Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 20 de 52
Filtrar
Más filtros

Banco de datos
Tipo del documento
Intervalo de año de publicación
1.
Biochemistry ; 52(41): 7217-7230, 2013 Oct 15.
Artículo en Inglés | MEDLINE | ID: mdl-24015903

RESUMEN

Differential interactions between influenza A virus protein hemagglutinin (HA) and α2→3 (avian) or α2→6 (human) sialylated glycan receptors play an important role in governing host specificity and adaptation of the virus. Previous analysis of HA-glycan interactions with trisaccharides showed that, in addition to the terminal sialic acid linkage, the conformation and topology of the glycans, while they are bound to HA, are key factors in regulating these interactions. Here, the solution conformation and dynamics of two representative avian and human glycan pentasaccharide receptors [LSTa, Neu5Ac-α(2→3)-Gal-ß(1→3)-GlcNAc-ß(1→3)-Gal-ß(1→4)-Glc; LSTc, (Neu5Ac-α(2→6)-Gal-ß(1→4)-GlcNAc-ß(1→3)-Gal-ß(1→4)-Glc] have been explored using nuclear magnetic resonance and molecular dynamics simulation. Analyses demonstrate that, in solution, human and avian receptors sample distinct conformations, topologies, and dynamics. These unique features of avian and human receptors in solution could represent distinct molecular characteristics for recognition by HA, thereby providing the HA-glycan interaction specificity in influenza.


Asunto(s)
Virus de la Influenza A/metabolismo , Gripe Aviar/metabolismo , Gripe Humana/metabolismo , Ácido N-Acetilneuramínico/metabolismo , Polisacáridos/química , Receptores Virales/química , Animales , Aves , Glicoproteínas Hemaglutininas del Virus de la Influenza/genética , Glicoproteínas Hemaglutininas del Virus de la Influenza/metabolismo , Humanos , Virus de la Influenza A/genética , Gripe Aviar/virología , Gripe Humana/virología , Modelos Moleculares , Polisacáridos/metabolismo , Receptores Virales/metabolismo
2.
Science ; 286(5439): 537-42, 1999 Oct 15.
Artículo en Inglés | MEDLINE | ID: mdl-10521350

RESUMEN

Although rapid sequencing of polynucleotides and polypeptides has become commonplace, it has not been possible to rapidly sequence femto- to picomole amounts of tissue-derived complex polysaccharides. Heparin-like glycosaminoglycans (HLGAGs) were readily sequenced by a combination of matrix-assisted laser desorption ionization mass spectrometry and a notation system for representation of polysaccharide sequences. This will enable identification of sequences that are critical to HLGAG biological activities in anticoagulation, cell growth, and differentiation.


Asunto(s)
Glicosaminoglicanos/química , Oligosacáridos/química , Análisis de Secuencia , Secuencia de Carbohidratos , Disacáridos/química , Heparina/química , Liasa de Heparina/metabolismo , Datos de Secuencia Molecular , Peso Molecular , Ácido Nitroso , Polisacárido Liasas/metabolismo , Espectrometría de Masa por Láser de Matriz Asistida de Ionización Desorción
4.
Curr Opin Chem Biol ; 4(6): 626-31, 2000 Dec.
Artículo en Inglés | MEDLINE | ID: mdl-11102866

RESUMEN

Heparin and heparan sulfate glycosaminoglycans are acidic complex polysaccharides found on the cell surface and in the extracellular matrix. Recent progress has uncovered a virtual explosion of important roles of these biopolymers in fundamental biological processes. Advances in the understanding of biosynthesis and structure and the development of novel analytical methods for composition and sequence analysis have provided remarkable insights into structure/function relationships of these complex and once elusive polysaccharides.


