2024MAY03: Our hosting provider has resolved some DB connectivity issues. We may experience some more outages as the issue is resolved. We apologize for the inconvenience. Dismiss and don't show again

Searching across hundreds of databases

Our searching services are busy right now. Your search will reload in five seconds.

X
Forgot Password

If you have forgotten your password you can enter your email here and get a temporary password sent to your email.

X
Forgot Password

If you have forgotten your password you can enter your email here and get a temporary password sent to your email.

This service exclusively searches for literature that cites resources. Please be aware that the total number of searchable documents is limited to those containing RRIDs and does not include all open-access literature.

Search

Type in a keyword to search

On page 1 showing 1 ~ 20 papers out of 169 papers

Cation-selective pathway of OmpF porin revealed by anomalous X-ray diffraction.

  • Balasundaresan Dhakshnamoorthy‎ et al.
  • Journal of molecular biology‎
  • 2010‎

The OmpF porin from the Escherichia coli outer membrane folds into a trimer of beta-barrels, each forming a wide aqueous pore allowing the passage of ions and small solutes. A long loop (L3) carrying multiple acidic residues folds into the beta-barrel pore to form a narrow "constriction zone". A strong and highly conserved charge asymmetry is observed at the constriction zone, with multiple basic residues attached to the wall of the beta-barrel (Lys16, Arg42, Arg82 and Arg132) on one side, and multiple acidic residues of L3 (Asp107, Asp113, Glu117, Asp121, Asp126, Asp127) on the other side. Several computational studies have suggested that a strong transverse electric field could exist at the constriction zone as a result of such charge asymmetry, giving rise to separate permeation pathways for cations and anions. To examine this question, OmpF was expressed, purified and crystallized in the P6(3) space group and two different data sets were obtained at 2.6 A and 3.0 A resolution with K(+) and Rb(+), respectively. The Rb(+)-soaked crystals were collected at the rubidium anomalous wavelength of 0.8149 A and cation positions were determined. A PEG molecule was observed in the pore region for both the K(+) and Rb(+)-soaked crystals, where it interacts with loop L3. The results reveal the separate pathways of anions and cations across the constriction zone of the OmpF pore.


X-ray crystal structure and specificity of the Plasmodium falciparum malaria aminopeptidase PfM18AAP.

  • Komagal Kannan Sivaraman‎ et al.
  • Journal of molecular biology‎
  • 2012‎

The malarial aminopeptidases have emerged as promising new drug targets for the development of novel antimalarial drugs. The M18AAP of Plasmodium falciparum malaria is a metallo-aminopeptidase that we show demonstrates a highly restricted specificity for peptides with an N-terminal Glu or Asp residue. Thus, the enzyme may function alongside other aminopeptidases in effecting the complete degradation or turnover of proteins, such as host hemoglobin, which provides a free amino acid pool for the growing parasite. Inhibition of PfM18AAP's function using antisense RNA is detrimental to the intra-erythrocytic malaria parasite and, hence, it has been proposed as a potential novel drug target. We report the X-ray crystal structure of the PfM18AAP aminopeptidase and reveal its complex dodecameric assembly arranged via dimer and trimer units that interact to form a large tetrahedron shape that completely encloses the 12 active sites within a central cavity. The four entry points to the catalytic lumen are each guarded by 12 large flexible loops that could control substrate entry into the catalytic sites. PfM18AAP thus resembles a proteasomal-like machine with multiple active sites able to degrade peptide substrates that enter the central lumen. The Plasmodium enzyme shows significant structural differences around the active site when compared to recently determined structures of its mammalian and human homologs, which provides a platform from which a rational approach to inhibitor design of new malaria-specific drugs can begin.


Molecular basis of bacterial defense against host lysozymes: X-ray structures of periplasmic lysozyme inhibitors PliI and PliC.

