatible, cleavage-compatible, and postcleavage steps. These high-resolution structures of TtAgo ternary complexes provide snapshots of distinct key steps in the catalytic cleavage pathway, opening opportunities for experimental probing into DNA target cleavage as a defense mechanism against plasmids and possibly other mobile elements (26, 27).  相似文献   

11.
12.
13.
Rickettsiae are responsible for some of the most devastating human infections. A high infectivity and severe illness after inhalation make some rickettsiae bioterrorism threats. We report that deletion of the exchange protein directly activated by cAMP (Epac) gene, Epac1, in mice protects them from an ordinarily lethal dose of rickettsiae. Inhibition of Epac1 suppresses bacterial adhesion and invasion. Most importantly, pharmacological inhibition of Epac1 in vivo using an Epac-specific small-molecule inhibitor, ESI-09, completely recapitulates the Epac1 knockout phenotype. ESI-09 treatment dramatically decreases the morbidity and mortality associated with fatal spotted fever rickettsiosis. Our results demonstrate that Epac1-mediated signaling represents a mechanism for host–pathogen interactions and that Epac1 is a potential target for the prevention and treatment of fatal rickettsioses.Rickettsiae are responsible for some of the most devastating human infections (14). It has been forecasted that temperature increases attributable to global climate change will lead to more widespread distribution of rickettsioses (5). These tick-borne diseases are caused by obligately intracellular bacteria of the genus Rickettsia, including Rickettsia rickettsii, the causative agent of Rocky Mountain spotted fever (RMSF) in the United States and Latin America (2, 3), and Rickettsia conorii, the causative agent of Mediterranean spotted fever endemic to southern Europe, North Africa, and India (6). A high infectivity and severe illness after inhalation make some rickettsiae (including Rickettsia prowazekii, R. rickettsii, Rickettsia typhi, and R. conorii) bioterrorism threats (7). Although the majority of rickettsial infections can be controlled by appropriate broad-spectrum antibiotic therapy if diagnosed early, up to 20% of misdiagnosed or untreated (1, 3) and 5% of treated RMSF cases (8) result in a fatal outcome caused by acute disseminated vascular endothelial infection and damage (9). Fatality rates as high as 32% have been reported in hospitalized patients diagnosed with Mediterranean spotted fever (10). In addition, strains of R. prowazekii resistant to tetracycline and chloramphenicol have been developed in laboratories (11). Disseminated endothelial infection and endothelial barrier disruption with increased microvascular permeability are the central features of SFG rickettsioses (1, 2, 9). The molecular mechanisms involved in rickettsial infection remain incompletely elucidated (9, 12). A comprehensive understanding of rickettsial pathogenesis and the development of novel mechanism-based treatment are urgently needed.Living organisms use intricate signaling networks for sensing and responding to changes in the external environment. cAMP, a ubiquitous second messenger, is an important molecular switch that translates environmental signals into regulatory effects in cells (13). As such, a number of microbial pathogens have evolved a set of diverse virulence-enhancing strategies that exploit the cAMP-signaling pathways of their hosts (14). The intracellular functions of cAMP are predominantly mediated by the classic cAMP receptor, protein kinase A (PKA), and the more recently discovered exchange protein directly activated by cAMP (Epac) (15). Thus, far, two isoforms, Epac1 and Epac2, have been identified in humans (16, 17). Epac proteins function by responding to increased intracellular cAMP levels and activating the Ras superfamily small GTPases Ras-proximate 1 and 2 (Rap1 and Rap2). Accumulating evidence demonstrates that the cAMP/Epac1 signaling axis plays key regulatory roles in controlling various cellular functions in endothelial cells in vitro, including cell adhesion (1821), exocytosis (22), tissue plasminogen activator expression (23), suppressor of cytokine signaling 3 (SOCS-3) induction (2427), microtubule dynamics (28, 29), cell–cell junctions, and permeability and barrier functions (3037). Considering the critical importance of endothelial cells in rickettsioses, we examined the functional roles of Epac1 in rickettsial pathogenesis in vivo, taking advantage of the recently generated Epac1 knockout mouse (38) and Epac-specific inhibitors (39, 40) generated from our laboratory. Our studies demonstrate that Epac1 plays a key role in rickettsial infection and represents a therapeutic target for fatal rickettsioses.  相似文献   

14.
15.