Asunto(s)
Anticoagulantes/química , Heparina/biosíntesis , Heparina/química , Heparitina Sulfato/biosíntesis , Heparitina Sulfato/química , Animales , Glicosaminoglicanos/biosíntesis , Glicosaminoglicanos/química , Humanos
5.
Angiogenesis ; 1(1): 45-54, 1997.
Artículo en Inglés | MEDLINE | ID: mdl-14517393

RESUMEN

The activity of several angiogenic factors, like fibroblast growth factors (FGF), is modulated by heparin-like glycosaminoglycans (HLGAGs), which are acidic polysaccharides present in the extracellular matrix and at the cell surface. FGF binds to HLGAG in the matrix, where it is sequestered in a protected and inactive form, and at the cell surface, where it activates its cognate signaling receptor. Here we review recent progress in elucidating how HLGAG regulates FGF-induced signal transduction. Data from crystal structures of FGF complexed to active and inactive oligosaccharides is analyzed in the context of current models for HLGAG modulation of FGF activity. We propose that FGF can dimerize in several different modes, stabilized by HLGAGs. Individual dimer modes may represent active or inactive FGF and it is possible that different HLGAGs preferentially stabilize different FGF dimer modes. Understanding HLGAG-FGF interactions can provide leverage for new approaches to therapeutic control of angiogenesis.

6.
Angiogenesis ; 1(2): 174-184, 1998.
Artículo en Inglés | MEDLINE | ID: mdl-14517383

RESUMEN

Lipid hydroperoxides (LHP) at high concentrations are cytotoxic, but at sublethal concentration, they induce synthesis of cytokine vascular growth factors. Intracorneal injections of 30 µg LHP placed 5 mm from the superior limbus stimulated early vasodilation of limbal vasculature and a rapidly developing, sustained neovascularization. Under these conditions, vessels grew at the rate of 0.3 mm/day to a total length of 7.5 mm, 25 days after injection. Cholesterol peroxides were less effective. Developing vessels were oriented towards the stimulus. Around the developing vessel there was dissolution of the stromal extracellular matrix. The most distal endothelial cells displayed prominent endoplasmic reticulum, a lack of basement membrane or tight junction complexes and leakage of fluorescein dye. Both the injection site and superior quadrant showed increased levels of tumor necrosis factor (TNF)-alpha and vascular endothelial growth factor after exposure to LHP. The neovascular response was inhibited by simultaneous administration of TNF-alpha antibody or pentoxifylline, an inhibitor of TNF-alpha synthesis. This corneal model of peroxide-induced neovascularization should prove useful for temporal studies of events in the initiation and propagation of signals leading to neovascularization, and for evaluating effects of treatment on neovascular growth.

7.
Br J Cancer ; 96(9): 1315-9, 2007 May 07.
Artículo en Inglés | MEDLINE | ID: mdl-17406364

RESUMEN

Nanotechnology is increasingly finding use in the management of cancer. Nanoscale devices have impacted cancer biology at three levels: early detection using, for example, nanocantilevers or nanoparticles; tumour imaging using radiocontrast nanoparticles or quantum dots; and drug delivery using nanovectors and hybrid nanoparticles. This review addresses some of the major milestones in the integration of nanotechnology and cancer biology, and the future of nanoscale approaches for cancer management.


Asunto(s)
Nanotecnología/métodos , Neoplasias/terapia , Antineoplásicos/uso terapéutico , Técnicas Biosensibles , Humanos , Microscopía Electrónica de Rastreo , Neoplasias/tratamiento farmacológico
8.
J Biol Chem ; 273(1): 248-55, 1998 Jan 02.
Artículo en Inglés | MEDLINE | ID: mdl-9417072

RESUMEN

Heparinases are bacterial enzymes that are powerful tools to study the physiological roles of heparin-like complex polysaccharides. In addition, heparinases have significant therapeutic applications. We had proposed earlier that cysteine 135 and histidine 203 together form the catalytic domain in heparinase I. We had also identified a heparin binding domain in heparinase I containing two positively charged clusters HB-1 and HB-2 in a primary heparin binding site and other positively charged residues in the vicinity of cysteine 135. In this study, through systematic site-directed mutagenesis studies, we show that the alteration of the positive charge of the HB-1 region has a pronounced effect on heparinase I activity. More specifically, site-directed mutagenesis of K199A (contained in HB-1) results in a 15-fold reduction in catalytic activity, whereas a K198A mutation (also in HB-1) results in only a 2- to 3-fold reduction in heparinase I activity. A K132A mutation, in close proximity to cysteine 135, also resulted in reduced (8-fold) activity. Heparin affinity chromatography experiments indicated moderately lowered binding affinities for the K132A, K198A, and the K199A mutant enzymes. The above results, taken together with our previous observations, lead us to propose that the positively charged heparin binding domain provides the necessary microenvironment for the catalytic domain of heparinase I. The dominant effect of lysine 199 suggests an additional, more direct, role in catalysis for this residue.


Asunto(s)
Flavobacterium/enzimología , Liasa de Heparina/metabolismo , Sitios de Unión , Cromatografía de Afinidad , Cromatografía Líquida de Alta Presión , Cromatografía por Intercambio Iónico , Heparina/metabolismo , Liasa de Heparina/química , Liasa de Heparina/genética , Mutagénesis Sitio-Dirigida
9.
Biochem Biophys Res Commun ; 229(3): 770-7, 1996 Dec 24.
Artículo en Inglés | MEDLINE | ID: mdl-8954971

RESUMEN

Heparinases I, II and III from F. heparinum cleave heparin-like molecules, with a high degree of substrate specificity, at the glucosamine-uronate linkage by elimination, leaving an unsaturated C4-C5 bond in the uronic acid. The primary sequences of these enzymes have been reported earlier. In this study we perform a comparative analysis of the properties and primary sequences of heparinase I, II and III. Alignment of the primary sequences revealed little sequence homology (15% residue identity in a LALIGN alignment) at both DNA and amino acid levels. There are three basic clusters in heparinase II satisfying the heparin binding consensus sequence with one of the sequences sharing homology with a consensus sequence in the heparin binding site of heparinase I and two basic clusters in heparinase III. Similar to heparinase I, there are two putative 'EF-hand' calcium coordinating motifs in heparinase III, while heparinase II does not contain any such motifs. Recombinant heparinases II and III's degradation of the substrate and the subsequent separation of the oligosaccharide products by POROS anion exchange chromatography were identical to those obtained from native heparinases II and III from F. heparinum.


Asunto(s)
Flavobacterium/enzimología , Polisacárido Liasas/genética , Cromatografía Líquida de Alta Presión , Liasa de Heparina , Proteínas Recombinantes/genética , Análisis de Secuencia
10.
J Biol Chem ; 273(17): 10160-7, 1998 Apr 24.
Artículo en Inglés | MEDLINE | ID: mdl-9553064

RESUMEN

The three heparinases derived from Flavobacterium heparinum are powerful tools for studying heparin-like glycosaminoglycans in major biological processes, including angiogenesis and development. Heparinase II is unique among the three enzymes because it is able to catalytically cleave both heparin and heparan sulfate-like regions of heparin-like glycosaminoglycans. Toward understanding the catalytic mechanism of heparin-like glycosaminoglycan degradation by heparinase II, we set out to investigate the role of the histidines of heparinase II in catalysis. We observe concentration-dependent inactivation of heparinase II in the presence of the reversible histidine-modifying reagent diethylpyrocarbonate (DEPC). With heparin as the substrate, the rate constant of inactivation was found to be 0.16 min-1 mM-1; with heparan sulfate as the substrate, the rate constant was determined to be 0.24 min-1 mM-1. Heparinase II activity is restored following hydroxylamine treatment. This, along with other experiments, strongly suggests that the inactivation of heparinase II by DEPC is specific for histidine residues and that three histidines are modified by DEPC. Substrate protection experiments show that heparinase II preincubation with heparin followed by the addition of DEPC resulted in a loss of enzymatic activity toward heparan sulfate but not heparin. However, heparinase II preincubation with heparan sulfate was unable to protect heparinase II from DEPC inactivation for either of the substrates. Proteolytic mapping studies with Lys-C were consistent with the chemical modification experiments and identified histidines 238, 451, and 579 as being important for heparinase II activity. Further mapping studies identified histidine 451 as being essential for heparin degradation. Site-directed mutagenesis experiments on the 13 histidines of heparinase II corroborated the chemical modification and the peptide mapping studies, establishing the importance of histidines 238, 451 and 579 in heparinase II activity.


Asunto(s)
Flavobacterium/enzimología , Liasa de Heparina/metabolismo , Histidina/metabolismo , Sitios de Unión , Dietil Pirocarbonato/química , Activación Enzimática , Heparina/metabolismo , Liasa de Heparina/química , Liasa de Heparina/genética , Heparitina Sulfato/metabolismo , Histidina/química , Concentración de Iones de Hidrógeno , Hidrólisis , Sondas Moleculares , Mutagénesis Sitio-Dirigida
11.
Proc Natl Acad Sci U S A ; 95(8): 4182-7, 1998 Apr 14.
Artículo en Inglés | MEDLINE | ID: mdl-9539710

RESUMEN

Heparinase I from Flavobacterium heparinum has important uses for elucidating the complex sequence heterogeneity of heparin-like glycosaminoglycans (HLGAGs). Understanding the biological function of HLGAGs has been impaired by the limited methods for analysis of pure or mixed oligosaccharide fragments. Here, we use methodologies involving MS and capillary electrophoresis to investigate the sequence of events during heparinase I depolymerization of HLGAGs. In an initial step, heparinase I preferentially cleaves exolytically at the nonreducing terminal linkage of the HLGAG chain, although it also cleaves internal linkages at a detectable rate. In a second step, heparinase I has a strong preference for cleaving the same substrate molecule processively, i.e., to cleave the next site toward the reducing end of the HLGAG chain. Computer simulation showed that the experimental results presented here from analysis of oligosaccharide degradation were consistent with literature data for degradation of polymeric HLGAG by heparinase I. This study presents direct evidence for a predominantly exolytic and processive mechanism of depolymerization of HLGAG by heparinase I.


Asunto(s)
Glicosaminoglicanos/química , Glicosaminoglicanos/metabolismo , Liasa de Heparina/metabolismo , Heparina/química , Oligosacáridos/química , Secuencia de Carbohidratos , Glicopéptidos/química , Glicopéptidos/metabolismo , Heparina/metabolismo , Conformación Molecular , Datos de Secuencia Molecular , Espectrometría de Masa por Láser de Matriz Asistida de Ionización Desorción , Especificidad por Sustrato
12.
Proc Natl Acad Sci U S A ; 95(8): 4176-81, 1998 Apr 14.
Artículo en Inglés | MEDLINE | ID: mdl-9539709

RESUMEN

Difficulties in determining composition and sequence of glycosaminoglycans, such as those related to heparin, have limited the investigation of these biologically important molecules. Here, we report methodology, based on matrix-assisted laser desorption ionization MS and capillary electrophoresis, to follow the time course of the enzymatic degradation of heparin-like glycosaminoglycans through the intermediate stages to the end products. MS allows the determination of the molecular weights of the sulfated carbohydrate intermediates and their approximate relative abundances at different time points of the experiment. Capillary electrophoresis subsequently is used to follow more accurately the abundance of the components and also to measure sulfated disaccharides for which MS is not well applicable. For those substrates that produce identical or isomeric intermediates, the reducing end of the carbohydrate chain was converted to the semicarbazone. This conversion increases the molecular weight of all products retaining the reducing terminus by the "mass tag" (in this case 56 Da) and thus distinguishes them from other products. A few picomoles of heparin-derived, sulfated hexa- to decasaccharides of known structure were subjected to heparinase I digestion and analyzed. The results indicate that the enzyme acts primarily exolytically and in a processive mode. The methodology described should be equally useful for other enzymes, including those modified by site-directed mutagenesis, and may lead to the development of an approach to the sequencing of complex glycosaminoglycans.


Asunto(s)
Glicosaminoglicanos/química , Glicosaminoglicanos/metabolismo , Liasa de Heparina/metabolismo , Heparina/química , Secuencia de Carbohidratos , Disacáridos/química , Electroforesis Capilar , Heparina/metabolismo , Datos de Secuencia Molecular , Peso Molecular , Oligosacáridos/química , Espectrometría de Masa por Láser de Matriz Asistida de Ionización Desorción
13.
J Biol Chem ; 274(7): 4082-8, 1999 Feb 12.
Artículo en Inglés | MEDLINE | ID: mdl-9933601

RESUMEN

The heparinases from Flavobacterium heparinum are lyases that specifically cleave heparin-like glycosaminoglycans. Previously, amino acids located in the active site of heparinase I have been identified and mapped. In an effort to further understand the mechanism by which heparinase I cleaves its polymer substrate, we sought to understand the role of calcium, as a necessary cofactor, in the enzymatic activity of heparinase I. Specifically, we undertook a series of biochemical and biophysical experiments to answer the question of whether heparinase I binds to calcium and, if so, which regions of the protein are involved in calcium binding. Using the fluorescent calcium analog terbium, we found that heparinase I tightly bound divalent and trivalent cations. Furthermore, we established that this interaction was specific for ions that closely approximate the ionic radius of calcium. Through the use of the modification reagents N-ethyl-5-phenylisoxazolium-3'-sulfonate (Woodward's reagent K) and 1-(3-dimethylaminopropyl)-3-ethylcarbodiimide hydrochloride, we showed that the interaction between heparinase I and calcium was essential for proper functioning of the enzyme. Preincubation with either calcium alone or calcium in the presence of heparin was able to protect the enzyme from inactivation by these modifying reagents. In addition, through mapping studies of Woodward's reagent K-modified heparinase I, we identified two putative calcium-binding sites, CB-1 (Glu207-Ala219) and CB-2 (Thr373-Arg384), in heparinase I that not only are specifically modified by Woodward's reagent K, leading to loss of enzymatic activity, but also conform to the calcium-coordinating consensus motif.


Asunto(s)
Calcio/metabolismo , Flavobacterium/enzimología , Liasa de Heparina/metabolismo , Secuencia de Aminoácidos , Sitios de Unión , Liasa de Heparina/genética , Concentración de Iones de Hidrógeno , Metales de Tierras Raras/metabolismo , Datos de Secuencia Molecular , Morfolinas/metabolismo , Mutagénesis Sitio-Dirigida , Terbio/metabolismo
14.
Proc Natl Acad Sci U S A ; 96(5): 1892-7, 1999 Mar 02.
Artículo en Inglés | MEDLINE | ID: mdl-10051565

RESUMEN

Fibroblast growth factor (FGF) 1 and FGF-2 are prototypic members of the FGF family, which to date comprises at least 18 members. Surprisingly, even though FGF-1 and FGF-2 share more than 80% sequence similarity and an identical structural fold, these two growth factors are biologically very different. FGF-1 and FGF-2 differ in their ability to bind isoforms of the FGF receptor family as well as the heparin-like glycosaminoglycan (HLGAG) component of proteoglycans on the cell surface to initiate signaling in different cell types. Herein, we provide evidence for one mechanism by which these two proteins could differ biologically. Previously, it has been noted that FGF-1 and FGF-2 can oligomerize in the presence of HLGAGs. Therefore, we investigated whether FGF-1 and FGF-2 oligomerize by the same mechanism or by a different one. Through a combination of matrix-assisted laser desorption ionization mass spectrometry and chemical crosslinking, we show here that, under identical conditions, FGF-1 and FGF-2 differ in the degree and kind of oligomerization. Furthermore, an extensive analysis of FGF-1 and FGF-2 uncomplexed and HLGAG complexed crystal structures enables us to readily explain why FGF-2 forms sequential oligomers whereas FGF-1 forms only dimers. FGF-2, which possesses an interface capable of protein association, forms a translationally related oligomer, whereas FGF-1, which does not have this interface, forms only a symmetrically related dimer. Taken together, these data show that FGF-1 and FGF-2, despite their sequence homology, differ in their mechanism of oligomerization.


