首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Chloride ion–pumping rhodopsin (ClR) in some marine bacteria utilizes light energy to actively transport Cl into cells. How the ClR initiates the transport is elusive. Here, we show the dynamics of ion transport observed with time-resolved serial femtosecond (fs) crystallography using the Linac Coherent Light Source. X-ray pulses captured structural changes in ClR upon flash illumination with a 550 nm fs-pumping laser. High-resolution structures for five time points (dark to 100 ps after flashing) reveal complex and coordinated dynamics comprising retinal isomerization, water molecule rearrangement, and conformational changes of various residues. Combining data from time-resolved spectroscopy experiments and molecular dynamics simulations, this study reveals that the chloride ion close to the Schiff base undergoes a dissociation–diffusion process upon light-triggered retinal isomerization.

Chloride ion (Cl) concentration in some bacterial cells is regulated by rhodopsin proteins, generally known as halorhodopsin, or hR. These proteins use light energy to pump Cl into cells (1, 2). Light is harvested by a molecule of retinal, covalently linked to an essential lysine residue in the seventh transmembrane helix of GPCR-like (G protein–coupled receptor) proteins. Light activation causes retinal to isomerize from the all-trans to the 13-cis configuration. This change triggers subsequent conformational changes throughout the rhodopsin molecule and releases chloride into the cytoplasm. Retinal thermally relaxes to the all-trans configuration within milliseconds and is then ready for the next photocycle. Cl ions are transported from the extracellular (EC) side to the cytoplasmic (CP) side during each photocycle (3, 4).Light-driven ion-pumping rhodopsin can be used to develop artificial solar energy harvesting and optogenetics (58), but the molecular mechanism must be understood in detail for such applications. Despite the importance of hR, our current experimental data concerning the structure and dynamics of the protein remain very limited. A related protein, proton (H+)-pumping bacteriorhodopsin (bR) discovered in the early 1970s, has been extensively studied by multiple methods, including time-resolved spectroscopy, crystallography, mutagenesis, and computer simulation (912). In particular, recent studies using time-resolved serial femtosecond crystallography (TR-SFX) methods performed at X-ray free-electron laser (XFEL) facilities allow three-dimensional (3D) visualization of retinal isomerization and associated local conformational changes. These changes are accompanied by movement of protons from a donor aspartate group to an acceptor aspartate (1315). However, the central component of this process, the transported H+, is difficult to observe by X-ray crystallography and could not be directly traced in bR TR-SFX studies. Recently, a breakthrough was reported in a study on the sodium-pumping rhodopsin KR2 (K. eikastus rhodopsin 2), in which electron density signals of Na+ uptake were observed at Δt = 1 ms after laser illumination (16).Cl, a strong X-ray scatterer, can be directly observed from electron density maps. These maps provide first-hand information on the movement of ions as being transported within short timescales after light activation. Furthermore, hR and bR presumably share a common molecular mechanism despite transporting ions in opposite directions. A close relationship is strongly implied by the interconversion of the function of two rhodopsins. Outward H+-pumping bR can be converted to an inward Cl pump by changing a single residue (D85T) (17), while hR from the cyanobacterium, Mastigocladopsis repens, is reported to pump protons after a single mutation (T74D) (18). The chloride pump can therefore serve as a system analogous to the proton transporter and provide valuable information that is difficult to obtain directly from bR.In this study, we focus on chloride ion–pumping rhodopsin (ClR) from the marine flavobacterium Nonlabens marinus S1-08T (19). The conserved DTD motif (Asp85-Thr89-Asp96) of the bR family, residues 85, 89, and 96, is replaced by an NTQ motif (Asn98- Thr102-Gln109) in ClR (Fig. 1). The sequence identity of ClR and canonical bR from Halobacterium salinarum is only 27%, but the two proteins, nevertheless, have highly similar structures, including the disposition of the retinal chromophore. ClR structures at cryogenic and room temperatures clearly reveal an architecture composed of seven transmembrane helices (TM A to G) (2, 20, 21). The retinal is covalently linked to the Nζ atom of the Lys235 located on TM-G. Anomalous diffraction signals of the Br identify a stable binding site near the protonated Schiff base (PSB) and a plausible exit site on the CP side (Fig. 1A). Buried water molecules and locations of cavities inside ClR suggest a pathway for Cl uptake on the EC side, but the molecular mechanism for light-triggered Cl pumping remains obscure. Upon light activation, the Cl tightly held near the PSB must break free from its hydrogen bonding network (Fig. 1B). It then passes through a hydrophobic region to reach the CP side (Fig. 1C). Crystal structures of ClR were previously determined with crystals under continuous illumination of visible laser light. Intriguingly, these steady-state models revealed unexpected movement of the retinal, without indication of photo-isomerization (22). Steady-state measurements, which show averages of mixed states, are thus of limited use in deciphering the molecular mechanism of light-driven Cl pumping.Open in a separate windowFig. 1.Structure of ClR and a plausible pathway of Cl transport. (A) Cross-sections of ClR with the backbone structure shown in cartoon representation. Transmembrane helices are marked using letters A through G, and the C-terminal helix H in the cytoplasm is also indicated. Surfaces are clipped to show the cross-section colored in yellow and the model being sliced and then opened about the axis near the helix E. Water molecules and Cl ions are shown as red- and green-colored spheres. Blue curves indicate the path of ion entering ClR and the principal pumping direction after passing retinal. (B) Key residues near the Cl ion and retinal, together with the NTQ motif shown in stick representation. (C) Residues that form a hydrophobic region between the retinal and the cytoplasm are highlighted in ball-and-stick representation. The red arrow points to a major barrier that Cl needs to overcome. ClR backbone is shown in cartoon representation, with residues colored based on hydrophobicity (the blue to red spectrum corresponds to the hydrophobicity scale from hydrophilic to hydrophobic).  相似文献   

2.
The role of IgE in allergic disease mechanisms is performed principally through its interactions with two receptors, FcεRI on mast cells and basophils, and CD23 (FcεRII) on B cells. The former mediates allergic hypersensitivity, the latter regulates IgE levels, and both receptors, also expressed on antigen-presenting cells, contribute to allergen uptake and presentation to the immune system. We have solved the crystal structure of the soluble lectin-like "head" domain of CD23 (derCD23) bound to a subfragment of IgE-Fc consisting of the dimer of Cε3 and Cε4 domains (Fcε3-4). One CD23 head binds to each heavy chain at the interface between the two domains, explaining the known 2:1 stoichiometry and suggesting mechanisms for cross-linking membrane-bound trimeric CD23 by IgE, or membrane IgE by soluble trimeric forms of CD23, both of which may contribute to the regulation of IgE synthesis by B cells. The two symmetrically located binding sites are distant from the single FcεRI binding site, which lies at the opposite ends of the Cε3 domains. Structural comparisons with both free IgE-Fc and its FcεRI complex reveal not only that the conformational changes in IgE-Fc required for CD23 binding are incompatible with FcεRI binding, but also that the converse is true. The two binding sites are allosterically linked. We demonstrate experimentally the reciprocal inhibition of CD23 and FcεRI binding in solution and suggest that the mutual exclusion of receptor binding allows IgE to function independently through its two receptors.  相似文献   