  • S Leysen‎ et al.
  • Journal of molecular biology‎
  • 2011‎

Lysozymes play a key role in the innate immune system of vertebrates and invertebrates by hydrolyzing peptidoglycan, a vital component of the bacterial cell wall. Gram-negative bacteria produce various types of lysozyme inhibitors that allow them to survive the bactericidal action of lysozyme when their outer membrane is permeabilized. So far, three lysozyme inhibitor families have been described: the Ivy (inhibitor of vertebrate lysozyme) family, the MliC/PliC (membrane-associated/periplasmic lysozyme inhibitor of C-type lysozyme) family, and the PliI (periplasmic lysozyme inhibitor of I-type lysozyme) family. Here, we report high-resolution crystal structures of Salmonella typhimurium PliC (PliC-St) and Aeromonas hydrophila PliI (PliI-Ah). The structure of PliI-Ah is the first in the recently discovered PliI family of lysozyme inhibitors, while the structure of PliC-St is the first structure of a periplasmic lysozyme inhibitor from the PliC/MliC family. Using small-angle X-ray scattering, we demonstrate that both PliC-St and PliI-Ah form stable dimers in solution. The functional dimer architecture of PliC-St is very different from that of the recently described MliC from Pseudomonas aeruginosa (MliC-Pa), despite the close resemblance of their monomers. Furthermore, PliI-Ah has distinctly different monomer and dimer folds compared to PliC, MliC, and Ivy proteins. Site-directed mutagenesis suggests that the inhibitory action of PliI-Ah proceeds via an insertion of a loop containing the conserved SGxY motif into the active center of I-type lysozymes. This motif is related to the functional SGxxY motif found in the MliC/PliC family.


Kinetic and X-ray structural evidence for negative cooperativity in substrate binding to nicotinate mononucleotide adenylyltransferase (NMAT) from Bacillus anthracis.

  • Valerie C Sershon‎ et al.
  • Journal of molecular biology‎
  • 2009‎

Biosynthesis of NAD(P) in bacteria occurs either de novo or through one of the salvage pathways that converge at the point where the reaction of nicotinate mononucleotide (NaMN) with ATP is coupled to the formation of nicotinate adenine dinucleotide (NaAD) and inorganic pyrophosphate. This reaction is catalyzed by nicotinate mononucleotide adenylyltransferase (NMAT), which is essential for bacterial growth, making it an attractive drug target for the development of new antibiotics. Steady-state kinetic and direct binding studies on NMAT from Bacillus anthracis suggest a random sequential Bi-Bi kinetic mechanism. Interestingly, the interactions of NaMN and ATP with NMAT were observed to exhibit negative cooperativity, i.e. Hill coefficients <1.0. Negative cooperativity in binding is supported by the results of X-ray crystallographic studies. X-ray structures of the B. anthracis NMAT apoenzyme, and the NaMN- and NaAD-bound complexes were determined to resolutions of 2.50 A, 2.60 A and 1.75 A, respectively. The X-ray structure of the NMAT-NaMN complex revealed only one NaMN molecule bound in the biological dimer, supporting negative cooperativity in substrate binding. The kinetic, direct-binding, and X-ray structural studies support a model in which the binding affinity of substrates to the first monomer of NMAT is stronger than that to the second, and analysis of the three X-ray structures reveals significant conformational changes of NMAT along the enzymatic reaction coordinate. The negative cooperativity observed in B. anthracis NMAT substrate binding is a unique property that has not been observed in other prokaryotic NMAT enzymes. We propose that regulation of the NAD(P) biosynthetic pathway may occur, in part, at the reaction catalyzed by NMAT.


Small-angle X-ray scattering and single-molecule FRET spectroscopy produce highly divergent views of the low-denaturant unfolded state.

  • Tae Yeon Yoo‎ et al.
  • Journal of molecular biology‎
  • 2012‎

The results of more than a dozen single-molecule Förster resonance energy transfer (smFRET) experiments suggest that chemically unfolded polypeptides invariably collapse from an expanded random coil to more compact dimensions as the denaturant concentration is reduced. In sharp contrast, small-angle X-ray scattering (SAXS) studies suggest that, at least for single-domain proteins at non-zero denaturant concentrations, such compaction may be rare. Here, we explore this discrepancy by studying protein L, a protein previously studied by SAXS (at 5 °C), which suggested fixed unfolded-state dimensions from 1.4 to 5 M guanidine hydrochloride (GuHCl), and by smFRET (at 25 °C), which suggested that, in contrast, the chain contracts by 15-30% over this same denaturant range. Repeating the earlier SAXS study under the same conditions employed in the smFRET studies, we observe little, if any, evidence that the unfolded state of protein L contracts as the concentration of GuHCl is reduced. For example, scattering profiles (and thus the shape and dimensions) collected within ∼4 ms after dilution to as low as 0.67 M GuHCl are effectively indistinguishable from those observed at equilibrium at higher denaturant. Our results thus argue that the disagreement between SAXS and smFRET is statistically significant and that the experimental evidence in favor of obligate polypeptide collapse at low denaturant cannot be considered conclusive yet.


