scispace - formally typeset
Search or ask a question

Showing papers in "Protein Science in 1994"


Journal ArticleDOI
TL;DR: In this paper, the authors derived the distribution functions of light-ray operators and derived the evolution equations for these distribution functions on the basis of the renormalization group equation of the considered operators.
Abstract: The widely used nonperturbative wave functions and distribution functions of QCD are determined as matrix elements of light-ray operators. These operators appear as large momentum limit of nonlocal hadron operators or as summed up local operators in light-cone expansions. Nonforward one-particle matrix elements of such operators lead to new distribution amplitudes describing both hadrons simultaneously. These distribution functions depend besides other variables on two scaling variables. They are applied for the description of exclusive virtual Compton scattering in the Bjorken region near forward direction and the two meson production process. The evolution equations for these distribution amplitudes are derived on the basis of the renormalization group equation of the considered operators. This includes that also the evolution kernels follow from the anomalous dimensions of these operators. Relations between different evolution kernels (especially the Altarelli-Parisi and the Brodsky-Lepage) kernels are derived and explicitly checked for the existing two-loop calculations of QCD. Technical basis of these results are support and analytically properties of the anomalous dimensions of light-ray operators obtained with the help of the $\alpha$-representation of Green's functions.

967 citations


Journal ArticleDOI
TL;DR: Three thousand eight hundred ninety‐nine β‐turns have been identified and classified using a nonhomologous data set of 205 protein chains to derive β‐ turn positional potentials for turn types I' and II' for the first time and to provide updated potentialS for formation of the more common types I, II, and VIII.
Abstract: Three thousand eight hundred ninety-nine beta-turns have been identified and classified using a nonhomologous data set of 205 protein chains. These were used to derive beta-turn positional potentials for turn types I' and II' for the first time and to provide updated potentials for formation of the more common types I, II, and VIII. Many of the sequence preferences for each of the 4 positions in turns can be rationalized in terms of the formation of stabilizing hydrogen bonds, preferences for amino acids to adopt a particular conformation in phi, psi space, and the involvement of turn types I' and II' in beta-hairpins. Only 1,632 (42%) of the turns occur in isolation; the remainder have at least 1 residue in common with another turn and have hence been classified as multiple turns. Several types of multiple turn have been identified and analyzed.

884 citations


Journal ArticleDOI
TL;DR: To reduce redundancy in the Protein Data Bank of 3D protein structures, which is caused by many homologous proteins in the data bank, a representative set of structures is selected to reduce time and effort in statistical analyses.
Abstract: To reduce redundancy in the Protein Data Bank of 3D protein structures, which is caused by many homologous proteins in the data bank, we have selected a representative set of structures. The selection algorithm was designed to (1) select as many nonhomologous structures as possible, and (2) to select structures of good quality. The representative set may reduce time and effort in statistical analyses.

800 citations


Journal ArticleDOI
TL;DR: The electrostatic contribution to the free energy of folding was calculated fo 21 salt bridges in 9 r X‐ray crystal structures using a continuum electrostatic approach with the DELPHI computer‐program package, which found the majority were found to be electrostatically destabilizing.
Abstract: The electrostatic contribution to the free energy of folding was calculated for 21 salt bridges in 9 protein X-ray crystal structures using a continuum electrostatic approach with the DELPHI computer-program package. The majority (17) were found to be electrostatically destabilizing; the average free energy change, which is analogous to mutation of salt bridging side chains to hydrophobic isosteres, was calculated to be 3.5 kcal/mol. This is fundamentally different from stability measurements using pKa shifts, which effectively measure the strength of a salt bridge relative to 1 or more charged hydrogen bonds. The calculated effect was due to a large, unfavorable desolvation contribution that was not fully compensated by favorable interactions within the salt bridge and between salt-bridge partners and other polar and charged groups in the folded protein. Some of the salt bridges were studied in further detail to determine the effect of the choice of values for atomic radii, internal protein dielectric constant, and ionic strength used in the calculations. Increased ionic strength resulted in little or no change in calculated stability for 3 of 4 salt bridges over a range of 0.1-0.9 M. The results suggest that mutation of salt bridges, particularly those that are buried, to "hydrophobic bridges" (that pack at least as well as wild type) can result in proteins with increased stability. Due to the large penalty for burying uncompensated ionizable groups, salt bridges could help to limit the number of low free energy conformations of a molecule or complex and thus play a role in determining specificity (i.e., the uniqueness of a protein fold or protein-ligand binding geometry).