3.
Pore-forming toxins (PFTs) are potent cytolytic agents secreted by pathogenic bacteria that protect microbes against the cell-mediated immune system (by targeting phagocytic cells), disrupt epithelial barriers, and liberate materials necessary to sustain growth and colonization. Produced by gram-positive and gram-negative bacteria alike, PFTs are released as water-soluble monomeric or dimeric species, bind specifically to target membranes, and assemble transmembrane channels leading to cell damage and/or lysis. Structural and biophysical analyses of individual steps in the assembly pathway are essential to fully understanding the dynamic process of channel formation. To work toward this goal, we solved by X-ray diffraction the 2.9-Å structure of the 450-kDa heptameric Vibrio cholerae cytolysin (VCC) toxin purified and crystallized in the presence of detergent. This structure, together with our previously determined 2.3-Å structure of the VCC water-soluble monomer, reveals in detail the architectural changes that occur within the channel region and accessory lectin domains during pore formation including substantial rearrangements of hydrogen-bonding networks in the pore-forming amphipathic loops. Interestingly, a ring of tryptophan residues forms the narrowest constriction in the transmembrane channel reminiscent of the phenylalanine clamp identified in anthrax protective antigen [Krantz BA, et al. (2005) Science 309:777–781]. Our work provides an example of a β-barrel PFT (β-PFT) for which soluble and assembled structures are available at high-resolution, providing a template for investigating intermediate steps in assembly.  相似文献   

4.
Initiation of the innate immune response requires agonist recognition by pathogen-recognition receptors such as the Toll-like receptors (TLRs). Toll/interleukin-1 receptor (TIR) domain-containing adaptors are critical in orchestrating the signal transduction pathways after TLR and interleukin-1 receptor activation. Myeloid differentiation primary response gene 88 (MyD88) adaptor-like (MAL)/TIR domain-containing adaptor protein (TIRAP) is involved in bridging MyD88 to TLR2 and TLR4 in response to bacterial infection. Genetic studies have associated a number of unique single-nucleotide polymorphisms in MAL with protection against invasive microbial infection, but a molecular understanding has been hampered by a lack of structural information. The present study describes the crystal structure of MAL TIR domain. Significant structural differences exist in the overall fold of MAL compared with other TIR domain structures: A sequence motif comprising a β-strand in other TIR domains instead corresponds to a long loop, placing the functionally important "BB loop" proline motif in a unique surface position in MAL. The structure suggests possible dimerization and MyD88-interacting interfaces, and we confirm the key interface residues by coimmunoprecipitation using site-directed mutants. Jointly, our results provide a molecular and structural basis for the role of MAL in TLR signaling and disease protection.  相似文献   

5.
Prothrombin is the zymogen precursor of the clotting enzyme thrombin, which is generated by two sequential cleavages at R271 and R320 by the prothrombinase complex. The structure of prothrombin is currently unknown. Prethrombin-1 differs from prothrombin for the absence of 155 residues in the N-terminal domain and is composed of a single polypeptide chain containing fragment 2 (residues 156–271), A chain (residues 272–320), and B chain (residues 321–579). The X-ray crystal structure of prethrombin-1 solved at 2.2-Å resolution shows an overall conformation significantly different (rmsd = 3.6 ) from that of its active form meizothrombin desF1 carrying a cleavage at R320. Fragment 2 is rotated around the y axis by 29° and makes only few contacts with the B chain. In the B chain, the oxyanion hole is disrupted due to absence of the I16-D194 ion pair and the Na+ binding site and adjacent primary specificity pocket are highly perturbed. A remarkable feature of the structure is that the autolysis loop assumes a helical conformation enabling W148 and W215, located 17 Å apart in meizothrombin desF1, to come within 3.3 Å of each other and completely occlude access to the active site. These findings suggest that the zymogen form of thrombin possesses conformational plasticity comparable to that of the mature enzyme and have significant implications for the mechanism of prothrombin activation and the zymogen → protease conversion in trypsin-like proteases.  相似文献   

6.
Crystal structure of a Trypanosoma brucei metacaspase   总被引:1,自引:0,他引:1  
Metacaspases are distantly related caspase-family cysteine peptidases implicated in programmed cell death in plants and lower eukaryotes. They differ significantly from caspases because they are calcium-activated, arginine-specific peptidases that do not require processing or dimerization for activity. To elucidate the basis of these differences and to determine the impact they might have on the control of cell death pathways in lower eukaryotes, the previously undescribed crystal structure of a metacaspase, an inactive mutant of metacaspase 2 (MCA2) from Trypanosoma brucei, has been determined to a resolution of 1.4 Å. The structure comprises a core caspase fold, but with an unusual eight-stranded β-sheet that stabilizes the protein as a monomer. Essential aspartic acid residues, in the predicted S1 binding pocket, delineate the arginine-specific substrate specificity. In addition, MCA2 possesses an unusual N terminus, which encircles the protein and traverses the catalytic dyad, with Y31 acting as a gatekeeper residue. The calcium-binding site is defined by samarium coordinated by four aspartic acid residues, whereas calcium binding itself induces an allosteric conformational change that could stabilize the active site in a fashion analogous to subunit processing in caspases. Collectively, these data give insights into the mechanistic basis of substrate specificity and mode of activation of MCA2 and provide a detailed framework for understanding the role of metacaspases in cell death pathways of lower eukaryotes.  相似文献   

7.
Bacitracin is a metalloantibiotic agent that is widely used as a medicine and feed additive. It interferes with bacterial cell-wall biosynthesis by binding undecaprenyl-pyrophosphate, a lipid carrier that serves as a critical intermediate in cell wall production. Despite bacitracin’s broad use, the molecular details of its target recognition have not been elucidated. Here we report a crystal structure for the ternary complex of bacitracin A, zinc, and a geranyl-pyrophosphate ligand at a resolution of 1.1 Å. The antibiotic forms a compact structure that completely envelopes the ligand’s pyrophosphate group, together with flanking zinc and sodium ions. The complex adopts a highly amphipathic conformation that offers clues to antibiotic function in the context of bacterial membranes. Bacitracin’s efficient sequestration of its target represents a previously unseen mode for the recognition of lipid pyrophosphates, and suggests new directions for the design of next-generation antimicrobial agents.  相似文献   