X-ray Crystallographic and Molecular Dynamic Analyses of Drosophila melanogaster Embryonic Muscle Myosin Define Domains Responsible for Isoform-Specific Properties.

  • James T Caldwell‎ et al.
  • Journal of molecular biology‎
  • 2020‎

Drosophila melanogaster is a powerful system for characterizing alternative myosin isoforms and modeling muscle diseases, but high-resolution structures of fruit fly contractile proteins have not been determined. Here we report the first x-ray crystal structure of an insect myosin: the D melanogaster skeletal muscle myosin II embryonic isoform (EMB). Using our system for recombinant expression of myosin heavy chain (MHC) proteins in whole transgenic flies, we prepared and crystallized stable proteolytic S1-like fragments containing the entire EMB motor domain bound to an essential light chain. We solved the x-ray crystal structure by molecular replacement and refined the resulting model against diffraction data to 2.2 Å resolution. The protein is captured in two slightly different renditions of the rigor-like conformation with a citrate of crystallization at the nucleotide binding site and exhibits structural features common to myosins of diverse classes from all kingdoms of life. All atom molecular dynamics simulations on EMB in its nucleotide-free state and a derivative homology model containing 61 amino acid substitutions unique to the indirect flight muscle isoform (IFI) suggest that differences in the identity of residues within the relay and the converter that are encoded for by MHC alternative exons 9 and 11, respectively, directly contribute to increased mobility of these regions in IFI relative to EMB. This suggests the possibility that alternative folding or conformational stability within these regions contribute to the observed functional differences in Drosophila EMB and IFI myosins.


Fiber diffraction data indicate a hollow core for the Alzheimer's aβ 3-fold symmetric fibril.

  • Michele McDonald‎ et al.
  • Journal of molecular biology‎
  • 2012‎

Amyloid β protein (Aβ), the principal component of the extracellular plaques found in the brains of patients with Alzheimer's disease, forms fibrils well suited to structural study by X-ray fiber diffraction. Fiber diffraction patterns from the 40-residue form Aβ(1-40) confirm a number of features of a 3-fold symmetric Aβ model from solid-state NMR (ssNMR) but suggest that the fibrils have a hollow core not present in the original ssNMR models. Diffraction patterns calculated from a revised 3-fold hollow model with a more regular β-sheet structure are in much better agreement with the observed diffraction data than patterns calculated from the original ssNMR model. Refinement of a hollow-core model against ssNMR data led to a revised ssNMR model, similar to the fiber diffraction model.


Neutron diffraction studies of a class A beta-lactamase Toho-1 E166A/R274N/R276N triple mutant.

  • Stephen J Tomanicek‎ et al.
  • Journal of molecular biology‎
  • 2010‎

beta-Lactam antibiotics have been used effectively over several decades against many types of bacterial infectious diseases. However, the most common cause of resistance to the beta-lactam antibiotics is the production of beta-lactamase enzymes that inactivate beta-lactams by rapidly hydrolyzing the amide group of the beta-lactam ring. Specifically, the class A extended-spectrum beta-lactamases (ESBLs) and inhibitor-resistant enzymes arose that were capable of hydrolyzing penicillins and the expanded-spectrum cephalosporins and monobactams in resistant bacteria, which lead to treatment problems in many clinical settings. A more complete understanding of the mechanism of catalysis of these ESBL enzymes will impact current antibiotic drug discovery efforts. Here, we describe the neutron structure of the class A, CTX-M-type ESBL Toho-1 E166A/R274N/R276N triple mutant in its apo form, which is the first reported neutron structure of a beta-lactamase enzyme. This neutron structure clearly reveals the active-site protonation states and hydrogen-bonding network of the apo Toho-1 ESBL prior to substrate binding and subsequent acylation. The protonation states of the active-site residues Ser70, Lys73, Ser130, and Lys234 in this neutron structure are consistent with the prediction of a proton transfer pathway from Lys73 to Ser130 that is likely dependent on the conformation of Lys73, which has been hypothesized to be coupled to the protonation state of Glu166 during the acylation reaction. Thus, this neutron structure is in agreement with a proposed mechanism for acylation that identifies Glu166 as the general base for catalysis.


