Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 53
Filter
1.
Br J Ophthalmol ; 90(1): 81-6, 2006 Jan.
Article in English | MEDLINE | ID: mdl-16361673

ABSTRACT

BACKGROUND/AIM: X linked retinoschisis (XLRS) is caused by mutations in RS1 which encodes the discoidin domain protein retinoschisin, secreted by photoreceptors and bipolar cells. Missense mutations occur throughout the gene and some of these are known to interfere with protein secretion. This study was designed to investigate the functional consequences of missense mutations at different locations in retinoschisin. METHODS AND RESULTS: The authors developed a structural model of the retinoschisin discoidin domain and used this to predict the effects of missense mutations. They expressed disease associated mutations and found that those affecting conserved residues prevented retinoschisin secretion. Most of the remaining mutations cluster within a series of loops on the surface of the beta barrel structure and do not interfere with secretion, suggesting this region may be a ligand binding site. They also demonstrated that wild type retinoschisin octamerises and associates with the cell surface. A subgroup of secreted mutations reduce oligomerisation (C59S, C219G, C223R). CONCLUSIONS: It is suggested that there are three different molecular mechanisms which lead to XLRS: mutations interfering with secretion, mutations interfering with oligomerisation, and mutations that allow secretion and oligomerisation but interfere with retinoschisin function. The authors conclude that binding of oligomerised retinoschisin at the cell surface is important in its presumed role in cell adhesion.


Subject(s)
Eye Proteins/genetics , Mutation, Missense , Retinoschisis/genetics , Amino Acid Sequence , Animals , Blotting, Western , COS Cells , Chlorocebus aethiops , Dimerization , Eye Proteins/metabolism , Factor Va/genetics , Humans , Male , Models, Molecular , Molecular Sequence Data , Mutagenesis, Site-Directed , Photoreceptor Cells, Vertebrate/metabolism , Retinoschisis/metabolism , Sequence Alignment
2.
Structure ; 4(1): 11-4, 1996 Jan 15.
Article in English | MEDLINE | ID: mdl-8805517

ABSTRACT

With the recent demonstration that multiwavelength anomalous dispersion (MAD) can provide accurate experimental phases at high resolution, crystallographers have gained a tool with which to study solvation and flexibility in proteins, and a test-bed for the development of crystallographic methods.


Subject(s)
Proteins/chemistry , X-Ray Diffraction/methods , Carrier Proteins/chemistry , Crystallography , Mannose-Binding Lectins , Protein Conformation , Solvents/chemistry , Ytterbium/metabolism
3.
Structure ; 2(1): 45-57, 1994 Jan 15.
Article in English | MEDLINE | ID: mdl-8075982

ABSTRACT

BACKGROUND: Pertussis toxin is an exotoxin of the A-B class produced by Bordetella pertussis. The holotoxin comprises 952 residues forming six subunits (five different sequences, S1-S5). It plays an important role in the development of protective immunity to whooping cough, and is an essential component of new acellular vaccines. It is also widely used as a biochemical tool to ADP-ribosylate GTP-binding proteins in the study of signal transduction. RESULTS: The crystal structure of pertussis toxin has been determined at 2.9 A resolution. The catalytic A-subunit (S1) shares structural homology with other ADP-ribosylating bacterial toxins, although differences in the carboxy-terminal portion explain its unique activation mechanism. Despite its heterogeneous subunit composition, the structure of the cell-binding B-oligomer (S2, S3, two copies of S4, and S5) resembles the symmetrical B-pentamers of the cholera toxin and Shiga toxin families, but it interacts differently with the A-subunit. The structural similarity is all the more surprising given that there is almost no sequence homology between B-subunits of the different toxins. Two peripheral domains that are unique to the pertussis toxin B-oligomer show unexpected structural homology with a calcium-dependent eukaryotic lectin, and reveal possible receptor-binding sites. CONCLUSION: The structure provides insight into the pathogenic mechanisms of pertussis toxin and the evolution of bacterial toxins. Knowledge of the tertiary structure of the active site forms a rational basis for elimination of catalytic activity in recombinant molecules for vaccine use.


Subject(s)
Pertussis Toxin , Protein Structure, Secondary , Virulence Factors, Bordetella/chemistry , Amino Acid Sequence , Bordetella pertussis , Computer Graphics , Crystallography, X-Ray/methods , Macromolecular Substances , Models, Molecular , Molecular Sequence Data , Sequence Homology, Amino Acid , Synchrotrons , Virulence Factors, Bordetella/isolation & purification , X-Ray Diffraction/methods
4.
Structure ; 8(3): 253-64, 2000 Mar 15.
Article in English | MEDLINE | ID: mdl-10745005

ABSTRACT

BACKGROUND: Shiga-like toxins (SLTs) are produced by the pathogenic strains of Escherichia coli that cause hemorrhagic colitis and hemolytic uremic syndrome. These diseases in humans are generally associated with group II family members (SLT-II and SLT-IIc), whereas SLT-IIe (pig edema toxin) is central to edema disease of swine. The pentameric B-subunit component of the majority of family members binds to the cell-surface glycolipid globotriaosyl ceramide (Gb(3)), but globotetraosyl ceramide (Gb(4)) is the preferred receptor for SLT-IIe. A double-mutant of the SLT-IIe B subunit that reverses two sequence differences from SLT-II (GT3; Gln65-->Glu, Lys67-->Gln, SLT-I numbering) has been shown to bind more strongly to Gb(3) than to Gb(4). RESULTS: To understand the molecular basis of receptor binding and specificity, we have determined the structure of the GT3 mutant B pentamer, both in complex with a Gb(3) analogue (2.0 A resolution; R = 0.155, R(free) = 0.194) and in its native form (2.35 A resolution; R = 0.187, R(free) = 0.232). CONCLUSIONS: These are the first structures of a member of the medically important group II Shiga-like toxins to be reported. The structures confirm the previous observation of multiple binding sites on each SLT monomer, although binding site 3 is not occupied in the GT3 structure. Analysis of the binding properties of mutants suggests that site 3 is a secondary Gb(4)-binding site. The two mutated residues are located appropriately to interact with the extra betaGalNAc residue on Gb(4). Differences in the binding sites provide a molecular basis for understanding the tissue specificities and pathogenic mechanisms of members of the SLT family.


Subject(s)
Bacterial Toxins/metabolism , Glycolipids/metabolism , Receptors, Cell Surface/metabolism , Amino Acid Sequence , Bacterial Toxins/chemistry , Bacterial Toxins/genetics , Base Sequence , Binding Sites , Carbohydrate Conformation , Carbohydrate Sequence , DNA Primers , Models, Molecular , Molecular Sequence Data , Mutation , Protein Conformation , Sequence Homology, Amino Acid , Shiga Toxin 2 , Trisaccharides/metabolism
5.
Structure ; 7(2): 111-8, 1999 Feb 15.
Article in English | MEDLINE | ID: mdl-10368279

ABSTRACT

BACKGROUND: Plasminogen activator inhibitor 1 (PAI-1) is a serpin that has a key role in the control of fibrinolysis through proteinase inhibition. PAI-1 also has a role in regulating cell adhesion processes relevant to tissue remodeling and metastasis; this role is mediated by its binding to the adhesive glycoprotein vitronectin rather than by proteinase inhibition. Active PAI-1 is metastable and spontaneously transforms to an inactive latent conformation. Previous attempts to crystallize the active conformation of PAI-1 have failed. RESULTS: The crystal structure of a stable quadruple mutant of PAI-1(Asn150-->His, Lys154-->Thr, Gln319-->Leu, Met354-->Ile) in its active conformation has been solved at a nominal 3 A resolution. In two of four independent molecules within the crystal, the flexible reactive center loop is unconstrained by crystal-packing contacts and is disordered. In the other two molecules, the reactive center loop forms intimate loop-sheet interactions with neighboring molecules, generating an infinite chain within the crystal. The overall conformation resembles that seen for other active inhibitory serpins. CONCLUSIONS: The structure clarifies the molecular basis of the stabilizing mutations and the reduced affinity of PAI-1, on cleavage or in the latent form, for vitronectin. The infinite chain of linked molecules also suggests a new mechanism for the serpin polymerization associated with certain diseases. The results support the concept that the reactive center loop of an active serpin is flexible and has no defined conformation in the absence of intermolecular contacts. The determination of the structure of the active form constitutes an essential step for the rational design of PAI-1 inhibitors.


Subject(s)
Cell Adhesion/drug effects , Fibrinolysis/drug effects , Plasminogen Activator Inhibitor 1/chemistry , Protein Conformation , Binding Sites , Crystallography, X-Ray , Models, Molecular , Mutation/genetics , Plasminogen Activator Inhibitor 1/genetics , Protein Binding , Protein Structure, Secondary , Protein Structure, Tertiary , Serine Proteinase Inhibitors/chemistry , Serpins/chemistry , Vitronectin/metabolism
6.
Biochim Biophys Acta ; 752(1): 118-26, 1983 Jun 16.
Article in English | MEDLINE | ID: mdl-6849959

ABSTRACT

The present report describes the purification and characterization of a non-specific phospholipid-transfer protein from rat lung. The protein is the major phospholipid-transfer protein in lung which transfers phosphatidylcholine. The transfer protein was purified 1200-fold, with a final yield of 3%. The activity of the protein was monitored by measuring the transfer of [14C]phosphatidylcholine from radioactively labeled liposomes to mitochondria. The purified proteins transfers phosphatidylcholine, phosphatidylinositol, phosphatidylserine and phosphatidylethanolamine from radioactively labeled microsomes to either mitochondria or liposomes. The transfer of each phospholipid is proportional to its content in the donor membrane. The protein was purified from a pH 5.1 supernatant preparation by fractionation on DEAE-cellulose, Sephadex G-75 and hydroxyapatite. The molecular weight of the purified protein was estimated as 35 000 by SDS-polyacrylamide gel electrophoresis. The amino acid analysis revealed a high content of glutamic acid (including glutamine) and glycine. The specificity of the purified protein for transfer of phospholipids suggests that it may be the phospholipid-transfer activity which is highly enriched in isolated type II alveolar cells of rat lung.


Subject(s)
Carrier Proteins/isolation & purification , Lung/metabolism , Membrane Proteins , Phospholipid Transfer Proteins , Animals , Chemical Phenomena , Chemistry , In Vitro Techniques , Liposomes/metabolism , Microsomes, Liver/metabolism , Mitochondria, Liver/metabolism , Molecular Weight , Rats
7.
Biochim Biophys Acta ; 794(1): 9-17, 1984 Jun 06.
Article in English | MEDLINE | ID: mdl-6733132

ABSTRACT

A purified phospholipid-transfer protein from rat lung has been characterized in terms of the specificity of the protein for phosphatidylcholine molecules with different apolar moieties. The study demonstrated that the lung-phospholipid-transfer protein discriminates between dipalmitoylphosphatidylcholine and molecular species of phosphatidylcholine with unsaturated acyl chains. The initial rate of transfer of dipalmitoylphosphatidylcholine is 1.5-fold greater than the rate of transfer of dioleoylphosphatidylcholine, 1-palmitoyl-2- arachidonylphosphatidylcholine , or egg phosphatidylcholine under most assay conditions. Although the protein preferentially transfers dipalmitoylphosphatidylcholine, the incorporation of increasing mole percentages of dipalmitoylphosphatidylcholine into unilamellar phosphatidylcholine vesicles profoundly affects their effectiveness as donors for phosphatidylcholine transfer by the transfer protein. At 60 mol% dipalmitoylphosphatidylcholine, the rate of transfer is one-third that observed when vesicles are composed of 100% egg phosphatidylcholine. Decreases in membrane fluidity as estimated by fluorescence polarization of 1,6-diphenyl-1,3,5-hexatriene correlate with decreases in the effectiveness of the vesicles as donors in the phospholipid-transfer reaction. The conclusion from these studies is that the rate of transfer of phosphatidylcholine by the purified phospholipid-transfer protein from lung is determined by physical properties of membrane interfaces with which the protein interacts, as well as by the specificity of the phospholipid-transfer protein for different molecular species of phosphatidylcholine.


Subject(s)
Carrier Proteins/metabolism , Lung/metabolism , Membrane Fluidity , Membrane Proteins , Phospholipid Transfer Proteins , Phospholipids/metabolism , Animals , Intracellular Membranes/metabolism , Kinetics , Liposomes , Microscopy, Fluorescence , Mitochondria, Liver/metabolism , Phosphatidylcholines , Rats , Substrate Specificity , Thermodynamics
8.
J Mol Biol ; 200(3): 523-51, 1988 Apr 05.
Article in English | MEDLINE | ID: mdl-3135412

ABSTRACT

Streptomyces griseus trypsin (SGT) is a bacterial serine proteinase that is more homologous to mammalian than to other bacterial enzymes. The structure of SGT has been solved primarily by molecular replacement, though some low-resolution phase information was supplied by heavy-atom derivatives. The mammalian pancreatic serine proteinases bovine trypsin (BT) and alpha-chymotrypsin (CHT) were used as molecular replacement models. Because these proteins have low homology with SGT compared to the majority of other successful replacement models, new strategies were required for molecular replacement to succeed. The model of SGT has been refined at 1.7 A resolution to a final R-factor of 0.161 (1 A = 0.1 nm); the correlation coefficient between all observed and calculated structure factor amplitudes is 0.908. Solvent molecules located in the crystal structure play an important role in stabilizing buried charged and polar groups. An additional contribution to stability can be seen in the fact that the majority of the charged side-chains are involved in ionic interactions, sometimes linking the two domains of SGT. A comparison of SGT with BT shows that the greatest similarities are in the active-site and substrate-binding regions, consistent with their similar substrate specificities. The modeling of complexes of SGT with two inhibitors of BT, pancreatic trypsin inhibitor (PTI) and the third domain of Japanese quail ovomucoid (OMJPQ3), helps to explain why PTI inhibits SGT but OMJPQ3 does not. Like BT, but unlike other bacterial serine proteinases of known structure, SGT has a buried N terminus. SGT has also a well-defined Ca2+-binding site, but this site differs in location from that of BT.