Asunto(s)
Factor 2 de Crecimiento de Fibroblastos/química , Glicosaminoglicanos/química , Heparina/química , Conformación Proteica , Sitios de Unión , Reactivos de Enlaces Cruzados , Cristalografía por Rayos X , Factor 1 de Crecimiento de Fibroblastos , Factor 2 de Crecimiento de Fibroblastos/metabolismo , Ligandos , Sustancias Macromoleculares , Oligosacáridos/química , Proteínas Recombinantes/química , Proteínas Recombinantes/metabolismo , Espectrometría de Masa por Láser de Matriz Asistida de Ionización Desorción
15.
Biochemistry ; 39(14): 4012-9, 2000 Apr 11.
Artículo en Inglés | MEDLINE | ID: mdl-10747789

RESUMEN

The heparinases from Flavobacterium heparinum are powerful tools in understanding how heparin-like glycosaminoglycans function biologically. Heparinase III is the unique member of the heparinase family of heparin-degrading lyases that recognizes the ubiquitous cell-surface heparan sulfate proteoglycans as its primary substrate. Given that both heparinase I and heparinase II contain catalytically critical histidines, we examined the role of histidine in heparinase III. Through a series of diethyl pyrocarbonate modification experiments, it was found that surface-exposed histidines are modified in a concentration-dependent fashion and that this modification results in inactivation of the enzyme (k(inact) = 0.20 +/- 0.04 min(-)(1) mM(-)(1)). The DEPC modification was pH dependent and reversible by hydroxylamine, indicating that histidines are the sole residue being modified. As previously observed for heparinases I and II, substrate protection experiments slowed the inactivation kinetics, suggesting that the modified residue(s) was (were) in or proximal to the active site of the enzyme. Proteolytic mapping experiments, taken together with site-directed mutagenesis studies, confirm the chemical modification experiments and point to two histidines, histidine 295 and histidine 510, as being essential for heparinase III enzymatic activity.


Asunto(s)
Heparitina Sulfato/metabolismo , Polisacárido Liasas/metabolismo , Animales , Sitios de Unión/genética , Flavobacterium , Heparitina Sulfato/química , Heparitina Sulfato/genética , Histidina , Mutagénesis Sitio-Dirigida , Polisacárido Liasas/química , Polisacárido Liasas/genética , Especificidad por Sustrato
16.
Biochem J ; 303 ( Pt 2): 527-30, 1994 Oct 15.
Artículo en Inglés | MEDLINE | ID: mdl-7980413

RESUMEN

Pancreatic and venom phospholipases A2 have complex and distinct oligomerization behaviour. Pancreatic enzymes are monomeric in solution, but their quaternary structure at interfaces is unknown. On the other hand, certain crotalid venom phospholipases A2 are dimeric in solution, and different reports have proposed either the monomer or the dimer as the catalytically functional subunit. In this study, enzyme immobilization was used as a tool for determining the functional subunits of these enzymes. The dimeric Crotalus atrox phospholipase A2 was covalently attached to agarose beads, via either the amine or the carboxylic groups of the protein. In the first case immobilization led to an 80% loss of activity as compared with the soluble form, and measured by using micellar diheptanoylphosphocholine. Inclusion of micellar protectants in the coupling media did not improve the activity. Enzyme immobilized via carboxylic groups was 2-3-fold more active than the amine-coupled form. In a second approach, Crotalus atrox enzyme was immobilized with single-subunit attachment. The removal, with denaturating washes, of the non-covalently bound units involved in monomer-monomer interactions, caused a large decrease in specific activity of the support-bound enzyme. This suggests the dimeric form as the fully active one. Similar procedures were also carried out with pig pancreatic and Naja naja phospholipases A2. The results indicated that these enzymes are active as monomers.