Structure of hibiscus latent singapore virus by fiber diffraction: a nonconserved his122 contributes to coat protein stability.

  • Sunil Kumar Tewary‎ et al.
  • Journal of molecular biology‎
  • 2011‎

Hibiscus latent Singapore virus (HLSV) is a rigid rod-shaped plant virus and a new member of the Tobamovirus family. Unlike all other Tobamoviruses, the HLSV genome contains a unique poly(A) tract in its 3' untranslated region. The virion is composed of a monomeric coat protein (CP) unit of 18 kDa, arranged as a right-handed helix around the virus axis. We have determined the structure of HLSV at 3.5 Å by X-ray fiber diffraction and refined it to an R-factor of 0.096. While the overall structure of the HLSV CP resembles that of other Tobamoviruses, there are a few unique differences. There is a kink in the LR helix due to the presence of His122. Also, the adjacent Lys123 may further destabilize the helix by positive charge repulsion, making the kink more pronounced. The His122-Asp88 salt bridge provides significant stability to the loop adjacent to the RR helix. Carboxyl-carboxylate interactions that drive viral disassembly are also different in HLSV. The nucleotide recognition mechanisms for virus assembly between HLSV and ribgrass mosaic virus are similar, but different between tobacco mosaic virus and cucumber green mottle mosaic virus.


Crystal structure of the light-driven chloride pump halorhodopsin from Natronomonas pharaonis.

  • Tsutomu Kouyama‎ et al.
  • Journal of molecular biology‎
  • 2010‎

The light-driven chloride pump halorhodopsin from Natronomonas pharaonis (phR) crystallised into the monoclinic space group C2, with a phR trimer per the asymmetric unit. Diffraction data at 2.0-A resolution showed that the carotenoid bacterioruberin binds to crevices between adjacent protein subunits in the trimeric assembly. Besides seven transmembrane helices (A to G) that characterise archaeal rhodopsins, the phR protomer possesses an amphipathic alpha-helix (A') at the N-terminus. This helix, together with a long loop between helices B and C, forms a hydrophobic cap that covers the extracellular surface and prevents a rapid ion exchange between the active centre and the extracellular medium. The retinal bound to Lys256 in helix G takes on an all-trans configuration with the Schiff base being hydrogen-bonded to a water molecule. The Schiff base also interacts with Asp252 and a chloride ion, the latter being fixed by two polar groups (Thr126 and Ser130) in helix C. In the anion uptake pathway, four ionisable residues (Arg123, Glu234, Arg176 and His100) and seven water molecules are aligned to form a long hydrogen-bonding network. Conversely, the cytoplasmic half is filled mostly by hydrophobic residues, forming a large energetic barrier against the transport of anion. The height of this barrier would be lowered substantially if the cytoplasmic half functions as a proton/HCl antiporter. Interestingly, there is a long cavity extending from the main-chain carbonyl of Lys256 to Thr71 in helix B. This cavity, which is commonly seen in halobacterial light-driven proton pumps, is one possible pathway that is utilised for a water-mediated proton transfer from the cytoplasmic medium to the anion, which is relocated to the cytoplasmic channel during the photocycle.


Structural basis for exquisite specificity of affinity clamps, synthetic binding proteins generated through directed domain-interface evolution.