8.
9.
Enpp1 is a membrane-bound glycoprotein that regulates bone mineralization by hydrolyzing extracellular nucleotide triphosphates to produce pyrophosphate. Enpp1 dysfunction causes human diseases characterized by ectopic calcification. Enpp1 also inhibits insulin signaling, and an Enpp1 polymorphism is associated with insulin resistance. However, the precise mechanism by which Enpp1 functions in these cellular processes remains elusive. Here, we report the crystal structures of the extracellular region of mouse Enpp1 in complex with four different nucleotide monophosphates, at resolutions of 2.7–3.2 Å. The nucleotides are accommodated in a pocket formed by an insertion loop in the catalytic domain, explaining the preference of Enpp1 for an ATP substrate. Structural mapping of disease-associated mutations indicated the functional importance of the interdomain interactions. A structural comparison of Enpp1 with Enpp2, a lysophospholipase D, revealed marked differences in the domain arrangements and active-site architectures. Notably, the Enpp1 mutant lacking the insertion loop lost the nucleotide-hydrolyzing activity but instead gained the lysophospholipid-hydrolyzing activity of Enpp2. Our findings provide structural insights into how the Enpp family proteins evolved to exert their diverse cellular functions.  相似文献   

10.
Cyanobacteriochromes (CBCRs) are small, linear tetrapyrrole (bilin)-binding photoreceptors in the phytochrome superfamily that regulate diverse light-mediated adaptive processes in cyanobacteria. More spectrally diverse than canonical red/far-red–sensing phytochromes, CBCRs were thought to be restricted to sensing visible and near UV light until recently when several subfamilies with far-red–sensing representatives (frCBCRs) were discovered. Two of these frCBCRs subfamilies have been shown to incorporate bilin precursors with larger pi-conjugated chromophores, while the third frCBCR subfamily uses the same phycocyanobilin precursor found in the bulk of the known CBCRs. To elucidate the molecular basis of far-red light perception by this third frCBCR subfamily, we determined the crystal structure of the far-red–absorbing dark state of one such frCBCR Anacy_2551g3 from Anabaena cylindrica PCC 7122 which exhibits a reversible far-red/orange photocycle. Determined by room temperature serial crystallography and cryocrystallography, the refined 2.7-Å structure reveals an unusual all-Z,syn configuration of the phycocyanobilin (PCB) chromophore that is considerably less extended than those of previously characterized red-light sensors in the phytochrome superfamily. Based on structural and spectroscopic comparisons with other bilin-binding proteins together with site-directed mutagenesis data, our studies reveal protein–chromophore interactions that are critical for the atypical bathochromic shift. Based on these analyses, we propose that far-red absorption in Anacy_2551g3 is the result of the additive effect of two distinct red-shift mechanisms involving cationic bilin lactim tautomers stabilized by a constrained all-Z,syn conformation and specific interactions with a highly conserved anionic residue.

Cyanobacteria have developed elaborate, spectrally tuned photoreceptors and light-harvesting systems for adaptation and survival in a wide range of ecological niches (15). Many photoreceptor systems are modular components of much larger signaling proteins that integrate different sensor and effector modules into a single protein molecule to interface with diverse signal transduction pathways. Photoreceptors in the phytochrome superfamily utilize a specific lineage of GAF (cGMP phosphodiesterase, adenylyl cyclase and FhlA) domain that binds a thioether-linked linear tetrapyrrole (bilin) chromophore for light perception (611). Bilin-based photoreceptors play critical roles in plant development as well as in regulating cyanobacterial phototaxis, development, and light harvesting (2, 3, 1217). Protein structural changes following the primary photochemical event then alter the downstream enzymatic activities and/or protein–protein interactions via an interdomain allosteric mechanism (18).Phytochromes possess a tripartite photosensory region consisting of three N-terminal domains (PAS, GAF, and PHY), known as the photosensory core module, in which the PAS and GAF domains are tethered via a “figure-eight knot” (14, 19, 20). In prototypical phytochromes, the bilin chromophore embedded in the GAF domain adopts a protonated 5-Z,syn, 10-Z,syn, 15-Z,anti configuration in the dark-adapted state. Light absorption triggers photoisomerization of the 15,16 double bond to generate a 15E,anti photoproduct, which typically absorbs far-red light (9, 14, 21). A long extension from the adjacent PHY domain is responsible for stabilizing the far-red–absorbing Pfr state (14, 20). In cyanobacteria, the phytochrome superfamily has diversified to yield a large family of more streamlined sensors, designated cyanobacteriochromes (CBCRs) (2, 4, 2226). Unlike canonical phytochromes, CBCR photosensory modules consist of one or more GAF domains that are sufficient for covalent attachment of bilin and photoconversion. These small CBCR domains have also been used as light-sensing modules in a variety of synthetic biology applications (2732). In contrast to canonical red/far-red phytochromes, CBCRs are able to sense light from near UV to far-red, utilizing a common phycocyanobilin (PCB) chromophore precursor (2224, 26).The remarkable spectral diversity of CBCRs (SI Appendix, Fig. S1A) arises from extensive molecular evolution of the GAF domain scaffold. Many CBCRs leverage two thioether linkages to sense blue, violet, or near-UV light (8, 22, 23, 25, 3335). Such “two-Cys” CBCRs possess an additional thioether linkage to the C10 methine bridge of the bilin that splits the chromophore in half, significantly shortening the conjugated π-system. Rupture of this covalent bond can occur upon 15Z/15E photoisomerization, which restores bilin conjugation across C10 to generate a photostate absorbing at wavelengths from teal to red (8, 33, 36, 37). Dual cysteine CBCRs have evolved multiple times, yielding a wide range of photocycles with (ultra)violet, blue, teal, green, orange, and red states (22).Red/green CBCRs such as AnPixJg2 and NpR6012g4 have red-absorbing dark states similar to phytochromes that photoconvert to green-absorbing lit states. In this CBCR subfamily, the molecular mechanism responsible for photoproduct tuning relies on trapping the 15E bilin in a twisted geometry that results in blue-shifted absorption (10, 11). In contrast, green/red CBCRs exhibit a reversed photocycle: the green-absorbing 15Z dark state photoconverts to yield a red-absorbing 15E photoproduct. This subfamily uses a protochromic mechanism first reported for the light-regulated histidine kinase RcaE (SI Appendix, Fig. S1B) in which photoconversion triggers a proton transfer to an uncharged chromophore inducing a spectral red shift (2, 38).Until recently, the light-sensing range of CBCRs appeared limited to the visible spectrum, thereby implicating phytochromes to be exclusively responsible for far-red sensing in cyanobacteria. Indeed, far-red–dependent remodeling of the photosynthetic apparatus in multiple cyanobacterial species is mediated by the red/far-red phytochrome RfpA (3, 39). The discovery of two lineages of CBCRs with far-red-absorbing dark states (frCBCRs) was thus surprising (40). Upon far-red light absorption, these frCBCRs convert to either an orange- or red-absorbing photoproduct state. These frCBCRs evolved from green/red CBCRs as part of a greater green/red (GGR) lineage and independent from evolution of other frCBCRs within the XRG (extended red/green) lineage (35, 40, 41). Owing to their small size and spectral overlap with the therapeutic window of optimum tissue penetrance (700 to 800 nm) (4246), frCBCRs represent tantalizing scaffolds for development of FR-responsive optogenetic reagents for biomedical research and imaging applications (45, 4750).To understand the molecular basis of far-red spectral tuning of the frCBCR family that evolved within GGR lineage, we determined the crystal structures of the FR-absorbing dark state of the representative FR/O CBCR Anacy_2551g3 from Anabaena cylindrica PCC 7122 at both ambient and cryogenic temperatures. These structures revealed an all-Z,syn configuration of its PCB chromophore that differs from those found in all known CBCRs and phytochromes. Based upon these crystallographic results, spectra of site-directed mutants of Anacy_2551g3 and related frCBCRs in the GGR lineage, and comparisons with other bilin-binding proteins, we identify key protein–chromophore interactions that support two tuning mechanisms simultaneously at work for far-red light detection in this family of frCBCRs.  相似文献   

11.
The influenza A virus M2 channel (AM2) is crucial in the viral life cycle. Despite many previous experimental and computational studies, the mechanism of the activating process in which proton permeation acidifies the virion to release the viral RNA and core proteins is not well understood. Herein the AM2 proton permeation process has been systematically characterized using multiscale computer simulations, including quantum, classical, and reactive molecular dynamics methods. We report, to our knowledge, the first complete free-energy profiles for proton transport through the entire AM2 transmembrane domain at various pH values, including explicit treatment of excess proton charge delocalization and shuttling through the His37 tetrad. The free-energy profiles reveal that the excess proton must overcome a large free-energy barrier to diffuse to the His37 tetrad, where it is stabilized in a deep minimum reflecting the delocalization of the excess charge among the histidines and the cost of shuttling the proton past them. At lower pH values the His37 tetrad has a larger total charge that increases the channel width, hydration, and solvent dynamics, in agreement with recent 2D-IR spectroscopic studies. The proton transport barrier becomes smaller, despite the increased charge repulsion, due to backbone expansion and the more dynamic pore water molecules. The calculated conductances are in quantitative agreement with recent experimental measurements. In addition, the free-energy profiles and conductances for proton transport in several mutants provide insights for explaining our findings and those of previous experimental mutagenesis studies.The influenza type A virus is a highly pathogenic RNA virus that causes flu in birds and mammals (1). The influenza A M2 (AM2) protein (2) contains a homotetramer channel that transports protons across the viral membrane and acidifies the virion interior, enabling the dissociation of the viral matrix proteins, which is a crucial step in viral replication (3). The protein has been the target of antiviral drugs amantadine and rimantadine (4, 5). Much effort has been devoted to discovering the structure and proton transport (PT) mechanism of the AM2 channel, resulting in many crystal structures available in the protein data bank (614). Based on the crystal structures and electrophysiology experiments, several PT models have been suggested. These mechanisms can be divided into two main categories, delineated by the role of the four histidine residues (a.k.a. the His37 tetrad) that reside in the middle of the AM2 transmembrane domain (AM2/TM) (Fig. 1A), which has been experimentally shown (15) to account for the proton permeation behavior of the full AM2 protein. The “shutter” mechanism (16, 17) suggests that the His37 tetrad works as a gate. At low pH the gate opens due to the electrostatic repulsion between the biprotonated, positively charged histidine residues. The excess proton is then transferred through continuous water wire via the Grotthuss mechanism, without changing the protonation state of His37. In contrast, the “shuttle” mechanism (14, 1822) suggests that at acidic pH values the His37 tetrad changes protonation states (Fig. 1B) as it shuttles the proton through the His37–Trp41 quartet region (Trp41 lies just below His37 and is also thought to play an important role in the PT mechanism).Open in a separate windowFig. 1.(A) Equilibrated structures of the AM2 channel transmembrane domain (definition in main text). The backbones of three helices are shown in gray, and the side chains of pore-lining residues are shown as sticks. From N terminus to C terminus, Val27, Ser31, His37, and Trp41 are shown in blue, yellow, orange, and purple, respectively. The water oxygen densities from the simulations are shown as red shading. One helix is removed for the purpose of clarity. (B) The QM/MM simulations reveal that the excess proton (purple) is delocalized around the entry water cluster and histidine residues as it shuttles although the His37 tetrad. His37 is explicitly protonated in agreement with the shuttle mechanism.The structure and dynamics of water in the AM2 protein have also gained attention, because it is the essential medium for proton permeation. The high-resolution crystal structure [Protein Data Bank (PDB) code 3LBW] crystallized at pH 6.5 revealed layers of well-ordered water clusters above the His37 tetrad (12). However, the water dynamics in the AM2 protein probed using 2D infrared (2D-IR) spectroscopy revealed that the well-ordered “ice-like” pore water dynamics at pH 8.0 change to more mobile and “liquid-like” dynamics (on the timescale of a few picoseconds) at pH 3.2 (23). This result suggests an interesting pH-dependent behavior of the AM2 protein that is highly relevant for understanding its PT mechanism.Although many computational studies have investigated the features of AM2 that may influence its PT in recent years (12, 2440), only a few have explicitly simulated any aspect of the explicit PT process (24, 25, 29, 37, 38). This is because it is challenging to accurately model the charge delocalization and Grotthuss shuttling of the hydrated excess proton in a computationally tractable way. Among these studies, only one so far has provided a free-energy profile [potential of mean force (PMF)] for PT across the entire AM2/TM channel (25), which is the essential property for understanding the full proton permeation mechanism. However, this particular free-energy study was limited by the approximation that the His37 tetrad remains in a fixed protonation state during the proton conduction; therefore, it could not capture the more plausible shuttle mechanism (14, 1822). More recently, Carnevale et al. (29) used a quantum mechanics/molecular mechanics (QM/MM) approach to investigate PT in the specific region of the His37 tetrad of AM2. Although they could not achieve sufficient sampling to calculate a free-energy profile, this work did allow the His37 tetrad to change protonation states as necessitated by the shuttle mechanism. Their work in fact helped to lay the foundation for the study of PT in the His37 tetrad, using a QM/MM approach in that region of the channel.In the present study, a powerful multiscale combination of classical, reactive, and ab initio (QM/MM) molecular dynamics (MD) simulations is used to systematically investigate the proton solvation and transport mechanism through the full AM2 protein. By using a classical force field, the conformational ensemble of AM2/TM is characterized starting from two recent high-resolution structures in low- and intermediate-pH conditions [PDB codes 3C9J (9) and 3LBW (12), respectively]. The influence of pH on the protein and water dynamics is then investigated. It is found that as the pH is lowered, the channel adopts a more open conformation, the channel pore is more hydrated, and the pore water molecules are more mobile. Following this we use a synthesis of the reactive multistate empirical valence bond (MS-EVB) (41) and QM/MM approaches to calculate complete proton permeation free-energy profiles (PMFs) through the AM2/TM channel. Because the MS-EVB potential can be derived by force matching of ab initio MD simulation data (42), and because the force-matching algorithm provides a PMF for the reference potential, this unique combination of approaches yields a consistent multiscale computational framework for obtaining an ab initio-level quality PT free-energy profile. This multiscale combination of methods is thus used to calculate the PMFs for excess protons permeating through the entire AM2/TM channel, accounting for the charge delocalization and shuttling of hydrated excess protons through both water and the His37 tetrad explicitly. By virtue of this methodology, it is found that at lower pH values the free-energy barrier for a proton to diffuse to His37 is decreased and the proton conductance increased, explaining the unique pH-dependent activation mechanism of AM2. We also report the PMFs of several AM2 mutants, including the amantadine-resistant and transmissible mutants V27A and S31N. The calculated conductances of both the wild-type AM2 and its mutants are in close agreement with experimental results and provide microscopic explanations for their trends.  相似文献   

12.
Light-driven chloride-pumping rhodopsins actively transport anions, including various halide ions, across cell membranes. Recent studies using time-resolved serial femtosecond crystallography (TR-SFX) have uncovered the structural changes and ion transfer mechanisms in light-driven cation-pumping rhodopsins. However, the mechanism by which the conformational changes pump an anion to achieve unidirectional ion transport, from the extracellular side to the cytoplasmic side, in anion-pumping rhodopsins remains enigmatic. We have collected TR-SFX data of Nonlabens marinus rhodopsin-3 (NM-R3), derived from a marine flavobacterium, at 10-µs and 1-ms time points after photoexcitation. Our structural analysis reveals the conformational alterations during ion transfer and after ion release. Movements of the retinal chromophore initially displace a conserved tryptophan to the cytoplasmic side of NM-R3, accompanied by a slight shift of the halide ion bound to the retinal. After ion release, the inward movements of helix C and helix G and the lateral displacements of the retinal block access to the extracellular side of NM-R3. Anomalous signal data have also been obtained from NM-R3 crystals containing iodide ions. The anomalous density maps provide insight into the halide binding site for ion transfer in NM-R3.

Microbial ion-pumping rhodopsins are integral membrane proteins that actively transport ions across membranes upon light stimulation (1). Bacteriorhodopsin (bR) and halorhodopsin (HR) are well-known microbial ion-pumping rhodopsins found in halophilic archaea (2, 3). bR is a light-driven outward proton pump and HR is a light-driven inward anion pump, specific for chloride ion. Microbial ion-pumping rhodopsins possess common structural features consisting of seven α-helices with an all-trans retinal covalently bound to a lysine residue as the chromophore, despite the transport of different ions (4). The retinal undergoes photoisomerization from the all-trans to 13-cis configuration, which initiates the photocycle accompanied by several intermediates to export ions (4, 5). Its light-controllable function is suitable for optogenetics applications for manipulating cells, such as neurons, by changing the ion concentration inside or outside the membrane (6, 7). In fact, microbial rhodopsins, including channelrhodopsins and HRs, are employed as optogenetic tools (810).Nonlabens marinus rhodopsin-3 (NM-R3) is a light-driven chloride pump recently discovered in a marine flavobacterium (11). It is a distinct chloride pump from HRs and shows low amino acid sequence homology with HRs (11). To date, HR-type chloride pumps have been found in haloarchaea, marine bacteria, and cyanobacteria, including Halobacterium salinarum, Natronomonas pharaonis, and Mastigocladopsins repens, with sequence identities of 20%, 21%, and 20% to NM-R3, respectively (3, 1215). Interestingly, NM-R3 has higher sequence identity (36%) to Krokinobacter rhodopsin 2 (KR2), a sodium pump found in Krokinobacter eikastus (16). NM-R3 possesses a unique NTQ motif (Asn98, Thr102, Gln109) in the third helix (helix C), which corresponds to key residues (DTD motif, Asp85, Thr89, Asp96) for proton transport in bR (11, 17, 18) (SI Appendix, Table S1). Asp85 acts as the primary proton acceptor of bR from the protonated Schiff-base (PSB), with assistance from Thr89 and Asp96, which is the proton donor (5, 17, 18). HRs from haloarchaea have a highly conserved TSA (Thr, Ser, Ala) motif, while the Ala residue is replaced by Asp in HR from cyanobacteria (19). In the X-ray crystal structure of NM-R3 (SI Appendix, Fig. S1A), a chloride ion located between the PSB and Asn98 (SI Appendix, Fig. S1B) is stabilized by the positive charge of the PSB (20). The position of this chloride ion is similar to those in the H. salinarum HR and N. pharaonis HR (NpHR) structures except for Thr and Ser, which correspond to Asn98 and Thr102 in NM-R3, respectively (2022). Several amino acid residues around the retinal, including Arg95, Trp99, Trp201, and Asp231, are highly conserved among ion-pumping rhodopsins. Previous spectroscopic studies suggested that NM-R3 displays a similar sequence of intermediates, with K-, L-, N-, and O-like species, as in other HRs (23) (Fig. 1A). Recently, intermediate structures of NM-R3 obtained by low-temperature trapping X-ray crystallography and serial femtosecond crystallography (SFX) have been reported (24, 25). However, the detailed ion-pump mechanism still remains unclear, due to the lack of dynamic structures of anion transport at atomic resolution.Open in a separate windowFig. 1.TR-visible absorption spectroscopy for microcrystals. (A) Photocycle model of NM-R3 in the 1 M NaCl buffer solution (23). (B) TR difference spectra ΔA upon the 532-nm excitation. The difference was calculated by subtracting the spectrum of NM-R3. (C) Global fitting analysis with two exponentials. The A1 and A2 amplitude spectra correspond to the differences of [ΔAO – ΔA10 µs] and [ΔA200 ms − ΔAO], respectively. Here, ΔAO represents the difference spectrum of the O intermediate minus NM-R3. (D) The isomeric forms of the retinal chromophore in bacterial-type rhodopsins.Time-resolved serial femtosecond crystallography (TR-SFX) is a powerful tool for visualizing reactions and motions in proteins at the atomic level (2628). In SFX, myriads of microcrystals are continuously injected by a sample injector into an irradiation point of X-ray free electron lasers (XFELs) at room temperature, thus providing diffraction patterns before the onset of radiation damage by the intense X-ray pulse. Combined with a visible-light pump laser for reaction initiation, TR-SFX has been applied to light-driven ion pumps to observe the structural dynamics during the ion transfer. While TR-SFX has revealed femto-to-millisecond structural dynamics in light-driven cation pumps, including bR and KR2 (2931), TR-SFX studies of anion pumps have been limited to early-stage structures adopted at picoseconds after light illumination (32). In addition, although NM-R3 pumps a chloride ion (Cl) as a physiological substrate, it can also transport bromide (Br), iodide (I), and other anions from the extracellular side to the cytoplasmic side (23). I or Br serves as a marker for tracking the positions of ions, due to the greater number of electrons, whereas Cl is less distinguishable in X-ray crystallography. Therefore, TR-SFX experiments using I or Br are expected to directly visualize the process of ion transport.Here, we report the conformational alterations in NM-R3 during Br or I pumping, obtained by both TR-SFX and time-resolved spectroscopy of crystals. The resulting sequence of movements in NM-R3 demonstrates how the chloride pump transports anions with a large ionic radius and prevents the backflow of anions from the cytoplasmic side.  相似文献   

13.
Protein release factor 3 (RF3), a guanosine triphosphatase, binds to ribosome after release of the nascent peptide and promotes dissociation of the class I release factors during the termination of protein synthesis. Here we present the crystal structure of the 70S ribosome with RF3 in the presence of a nonhydrolyzable GTP analogue, guanosine 5′-β,γ-methylenetriphosphate (GDPCP), refined to 3.8 Å resolution. The structure shows that the subunits of the ribosome are rotated relative to each other compared to the canonical state, resulting in a P/E hybrid state for the transfer RNA. The substantial conformational rearrangements in the complex are described and suggest how RF3, by stabilizing the hybrid state of the ribosome, facilitates the dissociation of class I release factors.  相似文献   

14.
Improved understanding of the relationship among structure, dynamics, and function for the enzyme phenylalanine hydroxylase (PAH) can lead to needed new therapies for phenylketonuria, the most common inborn error of amino acid metabolism. PAH is a multidomain homo-multimeric protein whose conformation and multimerization properties respond to allosteric activation by the substrate phenylalanine (Phe); the allosteric regulation is necessary to maintain Phe below neurotoxic levels. A recently introduced model for allosteric regulation of PAH involves major domain motions and architecturally distinct PAH tetramers [Jaffe EK, Stith L, Lawrence SH, Andrake M, Dunbrack RL, Jr (2013) Arch Biochem Biophys 530(2):73–82]. Herein, we present, to our knowledge, the first X-ray crystal structure for a full-length mammalian (rat) PAH in an autoinhibited conformation. Chromatographic isolation of a monodisperse tetrameric PAH, in the absence of Phe, facilitated determination of the 2.9 Å crystal structure. The structure of full-length PAH supersedes a composite homology model that had been used extensively to rationalize phenylketonuria genotype–phenotype relationships. Small-angle X-ray scattering (SAXS) confirms that this tetramer, which dominates in the absence of Phe, is different from a Phe-stabilized allosterically activated PAH tetramer. The lack of structural detail for activated PAH remains a barrier to complete understanding of phenylketonuria genotype–phenotype relationships. Nevertheless, the use of SAXS and X-ray crystallography together to inspect PAH structure provides, to our knowledge, the first complete view of the enzyme in a tetrameric form that was not possible with prior partial crystal structures, and facilitates interpretation of a wealth of biochemical and structural data that was hitherto impossible to evaluate.Mammalian phenylalanine hydroxylase (PAH) (EC 1.14.16.1) is a multidomain homo-multimeric protein whose dysfunction causes the most common inborn error in amino acid metabolism, phenylketonuria (PKU), and milder forms of hyperphenylalaninemia (OMIM 261600) (1). PAH catalyzes the hydroxylation of phenylalanine (Phe) to tyrosine, using nonheme iron and the cosubstrates tetrahydrobiopterin and molecular oxygen (2, 3). A detailed kinetic mechanism has recently been derived from elegant single-turnover studies (4). PAH activity must be carefully regulated, because although Phe is an essential amino acid, high Phe levels are neurotoxic. Thus, Phe allosterically activates PAH by binding to a regulatory domain. Phosphorylation at Ser16 potentiates the effects of Phe, with phosphorylated PAH achieving full activation at lower Phe concentrations than the unphosphorylated protein (5, 6). Allosteric activation by Phe is accompanied by a major conformational change, as evidenced by changes in protein fluorescence and proteolytic susceptibility, and by stabilization of a tetrameric conformer (3).There are >500 disease-associated missense variants of human PAH; the amino acid substitutions are distributed throughout the 452-residue protein and among all its domains (Fig. 1A) (79). Of those disease-associated variants that have been studied in vitro (e.g., ref. 10), some confound the allosteric response, and some are interpreted as structurally unstable. We also suggest that the activities of some disease-associated variants may be dysregulated by an altered equilibrium among conformers having different intrinsic levels of activity, arguing by analogy to the enzyme porphobilinogen synthase (PBGS) and its porphyria-associated variants (11). Consistent with this notion, we have recently established that PAH can assemble into architecturally distinct tetrameric conformers (12), and propose that these conformers differ in activity due to differences in active-site access. This idea has important implications for drug discovery, as it implies that small molecules could potentially modulate the conformational equilibrium of PAH, as has already been demonstrated for PBGS (e.g., ref. 13). Deciphering the relationship among PAH structure, dynamics, and function is a necessary first step in testing this hypothesis.Open in a separate windowFig. 1.The structure of PAH. (A) The annotated domain structure of mammalian PAH. (B) The 2.9 Å PAH crystal structure in orthogonal views, colored as in part A, subunit A is shown in ribbons; subunit B is as a Cα trace; subunit C is in sticks; and subunit D is in transparent spheres. In cyan, the subunits are labeled near the catalytic domain (Top); in red, they are labeled near the regulatory domain (Bottom). The dotted black circle illustrates the autoregulatory domain partially occluding the enzyme active site (iron, in orange sphere). (C) Comparison of the subunit structures of full-length PAH and those of the composite homology model; the subunit overlay aligns residues 144–410. The four subunits of the full-length PAH structure (the diagonal pairs of subunits are illustrated using either black or white) are aligned with the two subunits of 2PAH (cyan) and the one subunit of 1PHZ (orange). The catalytic domain is in spheres, the regulatory domain is in ribbons, and the multimerization domain is as a Cα trace. The arrow denotes where the ACT domain and one helix of 2PAH conflict.Numerous crystal structures are known for one- and two-domain constructs of mammalian PAH (14).

Table S1.

Mammalian PAH structures available in the PDB (August 2015)
Open in a separate windowEntries are colored by configuration of the active-site lid (approximately residues 130–150): “open,” “closed,” and “disordered.”*C, catalytic domain, M, multimerization domain; R, regulatory domain.wwPDB Chemical Component Dictionary (49).Disease-associated single-residue substituted variant (A313T).§Phosphorylated at Ser16.  相似文献   

15.
Pentameric ligand-gated ion channels mediate fast chemical transmission of nerve signals. The structure of a bacterial proton-gated homolog has been established in its open and locally closed conformations at acidic pH. Here we report its crystal structure at neutral pH, thereby providing the X-ray structures of the two end-points of the gating mechanism in the same pentameric ligand-gated ion channel. The large structural variability in the neutral pH structure observed in the four copies of the pentamer present in the asymmetric unit has been used to analyze the intrinsic fluctuations in this state, which are found to prefigure the transition to the open state. In the extracellular domain (ECD), a marked quaternary change is observed, involving both a twist and a blooming motion, and the pore in the transmembrane domain (TMD) is closed by an upper bend of helix M2 (as in locally closed form) and a kink of helix M1, both helices no longer interacting across adjacent subunits. On the tertiary level, detachment of inner and outer β sheets in the ECD reshapes two essential cavities at the ECD–ECD and ECD–TMD interfaces. The first one is the ligand-binding cavity; the other is close to a known divalent cation binding site in other pentameric ligand-gated ion channels. In addition, a different crystal form reveals that the locally closed and open conformations coexist as discrete ones at acidic pH. These structural results, together with site-directed mutagenesis, physiological recordings, and coarse-grained modeling, have been integrated to propose a model of the gating transition pathway.Pentameric ligand-gated ion channels (pLGICs) are a superfamily of membrane receptors that mediate fast chemical transmission of nerve signals in the central and peripheral nervous system (1). These allosteric receptors couple neurotransmitter (agonist) binding in the extracellular domain (ECD) to the opening of the ionic pore located in the transmembrane domain (TMD). There are two classes in pLGIC, with either cationic channels (acetylcholine — nAChR — and 5HT3 receptors) or anionic ones (glycine and GABA receptors). The molecular understanding of their allosteric transitions is a central issue in the pharmacology of pLGICs, as it would allow the rational design of novel orthosteric and allosteric ligands. Recently reported full-length structures of several prokaryotic and eukaryotic members of the family have provided significant insights into the conserved architecture of these receptors (25). Nevertheless, little is known about the structural events that link the binding/unbinding of agonist to the opening/closure of the channel gate. To understand the gating mechanism, the structures of the same receptor in different allosteric states are needed at atomic resolution. Here we report two crystal structures of wild-type GLIC, a proton-gated bacterial ion channel from Gloeobacter violaceus (6) at two different pHs, above and below pH50.GLIC structure was initially solved at acid pH in its open conformation (4), and later in a locally closed (LC) conformation (7) displayed by six different mutants with little variations. The LC conformation shares several structural features with the open form but shows a closed pore as a result of a concerted bending of its M2 helices. Independently, recent progress has been made in understanding the structural mechanism of modulation by effectors such as general anesthetics (8) or ethanol (9) and the permeation process (10), using the same bacterial model system. Two other atomic X-ray structures of members of the pLGIC family have been described: an ion channel from the bacteria Erwinia chrysantemi (ELIC), in a closed pore conformation (2) and a eukaryotic glutamate-activated chloride channel (GluCl) in an open pore conformation similar to GLIC’s (5). However, the lack of strong sequence identity between the available structures prevents us from reliably deriving a mechanism of the gating process because it is impossible to disentangle sequence effects from functionally relevant conformational changes.Functional studies indicate that pLGIC can exist, in addition to an open and a resting form, in several closed forms, including desensitized and intermediate forms (11). In this study, we report a X-ray structure of GLIC at neutral pH in an apparently “closed/resting-state.” The resolution is moderate (4.35 Å), but noncrystallographic averaging over the four copies of the pentamer present in the asymmetric unit results in a much better electron density map than suggested by the nominal resolution. At this neutral pH the pore is closed by a tilt of the upper part of helix M2, as in the LC form, with an additional hinge movement of the beginning of helix M1. This is very different from what is seen in ELIC. At the same time, the ECD has undergone significant structural changes that lead to a profound rearrangement of both the ECD–ECD and ECD–TMD interfaces. Strikingly, the large structural variations observed in the four copies of the pentamer at pH 7 imply a rather flat energy landscape in the absence of ligand, whereas a crystal form grown at pH 4, also presented here, shows that the LC form coexists with the open form, thus suggesting an energy landscape made of two wells in the presence of the ligand.  相似文献   

16.
Prostaglandin E2 (PGE2) is a key mediator in inflammatory response. The main source of inducible PGE2, microsomal PGE2 synthase-1 (mPGES-1), has emerged as an interesting drug target for treatment of pain. To support inhibitor design, we have determined the crystal structure of human mPGES-1 to 1.2 Å resolution. The structure reveals three well-defined active site cavities within the membrane-spanning region in each monomer interface of the trimeric structure. An important determinant of the active site cavity is a small cytosolic domain inserted between transmembrane helices I and II. This extra domain is not observed in other structures of proteins within the MAPEG (Membrane-Associated Proteins involved in Eicosanoid and Glutathione metabolism) superfamily but is likely to be present also in microsomal GST-1 based on sequence similarity. An unexpected feature of the structure is a 16-Å-deep cone-shaped cavity extending from the cytosolic side into the membrane-spanning region. We suggest a potential role for this cavity in substrate access. Based on the structure of the active site, we propose a catalytic mechanism in which serine 127 plays a key role. We have also determined the structure of mPGES-1 in complex with a glutathione-based analog, providing insight into mPGES-1 flexibility and potential for structure-based drug design.  相似文献   

17.
Neuroglobin (Ngb), a globular heme protein expressed in the brain of vertebrates, binds oxygen reversibly, with an affinity comparable to myoglobin (Mb). Despite low sequence identity, the overall 3D fold of Ngb and Mb is very similar. Unlike in Mb, in Ngb the sixth coordination position of the heme iron is occupied by the distal histidine, in the absence of an exogenous ligand. Endogenous ligation has been proposed as a unique mechanism for affinity regulation and ligand discrimination in heme proteins. This peculiarity might be related to the still-unknown physiological function of Ngb. Here, we present the x-ray structure of CO-bound ferrous murine Ngb at 1.7 A and a comparison with the 1.5-A structure of ferric bis-histidine Ngb. We have also used Fourier transform IR spectroscopy of WT and mutant CO-ligated Ngb to examine structural heterogeneity in the active site. Upon CO binding, the distal histidine retains (by and large) its position, whereas the heme group slides deeper into a preformed crevice, thereby reshaping the large cavity ( approximately 290 A(3)) connecting the distal and proximal heme sides with the bulk. The heme relocation is accompanied by a significant decrease of structural disorder, especially of the EF loop, which may be the signal whereby Ngb communicates hypoxic conditions. This unexpected structural change unveils a heme-sliding mechanism of affinity control that may be of significance to understanding Ngb's role in the pathophysiology of the brain.  相似文献   