Lytic polysaccharide monooxygenases (LPMOs) exhibit a mononuclear copper-containing active site and use dioxygen and a reducing agent to oxidatively cleave glycosidic linkages in polysaccharides. LPMOs represent a unique paradigm in carbohydrate turnover and exhibit synergy with hydrolytic enzymes in biomass depolymerization. To date, several features of copper binding to LPMOs have been elucidated, but the identity of the reactive oxygen species and the key steps in the oxidative mechanism have not been elucidated. Here, density functional theory calculations are used with an enzyme active site model to identify the reactive oxygen species and compare two hypothesized reaction pathways in LPMOs for hydrogen abstraction and polysaccharide hydroxylation; namely, a mechanism that employs a η1-superoxo intermediate, which abstracts a substrate hydrogen and a hydroperoxo species is responsible for substrate hydroxylation, and a mechanism wherein a copper-oxyl radical abstracts a hydrogen and subsequently hydroxylates the substrate via an oxygen-rebound mechanism. The results predict that oxygen binds end-on (η1) to copper, and that a copper-oxyl–mediated, oxygen-rebound mechanism is energetically preferred. The N-terminal histidine methylation is also examined, which is thought to modify the structure and reactivity of the enzyme. Density functional theory calculations suggest that this posttranslational modification has only a minor effect on the LPMO active site structure or reactivity for the examined steps. Overall, this study suggests the steps in the LPMO mechanism for oxidative cleavage of glycosidic bonds.Carbohydrates are the most diverse set of biomolecules, and thus, many enzyme classes have evolved to assemble, modify, and depolymerize carbohydrates, including glycosyltransferases, glycoside hydrolases, carbohydrate esterases, and polysaccharide lyases (1). Recently, a new enzymatic paradigm was discovered that employs copper-dependent oxidation to cleave glycosidic bonds in polysaccharides (213). These newly classified enzymes, termed lytic polysaccharide monooxygenases (LPMOs), broadly resemble other copper monooxygenases and some hydroxylation catalysts (1421).The discovery that LPMOs use an oxidative mechanism has attracted interest both because it is a unique paradigm for carbohydrate modification that employs a powerful C–H activation mechanism, and also because LPMOs synergize with hydrolytic enzymes in biomass conversion to sugars because they act directly on the crystalline polysaccharide surface without the requirement for depolymerization (4, 22, 23), making them of interest in biofuels production. LPMOs were originally characterized as Family 61 glycoside hydrolases (GH61s, reclassified as auxiliary activity 9, AA9) or Family 33 carbohydrate-binding modules (CBM33s, reclassified as AA10), which are structurally similar enzymes found in fungi and nonfungal organisms (22), respectively. In 2005, Vaaje-Kolstad et al. described the synergism (24) of a chitin-active CBM33 (chitin-binding protein, CBP21) with hydrolases, but the mechanism was not apparent. Harris et al. demonstrated that a GH61 boosts hydrolytic enzyme activity on lignocellulosic biomass (2). Vaaje-Kolstad et al. subsequently showed that CBP21 employs an oxidative mechanism to cleave glycosidic linkages in chitin (4).Following these initial discoveries, multiple features of LPMOs have been elucidated. LPMOs use copper (57) and produce either aldonic acids or 4-keto sugars at oxidized chain ends, believed to result from hydroxylation at the C1 or C4 carbon, respectively. Hydroxylation at the C1 carbon is proposed to spontaneously undergo elimination to a lactone followed by hydrolytic ring opening to an aldonic acid, whereas hydroxylation and elimination at C4 yields a 4-keto sugar at the nonreducing end (512). The active site is a mononuclear type(II) copper center ligated by a “histidine brace” (5, 12), comprising a bidentate N-terminal histidine ligand via the amino terminus and an imidazole ring nitrogen atom and another histidine residue also via a ring nitrogen atom. Hemsworth et al. reported a bacterial LPMO structure wherein the active site copper ion was photoreduced to Cu(I) (12), and Aachmann et al. demonstrated that Cu(I) binds with higher affinity than Cu(II) in CBP21 (13). A structural study of a fungal LPMO revealed an N-terminal methylation on a nitrogen atom in the imidazole ring of unknown function (5), but some LPMOs are active without this modification (6, 11). LPMOs require reducing agents for activity such as ascorbate (28, 1012), and cellobiose dehydrogenase (CDH), a common fungal secretome component, can potentiate LPMO activity in lieu of a small-molecule reducing agent (7, 8).Overall, many structural and mechanistic insights have been reported since the discoveries that LPMOs are oxidative enzymes (410). However, many questions remain regarding LPMO function (22, 25). Here, we examine the LPMO catalytic mechanism with density functional theory (DFT) calculations on an active site model (ASM) of a fungal LPMO. We seek to (i) understand the identity of the reactive oxygen species (ROS), (ii) compare two hypothesized catalytic mechanisms, and (iii) examine the role of N-terminal methylation in catalysis.  相似文献   

16.
The constitutive photomorphogenesis 9 (COP9) signalosome (CSN) plays key roles in many biological processes, such as repression of photomorphogenesis in plants and protein subcellular localization, DNA-damage response, and NF-κB activation in mammals. It is an evolutionarily conserved eight-protein complex with subunits CSN1 to CSN8 named following the descending order of molecular weights. Here, we report the crystal structure of the largest CSN subunit, CSN1 from Arabidopsis thaliana (atCSN1), which belongs to the Proteasome, COP9 signalosome, Initiation factor 3 (PCI) domain containing CSN subunit family, at 2.7 Å resolution. In contrast to previous predictions and distinct from the PCI-containing 26S proteasome regulatory particle subunit Rpn6 structure, the atCSN1 structure reveals an overall globular fold, with four domains consisting of helical repeat-I, linker helix, helical repeat-II, and the C-terminal PCI domain. Our small-angle X-ray scattering envelope of the CSN1–CSN7 complex agrees with the EM structure of the CSN alone (apo-CSN) and suggests that the PCI end of each molecule may mediate the interaction. Fitting of the CSN1 structure into the CSN–Skp1-Cul1-Fbox (SCF) EM structure shows that the PCI domain of CSN1 situates at the hub of the CSN for interaction with several other subunits whereas the linker helix and helical repeat-II of CSN1 contacts SCF using a conserved surface patch. Furthermore, we show that, in human, the C-terminal tail of CSN1, a segment not included in our crystal structure, interacts with IκBα in the NF-κB pathway. Therefore, the CSN complex uses multiple mechanisms to hinder NF-κB activation, a principle likely to hold true for its regulation of many other targets and pathways.The constitutive photomorphogenesis 9 (COP9) signalosome (CSN) is a more than 300-kDa complex that was first identified as a negative regulator of Constitutive Photomorphogenesis (COP) in plants (1, 2). In the subsequent years, the highly conserved protein complex was also found in fungi (3, 4), Caenorhabditis elegans (5), Drosophila melanogaster (6), and mammals (7, 8). The most studied function of the CSN complex in eukaryotes is the regulation of protein degradation through two pathways, deneddylation (911) and deubiquitination (12, 13). In the deneddylation pathway, the CSN complex can influence the cullin-RING ligase activity by removing Nedd8, a ubiquitin-like protein, from a cullin (9, 14). On the other hand, the CSN complex can also suppress cullin activity through recruitment of the deubiquitination enzyme USP15 (12) or Ubp12p, the Schizosaccharomyces pombe ortholog of human USP15 (13). Other functions of the CSN complex identified in mammalian cells include regulating the phosphorylation of ubiquitin–proteasome pathway substrates through CSN-associated kinases (7, 1518). Overall, the CSN complex appears to be a key player in protein subcellular localization (19, 20), DNA-damage response (21), NF-κB activation (22), development, and cell cycle control (23, 24). Thus, the functions of the CSN complex are beyond the regulation of light-dependent reaction in plants.The CSN complex in most of the species contains eight subunits named CSN1 to CSN8, in order of decreasing size. All eight subunits share homologous sequences with “lid” components of the 26S proteasome regulatory particle and the eukaryotic translation initiation factor 3 (eIF3) (7, 25). Among these eight subunits, CSN6 and catalytic CSN5 contain a conserved MPN-domain (MOV34, Pad1N-terminal) (26), and the rest of the CSN subunits bear a PCI-domain (Proteasome, COP9 signalosome, Initiation factor eIF3). The MPN-domain within CSN5 contains a metal chelating site and forms the catalytic region of the isopeptidase for deneddylation (27). Recently, the crystal structures of the CSN6–MPN domain and the CSN5 subunit have been revealed (28, 29). Interestingly, amino acids 97–131, a flexible segment within the CSN5–MPN domain, were proven to be essential in regulating the isopeptidase states of CSN5 (29). PCI is an ∼200-amino acid domain, beginning with an N-terminal helical bundle arrangement and ending with a globular winged-helix subdomain (30, 31). A number of interactions between PCI domains of CSN subunits have been identified by the yeast two-hybrid system (32, 33). Dessau et al. reported the crystallographic data of the PCI domain of Arabidopsis thaliana subunit 7, and their in vitro studies also suggested that the PCI domain mediates and stabilizes protein–protein interactions within the complex (34).Although many speculated on how the CSN subunits interact with each other and come into a functional unit, the architecture of the CSN complex was accessed by electron microscopy (EM) (35, 36) and native mass spectrometry approaches (37). These studies confirmed structural similarities among CSN, the proteasome lid, and eIF3. Furthermore, the CSN appears to contain two dominant subcomplexes, CSN1/2/3/8 and CSN 4/5/6/7 (37), which correspond to the large and the small segments, respectively, in an EM study of the CSN alone (apo-CSN) (36). An EM study of the CSN in complex with an Skp1-Cul1-Fbox (SCF) E3 ligase was also reported, showing reciprocal regulation between CSN and SCF (38). To date, unfortunately, there is no high-resolution mapping on these subunit interactions.To further define the CSN structure and to study its functional significance, we feel the need to obtain structures of CSN subunits at an atomic level. In our study, we used Arabidopsis thaliana CSN1 (atCSN1) as a guide to facilitate our understandings of the PCI-containing CSN subunits. The atCSN1, encoded in the chromosome 3, has 441 amino acids that are 45% identical in sequence to Homo sapiens CSN1. Among all of the subunits of the complex, CSN1 is known to be the longest and to play a crucial role in complex integrity (3941). Here, we report the crystal structure of atCSN1 and describe its integration within the complex as well as its interaction with IκBα in the NF-κB signaling pathway.  相似文献   

17.