  • Jin Huang‎ et al.
  • Journal of molecular biology‎
  • 2009‎

We have established a new protein-engineering strategy termed "directed domain-interface evolution" that generates a binding site by linking two protein domains and then optimizing the interface between them. Using this strategy, we have generated synthetic two-domain "affinity clamps" using PDZ and fibronectin type III (FN3) domains as the building blocks. While these affinity clamps all had significantly higher affinity toward a target peptide than the underlying PDZ domain, two distinct types of affinity clamps were found in terms of target specificity. One type conserved the specificity of the parent PDZ domain, and the other increased the specificity dramatically. Here, we characterized their specificity profiles using peptide phage-display libraries and scanning mutagenesis, which suggested a significantly enlarged recognition site of the high-specificity affinity clamps. The crystal structure of a high-specificity affinity clamp showed extensive contacts with a portion of the peptide ligand that is not recognized by the parent PDZ domain, thus rationalizing the improvement of the specificity of the affinity clamp. A comparison with another affinity clamp structure showed that, although both had extensive contacts between PDZ and FN3 domains, they exhibited a large offset in the relative position of the two domains. Our results indicate that linked domains could rapidly fuse and evolve as a single functional module, and that the inherent plasticity of domain interfaces allows for the generation of diverse active-site topography. These attributes of directed domain-interface evolution provide facile means to generate synthetic proteins with a broad range of functions.


Structural and functional characterization of plant aminoaldehyde dehydrogenase from Pisum sativum with a broad specificity for natural and synthetic aminoaldehydes.

  • Martina Tylichová‎ et al.
  • Journal of molecular biology‎
  • 2010‎

Aminoaldehyde dehydrogenases (AMADHs, EC 1.2.1.19) belong to the large aldehyde dehydrogenase (ALDH) superfamily, namely, the ALDH9 family. They oxidize polyamine-derived omega-aminoaldehydes to the corresponding omega-amino acids. Here, we report the first X-ray structures of plant AMADHs: two isoenzymes, PsAMADH1 and PsAMADH2, from Pisum sativum in complex with beta-nicotinamide adenine dinucleotide (NAD(+)) at 2.4 and 2.15 A resolution, respectively. Both recombinant proteins are dimeric and, similarly to other ALDHs, each monomer is composed of an oligomerization domain, a coenzyme binding domain and a catalytic domain. Each subunit binds NAD(+) as a coenzyme, contains a solvent-accessible C-terminal peroxisomal targeting signal (type 1) and a cation bound in the cavity close to the NAD(+) binding site. While the NAD(+) binding mode is classical for PsAMADH2, that for PsAMADH1 is unusual among ALDHs. A glycerol molecule occupies the substrate binding site and mimics a bound substrate. Structural analysis and substrate specificity study of both isoenzymes in combination with data published previously on other ALDH9 family members show that the established categorization of such enzymes into distinct groups based on substrate specificity is no more appropriate, because many of them seem capable of oxidizing a large spectrum of aminoaldehyde substrates. PsAMADH1 and PsAMADH2 can oxidize N,N,N-trimethyl-4-aminobutyraldehyde into gamma-butyrobetaine, which is the carnitine precursor in animal cells. This activity highly suggests that in addition to their contribution to the formation of compatible osmolytes such as glycine betaine, beta-alanine betaine and gamma-aminobutyric acid, AMADHs might participate in carnitine biosynthesis in plants.


Crystal structure of a subtilisin homologue, Tk-SP, from Thermococcus kodakaraensis: requirement of a C-terminal beta-jelly roll domain for hyperstability.