Subject(s)
Streptomyces griseus/enzymology , Trypsin , Amino Acid Sequence , Binding Sites , Calcium/metabolism , Hydrogen Bonding , Molecular Sequence Data , Protein Conformation , X-Ray Diffraction
9.
J Mol Biol ; 221(3): 729-31, 1991 Oct 05.
Article in English | MEDLINE | ID: mdl-1942026

ABSTRACT

The B-subunit of verotoxin-1, which is believed to form a pentamer (monomer Mr = 7691), has been crystallized by vapor diffusion over a wide range of conditions. The best crystals, obtained with polyethylene glycol 8000 as the precipitant, belong to the orthorhombic space group P2(1)2(1)2(1), with cell dimensions a = 59.2 A, b = 102.7 A, c = 56.3 A. The cell dimensions are consistent with one B-subunit pentamer per asymmetric unit, and the crystals diffract to at least 2.0 A resolution. Data collected using synchrotron radiation at a wavelength of 2.070 A may allow the structure to be solved using the anomalous signal from three sulfur atoms in the monomer, combined with averaging over the non-crystallographic symmetry.


Subject(s)
Bacterial Toxins/chemistry , Hydrogen-Ion Concentration , Shiga Toxin 1 , X-Ray Diffraction
10.
J Mol Biol ; 219(4): 671-92, 1991 Jun 20.
Article in English | MEDLINE | ID: mdl-2056534

ABSTRACT

The molecular structure of porcine pepsinogen at 1.8 A resolution has been determined by a combination of molecular replacement and multiple isomorphous phasing techniques. The resulting structure was refined by restrained-parameter least-squares methods. The final R factor [formula: see text] is 0.164 for 32,264 reflections with I greater than or equal to sigma (I) in the resolution range of 8.0 to 1.8 A. The model consists of 2785 protein atoms in 370 residues, a phosphoryl group on Ser68 and 238 ordered water molecules. The resulting molecular stereochemistry is consistent with a well-refined crystal structure with co-ordinate accuracy in the range of 0.10 to 0.15 A for the well-ordered regions of the molecule (B less than 15 A2). For the enzyme portion of the zymogen, the root-mean-square difference in C alpha atom co-ordinates with the refined porcine pepsin structure is 0.90 A (284 common atoms) and with the C alpha atoms of penicillopepsin it is 1.63 A (275 common atoms). The additional 44 N-terminal amino acids of the prosegment (Leu1p to Leu44p, using the letter p after the residue number to distinguish the residues of the prosegment) adopt a relatively compact structure consisting of a long beta-strand followed by two approximately orthogonal alpha-helices and a short 3(10)-helix. Intimate contacts, both electrostatic and hydrophobic interactions, are made with residues in the pepsin active site. The N-terminal beta-strand, Leu1p to Leu6p, forms part of the six-stranded beta-sheet common to the aspartic proteinases. In the zymogen the first 13 residues of pepsin, Ile1 to Glu13, adopt a completely different conformation from that of the mature enzyme. The C alpha atom of Ile1 must move approximately 44 A in going from its position in the inactive zymogen to its observed position in active pepsin. Electrostatic interactions of Lys36pN and hydrogen-bonding interactions of Tyr37pOH, and Tyr90H with the two catalytic aspartate groups, Asp32 and Asp215, prevent substrate access to the active site of the zymogen. We have made a detailed comparison of the mammalian pepsinogen fold with the fungal aspartic proteinase fold of penicillopepsin, used for the molecular replacement solution. A structurally derived alignment of the two sequences is presented.


Subject(s)
Pepsinogens/chemistry , Amino Acid Sequence , Animals , Aspartic Acid Endopeptidases/chemistry , Enzyme Precursors/chemistry , Hydrogen Bonding , Molecular Sequence Data , Molecular Structure , Pepsin A/chemistry , Protein Conformation , Sequence Alignment , Swine , X-Ray Diffraction
11.
J Mol Biol ; 258(4): 661-71, 1996 May 17.
Article in English | MEDLINE | ID: mdl-8637000

ABSTRACT

Pertussis toxin is a major virulence factor of Bordetella pertussis, the causative agent of whooping cough. The protein is a hexamer containing a catalytic subunit (S1) that is tightly associated with a pentameric cell-binding component (B-oligomer). In vitro experiments have shown that ATP and a number of detergents and phospholipids assist in activating the holotoxin by destabilizing the interaction between S1 and the B-oligomer. Similar processes may play a role in the activation of pertussis toxin in vivo. In this paper we present the crystal structure of the pertussis toxin-ATP complex and discuss the structural basis for the ATP-induced activation. In addition, we propose a physiological role for the ATP effect in the process by which the toxin enters the cytoplasm of eukaryotic cells. The key features of this proposal are that ATP binding signals the arrival of the toxin in the endoplasmic reticulum and, at the same time, triggers dissociation of the holotoxin prior to membrane translocation.


Subject(s)
Adenosine Triphosphate/chemistry , Pertussis Toxin , Virulence Factors, Bordetella/chemistry , Adenosine Triphosphate/pharmacology , Binding Sites , Biological Transport , Crystallography , Endoplasmic Reticulum/metabolism , Models, Molecular , Protein Conformation , Structure-Activity Relationship , Synchrotrons , Virulence Factors, Bordetella/metabolism , Virulence Factors, Bordetella/pharmacology
12.
J Mol Biol ; 241(2): 269-72, 1994 Aug 12.
Article in English | MEDLINE | ID: mdl-8057365

ABSTRACT

Wild-type and mutant forms of murine interleukin-5 (mIL-5) have been expressed in the baculovirus expression system, purified, and used in crystallization trials. Attempts to obtain diffraction quality crystals of wild-type protein were unsuccessful. The substitution of glutamine for Asn75 preserved biological activity, while removing one of two predicted N-linked glycosylation sites, and the resulting protein was crystallized from polyethylene glycol 8000 at pH 7.8 in two crystal forms. The orthorhombic crystals, which belong to space group P2(1)2(1)2 with cell dimensions a = 55.9 A, b = 83.0 A and c = 52.3 A, diffract to beyond 2.5 A resolution. The second crystal form belongs to a trigonal space group, either P3(1)21 or P3(2)21, with cell dimensions a = b = 62.1 A, c = 129.9 A, and diffracts to about 3.8 A resolution. Each crystal form probably contains one mIL-5 dimer per asymmetric unit.


Subject(s)
Interleukin-5/chemistry , Animals , Asparagine/chemistry , Baculoviridae , Crystallization , Crystallography, X-Ray , Glutamine/chemistry , Interleukin-5/genetics , Interleukin-5/isolation & purification , Isoelectric Point , Mice , Mutagenesis, Site-Directed , X-Ray Diffraction
13.
J Mol Biol ; 194(3): 573-5, 1987 Apr 05.
Article in English | MEDLINE | ID: mdl-3625777

ABSTRACT

Crystals of glyceraldehyde phosphate dehydrogenase from the glycosome of Trypanosoma brucei brucei have been grown, and a partial data set has been collected using synchrotron radiation. The crystals diffract initially to 2.3 A resolution. The space group is P2(1)2(1)2, with cell dimensions a = 135 A, b = 255 A, c = 115 A, so there are probably at least two tetramers in the asymmetric unit.


Subject(s)
Glyceraldehyde-3-Phosphate Dehydrogenases , Animals , Trypanosoma brucei brucei/enzymology , X-Ray Diffraction
14.
J Mol Biol ; 195(2): 397-418, 1987 May 20.
Article in English | MEDLINE | ID: mdl-3477645

ABSTRACT

The molecular structure of the complex between bovine pancreatic alpha-chymotrypsin (EC 3.4.4.5) and the third domain of the Kazal-type ovomucoid from Turkey (OMTKY3) has been determined crystallographically by the molecular replacement method. Restrained-parameter least-squares refinement of the molecular model of the complex has led to a conventional agreement factor R of 0.168 for the 19,466 reflections in the 1.8 A (1 A = 0.1 nm) resolution shell [I greater than or equal to sigma (I)]. The reactive site loop of OMTKY3, from Lys13I to Arg21I (I indicates inhibitor), is highly complementary to the surface of alpha-chymotrypsin in the complex. A total of 13 residues on the inhibitor make 113 contacts of less than 4.0 A with 21 residues of the enzyme. A short contact (2.95 A) from O gamma of Ser195 to the carbonyl-carbon atom of the scissile bond between Leu18I and Glu19I is present; in spite of it, this peptide remains planar and undistorted. Analysis of the interactions of the inhibitor with chymotrypsin explains the enhanced specificity that chymotrypsin has for P'3 arginine residues. There is a water-mediated ion pair between the guanidinium group on this residue and the carboxylate of Asp64. Comparison of the structure of the alpha-chymotrypsin portion of this complex with the several structures of alpha and gamma-chymotrypsin in the uncomplexed form shows a high degree of structural equivalence (root-mean-square deviation of the 234 common alpha-carbon atoms averages 0.38 A). Significant differences occur mainly in two regions Lys36 to Phe39 and Ser75 to Lys79. Among the 21 residues that are in contact with the ovomucoid domain, only Phe39 and Tyr146 change their conformations significantly as a result of forming the complex. Comparison of the structure of the OMTKY3 domain in this complex to that of the same inhibitor bound to a serine proteinase from Streptomyces griseus (SGPB) shows a central core of 44 amino acids (the central alpha-helix and flanking small 3-stranded beta-sheet) that have alpha-carbon atoms fitting to within 1.0 A (root-mean-square deviation of 0.45 A) whereas the residues of the reactive-site loop differ in position by up to 1.9 A (C alpha of Leu18I). The ovomucoid domain has a built-in conformational flexibility that allows it to adapt to the active sites of different enzymes. A comparison of the SGPB and alpha-chymotrypsin molecules is made and the water molecules bound at the inhibitor-enzyme interface in both complexes are analysed for similarities and differences.


Subject(s)
Chymotrypsin/antagonists & inhibitors , Egg Proteins/pharmacology , Ovomucin/pharmacology , Animals , Crystallography , Hydrogen Bonding , Models, Molecular , Molecular Conformation , Serine Endopeptidases , Temperature , Turkeys
15.
J Mol Biol ; 299(4): 1005-17, 2000 Jun 16.
Article in English | MEDLINE | ID: mdl-10843854

ABSTRACT

Fibers of pilin monomers (pili) form the dominant adhesin of Pseudomonas aeruginosa, and they play an important role in infections by this opportunistic bacterial pathogen. Blocking adhesion is therefore a target for vaccine development. The receptor-binding site is located in a C-terminal disulphide-bonded loop of each pilin monomer, but functional binding sites are displayed only at the tip of the pilus. A factor complicating vaccination is that different bacterial strains produce distinct, and sometimes highly divergent, pilin variants. It is surprising that all strains still appear to bind a common receptor, asialo-GM1. Here, we present the 1.63 A crystal structure of pilin from P. aeruginosa strain PAK. The structure shows that the proposed receptor-binding site is formed by two beta-turns that create a surface dominated by main-chain atoms. Receptor specificity could therefore be maintained, whilst allowing side-chain variation, if the main-chain conformation is conserved. The location of the binding site relative to the proposed packing of the pilus fiber raises new issues and suggests that the current fiber model may have to be reconsidered. Finally, the structure of the C-terminal disulphide-bonded loop will provide the template for the structure-based design of a consensus sequence vaccine.


Subject(s)
Membrane Proteins/chemistry , Membrane Proteins/metabolism , Pseudomonas aeruginosa/chemistry , Pseudomonas aeruginosa/classification , Amino Acid Sequence , Antigens, Bacterial/chemistry , Antigens, Bacterial/metabolism , Bacterial Vaccines/chemistry , Binding Sites , Crystallization , Crystallography, X-Ray , Disulfides/metabolism , Fimbriae Proteins , G(M1) Ganglioside/metabolism , Glycosylation , Membrane Proteins/classification , Models, Molecular , Molecular Sequence Data , Peptide Fragments/chemistry , Peptide Fragments/metabolism , Protein Structure, Secondary , Sequence Alignment , Substrate Specificity , Vaccines, Synthetic/chemistry
16.
J Mol Biol ; 293(3): 449-55, 1999 Oct 29.
Article in English | MEDLINE | ID: mdl-10543942

ABSTRACT

The function of the serpins as proteinase inhibitors depends on their ability to insert the cleaved reactive centre loop as the fourth strand in the main A beta-sheet of the molecule upon proteolytic attack at the reactive centre, P1-P1'. This mechanism is vulnerable to mutations which result in inappropriate intra- or intermolecular loop insertion in the absence of cleavage. Intermolecular loop insertion is known as serpin polymerisation and results in a variety of diseases, most notably liver cirrhosis resulting from mutations of the prototypical serpin alpha1-antitrypsin. We present here the 2.6 A structure of a polymer of alpha1-antitrypsin cleaved six residues N-terminal to the reactive centre, P7-P6 (Phe352-Leu353). After self insertion of P14 to P7, intermolecular linkage is affected by insertion of the P6-P3 residues of one molecule into the partially occupied beta-sheet A of another. This results in an infinite, linear polymer which propagates in the crystal along a 2-fold screw axis. These findings provide a framework for understanding the uncleaved alpha1-antitrypsin polymer and fibrillar and amyloid deposition of proteins seen in other conformational diseases, with the ordered array of polymers in the crystal resulting from slow accretion of the cleaved serpin over the period of a year.