626 citations


Journal ArticleDOI
TL;DR: Helix‐capping interactions were found to contribute to helix stability, even when the substitution site was not at the end of the peptide, and the implication for protein folding is that formation of peptide H‐bonds can largely offset the unfavorable entropy change caused by fixing the peptides backbone.
Abstract: Helix propensities of the amino acids have been measured in alanine-based peptides in the absence of helix-stabilizing side-chain interactions. Fifty-eight peptides have been studied. A modified form of the Lifson-Roig theory for the helix-coil transition, which includes helix capping (Doig AJ, Chakrabartty A, Klingler TM, Baldwin RL, 1994, Biochemistry 33:3396-3403), was used to analyze the results. Substitutions were made at various positions of homologous helical peptides. Helix-capping interactions were found to contribute to helix stability, even when the substitution site was not at the end of the peptide. Analysis of our data with the original Lifson-Roig theory, which neglects capping effects, does not produce as good a fit to the experimental data as does analysis with the modified Lifson-Roig theory. At 0 “C, Ala is a strong helix former, Leu and Arg are helix-indifferent, and all other amino acids are helix breakers of varying severity. Because Ala has a small side chain that cannot interact significantly with other side chains, helix formation by Ala is stabilized predominantly by the backbone (“peptide H-bonds”). The implication for protein folding is that formation of peptide H-bonds can largely offset the unfavorable entropy change caused by fixing the peptide backbone. The helix propensities of most amino acids oppose folding; consequently, the majority of isolated helices derived from proteins are unstable, unless specific side-chain interactions stabilize them.

587 citations


Journal ArticleDOI
TL;DR: The CTB5:GM1 pentasaccharide complex described here provides a detailed view of aprotein:ganglioside specific binding interaction, and is of interest not only for understanding cholera pathogenesis and for the design of drugs and development of vaccines but also for modeling other protein:gangLioside interactions such as those involved in GM1‐mediated signal transduction.
Abstract: Cholera toxin (CT) is an AB5 hexameric protein responsible for the symptoms produced by Vibrio cholerae infection. In the first step of cell intoxication, the B-pentamer of the toxin binds specifically to the branched pentasaccharide moiety of ganglioside GM1 on the surface of target human intestinal epithelial cells. We present here the crystal structure of the cholera toxin B-pentamer complexed with the GM1 pentasaccharide. Each receptor binding site on the toxin is found to lie primarily within a single B-subunit, with a single solvent-mediated hydrogen bond from residue Gly 33 of an adjacent subunit. The large majority of interactions between the receptor and the toxin involve the 2 terminal sugars of GM1, galactose and sialic acid, with a smaller contribution from the N-acetyl galactosamine residue. The binding of GM1 to cholera toxin thus resembles a 2-fingered grip: the Gal(beta 1-3)GalNAc moiety representing the "forefinger" and the sialic acid representing the "thumb." The residues forming the binding site are conserved between cholera toxin and the homologous heat-labile enterotoxin from Escherichia coli, with the sole exception of His 13. Some reported differences in the binding affinity of the 2 toxins for gangliosides other than GM1 may be rationalized by sequence differences at this residue. The CTB5:GM1 pentasaccharide complex described here provides a detailed view of a protein:ganglioside specific binding interaction, and as such is of interest not only for understanding cholera pathogenesis and for the design of drugs and development of vaccines but also for modeling other protein:ganglioside interactions such as those involved in GM1-mediated signal transduction.

564 citations


Journal ArticleDOI
TL;DR: This structural motif appears to be one of the smallest stable globular domains found in proteins and is commonly used in toxins and inhibitors that act by blocking the function of larger protein receptors such as ion channels or proteases.
Abstract: A common structural motif consisting of a cystine knot and a small triple-stranded beta-sheet has been defined from comparison of the 3-dimensional structures of the polypeptides omega-conotoxin GVIA (Conus geographus), kalata BI (Oldenlandia affinis DC), and CMTI-I (Curcurbita maxima). These 3 polypeptides have diverse biological activities and negligible amino acid sequence identity, but each contains 3 disulfide bonds that give rise to a cystine knot. This knot consists of a ring formed by the first 2 bonds (1-4 and 2-5) and the intervening polypeptide backbone, through which the third disulfide (3-6) passes. The other component of this motif is a triple-stranded, anti-parallel beta-sheet containing a minimum of 10 residues, XXC2, XC5X, XXC6X (where the numbers on the half-cysteine residues refer to their positions in the disulfide pattern). The presence in these polypeptides of both the cysteine knot and antiparallel beta-sheet suggests that both structural features are required for the stability of the motif. This structural motif is also present in other protease inhibitors and a spider toxin. It appears to be one of the smallest stable globular domains found in proteins and is commonly used in toxins and inhibitors that act by blocking the function of larger protein receptors such as ion channels or proteases.

512 citations


Journal ArticleDOI
TL;DR: It is shown that the diffusion‐collision model of protein folding is supported by a growing body of experimental and theoretical evidence, and future directions for developing the model and its applications are outlined.
Abstract: The diffusion-collision model of protein folding is assessed. A description is given of the qualitative aspects and quantitative results of the diffusion-collision model and their relation to available experimental data. We consider alternative mechanisms for folding and point out their relationship to the diffusion-collision model. We show that the diffusion-collision model is supported by a growing body of experimental and theoretical evidence, and we outline future directions for developing the model and its applications.

455 citations


Journal ArticleDOI
TL;DR: Results suggest that surface hydrophobicity can be used to identify regions of a protein's surface most likely to interact with a binding ligand, and may be useful for identifying small sets of well‐defined loci for possible ligand attachment.
Abstract: The role of hydrophobicity as a determinant of protein-protein interactions is examined. Surfaces of apo-protein targets comprising 9 classes of enzymes, 7 antibody fragments, hirudin, growth hormone, and retinol-binding protein, and their associated ligands with available X-ray structures for their complexed forms, are scanned to determine clusters of surface-accessible amino acids. Clusters of surface residues are ranked on the basis of the hydrophobicity of their constituent amino acids. The results indicate that the location of the co-crystallized ligand is commonly found to correspond with one of the strongest hydrophobic clusters on the surface of the target molecule. In 25 of 38 cases, the correspondence is exact, with the position of the most hydrophobic cluster coinciding with more than one-third of the surface buried by the bound ligand. The remaining 13 cases demonstrate this correspondence within the top 6 hydrophobic clusters. These results suggest that surface hydrophobicity can be used to identify regions of a protein's surface most likely to interact with a binding ligand. This fast and simple procedure may be useful for identifying small sets of well-defined loci for possible ligand attachment.

385 citations


Journal ArticleDOI
TL;DR: Comparison of this structure with the previously determined “open” form of this lipase, in which the active site is accessible to the solvent and presumably the substrate, shows that the transition between these 2 states requires only movement of the flap.
Abstract: The structure of Candida rugosa lipase in a new crystal form has been determined and refined at 2.1 A resolution. The lipase molecule was found in an inactive conformation, with the active site shielded from the solvent by a part of the polypeptide chain-the flap. Comparison of this structure with the previously determined "open" form of this lipase, in which the active site is accessible to the solvent and presumably the substrate, shows that the transition between these 2 states requires only movement of the flap. The backbone NH groups forming the putative oxyanion hole do not change position during this rearrangement, indicating that this feature is preformed in the inactive state. The 2 lipase conformations probably correspond to states at opposite ends of the pathway of interfacial activation. Quantitative analysis indicates a large increase of the hydrophobic surface in the vicinity of the active site. The flap undergoes a flexible rearrangement during which some of its secondary structure refolds. The interactions of the flap with the rest of the protein change from mostly hydrophobic in the inactive form to largely hydrophilic in the "open" conformation. Although the flap movement cannot be described as a rigid body motion, it has very definite hinge points at Glu 66 and at Pro 92. The rearrangement is accompanied by a cis-trans isomerization of this proline, which likely increases the energy required for the transition between the 2 states, and may play a role in the stabilization of the active conformation at the water/lipid interface. Carbohydrate attached at Asn 351 also provides stabilization for the open conformation of the flap.

355 citations


Journal ArticleDOI
TL;DR: Investigation of the mechanism by which HNP‐2, one of 4 human neutrophil defensins, permeabilizes large unilamellar vesicles formed from the anionic lipid palmitoyloleoylphosphatidylglycerol (POPG) suggests that H NP‐2 can form pores that have a maximum diameter of approximately 25 Å, and introduces a simple method for determining whether leakage from vesicle contents is graded or all
Abstract: Defensins comprise a family of broad-spectrum antimicrobial peptides that are stored in the cytoplasmic granules of mammalian neutrophils and Paneth cells of the small intestine. Neutrophil defensins are known to permeabilize cell membranes of susceptible microorganisms, but the mechanism of permeabilization is uncertain. We report here the results of an investigation of the mechanism by which HNP-2, one of 4 human neutrophil defensins, permeabilizes large unilamellar vesicles formed from the anionic lipid palmitoyloleoylphosphatidylglycerol (POPG). As observed by others, we find that HNP-2 (net charge = +3) cannot bind to vesicles formed from neutral lipids. The binding of HNP-2 to vesicles containing varying amounts of POPG and neutral (zwitterionic) palmitoyloleoylphosphatidylcholine (POPC) demonstrates that binding is initiated through electrostatic interactions. Because vesicle aggregation and fusion can confound studies of the interaction of HNP-2 with vesicles, those processes were explored systematically by varying the concentrations of vesicles and HNP-2, and the POPG:POPC ratio. Vesicles (300 microM POPG) readily aggregated at HNP-2 concentrations above 1 microM, but no mixing of vesicle contents could be detected for concentrations as high as 2 microM despite the fact that intervesicular lipid mixing could be demonstrated. This indicates that if fusion of vesicles occurs, it is hemi-fusion, in which only the outer monolayers mix at bilayer contact sites. Under conditions of limited aggregation and intervesicular lipid mixing, the fractional leakage of small solutes is a sigmoidal function of peptide concentration. For 300 microM POPG vesicles, 50% of entrapped solute is released by 0.7 microM HNP-2. We introduce a simple method for determining whether leakage from vesicles is graded or all-or-none. We show by means of this fluorescence "requenching" method that native HNP-2 induces vesicle leakage in an all-or-none manner, whereas reduced HNP-2 induces partial, or graded, leakage of vesicle contents. At HNP-2 concentrations that release 100% of small (approximately 400 Da) markers, a fluorescent dextran of 4,400 Da is partially retained in the vesicles, and a 18,900-Da dextran is mostly retained. These results suggest that HNP-2 can form pores that have a maximum diameter of approximately 25 A. A speculative multimeric model of the pore is presented based on these results and on the crystal structure of a human defensin.

Journal ArticleDOI
TL;DR: The determination of the 3‐dimensional structure of the influenza virus neuraminidase in 1983 has served as a platform for understanding interactions between antibodies and protein antigens, for investigating antigenic variation in influenza viruses, and for devising new inhibitors of the enzyme.
Abstract: The determination of the 3-dimensional structure of the influenza virus neuraminidase in 1983 has served as a platform for understanding interactions between antibodies and protein antigens, for investigating antigenic variation in influenza viruses, and for devising new inhibitors of the enzyme. That work is reviewed here, together with more recent developments that have resulted in one of the inhibitors entering clinical trials as an anti-influenza virus drug.

Journal ArticleDOI
TL;DR: The results indicate that the ionic nature of GdnHCl masks electrostatic interactions in these model proteins, a phenomenon that was absent when the uncharged urea was used.
Abstract: The objective of this study was to address the question of whether or not urea and guanidine hydrochloride (GdnHCl) give the same estimates of the stability of a particular protein We previously suspected that the estimates of protein stability from GdnHCl and urea denaturation data might differ depending on the electrostatic interactions stabilizing the proteins Therefore, 4 coiled-coil analogs were designed, where the number of intrachain and interchain electrostatic attractions (A) were systematically changed to repulsions (R): 20A, 15A5R, 10A10R, and 20R The GdnHCl denaturation data showed that the 4 coiled-coil analogs, which had electrostatic interactions ranging from 20 attractions to 20 repulsions, had very similar [GdnHCl]1/2 values (average of congruent to 35 M) and, as well, their delta delta Gu values were very close to 0 (02 kcal/mol) In contrast, urea denaturation showed that the [urea]1/2 values proportionately decreased with the stepwise change from 20 electrostatic attractions to 20 repulsions (20A, 74 M; 15A5R, 54 M; 10A10R, 32 M; and 20R, 14 M), and the delta delta Gu values correspondingly increased with the increasing differences in electrostatic interactions (20A-15A5R, 15 kcal/mol; 20A-10A10R, 37 kcal/mol; and 20A-20R, 58 kcal/mol) These results indicate that the ionic nature of GdnHCl masks electrostatic interactions in these model proteins, a phenomenon that was absent when the unchanged urea was used Thus, GdnHCl and urea denaturations may give vastly different estimates of protein stability, depending on how important electrostatic interactions are to the protein

Journal ArticleDOI
TL;DR: A database of protein structure alignments as well as methods and tools that use this database to improve comparative protein modeling are described, including the derivation of the probability density function for comparative modeling of the cis/trans isomerism of the proline residues.
Abstract: We describe a database of protein structure alignments as well as methods and tools that use this database to improve comparative protein modeling. The current version of the database contains 105 alignments of similar proteins or protein segments. The database comprises 416 entries, 78,495 residues, 1,233 equivalent entry pairs, and 230,396 pairs of equivalent alignment positions. At present, the main application of the database is to improve comparative modeling by satisfaction of spatial restraints implemented in the program MODELLER (Sali A, Blundell TL, 1993, J Mol Biol 234:779-815). To illustrate the usefulness of the database, the restraints on the conformation of a disulfide bridge provided by an equivalent disulfide bridge in a related structure are derived from the alignments; the prediction success of the disulfide dihedral angle classes is increased to approximately 80%, compared to approximately 55% for modeling that relies on the stereochemistry of disulfide bridges alone. The second example of the use of the database is the derivation of the probability density function for comparative modeling of the cis/trans isomerism of the proline residues; the prediction success is increased from 0% to 82.9% for cis-proline and from 93.3% to 96.2% for trans-proline. The database is available via electronic mail.

Journal ArticleDOI
TL;DR: It is proposed that Glu 78 is the nucleophile and that GLU 172 is the acid‐base catalyst in the reaction, on the basis of the work investigated.
Abstract: Using site-directed mutagenesis we have investigated the catalytic residues in a xylanase from Bacillus circulans. Analysis of the mutants E78D and E172D indicated that mutations in these conserved residues do not grossly alter the structure of the enzyme and that these residues participate in the catalytic mechanism. We have now determined the crystal structure of an enzyme-substrate complex to 108 A resolution using a catalytically incompetent mutant (E172C). In addition to the catalytic residues, Glu 78 and Glu 172, we have identified 2 tyrosine residues, Tyr 69 and Tyr 80, which likely function in substrate binding, and an arginine residue, Arg 112, which plays an important role in the active site of this enzyme. On the basis of our work we would propose that Glu 78 is the nucleophile and that Glu 172 is the acid-base catalyst in the reaction.