  • Tita Foophow‎ et al.
  • Journal of molecular biology‎
  • 2010‎

Tk-SP is a hyperthermostable subtilisin-like serine protease from Thermococcus kodakaraensis and is autoprocessed from its precursor (Pro-Tk-SP) with N- and C-propeptides. The crystal structure of the active-site mutant of Pro-Tk-SP lacking C-propeptide, ProN-Tk-S359A, was determined at 2.0 A resolution. ProN-Tk-S359A consists of the N-propeptide, subtilisin, and beta-jelly roll domains. Two Ca(2+) ions bind to the beta-jelly roll domain. The overall structure of ProN-Tk-S359A without the beta-jelly roll domain is similar to that of the bacterial propeptide:subtilisin complex, except that it does not contain Ca(2+) ions. To analyze the role of the beta-jelly roll domain of Tk-SP, we constructed a series of the active-site mutants of Tk-SP with (Tk-S359A/C) and without (Tk-S359A/CDeltaJ) beta-jelly roll domain. Both Tk-S359C and Tk-S359CDeltaJ exhibited protease activities in gel assay, indicating that the beta-jelly roll domain is not required for folding or activity. However, the T(m) value of Tk-S359ADeltaJ determined by far-UV CD spectroscopy in the presence of 10-mM CaCl(2) was lower than that of Tk-S359A by 29.4 degrees C. The T(m) value of Tk-S359A was decreased by 29.5 degrees C by the treatment with 10 mM ethylenediaminetetraacetic acid, indicating that the beta-jelly roll domain contributes to the stabilization of Tk-S359A only in a Ca(2+)-bound form. Tk-SP highly resembles subtilisin-like serine proteases from Pyrococcus furiosus, Thermococcus gammatolerans, and Thermococcus onnurineus in size and amino acid sequence. We propose that attachment of a beta-jelly roll domain to the C-terminus is one of the strategies of the proteins from hyperthermophiles to adapt to high-temperature environment.


Crystal structures of human HMG-CoA synthase isoforms provide insights into inherited ketogenesis disorders and inhibitor design.

  • Naeem Shafqat‎ et al.
  • Journal of molecular biology‎
  • 2010‎

3-Hydroxy-3-methylglutaryl coenzyme A (CoA) synthase (HMGCS) catalyzes the condensation of acetyl-CoA and acetoacetyl-CoA into 3-hydroxy-3-methylglutaryl CoA. It is ubiquitous across the phylogenetic tree and is broadly classified into three classes. The prokaryotic isoform is essential in Gram-positive bacteria for isoprenoid synthesis via the mevalonate pathway. The eukaryotic cytosolic isoform also participates in the mevalonate pathway but its end product is cholesterol. Mammals also contain a mitochondrial isoform; its deficiency results in an inherited disorder of ketone body formation. Here, we report high-resolution crystal structures of the human cytosolic (hHMGCS1) and mitochondrial (hHMGCS2) isoforms in binary product complexes. Our data represent the first structures solved for human HMGCS and the mitochondrial isoform, allowing for the first time structural comparison among the three isoforms. This serves as a starting point for the development of isoform-specific inhibitors that have potential cholesterol-reducing and antibiotic applications. In addition, missense mutations that cause mitochondrial HMGCS deficiency have been mapped onto the hHMGCS2 structure to rationalize the structural basis for the disease pathology.


Coupling of domain swapping to kinetic stability in a thioredoxin mutant.

  • Abel Garcia-Pino‎ et al.
  • Journal of molecular biology‎
  • 2009‎

The thioredoxin (Trx) fold is a small monomeric domain that is ubiquitous in redox-active enzymes. Trxs are characterized by a typical WCGPC active-site sequence motif. A single active-site mutation of the tryptophan to an alanine in Staphylococcus aureus Trx converts the oxidized protein into a biologically inactive domain-swapped dimer. While the monomeric protein unfolds reversibly in a two-state manner, the oxidized dimeric form is kinetically stable and converts to the monomeric form upon refolding. After reduction, the half-life of the dimer decreases many orders of magnitude to approximately 4.3 h, indicating that the active-site disulfide between Cys29 and Cys32 is an important determinant for the kinetics of unfolding. We propose kinetic stability as a possible evolutionary strategy in the evolution of multimeric proteins from their monomeric ancestors by domain swapping, which, for this biologically inactive Trx mutant, turned out to be an evolutionary dead end.


Crystal structure of dicamba monooxygenase: a Rieske nonheme oxygenase that catalyzes oxidative demethylation.