Global analysis of gene expression via RNA sequencing was conducted for trisomics for the left arm of chromosome 2 (2L) and compared with the normal genotype. The predominant response of genes on 2L was dosage compensation in that similar expression occurred in the trisomic compared with the diploid control. However, the male and female trisomic/normal expression ratio distributions for 2L genes differed in that females also showed a strong peak of genes with increased expression and males showed a peak of reduced expression relative to the opposite sex. For genes in other autosomal regions, the predominant response to trisomy was reduced expression to the inverse of the altered chromosomal dosage (2/3), but a minor peak of increased expression in females and further reduced expression in males were also found, illustrating a sexual dimorphism for the response to aneuploidy. Moreover, genes with sex-biased expression as revealed by comparing amounts in normal males and females showed responses of greater magnitude to trisomy 2L, suggesting that the genes involved in dosage-sensitive aneuploid effects also influence sex-biased expression. Each autosomal chromosome arm responded to 2L trisomy similarly, but the ratio distributions for X-linked genes were distinct in both sexes, illustrating an X chromosome-specific response to aneuploidy.Changes in chromosomal dosage have long been known to affect the phenotype or viability of an organism (14). Altering the dosage of individual chromosomes typically has a greater impact than varying the whole genome (57). This general rule led to the concept of “genomic balance” in that dosage changes of part of the genome produce a nonoptimal relationship of gene products. The interpretation afforded these observations was that genes on the aneuploid chromosome produce a dosage effect for the amount of gene product present in the cell (8).However, when gene expression studies were conducted on aneuploids, it became known that transacting modulations of gene product amounts were also more prevalent with aneuploidy than with whole-genome changes (914). Assays of enzyme activities, protein, and RNA levels revealed that any one chromosomal segment could modulate in trans the expression of genes throughout the genome (915). These modulations could be positively or negatively correlated with the changed chromosomal segment dosage, but inverse correlations were the most common (1013). For genes on the varied segment, not only were dosage effects observed, but dosage compensation was also observed, which results from a cancelation of gene dosage effects by inverse effects operating simultaneously on the varied genes (9, 10, 1418). This circumstance results in “autosomal” dosage compensation (14, 1618). Studies of trisomic X chromosomes examining selected endogenous genes or global RNA sequencing (RNA-seq) studies illustrate that the inverse effect can also account for sex chromosome dosage compensation in Drosophila (15, 1921). In concert, autosomal genes are largely inversely affected by trisomy of the X chromosome (15, 19, 21).The dosage effects of aneuploidy can be reduced to the action of single genes whose functions tend to be involved in heterogeneous aspects of gene regulation but which have in common membership in macromolecular complexes (8, 2224). This fact led to the hypothesis that genomic imbalance effects result from the altered stoichiometry of subunits that affects the function of the whole and that occurs from partial but not whole-genome dosage change (8, 2225). Genomic balance also affects the evolutionary trajectory of duplicate genes differently based on whether the mode of duplication is partial or whole-genome (22, 23).Here we used RNA-seq to examine global patterns of gene expression in male and female larvae trisomic for the left arm of chromosome 2 (2L). The results demonstrate the strong prevalence of aneuploidy dosage compensation and of transacting inverse effects. Furthermore, because both trisomic males and females could be examined, a sexual dimorphism of the aneuploid response was discovered. Also, the response of the X chromosome to trisomy 2L was found to be distinct from that of the autosomes, illustrating an X chromosome-specific effect. Genes with sex-biased expression, as determined by comparing normal males and females, responded more strongly to trisomy 2L. Collectively, the results illustrate the prevalence of the inverse dosage effect in trisomic Drosophila and suggest that the X chromosome has evolved a distinct response to genomic imbalance as would be expected under the hypothesis that X chromosome dosage compensation uses the inverse dosage effect as part of its mechanism (15).  相似文献   

18.
19.
20.
Kinesins perform mechanical work to power a variety of cellular functions, from mitosis to organelle transport. Distinct functions shape distinct enzymologies, and this is illustrated by comparing kinesin-1, a highly processive transport motor that can work alone, to Eg5, a minimally processive mitotic motor that works in large ensembles. Although crystallographic models for both motors reveal similar structures for the domains involved in mechanochemical transduction—including switch-1 and the neck linker—how movement of these two domains is coordinated through the ATPase cycle remains unknown. We have addressed this issue by using a novel combination of transient kinetics and time-resolved fluorescence, which we refer to as “structural kinetics,” to map the timing of structural changes in the switch-1 loop and neck linker. We find that differences between the structural kinetics of Eg5 and kinesin-1 yield insights into how these two motors adapt their enzymologies for their distinct functions.There are more than 42 kinesin genes in the human genome, representing 14 distinct classes (1). All are members of the P-loop NTPase superfamily of nucleotide triphosphate hydrolases (24). Like other NTPases, kinesins share a conserved Walker motif nucleotide-binding fold (2, 4) that consists of a central twisted β-sheet and three nucleotide-binding loops, which are termed switch-1, switch-2, and the P-loop. Kinesins also share a common microtubule (MT) binding interface, which isomerizes between states that either bind MTs weakly or strongly, and a mechanical element, termed the neck linker (NL). The NL has been proposed to isomerize between two conformations: one that is flexible and termed undocked, and the other that is ordered and termed docked, where it interacts with a cleft in the motor domain formed by the twisted β-sheet and is oriented along the MT axis (57). NL isomerization (5, 8) is hypothesized to be the force-generating transition in kinesin motors (6, 7, 911), and its position has also been proposed to coordinate the ATPase cycles of processive kinesin dimers by regulating nucleotide binding and hydrolysis (11).Spectroscopic and structural studies have led to a model to explain how kinesins generate force (57, 9, 10, 1215) (summarized in SI Appendix, Fig. S1), which proposes that the conformations of the nucleotide binding site, the MT-binding interface, and the NL are all determined by the state of the catalytic site. It predicts that when unbound to the MT, the motor contains ADP in its catalytic site and its NL is undocked. MT binding accelerates ADP dissociation, thereby allowing ATP to bind, the NL to dock, and mechanical work to be performed. ATP hydrolysis and phosphate release are then followed by dissociation from the MT to complete the cycle (5, 710, 14). This model also argues that: (i) NL docking of the MT-attached motor domain moves the tethered, trailing head into a forward position, where it undergoes a biased diffusional search to attach to the next MT-binding site (11, 14); (ii) switch-1, which coordinates the γ-phosphate of ATP, alternates between two conformations, referred to as “open” and “closed,” and the NL alternates between docked and undocked (5, 6, 10, 1315); and (iii) coordination between the conformations of switch-1 and the NL regulates the timing of the ATPase cycles of the two motor domains in processive kinesin dimers (11). However, the model fails to explain several features of kinesins. For example, it predicts that ATP does not bind to kinesin when the NL is docked. This prediction is inconsistent with studies of both Eg5 and kinesin-1, which suggest ATP binds more readily when the NL is docked (11, 16, 17). The model also predicts that the NL should be docked after ATP binding. However, electron paramagnetic resonance (EPR) probes attached to the NL show a significant population of both mobile and immobile NL states in the presence of both pre- and posthydrolytic ATP analogs (5). Furthermore, the model cannot explain the load dependence of stall, detachment, and back stepping, all of which require a branched pathway (11).To resolve these uncertainties, we have measured the kinetics of the structural changes that occur in switch-1 and the NL with nucleotide binding while the motor is bound to the MT in an experimental design that we refer to as “structural kinetics.” We carried out these experiments using an novel spectroscopic approach, termed transient time-resolved fluorescence resonance energy transfer, (TR)2FRET, that allows us to monitor the kinetics and thermodynamics of both the undocked/docked transition in the NL and the open/closed transition in switch-1 that accompany the process of nucleotide binding. These experiments explain differences in the enzymologies of kinesin-1 and Eg5 and suggest an interesting role for the L5 loop in controlling the timing of conformational changes in the Eg5 switch-1 and NL.  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