Journal ArticleDOI
TL;DR: The crystal structure of ternary and binary substrate complexes of the catalytic subunit of CAMP‐dependent protein kinase has been refined and it is concluded that this structural motif is a highly mobile segment of the protein.
Abstract: The crystal structure of ternary and binary substrate complexes of the catalytic subunit of cAMP-dependent protein kinase has been refined at 2.2 and 2.25 A resolution, respectively. The ternary complex contains ADP and a 20-residue substrate peptide, whereas the binary complex contains the phosphorylated substrate peptide. These 2 structures were refined to crystallographic R-factors of 17.5 and 18.1%, respectively. In the ternary complex, the hydroxyl oxygen OG of the serine at the P-site is 2.7 A from the OD1 atom of Asp 166. This is the first crystallographic evidence showing the direct interaction of this invariant carboxylate with a peptide substrate, and supports the predicted role of Asp 166 as a catalytic base and as an agent to position the serine -OH for nucleophilic attack. A comparison of the substrate and inhibitor ternary complexes places the hydroxyl oxygen of the serine 2.7 A from the gamma-phosphate of ATP and supports a direct in-line mechanism for phosphotransfer. In the binary complex, the phosphate on the Ser interacts directly with the epsilon N of Lys 168, another conserved residue. In the ternary complex containing ATP and the inhibitor peptide, Lys 168 interacts electrostatically with the gamma-phosphate of ATP (Zheng J, Knighton DR, Ten Eyck LF, Karlsson R, Xuong NH, Taylor SS, Sowadski JM, 1993, Biochemistry 32:2154-2161). Thus, Lys 168 remains closely associated with the phosphate in both complexes. A comparison of this binary complex structure with the recently solved structure of the ternary complex containing ATP and inhibitor peptide also reveals that the phosphate atom traverses a distance of about 1.5 A following nucleophilic attack by serine and transfer to the peptide. No major conformational changes of active site residues are seen when the substrate and product complexes are compared, although the binary complex with the phosphopeptide reveals localized changes in conformation in the region corresponding to the glycine-rich loop. The high B-factors for this loop support the conclusion that this structural motif is a highly mobile segment of the protein.

Journal ArticleDOI
TL;DR: The buried water molecules and internal cavities in a set of 75 high‐resolution, nonhomologous, monomeric protein structures are analyzed to show the total volume of a protein's large cavities is proportional to its molecular weight and is not dependent on structural class.
Abstract: We have analyzed the buried water molecules and internal cavities in a set of 75 high-resolution, nonhomologous, monomeric protein structures. The number of hydrogen bonds formed between each water molecule and the protein varies from 0 to 4, with 3 being most common. Nearly half of the water molecules are found in pairs or larger clusters. Approximately 90% are shown to be associated with large cavities within the protein, as determined by a novel program, PRO_ACT. The total volume of a protein's large cavities is proportional to its molecular weight and is not dependent on structural class. The largest cavities in proteins are generally elongated rather than globular. There are many more empty cavities than hydrated cavities. The likelihood of a cavity being occupied by a water molecule increases with cavity size and the number of available hydrogen bond partners, with each additional partner typically stabilizing the occupied state by 0.6 kcal/mol.

Journal ArticleDOI
TL;DR: An automatic method to identify protein domains from sequence comparisons is developed and 2 examples of previously unrecognized domain arrangements discovered with the help of, ProDom are provided.
Abstract: The structure of many proteins consists of a combination of discrete modules that have been shuffled during evolution. Such modules can frequently be recognized from the analysis of homology. Here we present a systematic analysis of the modular organization of all sequenced proteins. To achieve this we have developed an automatic method to identify protein domains from sequence comparisons. Homologous domains can then be clustered into consistent families. The method was applied to all 21,098 nonfragment protein sequences in SWISS-PROT 21.0, which was automatically reorganized into a comprehensive protein domain database, ProDom. We have constructed multiple sequence alignments for each domain family in ProDom, from which consensus sequences were generated. These nonreduntant domain consensuses are useful for fast homology searches. Domain organization in ProDom is exemplified for proteins of the phosphoenolpyruvate:sugar phosphotransferase system (PEP:PTS) and for bacterial 2-component regulators. We provide 2 examples of previously unrecognized domain arrangements discovered with the help of ProDom.

Journal ArticleDOI
TL;DR: The practical use, interpretation, and utility of estimation of conformational stability of dimers by equilibrium denaturation methods are discussed, which gives a rationale for the formation of oligomers.
Abstract: The conformational stability of dimeric globular proteins can be measured by equilibrium denaturation studies in solvents such as guanidine hydrochloride or urea. Many dimeric proteins denature with a 2-state equilibrium transition, whereas others have stable intermediates in the process. For those proteins showing a single transition of native dimer to denatured monomer, the conformational stabilities, delta Gu (H2O), range from 10 to 27 kcal/mol, which is significantly greater than the conformational stability found for monomeric proteins. The relative contribution of quaternary interactions to the overall stability of the dimer can be estimated by comparing delta Gu (H2O) from equilibrium denaturation studies to the free energy associated with simple dissociation in the absence of denaturant. In many cases the large stabilization energy of dimers is primarily due to the intersubunit interactions and thus gives a rationale for the formation of oligomers. The magnitude of the conformational stability is related to the size of the polypeptide in the subunit and depends upon the type of structure in the subunit interface. The practical use, interpretation, and utility of estimation of conformational stability of dimers by equilibrium denaturation methods are discussed.

Journal ArticleDOI
TL;DR: Progressive sequence alignments suggest that the 13 isoprenyl diphosphate synthases evolved from a common ancestor into 3 distinct clusters.
Abstract: Isoprenyl diphosphate synthases are ubiquitous enzymes that catalyze the basic chain‐elongation reaction in the isoprene biosynthetic pathway. Pairwise sequence comparisons were made for 6 farnesyl diphosphate synthases, 6 geranylgeranyl diphosphate synthases, and a hexaprenyl diphosphate synthase. Five regions with highly conserved residues, two of which contain aspartate‐rich DDXX(XX)D motifs found in many prenyltransferases, were identified. A consensus secondary structure for the group, consisting mostly of α‐helices, was predicted for the multiply aligned sequences from amino acid compositions, computer assignments of local structure, and hydropathy indices. Progressive sequence alignments suggest that the 13 isoprenyl diphosphate synthases evolved from a common ancestor into 3 distinct clusters. The most distant separation is between yeast hexaprenyl diphosphate synthetase and the other enzymes. Except for the chromoplastic geranylgeranyl diphosphate synthase from Capsicum annuum, the remaining farnesyl and geranylgeranyl diphosphate synthases segregate into prokaryotic/archaebacterial and eukaryotic families. Copyright

Journal ArticleDOI
TL;DR: The 3-dimensional structures of malate and lactate dehydrogenases are similar, despite their low amino acid sequence identity as discussed by the authors, and the mechanism of catalysis is similar to that of an enzyme with which it shares a similar 3D structure.
Abstract: Malate dehydrogenases are widely distributed and alignment of the amino acid sequences show that the enzyme has diverged into 2 main phylogenetic groups. Multiple amino acid sequence alignments of malate dehydrogenases also show that there is a low degree of primary structural similarity, apart from in several positions crucial for nucleotide binding, catalysis, and the subunit interface. The 3-dimensional structures of several malate dehydrogenases are similar, despite their low amino acid sequence identity. The coenzyme specificity of malate dehydrogenase may be modulated by substitution of a single residue, as can the substrate specificity. The mechanism of catalysis of malate dehydrogenase is similar to that of lactate dehydrogenase, an enzyme with which it shares a similar 3-dimensional structure. Substitution of a single amino acid residue of a lactate dehydrogenase changes the enzyme specificity to that of a malate dehydrogenase, but a similar substitution in a malate dehydrogenase resulted in relaxation of the high degree of specificity for oxaloacetate. Knowledge of the 3-dimensional structures of malate and lactate dehydrogenases allows the redesign of enzymes by rational rather than random mutation and may have important commercial implications.

Journal ArticleDOI
TL;DR: The results strongly suggest that large local motions proximate to the scissile bond are required for proteolysis, and it is this ability to unfold locally without perturbing the overall protein conformation that is the prime determinant for limited proteolyses.
Abstract: Previous analyses of limited proteolytic sites within native, folded protein structures have shown that a significant conformational change is required in order to facilitate binding into the active site of the attacking proteinase. For the serine proteinases, the optimum conformation to match the proteinase binding-site geometry has been well characterized crystallographically by the conserved main-chain geometry of the reactive site loops of their protein inhibitors. A good substrate must adopt a conformation very similar to this "target" main-chain conformation prior to cleavage. Using a "loop-closure" modeling approach, we have tested the ability of a set of tryptic-limited proteolytic sites to achieve this target conformation and further tested their suitability for cleavage. The results show that in most cases, significant changes in the conformation of at least 12 residues are required. All the putative tryptic cleavage sites in 1 protein, elastase, were also modeled and tested to compare the results to the actual nicksite in that protein. These results strongly suggest that large local motions proximate to the scissile bond are required for proteolysis, and it is this ability to unfold locally without perturbing the overall protein conformation that is the prime determinant for limited proteolysis.

Journal ArticleDOI
TL;DR: The collectins are a group of mammalian lectins containing collagen‐like regions that share a very similar modular domain composition and overall 3‐dimensional structure and appear to play similar biological roles in the preimmune defense against microorganisms in both serum and lung surfactant.
Abstract: The collectins are a group of mammalian lectins containing collagen-like regions. They include mannan binding protein, bovine conglutinin, lung surfactant protein A, lung surfactant protein D, and a newly discovered bovine protein named collectin-43. These proteins share a very similar modular domain composition and overall 3-dimensional structure. They also appear to play similar biological roles in the preimmune defense against micro-organisms in both serum and lung surfactant. The close evolutionary relationship between the collectins is further emphasized by a common pattern of exons in their genomic structures and the presence of a gene cluster on chromosome 10 in humans that contains the genes known for the human collectins. Studies on the structure/function relationships within the collectins could provide insight into the properties of a growing number of proteins also containing collagenous regions such as C1q, the hibernation protein, the alpha- and beta-ficolins, as well as the membrane acetylcholinesterase and the macrophage scavenger receptor.

Journal ArticleDOI
TL;DR: The crystal structure of the P1/Mahoney poliovirus empty capsid suggests plausible (and ultimately testable) models for the initiation of encapsidation, for the RNA‐dependent autocatalytic cleavage of VP0, and for the role of the cleavage in establishing the ordered N‐terminal network and in generating stable virions.
Abstract: The crystal structure of the P1/Mahoney poliovirus empty capsid has been determined at 2.9 A resolution. The empty capsids differ from mature virions in that they lack the viral RNA and have yet to undergo a stabilizing maturation cleavage of VP0 to yield the mature capsid proteins VP4 and VP2. The outer surface and the bulk of the protein shell are very similar to those of the mature virion. The major differences between the 2 structures are focused in a network formed by the N-terminal extensions of the capsid proteins on the inner surface of the shell. In the empty capsids, the entire N-terminal extension of VP1, as well as portions corresponding to VP4 and the N-terminal extension of VP2, are disordered, and many stabilizing interactions that are present in the mature virion are missing. In the empty capsid, the VP0 scissile bond is located some 20 A away from the positions in the mature virion of the termini generated by VP0 cleavage. The scissile bond is located on the rim of a trefoil-shaped depression in the inner surface of the shell that is highly reminiscent of an RNA binding site in bean pod mottle virus. The structure suggests plausible (and ultimately testable) models for the initiation of encapsidation, for the RNA-dependent autocatalytic cleavage of VP0, and for the role of the cleavage in establishing the ordered N-terminal network and in generating stable virions.

Journal ArticleDOI
TL;DR: In addition to helping form the active site funnel in superoxide dismutase, the structural arginines found in this study play such diverse roles as stapling together 3 strands of backbone from different regions of the primary sequence, and tying α‐helix to α‐ helix, βturn to β‐turn, and subunit to subunit.
Abstract: We propose that arginine side chains often play a previously unappreciated general structural role in the maintenance of tertiary structure in proteins, wherein the positively charged guanidinium group forms multiple hydrogen bonds to backbone carbonyl oxygens. Using as a criterion for a "structural" arginine one that forms 4 or more hydrogen bonds to 3 or more backbone carbonyl oxygens, we have used molecular graphics to locate arginines of interest in 4 proteins: Arg 180 in Thermus thermophilus manganese superoxide dismutase, Arg 254 in human carbonic anhydrase II, Arg 31 in Streptomyces rubiginosus xylose isomerase, and Arg 313 in Rhodospirillum rubrum ribulose-1,5-bisphosphate carboxylase/oxygenase. Arg 180 helps to mold the active site channel of superoxide dismutase, whereas in each of the other enzymes the structural arginine is buried in the "mantle" (i.e., inside, but near the surface) of the protein interior well removed from the active site, where it makes 5 hydrogen bonds to 4 backbone carbonyl oxygens. Using a more relaxed criterion of 3 or more hydrogen bonds to 2 or more backbone carbonyl oxygens, arginines that play a potentially important structural role were found in yeast enolase, Bacillus stearothermophilus glyceraldehyde-3-phosphate dehydrogenase, bacteriophage T4 and human lysozymes, Enteromorpha prolifera plastocyanin, HIV-1 protease, Trypanosoma brucei brucei and yeast triosephosphate isomerases, and Escherichia coli trp aporepressor (but not trp repressor or the trp repressor/operator complex).(ABSTRACT TRUNCATED AT 250 WORDS)

Journal ArticleDOI
TL;DR: Models of the N domains of HGF/SF, HGF1/MSP, and plasminogen, characterized by the presence of 4 conserved Cys residues forming a loop in a loop, have been modeled based on disulfide‐bond constraints and suggest a mechanism for the formation of a noncovalent H GF/SF homodimer that may be responsible for the activation of the Met receptor.
Abstract: Plasminogen-related growth factors, a new family of polypeptide growth factors with the basic domain organization and mechanism of activation of the blood proteinase plasminogen, include hepatocyte growth factor/scatter factor (HGF/SF), a potent effector of the growth, movement, and differentiation of epithelia and endothelia, and hepatocyte growth factor-like/macrophage stimulating protein (HGF1/MSP), an effector of macrophage chemotaxis and phagocytosis. Phylogeny of the serine proteinase domains and analysis of intron-exon boundaries and kringle sequences indicate that HGF/SF, HGF1/MSP, plasminogen, and apolipoprotein (a) have evolved from a common ancestral gene that consisted of an N-terminal domain corresponding to plasminogen activation peptide (PAP), 3 copies of the kringle domain, and a serine proteinase domain. Models of the N domains of HGF/SF, HGF1/MSP, and plasminogen, characterized by the presence of 4 conserved Cys residues forming a loop in a loop, have been modeled based on disulfide-bond constraints. There is a distinct pattern of charged and hydrophobic residues in the helix-strand-helix motif proposed for the PAP domain of HGF/SF; these may be important for receptor interaction. Three-dimensional structures of the 4 kringle and the serine proteinase domains of HGF/SF were constructed by comparative modeling using the suite of programs COMPOSER and were energy minimized. Docking of a lysine analogue indicates a putative lysine-binding pocket within kringle 2 (and possibly another in kringle 4). The models suggest a mechanism for the formation of a noncovalent HGF/SF homodimer that may be responsible for the activation of the Met receptor. These data provide evidence for the divergent evolution and structural similarity of plasminogen, HGF/SF, and HGF1/MSP, and highlight a new strategy for growth factor evolution, namely the adaptation of a proteolytic enzyme to a role in receptor activation.

Journal ArticleDOI
TL;DR: Inteins (protein introns) are internal portions of protein sequences that are posttranslationally excised while the flanking regions are spliced together, making an additional protein product.
Abstract: Inteins (protein introns) are internal portions of protein sequences that are posttranslationally excised while the flanking regions are spliced together, making an additional protein product. Inteins have been found in a number of homologous genes in yeast, mycobacteria, and extreme thermophile archaebacteria. The inteins are probably multifunctional, autocatalyzing their own splicing, and some were also shown to be DNA endonucleases. The splice junction regions and two regions similar to homing endonucleases were thought to be the only common sequence features of inteins. This work analyzed all published intein sequences with recently developed methods for detecting weak, conserved sequence features. The methods complemented each other in the identification and assessment of several patterns characterizing the intein sequences. New intein conserved features are discovered and the known ones are quantitatively described and localized. The general sequence description of all the known inteins is derived from the motifs and their relative positions. The intein sequence description is used to search the sequence databases for intein-like proteins. A sequence region in a mycobacterial open reading frame possessing all of the intein motifs and absent from sequences homologous to both of its flanking sequences is identified as an intein. A newly discovered putative intein in red algae chloroplasts is found not to contain the endonuclease motifs present in all other inteins. The yeast HO endonuclease is found to have an overall intein-like structure and a few viral polyprotein cleavage sites are found to be significantly similar to the inteins amino-end splice junction motif. The intein features described may serve for detection of intein sequences.

Journal ArticleDOI
TL;DR: The structure of toxic monomeric diphtheria toxin (DT) was determined at 2.3 Å resolution by molecular replacement based on the domain structures in dimeric DT and refined to an R factor of 20.7%.
Abstract: The structure of toxic monomeric diphtheria toxin (DT) was determined at 2.3 A resolution by molecular replacement based on the domain structures in dimeric DT and refined to an R factor of 20.7%. The model consists of 2 monomers in the asymmetric unit (1,046 amino acid residues), including 2 bound adenylyl 3'-5' uridine 3' monophosphate molecules and 396 water molecules. The structures of the 3 domains are virtually identical in monomeric and dimeric DT; however, monomeric DT is compact and globular as compared to the "open" monomer within dimeric DT (Bennett MJ, Choe S, Eisenberg D, 1994b, Protein Sci 3:0000-0000). Detailed differences between monomeric and dimeric DT are described, particularly (1) changes in main-chain conformations of 8 residues acting as a hinge to "open" or "close" the receptor-binding (R) domain, and (2) a possible receptor-docking site, a beta-hairpin loop protruding from the R domain containing residues that bind the cell-surface DT receptor. Based on the monomeric and dimeric DT crystal structures we have determined and the solution studies of others, we present a 5-step structure-based mechanism of intoxication: (1) proteolysis of a disulfide-linked surface loop (residues 186-201) between the catalytic (C) and transmembrane (T) domains; (2) binding of a beta-hairpin loop protruding from the R domain to the DT receptor, leading to receptor-mediated endocytosis; (3) low pH-triggered open monomer formation and exposure of apolar surfaces in the T domain, which insert into the endosomal membrane; (4) translocation of the C domain into the cytosol; and (5) catalysis by the C domain of ADP-ribosylation of elongation factor 2.

Journal ArticleDOI
TL;DR: Observations suggest that surfactants, used in conjunction with ESI‐MS, can be useful for protein structure studies, if care is used in the interpretation of the results.
Abstract: Electrospray ionization mass spectrometry (ESI-MS) has proven to be a useful tool for examining noncovalent complexes between proteins and a variety of ligands. It has also been used to distinguish between denatured and refolded forms of proteins. Surfactants are frequently employed to enhance solubilization or to modify the tertiary or quaternary structure of proteins, but are usually considered incompatible with mass spectrometry. A broad range of ionic, nonionic, and zwitterionic surfactants was examined to characterize their effects on ESI-MS and on protein structure under ESI-MS conditions. Solution conditions studied include 4% acetic acid/50% acetonitrile/46% H2O and 100% aqueous. Of the surfactants examined, the nonionic saccharides, such as n-dodecyl-beta-D-glucopyranoside, at 0.1% to 0.01% (w/v) concentrations, performed best, with limited interference from chemical background and adduct formation. Under the experimental conditions used, ESI-MS performance in the presence of surfactants was found to be unrelated to critical micelle concentration. It is demonstrated that surfactants can affect both the tertiary and quaternary structures of proteins under conditions used for ESI-MS. However, several of the surfactants caused significant shifts in the charge-state distributions, which appeared to be independent of conformational effects. These observations suggest that surfactants, used in conjunction with ESI-MS, can be useful for protein structure studies, if care is used in the interpretation of the results.

Journal ArticleDOI
TL;DR: Simulation of the active site of acetylcholinesterase from Torpedo californica suggests that a bis‐quaternary decamethonium (DECA) ion, acquired during enzyme purification, resides in the gorge and appears to stabilize part of the gorge wall through electrostatic interactions.
Abstract: The active site of acetylcholinesterase (AChE) from Torpedo californica is located 20 A from the enzyme surface at the bottom of a narrow gorge. To understand the role of this gorge in the function of AChE, we have studied simulations of its molecular dynamics. When simulations were conducted with pure water filling the gorge, residues in the vicinity of the active site deviated quickly and markedly from the crystal structure. Further study of the original crystallographic data suggests that a bis-quaternary decamethonium (DECA) ion, acquired during enzyme purification, residues in the gorge. There is additional electron density within the gorge that may represent small bound cations. When DECA and 2 cations are placed within the gorge, the simulation and the crystal structure are dramatically reconciled. The small cations, more so than DECA, appear to stabilize part of the gorge wall through electrostatic interactions. This part of the gorge wall is relatively thin and may regulate substrate, product, and water movement through the active site.