  • Razvan Dumitru‎ et al.
  • Journal of molecular biology‎
  • 2009‎

Dicamba (3,6-dichloro-2-methoxybenzoic acid) is a widely used herbicide that is efficiently degraded by soil microbes. These microbes use a novel Rieske nonheme oxygenase, dicamba monooxygenase (DMO), to catalyze the oxidative demethylation of dicamba to 3,6-dichlorosalicylic acid (DCSA) and formaldehyde. We have determined the crystal structures of DMO in the free state, bound to its substrate dicamba, and bound to the product DCSA at 2.10-1.75 A resolution. The structures show that the DMO active site uses a combination of extensive hydrogen bonding and steric interactions to correctly orient chlorinated, ortho-substituted benzoic-acid-like substrates for catalysis. Unlike other Rieske aromatic oxygenases, DMO oxygenates the exocyclic methyl group, rather than the aromatic ring, of its substrate. This first crystal structure of a Rieske demethylase shows that the Rieske oxygenase structural scaffold can be co-opted to perform varied types of reactions on xenobiotic substrates.


Switching from an induced-fit to a lock-and-key mechanism in an aminoacyl-tRNA synthetase with modified specificity.

  • Emmanuelle Schmitt‎ et al.
  • Journal of molecular biology‎
  • 2009‎

Methionyl-tRNA synthetase (MetRS) specifically binds its methionine substrate in an induced-fit mechanism, with methionine binding causing large rearrangements. Mutated MetRS able to efficiently aminoacylate the methionine (Met) analog azidonorleucine (Anl) have been identified by saturation mutagenesis combined with in vivo screening procedures. Here, the crystal structure of such a mutated MetRS was determined in the apo form as well as complexed with Met or Anl (1.4 to 1.7 A resolution) to reveal the structural basis for the altered specificity. The mutations result in both the loss of important contacts with Met and the creation of new contacts with Anl, thereby explaining the specificity shift. Surprisingly, the conformation induced by Met binding in wild-type MetRS already occurs in the apo form of the mutant enzyme. Therefore, the mutations cause the enzyme to switch from an induced-fit mechanism to a lock-and-key one, thereby enhancing its catalytic efficiency.


Crystal structure of red chlorophyll catabolite reductase: enlargement of the ferredoxin-dependent bilin reductase family.

  • Masakazu Sugishima‎ et al.
  • Journal of molecular biology‎
  • 2009‎

The key steps in the degradation pathway of chlorophylls are the ring-opening reaction catalyzed by pheophorbide a oxygenase and sequential reduction by red chlorophyll catabolite reductase (RCCR). During these steps, chlorophyll catabolites lose their color and phototoxicity. RCCR catalyzes the ferredoxin-dependent reduction of the C20/C1 double bond of red chlorophyll catabolite. RCCR appears to be evolutionarily related to the ferredoxin-dependent bilin reductase (FDBR) family, which synthesizes a variety of phytobilin pigments, on the basis of sequence similarity, ferredoxin dependency, and the common tetrapyrrole skeleton of their substrates. The evidence, however, is not robust; the identity between RCCR and FDBR HY2 from Arabidopsis thaliana is only 15%, and the oligomeric states of these enzymes are different. Here, we report the crystal structure of A. thaliana RCCR at 2.4 A resolution. RCCR forms a homodimer, in which each subunit folds in an alpha/beta/alpha sandwich. The tertiary structure of RCCR is similar to those of FDBRs, strongly supporting that these enzymes evolved from a common ancestor. The two subunits are related by noncrystallographic 2-fold symmetry in which the alpha-helices near the edge of the beta-sheet unique in RCCR participate in intersubunit interaction. The putative RCC-binding site, which was derived by superimposing RCCR onto biliverdin-bound forms of FDBRs, forms an open pocket surrounded by conserved residues among RCCRs. Glu154 and Asp291 of A. thaliana RCCR, which stand opposite each other in the pocket, likely are involved in substrate binding and/or catalysis.


Effect of xenon binding to a hydrophobic cavity on the proton pumping cycle in bacteriorhodopsin.