18.
Orsay, the first virus discovered to naturally infect Caenorhabditis elegans or any nematode, has a bipartite, positive-sense RNA genome. Sequence analyses show that Orsay is related to nodaviruses, but molecular characterizations of Orsay reveal several unique features, such as the expression of a capsid–δ fusion protein and the use of an ATG-independent mechanism for translation initiation. Here we report the crystal structure of an Orsay virus-like particle assembled from recombinant capsid protein (CP). Orsay capsid has a T = 3 icosahedral symmetry with 60 trimeric surface spikes. Each CP can be divided into three regions: an N-terminal arm that forms an extended protein interaction network at the capsid interior, an S domain with a jelly-roll, β-barrel fold forming the continuous capsid, and a P domain that forms surface spike projections. The structure of the Orsay S domain is best aligned to T = 3 plant RNA viruses but exhibits substantial differences compared with the insect-infecting alphanodaviruses, which also lack the P domain in their CPs. The Orsay P domain is remotely related to the P1 domain in calicivirus and hepatitis E virus, suggesting a possible evolutionary relationship. Removing the N-terminal arm produced a slightly expanded capsid with fewer nucleic acids packaged, suggesting that the arm is important for capsid stability and genome packaging. Because C. elegans-Orsay serves as a highly tractable model for studying viral pathogenesis, our results should provide a valuable structural framework for further studies of Orsay replication and infection.In the past four decades, the nematode Caenorhabditis elegans has become an important model organism for studying various biological processes such as development, metabolism, aging, the cell cycle, and gene regulation (1). Many key biomedical discoveries about common diseases, including Alzheimer’s disease, diabetes type 2, and even depression, were first made in C. elegans (2). It has been estimated that around 70% of human genes have homologs in C. elegans, including the majority of human disease genes and pathways (3).It was not until very recently that Orsay, the first virus that naturally infects C. elegans in the wild, was discovered (4). In addition, two other nematode viruses, Santeuil and Le Blanc, both of which infect C. briggsae, have also been identified (4, 5). Infections by these newly identified viruses cause abnormal intestinal morphologies without obvious effect on longevity or brood size. Orsay viral particles have been observed intracellularly in both intestinal cells and the somatic gonad (4). Sequence analysis indicated that all three nematode viruses are related to viruses in the family Nodaviridae (4, 5). Nodaviruses are nonenveloped RNA viruses that can be further classified as alphanodaviruses and betanodaviruses, which are known to infect insects and fish, respectively (6).Orsay virus has a bipartite, positive-sense RNA genome ∼6.3 kb in size (4). The RNA1 segment of the Orsay virus encodes a predicted RNA-dependent RNA polymerase (RdRP) of 982 amino acids. RNA2 has two ORFs: one coding for a putative capsid protein (CP) and the other for a nonstructural protein δ of unknown function. The protein translation of the Orsay CP uses an AUG-independent, noncanonical initiation mechanism that is likely shared among all three nematode viruses (7). In addition to the viral CP, a CP–δ fusion protein, which is likely generated by ribosomal frameshifting, has been detected in infected cells as well as purified viruses (7). The Orsay CP and RdRP share up to 26–30% amino acid identity with those from betanodaviruses in the Nodaviridae, but the δ protein does not have any sequence homologs in GenBank.The combination of a simple multicellular host organism and a small naturally infecting virus makes C. elegans-Orsay a highly tractable model for studying host–virus interactions (8). To characterize the molecular properties of the newly discovered Orsay virus, we report the crystal structure of an Orsay virus-like particle (VLP), which shows a T = 3 icosahedral symmetry with a total of 60 trimeric spikes. Each CP molecule can be divided into three linear regions, namely the N-terminal arm, the S domain forming the continuous capsid shell, and the P domain, which forms trimeric surface protrusions. The jelly-roll β-barrel in the S domain closely resembles those from small plant RNA viruses. The P domain consists of a distorted β-barrel that is only remotely related to the P1 domain from a norovirus CP (9). The N-terminal arm of the Orsay CP forms an extensive network inside the capsid through a β-annulus structure similar to tomato bushy stunt virus (10). An N-terminally truncated mutant, CP42–391, produces a slightly expanded capsid with less RNA packaged inside, suggesting important roles for the N-terminal arm in capsid stabilization and genome packaging. Our results also demonstrate that the Orsay capsid is structurally distinct from alphanodaviruses. Together with primary sequence-based phylogeny, these results suggest that the nematode-infecting viruses and betanodaviruses have a common evolutionary history.  相似文献   

19.
The activity of P-type plasma membrane H(+)-ATPases is modulated by H(+) and cations, with K(+) and Ca(2+) being of physiological relevance. Using X-ray crystallography, we have located the binding site for Rb(+) as a K(+) congener, and for Tb(3+) and Ho(3+) as Ca(2+) congeners. Rb(+) is found coordinated by a conserved aspartate residue in the phosphorylation domain. A single Tb(3+) ion is identified positioned in the nucleotide-binding domain in close vicinity to the bound nucleotide. Ho(3+) ions are coordinated at two distinct sites within the H(+)-ATPase: One site is at the interface of the nucleotide-binding and phosphorylation domains, and the other is in the transmembrane domain toward the extracellular side. The identified binding sites are suggested to represent binding pockets for regulatory cations and a H(+) binding site for protons leaving the pump molecule. This implicates Ho(3+) as a novel chemical tool for identification of proton binding sites.  相似文献   

20.
Natural killer (NK) cells are a group of innate immune cells that carry out continuous surveillance for the presence of virally infected or cancerous cells. The natural cytotoxicity receptor (NCR) NKp30 is critical for the elimination of a large group of tumor cell types. Although several ligands have been proposed for NKp30, the lack of a conserved structural feature among these ligands and their uncertain physiological relevance has contributed to confusion in the field and hampered a full understanding of the receptor. To gain insights into NKp30 ligand recognition, we have determined the crystal structure of the extracellular domain of human NKp30. The structure displays an I-type Ig-like fold structurally distinct from the other natural cytotoxicity receptors NKp44 and NKp46. Using cytolytic killing assays against a range of tumor cell lines and subsequent peptide epitope mapping of a NKp30 blocking antibody, we have identified a critical ligand binding region on NKp30 involving its F strand. Using different solution binding studies, we show that the N-terminal domain of B7-H6 is sufficient for NKp30 recognition. Mutations on NKp30 further confirm that residues in the vicinity of the F strand, including part of the C strand and the CD loop, affect binding to B7-H6. The structural comparison of NKp30 with CD28 family receptor and ligand complexes also supports the identified ligand binding site. This study provides insights into NKp30 ligand recognition and a framework for a potential family of unidentified ligands.  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