Asunto(s)
Venenos de Crotálidos/enzimología , Venenos Elapídicos/enzimología , Páncreas/enzimología , Fosfolipasas A/química , Animales , Crotalus , Elapidae , Enzimas Inmovilizadas/química , Enzimas Inmovilizadas/metabolismo , Micelas , Fosfatidilinositoles/química , Fosfolipasas A/metabolismo , Fosfolipasas A2 , Unión Proteica , Conformación Proteica , Porcinos
17.
Eur J Biochem ; 223(2): 611-6, 1994 Jul 15.
Artículo en Inglés | MEDLINE | ID: mdl-8055932

RESUMEN

Pig phospholipase A2, pig iso-phospholipase A2 and bovine pancreatic phospholipase A2 were reacted in solution with 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide, in the presence of N-hydroxysulfosuccinimide, at pH 7. The influence of micellar protectants was analyzed. In the presence of n-hexadecylphosphocholine, the losses of activity in micellar diheptanoyl-lecithin were 80, 35, and 10% in bovine phospholipase A2, pig iso-phospholipase A2, and pig phospholipase A2, respectively. With 1-oleoylglycerophosphocholine, the bovine enzyme lost 40% activity, but the pig enzyme was activated sevenfold. The modified pig enzyme showed pre-micellar activation on monomeric diheptanoyl-lecithin, and either reduced or increased activities on mixed micelles of bile salt with egg phosphatidylcholine, depending on the composition of the micelles. This activation is consistent with previous protein-engineering studies of pig pancreatic phospholipase A2. In this study, we present new information concerning the specificity and interfacial recognition behaviour of this enzyme in relation to this activation.


Asunto(s)
Etildimetilaminopropil Carbodiimida/farmacología , Lisofosfatidilcolinas , Páncreas/enzimología , Fosfolipasas A/metabolismo , Animales , Calcio/metabolismo , Bovinos , Etildimetilaminopropil Carbodiimida/química , Glicerilfosforilcolina/análogos & derivados , Glicerilfosforilcolina/farmacología , Concentración de Iones de Hidrógeno , Cinética , Micelas , Fosfolipasas A/química , Fosfolipasas A2 , Especificidad por Sustrato , Succinimidas/farmacología , Porcinos
18.
Proc Natl Acad Sci U S A ; 95(21): 12232-7, 1998 Oct 13.
Artículo en Inglés | MEDLINE | ID: mdl-9770469

RESUMEN

Heparin-like glycosaminoglycans, acidic complex polysaccharides present on cell surfaces and in the extracellular matrix, regulate important physiological processes such as anticoagulation and angiogenesis. Heparin-like glycosaminoglycan degrading enzymes or heparinases are powerful tools that have enabled the elucidation of important biological properties of heparin-like glycosaminoglycans in vitro and in vivo. With an overall goal of developing an approach to sequence heparin-like glycosaminoglycans using the heparinases, we recently have elaborated a mass spectrometry methodology to elucidate the mechanism of depolymerization of heparin-like glycosaminoglycans by heparinase I. In this study, we investigate the mechanism of depolymerization of heparin-like glycosaminoglycans by heparinase II, which possesses the broadest known substrate specificity of the heparinases. We show here that heparinase II cleaves heparin-like glycosaminoglycans endolytically in a nonrandom manner. In addition, we show that heparinase II has two distinct active sites and provide evidence that one of the active sites is heparinase I-like, cleaving at hexosamine-sulfated iduronate linkages, whereas the other is presumably heparinase III-like, cleaving at hexosamine-glucuronate linkages. Elucidation of the mechanism of depolymerization of heparin-like glycosaminoglycans by the heparinases and mutant heparinases could pave the way to the development of much needed methods to sequence heparin-like glycosaminoglycans.


Asunto(s)
Glicosaminoglicanos/metabolismo , Heparina/química , Polisacárido Liasas/metabolismo , Sitios de Unión , Biopolímeros , Glicosaminoglicanos/química , Espectrometría de Masa por Láser de Matriz Asistida de Ionización Desorción
19.
J Biol Chem ; 273(36): 22904-12, 1998 Sep 04.
Artículo en Inglés | MEDLINE | ID: mdl-9722510

RESUMEN

Heparinase II (no EC number) is one of three lyases isolated from Flavobacterium heparinum that degrade heparin-like complex polysaccharides. Heparinase II is unique among the heparinases in that it has broad substrate requirements and possesses the ability to degrade both heparin and heparan sulfate-like regions of glycosaminoglycans. This study set out to investigate the role of cysteines in heparinase II activity. Through a series of chemical modification experiments, it was found that one of the three cysteines in heparinase II is surface-accessible and possesses unusual chemical reactivity toward cysteine-specific chemical modifying reagents. Substrate protection experiments suggest that this surface-accessible cysteine is proximate to the active site, since addition of substrate shields the cysteine from modifying reagents. The cysteine, present in an ionic environment, was mapped by radiolabeling with N-[3H]ethylmaleimide and identified as cysteine 348. Site-directed mutagenesis of cysteine 348 to an alanine resulted in loss of activity toward heparin but not heparan sulfate, indicating that cysteine 348 is required for heparinase II activity toward heparin but is not essential for the breakdown of heparan sulfate. Furthermore, we show in this study that cysteine 164 and cysteine 189 are functionally unimportant for heparinase II.


Asunto(s)
Flavobacterium/enzimología , Polisacárido Liasas/metabolismo , Sitios de Unión , Cisteína/química , Cisteína/metabolismo , Heparina/farmacología , Liasa de Heparina/metabolismo , Heparitina Sulfato/farmacología , Concentración de Iones de Hidrógeno , Metaloendopeptidasas/farmacología , Mutagénesis Sitio-Dirigida , Fragmentos de Péptidos , Polisacárido Liasas/efectos de los fármacos , Polisacárido Liasas/genética , Proteínas Recombinantes/metabolismo , Reactivos de Sulfhidrilo/farmacología
20.
Biochemistry ; 32(32): 8098-102, 1993 Aug 17.
Artículo en Inglés | MEDLINE | ID: mdl-8347610

RESUMEN

Phospholipase A2 from Naja naja kaouthia venom was covalently coupled onto agarose beads using two different chemistries. The effect of micellar competitive inhibitors in the coupling media was evaluated. Enzyme bound to N-hydroxysuccinimide-activated agarose, which is reactive primarily toward epsilon-amino groups, had 20% activity retention against micellar diheptanoylphosphatidylcholine (DiC7-PC). Enzyme bound through carboxylic groups, using a modification of the carbodiimide method, had 50% retention. Similar relative activities were observed, for both conjugates, in monomeric dihexanoyl-PC and in mixed micelles of Triton X-100 with dipalmitoyl-PC or dioleoylphosphatidylethanolamine. The soluble form of the enzyme showed premicellar activation against monomeric DiC7-PC, while the immobilized form showed interfacial recognition at concentrations around the critical micellar concentration. These results suggest that the enzyme activity lost upon immobilization is a result of the inherent chemical modification of the enzyme and that enzyme oligomerization and interfacial recognition are not cause-effect phenomena.


Asunto(s)
Venenos Elapídicos/química , Enzimas Inmovilizadas , Fosfolipasas A/química , Unión Competitiva , Activación Enzimática , Micelas , Octoxinol , Fosfatidilcolinas/metabolismo , Fosfolipasas A/antagonistas & inhibidores , Fosfolipasas A/metabolismo , Fosfolipasas A2 , Fosfolípidos/metabolismo , Polietilenglicoles , Unión Proteica , Sefarosa , Succinimidas
SELECCIÓN DE REFERENCIAS
DETALLE DE LA BÚSQUEDA