Subject(s)
Liver Cirrhosis/metabolism , Peptide Fragments/chemistry , alpha 1-Antitrypsin/chemistry , alpha 1-Antitrypsin/metabolism , Amino Acid Sequence , Crystallization , Crystallography, X-Ray , Humans , Liver Cirrhosis/genetics , Liver Cirrhosis/pathology , Models, Molecular , Molecular Sequence Data , Peptide Fragments/genetics , Peptide Fragments/metabolism , Plaque, Amyloid/metabolism , Plaque, Amyloid/pathology , Polymers , Protein Structure, Secondary , Recombinant Proteins/chemistry , Recombinant Proteins/metabolism , Time Factors , alpha 1-Antitrypsin/genetics
17.
J Mol Biol ; 305(4): 773-83, 2001 Jan 26.
Article in English | MEDLINE | ID: mdl-11162091

ABSTRACT

Plasminogen activator inhibitor type 1 (PAI-1) is a member of the serine protease inhibitor (serpin) superfamily. Its highly mobile reactive-center loop (RCL) is thought to account for both the rapid inhibition of tissue-type plasminogen activator (t-PA), and the rapid and spontaneous transition of the unstable, active form of PAI-1 into a stable, inactive (latent) conformation (t(1/2) at 37 degrees C, 2.2 hours). We determined the amino acid residues responsible for the inherent instability of PAI-1, to assess whether these properties are independent and, consequently, whether the structural basis for inhibition and latency transition is different. For that purpose, a hypermutated PAI-1 library that is displayed on phage was pre-incubated for increasing periods (20 to 72 hours) at 37 degrees C, prior to a stringent selection for rapid t-PA binding. Accordingly, four rounds of phage-display selection resulted in the isolation of a stable PAI-1 variant (st-44: t(1/2) 450 hours) with 11 amino acid mutations. Backcrossing by DNA shuffling of this stable mutant with wt PAI-1 was performed to eliminate non-contributing mutations. It was shown that the combination of mutations at positions 50, 56, 61, 70, 94, 150, 222, 223, 264 and 331 increases the half-life of PAI-1 245-fold. Furthermore, within the limits of detection the stable mutants isolated are functionally indistinguishable from wild-type PAI-1 with respect to the rate of inhibition of t-PA, cleavage by t-PA, and binding to vitronectin. These stabilizing mutations constitute largely reversions to the stable "serpin consensus sequence" and are located in areas implicated in PAI-1 stability (e.g. the vitronectin-binding domain and the proximal hinge). Collectively, our data provide evidence that the structural requirements for PAI-1 loop insertion during latency transition and target proteinase inhibition can be separated.


Subject(s)
Mutagenesis/genetics , Peptide Library , Plasminogen Activator Inhibitor 1/chemistry , Plasminogen Activator Inhibitor 1/metabolism , Tissue Plasminogen Activator/antagonists & inhibitors , Animals , Consensus Sequence , Half-Life , Humans , Kinetics , Mice , Models, Molecular , Plasminogen Activator Inhibitor 1/genetics , Plasminogen Activator Inhibitor 1/isolation & purification , Protein Binding , Protein Conformation , Recombinant Proteins/chemistry , Recombinant Proteins/isolation & purification , Recombinant Proteins/metabolism , Surface Plasmon Resonance , Thermodynamics , Tissue Plasminogen Activator/metabolism , Vitronectin/metabolism
18.
J Mol Biol ; 206(2): 365-79, 1989 Mar 20.
Article in English | MEDLINE | ID: mdl-2716052

ABSTRACT

The crystal structure of lipoamide dehydrogenase from Azotobacter vinelandii has been determined by a combination of molecular replacement and isomorphous replacement techniques yielding eventually a good-quality 2.8 A electron density map. Initially, the structure determination was attempted by molecular replacement procedures alone using a model of human glutathione reductase, which has 26% sequence identity with this bacterial dehydrogenase. The rotation function yielded the correct orientation of the model structure both when the glutathione reductase dimer and monomer were used as starting model. The translation function could not be solved, however. Consequently, data for two heavy-atom derivatives were collected using the Hamburg synchotron facilities. The derivatives had several sites in common, which was presumably a major reason why the electron density map obtained by isomorphous information alone was of poor quality. Application of solvent flattening procedures cleaned up the map considerably, however, showing clearly the outline of the lipoamide dehydrogenase dimer, which has a molecular weight of 100,000. Application of the "phased translation function", which combines the phase information of both isomorphous and molecular replacement, led to an unambiguous determination of the position of the model structure in the lipoamide dehydrogenase unit cell. The non-crystallographic 2-fold axis of the dimer was optimized by several cycles of constrained-restrained least-squares refinement and subsequently used for phase improvement by 2-fold density averaging. After ten cycles at 3.5 A, the resolution was gradually extended to 2.8 A in another 140 cycles. The 2.8 A electron density distribution obtained in this manner was of much improved quality and allowed building of an atomic model of A. vinelandii lipoamide dehydrogenase. It appears that in the orthorhombic crystals used each dimer is involved in contacts with eight surrounding dimers, leaving unexplained why the crystals are rather fragile. Contacts between subunits within one dimer, which are quite extensive, can be divided into two regions separated by a cavity. In one of the contact regions, the level of sequence identity with glutathione reductase is very low but it is quite high in the other. The folding of the polypeptide chain in each subunit is quite similar to that of glutathione reductase, as is the extended conformation of the co-enzyme FAD.(ABSTRACT TRUNCATED AT 400 WORDS)


Subject(s)
Azotobacter/enzymology , Dihydrolipoamide Dehydrogenase , Amino Acid Sequence , Amino Acids , Crystallization , Models, Molecular , Molecular Sequence Data , X-Ray Diffraction
19.
Protein Sci ; 4(10): 2087-99, 1995 Oct.
Article in English | MEDLINE | ID: mdl-8535245

ABSTRACT

Several sets of amino acid surface areas and transfer free energies were used to derive a total of nine sets of atomic solvation parameters (ASPs). We tested the accuracy of each of these sets of parameters in predicting the experimentally determined transfer free energies of the amino acid derivatives from which the parameters were derived. In all cases, the calculated and experimental values correlated well. We then chose three parameter sets and examined the effect of adding an energetic correction for desolvation based on these three parameter sets to the simple potential function used in our multiple start Monte Carlo docking method. A variety of protein-protein interactions and docking results were examined. In the docking simulations studied, the desolvation correction was only applied during the final energy calculation of each simulation. For most of the docking results we analyzed, the use of an octanol-water-based ASP set marginally improved the energetic ranking of the low-energy dockings, whereas the other ASP sets we tested disturbed the ranking of the low-energy dockings in many of the same systems. We also examined the correlation between the experimental free energies of association and our calculated interaction energies for a series of proteinase-inhibitor complexes. Again, the octanol-water-based ASP set was compatible with our standard potential function, whereas ASP sets derived from other solvent systems were not.


Subject(s)
Amino Acids , Endopeptidases/chemistry , Ovomucin/chemistry , Protease Inhibitors/chemistry , Proteins/chemistry , Serine Endopeptidases/chemistry , Animals , Computer Simulation , Ovomucin/metabolism , Regression Analysis , Serine Endopeptidases/metabolism , Streptomyces griseus/enzymology , Surface Properties , Thermodynamics , Turkeys
20.
Protein Sci ; 4(5): 885-99, 1995 May.
Article in English | MEDLINE | ID: mdl-7663344

ABSTRACT

The development of general strategies for the performance of docking simulations is prerequisite to the exploitation of this powerful computational method. Comprehensive strategies can only be derived from docking experiences with a diverse array of biological systems, and we have chosen the ubiquitin/diubiquitin system as a learning tool for this process. Using our multiple-start Monte Carlo docking method, we have reconstructed the known structure of diubiquitin from its two halves as well as from two copies of the uncomplexed monomer. For both of these cases, our relatively simple potential function ranked the correct solution among the lowest energy configurations. In the experiments involving the ubiquitin monomer, various structural modifications were made to compensate for the lack of flexibility and for the lack of a covalent bond in the modeled interaction. Potentially flexible regions could be identified using available biochemical and structural information. A systematic conformational search ruled out the possibility that the required covalent bond could be formed in one family of low-energy configurations, which was distant from the observed dimer configuration. A variety of analyses was performed on the low-energy dockings obtained in the experiment involving structurally modified ubiquitin. Characterization of the size and chemical nature of the interface surfaces was a powerful adjunct to our potential function, enabling us to distinguish more accurately between correct and incorrect dockings. Calculations with the structure of tetraubiquitin indicated that the dimer configuration in this molecule is much less favorable than that observed in the diubiquitin structure, for a simple monomer-monomer pair. Based on the analysis of our results, we draw conclusions regarding some of the approximations involved in our simulations, the use of diverse chemical and biochemical information in experimental design and the analysis of docking results, as well as possible modifications to our docking protocol.


Subject(s)
Monte Carlo Method , Ubiquitins/chemistry , Algorithms , Computer Graphics , Computer Simulation , Databases, Factual , Molecular Structure , Protein Binding , Protein Conformation , Software , Ubiquitins/metabolism
SELECTION OF CITATIONS
SEARCH DETAIL