  • Naoki Hayakawa‎ et al.
  • Journal of molecular biology‎
  • 2008‎

To understand the functional role of apolar cavities in bacteriorhodopsin, a light-driven proton pump found in Halobacterium salinarum, we investigated the crystal structure in pressurized xenon or krypton. Diffraction data from the P622 crystal showed that one Xe or Kr atom binds to a preexisting hydrophobic cavity buried between helices C and D, located at the same depth from the membrane surface as Asp96, a key residue in the proton uptake pathway. The occupation fraction of Xe or Kr was calculated as approximately 0.32 at a pressure of 1 MPa. In the unphotolyzed state, the binding of Xe or Kr caused no large deformation of the cavity. However, the proton pumping cycle was greatly perturbed when an aqueous suspension of purple membrane was pressurized with xenon gas; that is, the decay of the M state was accelerated significantly (~5 times at full occupancy), while the decay of an equilibrium state of N and O was slightly decelerated. A similar but much smaller perturbation in the reaction kinetics was observed upon pressurization with krypton gas. In a glycerol/water mixture, xenon-induced acceleration of M decay became less significant in proportion to the water activity. Together with the structure of the xenon-bound protein, these observations suggest that xenon binding helps water molecules permeate into apolar cavities in the proton uptake pathway, thereby accelerating the water-mediated proton transfer from Asp96 to the Schiff base.


Crystallographic analysis of the primary photochemical reaction of squid rhodopsin.

  • Midori Murakami‎ et al.
  • Journal of molecular biology‎
  • 2011‎

Visual signal transduction is initiated by the photoisomerization of 11-cis retinal upon rhodopsin ligation. Unlike vertebrate rhodopsin, which interacts with Gt-type G-protein to stimulate the cyclic GMP signaling pathway, invertebrate rhodopsin interacts with Gq-type G-protein to stimulate a signaling pathway that is based on inositol 1,4,5-triphosphate. Since the inositol 1,4,5-triphosphate signaling pathway is utilized by mammalian nonvisual pigments and a large number of G-protein-coupled receptors, it is important to elucidate how the activation mechanism of invertebrate rhodopsin differs from that of vertebrate rhodopsin. Previous crystallographic studies of squid and bovine rhodopsins have shown that there is a profound difference in the structures of the retinal-binding pockets of these photoreceptors. Here, we report the crystal structures of all-trans bathorhodopsin (Batho; the first photoreaction intermediate) and the artificial 9-cis isorhodopsin (Iso) of squid rhodopsin. Upon the formation of Batho, the central moiety of the retinal was observed to move largely towards the cytoplasmic side, while the Schiff base and the ionone ring underwent limited movements (i.e., the all-trans retinal in Batho took on a right-handed screwed configuration). Conversely, the 9-cis retinal in Iso took on a planar configuration. Our results suggest that the light energy absorbed by squid rhodopsin is mostly converted into the distortion energy of the retinal polyene chain and surrounding residues.


  1. SciCrunch.org Resources

    Welcome to the FDI Lab - SciCrunch.org Resources search. From here you can search through a compilation of resources used by FDI Lab - SciCrunch.org and see how data is organized within our community.

  2. Navigation

    You are currently on the Community Resources tab looking through categories and sources that FDI Lab - SciCrunch.org has compiled. You can navigate through those categories from here or change to a different tab to execute your search through. Each tab gives a different perspective on data.

  3. Logging in and Registering

    If you have an account on FDI Lab - SciCrunch.org then you can log in from here to get additional features in FDI Lab - SciCrunch.org such as Collections, Saved Searches, and managing Resources.

  4. Searching

    Here is the search term that is being executed, you can type in anything you want to search for. Some tips to help searching:

    1. Use quotes around phrases you want to match exactly
    2. You can manually AND and OR terms to change how we search between words
    3. You can add "-" to terms to make sure no results return with that term in them (ex. Cerebellum -CA1)
    4. You can add "+" to terms to require they be in the data
    5. Using autocomplete specifies which branch of our semantics you with to search and can help refine your search
  5. Save Your Search

    You can save any searches you perform for quick access to later from here.

  6. Query Expansion

    We recognized your search term and included synonyms and inferred terms along side your term to help get the data you are looking for.

  7. Collections

    If you are logged into FDI Lab - SciCrunch.org you can add data records to your collections to create custom spreadsheets across multiple sources of data.

  8. Facets

    Here are the facets that you can filter your papers by.

  9. Options

    From here we'll present any options for the literature, such as exporting your current results.

  10. Further Questions

    If you have any further questions please check out our FAQs Page to ask questions and see our tutorials. Click this button to view this tutorial again.

Publications Per Year

X

Year:

Count: