82 results on '"Schowen RL"'
Search Results
2. Mechanism of Decarboxylation of Pyruvic Acid in the Presence of Hydrogen Peroxide.
- Author
-
Lopalco A, Dalwadi G, Niu S, Schowen RL, Douglas J, and Stella VJ
- Subjects
- Chromatography, High Pressure Liquid methods, Decarboxylation, Hydrogen Peroxide analysis, Hydrogen-Ion Concentration, Pyruvic Acid analysis, Hydrogen Peroxide chemistry, Hydrogen Peroxide metabolism, Pyruvic Acid chemistry, Pyruvic Acid metabolism
- Abstract
The purpose of this work was to probe the rate and mechanism of rapid decarboxylation of pyruvic acid in the presence of hydrogen peroxide (H2O2) to acetic acid and carbon dioxide over the pH range 2-9 at 25 °C, utilizing UV spectrophotometry, high performance liquid chromatography (HPLC), and proton and carbon nuclear magnetic resonance spectrometry ((1)H, (13)C-NMR). Changes in UV absorbance at 220 nm were used to determine the kinetics as the reaction was too fast to follow by HPLC or NMR in much of the pH range. The rate constants for the reaction were determined in the presence of molar excess of H2O2 resulting in pseudo first-order kinetics. No buffer catalysis was observed. The calculated second-order rate constants for the reaction followed a sigmoidal shape with pH-independent regions below pH 3 and above pH 7 but increased between pH 4 and 6. Between pH 4 and 9, the results were in agreement with a change from rate-determining nucleophilic attack of the deprotonated peroxide species, HOO(-), on the α-carbonyl group followed by rapid decarboxylation at pH values below 6 to rate-determining decarboxylation above pH 7. The addition of H2O2 to ethyl pyruvate was also characterized., (Copyright © 2016 American Pharmacists Association®. Published by Elsevier Inc. All rights reserved.)
- Published
- 2016
- Full Text
- View/download PDF
3. Isotope Effects and Temperature Dependences in the Action of the Glucose Dehydrogenase of the Mesophilic Bacterium Bacillus megaterium.
- Author
-
Anandarajah K, Schowen KB, and Schowen RL
- Abstract
The glucose dehydrogenase of the mesophilic bacterium Bacillus megaterium (optimal growth around 35 °C) exhibits non-linear Eyring temperature dependences from 25 to 55 °C in its catalysis of the oxidation by hydride-transfer to NAD
+ of the β-anomers of 1- h -D-glucose and 1- d -D-glucose (rate constant kcat /KMβ ). A break around 300K separates a high-T region from a low-T region. In the high-T region, isotopic enthalpies of activation within a considerable experimental error are equal to zero. In the low-T region, the enthalpies of activation are roughly equal for the isotopic substrates but are different from zero. An alternative treatment with Eyring plots taken as effectively linear produces enthalpies of activation having the unusual feature of being larger for the H-substrate (26 kJ/mol) than for the D-substrate (21 kJ/mol). Compensation of the enthalpic effect by a more positive entropy for the H-substrate then reproduces the isotope effects. For oxidation by NADP+ of the same pair of isotopic glucose substrates, catalysis by the glucose dehydrogenase of Thermoplasma acidophilum , a thermophilic archaeon, leads to temperature dependences characterized by a high-T region and a low-T region separated by a gentle thermal transition (K. Anandarajah, K.B. Schowen, and R.L. Schowen, Z. phys. Chem. 2008, 222, 1333-1347). Tentative approaches to a mechanistic interpretation of both cases rely on models featuring configurational searches of the enzyme for tunneling states, followed by hydrogen-transfer tunneling, although explanations can be constructed also on the basis of simple transition-state stabilization without tunnelling.- Published
- 2013
- Full Text
- View/download PDF
4. Comparative kinetics of cofactor association and dissociation for the human and trypanosomal S-adenosylhomocysteine hydrolases. 3. Role of lysyl and tyrosyl residues of the C-terminal extension.
- Author
-
Cai S, Fang J, Li QS, Borchardt RT, Kuczera K, Middaugh CR, and Schowen RL
- Subjects
- Binding Sites, Humans, Kinetics, NAD chemistry, S-Adenosylhomocysteine metabolism, Trypanosoma metabolism, X-Rays, Adenosylhomocysteinase chemistry, Adenosylhomocysteinase metabolism, NAD metabolism
- Abstract
On the basis of the available X-ray structures of S-adenosylhomocysteine hydrolases (SAHHs), free energy simulations employing the MM-GBSA approach were applied to predict residues important to the differential cofactor binding properties of human and trypanosomal SAHHs (Hs-SAHH and Tc-SAHH), within 5 Å of the cofactor NAD(+)/NADH binding site. Among the 38 residues in this region, only four are different between the two enzymes. Surprisingly, the four nonidentical residues make no major contribution to differential cofactor binding between Hs-SAHH and Tc-SAHH. On the other hand, four pairs of identical residues are shown by free energy simulations to differentiate cofactor binding between Hs-SAHH and Tc-SAHH. Experimental mutagenesis was performed to test these predictions for a lysine residue and a tyrosine residue of the C-terminal extension that penetrates a partner subunit to form part of the cofactor binding site. The K431A mutant of Tc-SAHH (TcK431A) loses its cofactor binding affinity but retains the wild type's tetrameric structure, while the corresponding mutant of Hs-SAHH (HsK426A) loses both cofactor affinity and tetrameric structure [Ault-Riche, D. B., et al. (1994) J. Biol. Chem. 269, 31472-31478]. The tyrosine mutants HsY430A and TcY435A alter the NAD(+) association and dissociation kinetics, with HsY430A increasing the cofactor equilibrium dissociation constant from approximately 10 nM (Hs-SAHH) to ∼800 nM and TcY435A increasing the cofactor equilibrium dissociation constant from approximately 100 nM (Tc-SAHH) to ∼1 mM. Both changes result from larger increases in the off rate combined with smaller decreases in the on rate. These investigations demonstrate that computational free energy decomposition may be used to guide experimental studies by suggesting sensitive sites for mutagenesis. Our finding that identical residues in two orthologous proteins may give significantly different binding free energy contributions strongly suggests that comparative studies of homologous proteins should investigate not only different residues but also identical residues in these proteins.
- Published
- 2010
- Full Text
- View/download PDF
5. The rationale for targeting the NAD/NADH cofactor binding site of parasitic S-adenosyl-L-homocysteine hydrolase for the design of anti-parasitic drugs.
- Author
-
Cai S, Li QS, Fang J, Borchardt RT, Kuczera K, Middaugh CR, and Schowen RL
- Subjects
- Adenosylhomocysteinase antagonists & inhibitors, Binding Sites, Humans, Models, Molecular, NAD chemistry, Protein Binding, Protein Conformation, Substrate Specificity, Thermodynamics, Adenosylhomocysteinase chemistry, Adenosylhomocysteinase metabolism, Chagas Disease drug therapy, NAD metabolism, Trypanocidal Agents chemistry, Trypanocidal Agents pharmacology, Trypanosoma cruzi enzymology
- Abstract
Trypanosomal S-adenoyl-L-homocysteine hydrolase (Tc-SAHH), considered as a target for treatment of Chagas disease, has the same catalytic mechanism as human SAHH (Hs-SAHH) and both enzymes have very similar x-ray structures. Efforts toward the design of selective inhibitors against Tc-SAHH targeting the substrate binding site have not to date shown any significant promise. Systematic kinetic and thermodynamic studies on association and dissociation of cofactor NAD/H for Tc-SAHH and Hs-SAHH provide a rationale for the design of anti-parasitic drugs directed toward cofactor-binding sites. Analogues of NAD and their reduced forms show significant selective inactivation of Tc-SAHH, confirming that this design approach is rational.
- Published
- 2009
- Full Text
- View/download PDF
6. Evaluation of NAD(H) analogues as selective inhibitors for Trypanosoma cruzi S-adenosylhomocysteine hydrolase.
- Author
-
Li QS, Cai S, Fang J, Borchardt RT, Kuczera K, Middaugh CR, and Schowen RL
- Subjects
- Chagas Disease drug therapy, Humans, Trypanosoma cruzi drug effects, Adenosylhomocysteinase antagonists & inhibitors, Adenosylhomocysteinase metabolism, NAD analogs & derivatives, Trypanocidal Agents chemistry, Trypanocidal Agents pharmacology, Trypanosoma cruzi enzymology
- Abstract
S-Adenosylhomocysteine (AdoHcy) hydrolases (SAHHs) from human sources (Hs-SAHHs) bind the cofactor NAD(+) more tightly than several parasitic SAHHs by around 1000-fold. This property suggests the cofactor binding site of this essential enzyme as a potential anti-parasitic drug target, e.g., against SAHH from Trypansoma cruzi (Tc-SAHH). The on-rate and off-rate constants and the equilibrium dissociation constants were determined for NAD(+)/NADH analogues and suggested that NADH analogues were the most promising for selective inhibition of Tc-SAHH. None significantly inhibited Hs-SAHH while S-NADH and H-NADH (see Figure 1) reduced the catalytic activity of Tc-SAHH to < 10% in six minutes of exposure.
- Published
- 2009
- Full Text
- View/download PDF
7. Comparative kinetics of cofactor association and dissociation for the human and trypanosomal S-adenosylhomocysteine hydrolases. 2. The role of helix 18 stability.
- Author
-
Li QS, Cai S, Fang J, Borchardt RT, Kuczera K, Middaugh CR, and Schowen RL
- Subjects
- Adenosylhomocysteinase genetics, Animals, Calorimetry, Differential Scanning, Catalysis, Circular Dichroism, Enzyme Stability, Humans, Kinetics, Leishmania donovani enzymology, Mutation, Plasmodium falciparum enzymology, Protein Structure, Secondary, Temperature, Adenosylhomocysteinase chemistry, Adenosylhomocysteinase metabolism, NAD metabolism, Trypanosoma cruzi enzymology
- Abstract
The S-adenosyl- l-homocysteine (AdoHcy) hydrolases (SAHH) from Homo sapiens (Hs-SAHH) and from the parasite Trypanosoma cruzi (Tc-SAHH) are very similar in structure and catalytic properties but differ in the kinetics and thermodynamics of association and dissociation of the cofactor NAD (+). The binding of NAD (+) and NADH in SAHH appears structurally to be mediated by helix 18, formed by seven residues near the C-terminus of the adjacent subunit. Helix-propensity estimates indicate decreasing stability of helix 18 in the order Hs-SAHH > Tc-SAHH > Ld-SAHH (from Leishmania donovani) > Pf-SAHH (from Plasmodium falciparum), which would be consistent with the previous observations. Here we report the properties of Hs-18Pf-SAHH, the human enzyme with plasmodial helix 18, and Tc-18Hs-SAHH, the trypanosomal enzyme with human helix 18. Hs-18Tc-SAHH, the human enzyme with trypanosomal helix 18, was also prepared but differed insignificantly from Hs-SAHH. Association of NAD (+) with Hs-SAHH, Hs-18Pf-SAHH, Tc-18Hs-SAHH, and Tc-SAHH exhibited biphasic kinetics for all enzymes. A thermal maximum in rate, attributed to the onset of local structural alterations in or near the binding site, occurred at 35, 33, 30, and 15 degrees C, respectively. This order is consistent with some reversible changes within helix 18 but does require influence of other properties of the "host enzyme". Dissociation of NAD (+) from the same series of enzymes also exhibited biphasic kinetics with a transition to faster rates (a larger entropy of activation more than compensates for a larger enthalpy of activation) at temperatures of 41, 38, 36, and 29 degrees C, respectively. This order is also consistent with changes in helix 18 but again requiring influence of other properties of the "host enzyme". Global unfolding of all fully reconstituted holoenzymes occurred around 63 degrees C, confirming that the kinetic transition temperatures did not arise from a major disruption of the protein structure.
- Published
- 2008
- Full Text
- View/download PDF
8. Molecular dynamics simulations of domain motions of substrate-free S-adenosyl- L-homocysteine hydrolase in solution.
- Author
-
Hu C, Fang J, Borchardt RT, Schowen RL, and Kuczera K
- Subjects
- Animals, Computer Simulation, Dimerization, Humans, Motion, Protein Structure, Tertiary, Protein Subunits, Solutions, Adenosylhomocysteinase chemistry, Models, Molecular
- Abstract
S-Adenosyl-L-homocysteine hydrolase (SAHH) is an enzyme regulating intracellular methylation reactions. The homotetrameric SAHH exists in an open conformation in absence of substrate, while enzyme:inhibitor complexes crystallize in the closed conformation, in which the ligands are engulfed by the protein due to an 18 degrees domain reorientation within each of the four subunits. We present a microscopic description of the structure and dynamics of the substrate-free, NAD(+)-bound SAHH in solution, based on a 15-ns molecular dynamics simulation in explicit solvent. In the trajectory, the four cofactor-binding domains formed a relatively rigid core with structure very similar to the crystal conformation. The four substrate-binding domains, located at the protein exterior, also retained internal structures similar to the crystal, while undergoing large amplitude rigid-body reorientations. The trajectory domain motions exhibited two interesting properties. First, within each subunit the domains fluctuated between open and closed conformations, while at the tetramer level 80% of the domain motions were perpendicular to the direction of the open-to-closed structural transition. Second, the domain reorientations in solution could be represented as a sum of two components, faster, with 20-50 ps correlation time and 3-4 degrees amplitude, and slower, with 8-23 ns correlation time and amplitude of 14-22 degrees . The faster motion is similar to the 1.5 cm(-1) frequency hinge-bending vibrations found in our recent normal mode analysis (Wang et al., Biochemistry 2005;44:7228-7239). The slower motion agrees with fluorescence anisotropy decay measurements, which detected a 10-20 ns domain reorientation of ca. 26 degrees amplitude in the substrate-free enzyme (Wang et al., Biochemistry 2006;45:7778-7786). Our simulations are thus in excellent agreement with experimental data. The simulations allow us to assign the observed nanosecond fluorescence anisotropy signal to fluctuations in domain orientations, and indicate that the microscopic mechanism of the motion involves rotational diffusion within a cone of 10-20 degrees . Overall, our simulation results complement the existing experimental data and provide important new insights into SAHH domain motions in solution, which play a crucial role in the catalytic mechanism of SAHH., ((c) 2007 Wiley-Liss, Inc.)
- Published
- 2008
- Full Text
- View/download PDF
9. The photo-Favorskii reaction of p-hydroxyphenacyl compounds is initiated by water-assisted, adiabatic extrusion of a triplet biradical.
- Author
-
Givens RS, Heger D, Hellrung B, Kamdzhilov Y, Mac M, Conrad PG 2nd, Cope E, Lee JI, Mata-Segreda JF, Schowen RL, and Wirz J
- Subjects
- Benzyl Alcohols chemical synthesis, Benzyl Alcohols chemistry, Free Radicals chemistry, Free Radicals radiation effects, Molecular Structure, Phenylacetates chemical synthesis, Phenylacetates radiation effects, Photochemistry, Quantum Theory, Ultraviolet Rays, Water chemistry, Phenylacetates chemistry
- Published
- 2008
- Full Text
- View/download PDF
10. The antiviral drug ribavirin is a selective inhibitor of S-adenosyl-L-homocysteine hydrolase from Trypanosoma cruzi.
- Author
-
Cai S, Li QS, Borchardt RT, Kuczera K, and Schowen RL
- Subjects
- Adenosylhomocysteinase biosynthesis, Adenosylhomocysteinase isolation & purification, Animals, Antiviral Agents chemical synthesis, Antiviral Agents chemistry, Binding Sites, Chromatography, High Pressure Liquid methods, Drug Design, Enzyme Activation drug effects, Enzyme Inhibitors chemical synthesis, Enzyme Inhibitors chemistry, Humans, Kinetics, Molecular Conformation, NAD chemistry, NAD drug effects, Recombinant Proteins antagonists & inhibitors, Recombinant Proteins biosynthesis, Recombinant Proteins isolation & purification, Ribavirin chemical synthesis, Ribavirin chemistry, Spectrometry, Mass, Electrospray Ionization methods, Structure-Activity Relationship, Time Factors, Adenosylhomocysteinase antagonists & inhibitors, Antiviral Agents pharmacology, Enzyme Inhibitors pharmacology, Ribavirin pharmacology, Trypanosoma cruzi enzymology
- Abstract
Ribavirin (1,2,4-triazole-3-carboxamide riboside) is a well-known antiviral drug. Ribavirin has also been reported to inhibit human S-adenosyl-L-homocysteine hydrolase (Hs-SAHH), which catalyzes the conversion of S-adenosyl-L-homocysteine to adenosine and homocysteine. We now report that ribavirin, which is structurally similar to adenosine, produces time-dependent inactivation of Hs-SAHH and Trypanosoma cruzi SAHH (Tc-SAHH). Ribavirin binds to the adenosine-binding site of the two SAHHs and reduces the NAD(+) cofactor to NADH. The reversible binding step of ribavirin to Hs-SAHH and Tc-SAHH has similar K(I) values (266 and 194 microM), but the slow inactivation step is 5-fold faster with Tc-SAHH. Ribavirin may provide a structural lead for design of more selective inhibitors of Tc-SAHH as potential anti-parasitic drugs.
- Published
- 2007
- Full Text
- View/download PDF
11. Trehalose and calcium exert site-specific effects on calmodulin conformation in amorphous solids.
- Author
-
Li Y, Williams TD, Schowen RL, and Topp EM
- Subjects
- Binding Sites, Phase Transition, Protein Binding, Protein Conformation, Calcium metabolism, Calmodulin metabolism, Deuterium Exchange Measurement methods, Drug Compounding methods, Trehalose chemistry
- Abstract
We have adapted hydrogen/deuterium (H/D) exchange with electrospray ionization mass spectrometry (ESI-MS) to study protein conformation and excipient interactions in lyophilized solids. Using calmodulin (CaM, 17 kD) as a model protein, we demonstrate that trehalose and calcium exert site-specific effects on protein conformation. The effects of calcium are observed primarily in the calcium binding loops, while those of trehalose are observed primarily in non-terminal alpha-helical regions. To our knowledge, this is the first demonstration of site-specificity in the effects of excipients on protein structure in the solid state, and of the utility of H/D exchange with ESI-MS to characterize proteins in amorphous solids., ((c) 2007 Wiley Periodicals, Inc.)
- Published
- 2007
- Full Text
- View/download PDF
12. Characterizing protein structure in amorphous solids using hydrogen/deuterium exchange with mass spectrometry.
- Author
-
Li Y, Williams TD, Schowen RL, and Topp EM
- Subjects
- Amino Acid Sequence, Calmodulin chemistry, Calmodulin genetics, Calorimetry, Differential Scanning, Chromatography, High Pressure Liquid, Deuterium chemistry, Freeze Drying, Hydrogen chemistry, Molecular Sequence Data, Molecular Structure, Peptide Fragments chemistry, Peptide Fragments genetics, Powder Diffraction, Protein Structure, Secondary, Recombinant Proteins chemistry, Recombinant Proteins genetics, Spectrometry, Mass, Electrospray Ionization, Spectroscopy, Fourier Transform Infrared, Proteins chemistry
- Abstract
Elucidating protein structure in amorphous solids is central to the rational design of stable lyophilized protein drugs. Hydrogen/deuterium (H/D) exchange with electrospray ionization mass spectrometry was applied to lyophilized powders containing calmodulin (17 kDa) and exposed to D(2)O vapor at controlled relative humidity (RH) and temperature. H/D exchange was influenced by RH and by the inclusion of calcium chloride and/or trehalose in the solid. The effects were not exhibited uniformly along the protein backbone but occurred in a site-specific manner, with calcium primarily influencing the calcium-binding loops and trehalose primarily influencing the alpha-helices. The results demonstrate that the method can provide quantitative and site-specific structural information on proteins in amorphous solids and on changes in structure induced by protein cofactors and formulation excipients. Such information is not readily available with other techniques used to characterize proteins in the solid state, such as Fourier transform infrared, Raman, and near-infrared spectroscopy.
- Published
- 2007
- Full Text
- View/download PDF
13. Comparative kinetics of cofactor association and dissociation for the human and trypanosomal S-adenosylhomocysteine hydrolases. 1. Basic features of the association and dissociation processes.
- Author
-
Li QS, Cai S, Borchardt RT, Fang J, Kuczera K, Middaugh CR, and Schowen RL
- Subjects
- Adenosylhomocysteinase antagonists & inhibitors, Animals, Binding Sites, Kinetics, Temperature, Adenosylhomocysteinase metabolism, NAD metabolism, Trypanosoma cruzi enzymology
- Abstract
The S-adenosyl-l-homocysteine (AdoHcy) hydrolases catalyze the reversible conversion of AdoHcy to adenosine and homocysteine, making use of a catalytic cycle in which a tightly bound NAD+ oxidizes the 3-hydroxyl group of the substrate at the beginning of the cycle, activating the 4-CH bond for elimination of homocysteine, followed by Michael addition of water to the resulting intermediate and a final reduction by the tightly bound NADH to give adenosine. The equilibrium and kinetic properties of the association and dissociation of the cofactor NAD+ from the enzymes of Homo sapiens (Hs-SAHH) and Trypanosoma cruzi (Tc-SAHH) are qualitatively similar but quantitatively distinct. Both enzymes bind NAD+ in a complex scheme. The four active sites of the homotetrameric apoenzyme appear to divide into two numerically equal classes of active sites. One class of sites binds cofactor weakly and generates full activity very rapidly (in less than 1 min). The other class binds cofactor more strongly but generates activity only slowly (>30 min). In the case of Tc-SAHH, the final affinity for NAD+ is roughly micromolar and this affinity persists as the equilibrium affinity. In the case of Hs-SAHH, the slow-binding phase terminates in micromolar affinity also, but over a period of hours, the dissociation rate constant decreases until the final equilibrium affinity is in the nanomolar range. The slow binding of NAD+ by both enzymes exhibits saturation kinetics with respect to the cofactor concentration; however, binding to Hs-SAHH has a maximum rate constant around 0.06 s-1, while the rate constant for binding to Tc-SAHH levels out at 0.006 s-1. In contrast to the complex kinetics of association, both enzymes undergo dissociation of NAD+ from all four sites in a single first-order reaction. The equilibrium affinities of both Hs-SAHH and Tc-SAHH for NADH are in the nanomolar range. The dissociation rate constants and the slow-binding association rate constants for NAD+ show a complex temperature dependence with both enzymes; however, the cofactor always dissociates more rapidly from Tc-SAHH than from Hs-SAHH, the ratio being around 80-fold at 37 degrees C, and the cofactor binds more rapidly to Hs-SAHH than to Tc-SAHH above approximately 16 degrees C. These features present an opening for selective inhibition of Tc-SAHH over Hs-SAHH, demonstrated with the thioamide analogues of NAD+ and NADH. Both analogues bind to Hs-SAHH with approximately 40 nM affinities but much more weakly to Tc-SAHH (0.6-15 microM). Nevertheless, both analogues inactivated Tc-SAHH 60% (NAD+ analogue) or 100% (NADH analogue) within 30 min, while the degree of inhibition of Hs-SAHH approached 30% only after 12 h. The rate of loss of activity is equal to the rate of dissociation of the cofactor and thus 80-fold faster at 37 degrees C for Tc-SAHH.
- Published
- 2007
- Full Text
- View/download PDF
14. Isotopic and other studies on the molecular origins of substrate regulation of some pyruvate decarboxylases: a reconsideration.
- Author
-
Schowen RL
- Subjects
- Cysteine chemistry, Decarboxylation, Enzyme Activation, Isotopes, Kinetics, Models, Molecular, Pyruvate Decarboxylase metabolism, Pyruvates chemistry, Pyruvates metabolism, Saccharomyces cerevisiae Proteins metabolism, Models, Chemical, Pyruvate Decarboxylase chemistry, Saccharomyces cerevisiae Proteins chemistry
- Abstract
Solvent isotope effect and beta-secondary isotope effect studies of the steady-state reaction of the pyruvate decarboxylase of Saccharomyces cerevisiae (ScPDC), which gains its full activity in a process requiring some seconds, suggested a model that ascribed the isotope effects to a reversible carbonyl-addition reaction between the 'regulatory pyruvate' molecule and the sulfhydryl group of Cys 221, occurring at the beginning and end of each catalytic cycle. The slow ('hysteretic') regulatory activation was thought to consist only of binding the regulatory molecule properly. Since the proposal of this model, much important progress in understanding both the structural and dynamic chemistry of ScPDC has been made in several laboratories. The purpose of this article is to review this new information and to evaluate the above model and others in the light of currently available evidence.
- Published
- 2007
- Full Text
- View/download PDF
15. Effects of ligand binding and oxidation on hinge-bending motions in S-adenosyl-L-homocysteine hydrolase.
- Author
-
Wang M, Unruh JR, Johnson CK, Kuczera K, Schowen RL, and Borchardt RT
- Subjects
- Adenosylhomocysteinase genetics, Amino Acid Sequence, Amino Acid Substitution, Fluorescence Polarization, Humans, Ligands, Maleimides chemistry, Oxidation-Reduction, Protein Structure, Quaternary, Protein Structure, Tertiary, Adenosylhomocysteinase chemistry
- Abstract
Domain motions of S-adenosyl-l-homocysteine (AdoHcy) hydrolase have been detected by time-resolved fluorescence anisotropy measurements. Time constants for reorientational motions in the native enzyme were compared with those for enzymes where key residues were altered by site-directed mutation. Mutations M351P, H353A, and P354A were selected in a hinge region for motion between the open and closed forms of the enzyme, as identified in a previous normal-mode study [Wang et al. (2005) Domain motions and the open-to-closed conformational transition of an enzyme: A normal-mode analysis of S-adenosyl-l-homocysteine hydrolase, Biochemistry 44, 7228-7239]. In wild-type, substrate-free AdoHcy hydrolase (NAD(+) cofactor in each subunit), reorientational motions were detected on time scales of 10-20 and 80-90 ns. The faster motion is attributed to the domain motion, and the slower motion is attributed to the tumbling of the enzyme. The domain motion was also detected for the enzyme complexes E(NADH/3'-keto-adenosine) and E(NAD(+)/3'-deoxyadenosine) but was absent for the complex E(NADH/3'-keto-neplanocin A). The results indicate that AdoHcy hydrolase exists in equilibrium of open and closed structures, with the equilibrium shifted toward the more mobile open form for the substrate-free enzyme, E(NAD(+)), and for intermediates formed early in the catalytic cycle after substrate binding or formed late prior to product release, E(NAD(+)/ligand). However, the strong inhibitor neplanocin A upon binding undergoes oxidation, forming the complex E(NADH/3'-keto-neplanocin). For this complex, which is analogous to the enzyme complex with the central catalytic intermediate, the equilibrium was shifted toward the more rigid closed form. A similar pattern was observed for M351P and P354A mutants. In contrast, the domain motion could not be detected, either in the absence or presence of ligands or with the cofactor in either the oxidized or reduced state, for the H353A protein, suggesting that this mutation changes the hinge-bending dynamics of the enzyme.
- Published
- 2006
- Full Text
- View/download PDF
16. Effect of N-1 and N-2 residues on peptide deamidation rate in solution and solid state.
- Author
-
Li B, Schowen RL, Topp EM, and Borchardt RT
- Subjects
- Peptides chemistry, Peptides pharmacokinetics, Solubility, Solutions, Amides metabolism, Asparagine chemistry, Models, Chemical, Peptide Fragments chemistry, Peptide Fragments pharmacokinetics
- Abstract
The deamidation kinetics of 7 model peptides (VYPNGA, VYGNGA, VFGNGA, VIGNGA, VGGNGA, VGPNGA, and VGYNGA) were studied at 70 degrees C in pH 10 buffer solutions and at 70 degrees C and 50% relative humidity in lyophilized solid formulations containing polyvinyl pyrrolidone (PVP). The disappearance of the model peptides from solution and solid-state formulations followed apparent first-order kinetics, proceeding to completion in solution. In the solid state, the reactions showed plateaus with approximately 10% to 30% of the model peptides remaining; this was thought to be due to reversible complexation of the peptides and the PVP followed by slow dissociation of the complexes. The residues immediately N-terminal to asparagine (N-1, N-2) influenced the rate of deamidation significantly in the solid state but had minimal effect in solution. Increases in the volume and hydrophobicity of the N-1 and N-2 residues decreased the rate of deamidation in the solid state, but neither parameter alone adequately accounted for the observed effects. An empirical model using a linear combination of volume and hydrophobicity was developed; it showed that the influences of the volume and the hydrophobicity of the residues in the N-1 and N-2 positions are approximately equally important for the N-1 and N-2 residues.
- Published
- 2006
- Full Text
- View/download PDF
17. Effects of sucrose and mannitol on asparagine deamidation rates of model peptides in solution and in the solid state.
- Author
-
Li B, O'Meara MH, Lubach JW, Schowen RL, Topp EM, Munson EJ, and Borchardt RT
- Subjects
- Crystallization, Freeze Drying, Mannitol pharmacology, Models, Chemical, Protein Conformation, Protein Denaturation drug effects, Solutions, Sucrose pharmacology, Temperature, Time Factors, Water chemistry, Asparagine chemistry, Mannitol chemistry, Peptides chemistry, Sucrose chemistry
- Abstract
Asparagine (Asn) degradation kinetics in two model peptides, Gly-Gln-Asn-Gly-Gly (GQNGG) and Val-Tyr-Pro-Asn-Gly-Ala (VYPNGA), were studied at 50 degrees C in pH 7 buffer solutions in the presence and absence of 5% (w/v) sucrose or mannitol and at 50 degrees C and 30% relative humidity in solid samples lyophilized from these solutions. Solid formulations were characterized using Karl Fischer coulometric titration, thermal gravimetric analysis (TGA), differential scanning calorimetry (DSC), Fourier-transform infrared spectrometry (FTIR), and solid-state nuclear magnetic resonance (NMR) spectroscopy. GQNGG and VYPNGA showed similar pseudo first-order deamidation rates in solution in the absence of sucrose and mannitol. Adding 5% sucrose or mannitol decreased the rates by no more than 17%. The model peptides degraded 2- to 80-fold more slowly in the solid formulations of sucrose and mannitol than in 5% solutions of these carbohydrates. Ratios of deamidation rates of the model peptides depended upon the solid matrix. In the mannitol solid, the ratio of deamidation rates of GQNGG and VYPNGA (GQNGG:VYPNGA) was approximately 8, while in the sucrose solid, the model peptides deamidated at similar rates (GQNGG:VYPNGA congruent with 1). DSC showed the mannitol formulations to be largely amorphous immediately after lyophilization with some ordered, crystalline-like structure; the extent of ordered structure increased during storage as shown by FTIR and ssNMR. In contrast, the sucrose formulation was largely amorphous after lyophilization and remained so during storage. Together, the results showed that 5% sucrose or mannitol in solution does not significantly change the rates of Asn deamidation of the model peptides, while sucrose stabilizes the model peptides against deamidation more than mannitol in the solid state., ((c) 2005 Wiley-Liss, Inc. and the American Pharmacists Association)
- Published
- 2005
- Full Text
- View/download PDF
18. Domain motions and the open-to-closed conformational transition of an enzyme: a normal mode analysis of S-adenosyl-L-homocysteine hydrolase.
- Author
-
Wang M, Borchardt RT, Schowen RL, and Kuczera K
- Subjects
- Adenine chemistry, Adenosine chemistry, Adenosylhomocysteinase antagonists & inhibitors, Animals, Binding Sites, Catalysis, Crystallography, X-Ray, Dimerization, Enzyme Inhibitors chemistry, Humans, Models, Chemical, Models, Molecular, Protein Conformation, Protein Structure, Tertiary, Rats, Substrate Specificity, Adenine analogs & derivatives, Adenosine analogs & derivatives, Adenosylhomocysteinase chemistry, Adenosylhomocysteinase metabolism, Thermodynamics
- Abstract
The structure and fluctuations of the enzyme S-adenosyl-L-homocysteine hydrolase (SAHH) are analyzed in an effort to explain its biological function. Besides the previously identified open structure, characteristic of the substrate-free enzyme, we find two distinct structures in enzyme-inhibitor complexes, the closed and closed-twisted conformers. Both closed conformers differ from the open form by a hinge bending motion of two large domains within each subunit, which isolate the inhibitor bound in the active site from the bulk solvent. The closed-twisted form further differs from the closed form by a rigid body twist of the two-subunit dimers. The local structural fluctuations of SAHH are analyzed by performing block normal mode analysis of the tetrameric enzyme in its three forms. For the open form, we find that the four lowest-frequency normal modes, corresponding to the collective motions of the protein with the largest amplitudes, are essentially combinations of the hinge bending deformations of the individual subunits. Thus, the mechanical properties of the open structure of SAHH lead to the presence of structural fluctuations in the direction of the open-to-closed conformational transition. A candidate for such a motion has been observed in previous fluorescence depolarization studies of the enzyme. Both structural and normal mode analyses indicate that residues 180-190 and 350-356 form hinge regions, connecting large domains which tend to move as rigid bodies in response to interactions with substrate, intermediates, and the product of the enzymatic reactions. We propose that these hinge regions play a crucial role in the enzymatic mechanism of SAHH. In contrast to the open form, normal mode calculations for the closed conformations show strong coupling of the hinge bending motions of the individual subunits to each other and to other low-frequency vibrations. Thus, information about structural changes related to reaction progress in one active site may be mechanically transmitted to other subunits of the protein, explaining the cooperativity found in the enzyme kinetics.
- Published
- 2005
- Full Text
- View/download PDF
19. Effects of acidic N + 1 residues on asparagine deamidation rates in solution and in the solid state.
- Author
-
Li B, Gorman EM, Moore KD, Williams T, Schowen RL, Topp EM, and Borchardt RT
- Subjects
- Asparagine analysis, Aspartic Acid chemistry, Drug Stability, Hydrogen-Ion Concentration, Oligopeptides analysis, Oligopeptides chemistry, Pharmaceutical Solutions analysis, Pharmaceutical Solutions chemistry, Amides chemistry, Asparagine chemistry
- Abstract
The deamidation kinetics of four model peptides (AcGQNGG, AcGQNDG, AcGQNEG, and AcGQNQG) were studied in solution (70 degrees C, pH 5-10) and in lyophilized solids [70 degrees C, 50% relative humidity, "effective pH" ('pH') 5-10] containing polyvinyl pyrrolidone. AcGQNGG, AcGQNEG, and AcGQNQG degraded exclusively through Asn deamidation, whereas AcGQNDG also displayed Asp isomerization, and Asp-Gly peptide bond cleavage. The pH/'pH'-rate profiles were consistent with a shift in the rate-determining step of Asn deamidation from carbonyl addition to expulsion of ammonia with increasing pH. In solution, AcGQNGG deamidated up to 38-fold faster than the other peptides, indicating the importance of steric effects of the N + 1 residue. AcGQNGG and AcGQNQG had up to 60 times slower rates of deamidation in the solid state than in solution. In contrast, the deamidation rates of AcGQNEG and AcGQNDG in the solid state were similar to those in solution. N + 1 Glu or Asp residue may enhance local hydration, so that the deamidation of Asn in the solid formulations actually proceeds in a solution-like environment., (Copyright 2005 Wiley-Liss, Inc. and the American Pharmacists Association.)
- Published
- 2005
- Full Text
- View/download PDF
20. The crystal structures of Klebsiella pneumoniae acetolactate synthase with enzyme-bound cofactor and with an unusual intermediate.
- Author
-
Pang SS, Duggleby RG, Schowen RL, and Guddat LW
- Subjects
- Amino Acid Sequence, Binding Sites, Crystallization, Flavin-Adenine Dinucleotide chemistry, Models, Molecular, Molecular Sequence Data, Protein Structure, Secondary, Thiamine Pyrophosphate physiology, Acetolactate Synthase chemistry, Klebsiella pneumoniae enzymology
- Abstract
Acetohydroxyacid synthase (AHAS) and acetolactate synthase (ALS) are thiamine diphosphate (ThDP)-dependent enzymes that catalyze the decarboxylation of pyruvate to give a cofactor-bound hydroxyethyl group, which is transferred to a second molecule of pyruvate to give 2-acetolactate. AHAS is found in plants, fungi, and bacteria, is involved in the biosynthesis of the branched-chain amino acids, and contains non-catalytic FAD. ALS is found only in some bacteria, is a catabolic enzyme required for the butanediol fermentation, and does not contain FAD. Here we report the 2.3-A crystal structure of Klebsiella pneumoniae ALS. The overall structure is similar to AHAS except for a groove that accommodates FAD in AHAS, which is filled with amino acid side chains in ALS. The ThDP cofactor has an unusual conformation that is unprecedented among the 26 known three-dimensional structures of nine ThDP-dependent enzymes, including AHAS. This conformation suggests a novel mechanism for ALS. A second structure, at 2.0 A, is described in which the enzyme is trapped halfway through the catalytic cycle so that it contains the hydroxyethyl intermediate bound to ThDP. The cofactor has a tricyclic structure that has not been observed previously in any ThDP-dependent enzyme, although similar structures are well known for free thiamine. This structure is consistent with our proposed mechanism and probably results from an intramolecular proton transfer within a tricyclic carbanion that is the true reaction intermediate. Modeling of the second molecule of pyruvate into the active site of the enzyme with the bound intermediate is consistent with the stereochemistry and specificity of ALS.
- Published
- 2004
- Full Text
- View/download PDF
21. Polyvinylpyrrolidone-drug conjugate: synthesis and release mechanism.
- Author
-
D'Souza AJ, Schowen RL, and Topp EM
- Subjects
- Chemistry, Pharmaceutical, Povidone chemical synthesis, Povidone pharmacokinetics
- Abstract
Covalent conjugates of polyvinylpyrrolidone (PVP) with para-nitroaniline (PNA) were synthesized as a model PVP-drug conjugate, and PNA release was evaluated in vitro. Pyrrolidone ring opening with subsequent t-BOC protection of the pyrrolidone nitrogen and reaction with PNA in methylene chloride (CH2Cl2) produced a PVP-PNA conjugate with 3% of the pyrrolidone groups modified. Rates of PNA release from N-deprotected conjugates were twofold greater than those that were N-protected, indicating participation of the pyrrolidone N in release. Additional studies with monomeric analogs supported intramolecular base catalysis rather than lactam formation as the mechanism of this involvement. The approach serves as a prototype for the conjugation of other drugs with primary and secondary amine functional groups with PVP, including peptides and proteins.
- Published
- 2004
- Full Text
- View/download PDF
22. How an enzyme surmounts the activation energy barrier.
- Author
-
Schowen RL
- Subjects
- Catalysis, Chorismate Mutase chemistry, Chorismic Acid chemistry, Chorismic Acid metabolism, Cyclohexanecarboxylic Acids chemistry, Cyclohexanecarboxylic Acids metabolism, Cyclohexenes, Enzyme Activation, Hydrophobic and Hydrophilic Interactions, Molecular Conformation, Static Electricity, Thermodynamics, Chorismate Mutase metabolism
- Published
- 2003
- Full Text
- View/download PDF
23. Trypanosoma cruzi: molecular cloning and characterization of the S-adenosylhomocysteine hydrolase.
- Author
-
Parker NB, Yang X, Hanke J, Mason KA, Schowen RL, Borchardt RT, and Yin DH
- Subjects
- Adenosylhomocysteinase chemistry, Adenosylhomocysteinase metabolism, Amino Acid Sequence, Animals, Base Sequence, Chromatography, DEAE-Cellulose, Chromatography, Gel, Chromatography, Ion Exchange, Consensus Sequence, Conserved Sequence, Cosmids chemistry, Cosmids genetics, Electrophoresis, Polyacrylamide Gel, Gene Expression Regulation, Enzymologic, Gene Library, Humans, Leishmania donovani enzymology, Leishmania donovani genetics, Molecular Sequence Data, Molecular Weight, NAD metabolism, Rats, Sequence Alignment, Sequence Homology, Nucleic Acid, Trypanosoma cruzi genetics, Adenosylhomocysteinase genetics, Cloning, Molecular, Trypanosoma cruzi enzymology
- Abstract
S-Adenosylhomocysteine (AdoHcy) hydrolase has emerged as an attractive target for antiparasitic drug design because of its role in the regulation of all S-adenosylmethionine-dependent transmethylation reactions, including those reactions crucial for parasite replication. From a genomic DNA library of Trypanosoma cruzi, we have isolated a gene that encodes a polypeptide containing a highly conserved AdoHcy hydrolase consensus sequence. The recombinant T. cruzi enzyme was overexpressed in Escherichia coli and purified as a homotetramer. At pH 7.2 and 37 degrees C, the purified enzyme hydrolyzes AdoHcy to adenosine and homocysteine with a first-order rate constant of 1 s(-1) and synthesizes AdoHcy from adenosine and homocysteine with a pseudo-first-order rate constant of 3 s(-1) in the presence of 1 mM homocysteine. The reversible catalysis depends on the binding of NAD(+) to the enzyme. In spite of the significant structural homology between the parasitic and human AdoHcy hydrolase, the K(d) of 1.3 microM for NAD(+) binding to the T. cruzi enzyme is approximately 11-fold higher than the K(d) (0.12 microM) for NAD(+) binding to the human enzyme.
- Published
- 2003
- Full Text
- View/download PDF
24. Racemization of an asparagine residue during peptide deamidation.
- Author
-
Li B, Borchardt RT, Topp EM, VanderVelde D, and Schowen RL
- Subjects
- Models, Chemical, Nuclear Magnetic Resonance, Biomolecular, Stereoisomerism, Succinimides chemistry, Asparagine chemistry, Oligopeptides chemistry
- Abstract
The Asn residue in the pentapeptide Asn-Gln-Asn-Glu-Gly undergoes racemization at the Calpha center in the course of deamidation of this residue through a succinimide intermediate. The succinimide intermediate is known to racemize at the corresponding center, leading to racemized products of deamidation. Return of this intermediate to reactant Asn is very unlikely in dilute solution where attack of product ammonia on the succinimide is precluded, and Asn racemization has not been previously observed. We give evidence that the observed racemization occurs at the tetrahedral-intermediate stage preceding the succinimide intermediate. The observation is significant for protein stability in vivo and in vitro and has importance in medicine, food chemistry, and archaeological/palaeological dating.
- Published
- 2003
- Full Text
- View/download PDF
25. A mechanistic and kinetic study of the E-ring hydrolysis and lactonization of a novel phosphoryloxymethyl prodrug of camptothecin.
- Author
-
Hanson BA, Schowen RL, and Stella VJ
- Subjects
- Camptothecin chemistry, Hydrolysis, Lactones chemistry, Mechanics, Prodrugs chemistry, Camptothecin pharmacokinetics, Lactones pharmacokinetics, Prodrugs pharmacokinetics
- Abstract
Purpose: This study was done to determine the E-ring hydrolysis and lactonization mechanism of a water-soluble 20-phosphoryloxymethyl (POM) prodrug of camptothecin (P-CPT). Specifically, the role of the phosphate group in facilitating E-ring hydrolysis was examined., Methods: Resolution between the lactone and carboxylate forms of P-CPT and camptothecin (CPT) was achieved with a RPHPLC assay using UV-visible detection. E-ring P-CPT hydrolysis and lactonization kinetics were followed using 20 mM acetate or phosphate buffer (micro = 0.15 NaCl) over the pH range of 4 to 8 at 25.0 degrees C. A kinetic solvent isotope effect (KSIE) study was used to further probe the mechanism of E-ring hydrolysis., Results: The hydrolysis and lactonization reactions followed pseudo-first-order kinetics in the approach to equilibrium. The equilibrium ratio of the open and closed forms of P-CPT was dependent on pH, with the closed form dominant at low pH and the open form dominant at high pH. Buffer concentration changes had little to no effect on the rate of P-CPT E-ring hydrolysis. The KSIE study provided an overall isotope effect of 2.47 and a proton inventory KSIE consistent with an intramolecular general base catalysis., Conclusions: P-CPT has a pH-dependent equilibrium between the lactone and carboxylate forms similar but not identical to that of CPT. The results suggest a hydrolysis reaction mechanism that involves a single site hydrogen exchange facilitated intramolecularly by the dianionic phosphate moiety of P-CPT via either general base catalysis of the lactone ring attack by water or breakdown of the tetrahedral intermediate.
- Published
- 2003
- Full Text
- View/download PDF
26. Asparagine deamidation in recombinant human lymphotoxin: hindrance by three-dimensional structures.
- Author
-
Xie M, Shahrokh Z, Kadkhodayan M, Henzel WJ, Powell MF, Borchardt RT, and Schowen RL
- Subjects
- Circular Dichroism, Crystallization, Drug Stability, Humans, Hydrogen-Ion Concentration, Kinetics, Mass Spectrometry, Models, Molecular, Peptide Mapping, Protein Conformation, Protein Denaturation, Spectrometry, Fluorescence, Time Factors, Asparagine chemistry, Lymphotoxin-alpha chemistry
- Abstract
The chemical stability of recombinant human lymphotoxin (rhLT) was evaluated at pH 7, 9, and 11 and 40 degrees C using quantitative tryptic map and urea-IEF methods. Degradation products were characterized by mass spectrometry. The stability of denatured rhLT protein was also evaluated to elucidate the effects of three-dimensional structures on Asn deamidation in rhLT. Two sites that underwent Asn deamidation were identified in rhLT, Asn(19) and Asn(40)-Asn(41). At pH 11 and 40 degrees C, deamidation at Asn(19) and Asn(40)-Asn(41) had half-lives of 14 +/- 4 and 80 +/- 24 days, respectively. Upon denaturation, 31- and ninefold acceleration in the degradation rates was observed at the Asn(19) and Asn(40)-Asn(41) sites, respectively. The rate of Asn(19) degradation in denatured rhLT was comparable to that of the model peptide possessing the same primary sequence as the Asn(19)-containing region in rhLT. Analysis of the rhLT crystal structure revealed that both Asn deamidation sites were located in beta-turn structures with extensive hydrogen-bonding networks created with nearby residues in the tertiary structures. The results suggested that these tertiary and secondary structures, if held true in solution, were probably responsible for the stabilization of Asn in the native rhLT protein by reducing flexibility, thus preventing adoption of the favorable conformation required for cyclic-imide formation., (Copyright 2003 Wiley-Liss, Inc. and the American Pharmaceutical Association J Pharm Sci 92:869-880, 2003)
- Published
- 2003
- Full Text
- View/download PDF
27. Reaction of a peptide with polyvinylpyrrolidone in the solid state.
- Author
-
D'Souza AJ, Schowen RL, Borchardt RT, Salsbury JS, Munson EJ, and Topp EM
- Subjects
- Chromatography, High Pressure Liquid methods, Peptides chemistry, Povidone chemistry, Peptides analysis, Peptides metabolism, Povidone analysis, Povidone metabolism
- Abstract
During stability studies at high temperature (70 degrees C) and low relative humidity ( approximately 0%), the recovery of an asparagine containing hexapeptide (VYPNGA) and its known deamidation products from solid polyvinylpyrrolidone (PVP) matrices was incomplete. To determine the causes of this mass loss, formulations were prepared by lyophilizing solutions containing PVP, glycerol, and the Asn-hexapeptide in pH 7.5 phosphate buffer, followed by storage at 70 degrees C and 0% relative humidity. Asn-hexapeptide loss was mono-exponential and reached a plateau at about 30% remaining. Total recovery of the peptide and its known deamidation products was approximately 30% of peptide load. Size exclusion chromatography with fluorescence detection indicated the formation of a PVP-peptide adduct that was stable in the presence of 6 M guanidine hydrochloride. Similar stability studies using N-acetyl phenylalanine, phenylalanine ethyl ester, and N-acetyl tyrosine ethyl ester demonstrated that the reaction involves the peptide N-terminus. The adduct was disrupted in the presence of carboxypeptidase-A, suggesting the formation of an amide bond between the peptide and PVP. (15)N solid-state nuclear magnetic resonance spectroscopy using (15)N-labeled valine as a model of the peptide N-terminus showed different populations of (15)N, suggesting that noncovalent peptide-polymer interactions precede amide bond formation., (Copyright 2003 Wiley-Liss, Inc. and the American Pharmaceutical Association)
- Published
- 2003
- Full Text
- View/download PDF
28. Catalytic strategy of S-adenosyl-L-homocysteine hydrolase: transition-state stabilization and the avoidance of abortive reactions.
- Author
-
Yang X, Hu Y, Yin DH, Turner MA, Wang M, Borchardt RT, Howell PL, Kuczera K, and Schowen RL
- Subjects
- Adenosine chemistry, Adenosylhomocysteinase, Binding Sites, Buffers, Catalysis, Crystallization, Crystallography, X-Ray, Enzyme Inhibitors chemistry, Enzyme Stability, Humans, Hydrolases antagonists & inhibitors, Hydrolysis, NAD chemistry, Oxidation-Reduction, Pregnancy Proteins chemistry, Protein Conformation, Protein Structure, Tertiary, Substrate Specificity, Water chemistry, Adenosine analogs & derivatives, Hydrolases chemistry
- Abstract
S-Adenosylhomocysteine hydrolase (AdoHcy hydrolase) crystallizes from solutions containing the intermediate analogue neplanocin A with the analogue bound in its 3'-keto form at the active sites of all of its four subunits and the four tightly bound cofactors in their reduced (NADH) state. The enzyme is in the closed conformation, which corresponds to the structure in which the catalytic chemistry occurs. Examination of the structure in the light of available, very detailed kinetic studies [Porter, D. J., Boyd, F. L. (1991) J. Biol. Chem. 266, 21616-21625. Porter, D. J., Boyd, F. L. (1992) J. Biol. Chem. 267, 3205-3213. Porter, D. J. (1998) J. Biol. Chem. 268, 66-73] suggests elements of the catalytic strategy of AdoHcy hydrolase for acceleration of the reversible conversion of AdoHcy to adenosine (Ado) and homocysteine (Hcy). The enzyme, each subunit of which possesses a substrate-binding domain that in the absence of substrate is in rapid motion relative to the tetrameric core of the enzyme, first binds substrate and ceases motion. Probably concurrently with oxidation of the substrate to its 3'-keto form, the closed active site is "sealed off" from the environment, as indicated by a large (10(8)(-)(9)-fold) reduction in the rate of departure of ligands, a feature that prevents exposure of the labile 3'-keto intermediates to the aqueous environment. Elimination of the 5'-substituent (Hcy in the hydrolytic direction, water in the synthetic direction) generates the central intermediate 4',5'-didehydro-5'-deoxy-3'-ketoadenosine. Abortive 3'-reduction of the central intermediate is prevented by a temporary suspension of all or part of the redox catalytic power of the enzyme during the existence of the central intermediate. The abortive reduction is 10(4)-fold slower than the productive reductions at the ends of the catalytic cycle and has a rate constant similar to those of nonenzymic intramolecular model reactions. The mechanism for suspending the redox catalytic power appears to be a conformationally induced increase in the distance across which hydride transfer must occur between cofactor and substrate, the responsible conformational change again being that which "seals" the active site. The crystal structure reveals a well-defined chain of three water molecules leading from the active site to the subunit surface, which may serve as a relay for proton exchange between solvent and active site in the closed form of the enzyme, permitting maintenance of active-site functional groups in catalytically suitable protonation states.
- Published
- 2003
- Full Text
- View/download PDF
29. The trapping of a spontaneously "flipped-out" base from double helical nucleic acids by host-guest complexation with beta-cyclodextrin: the intrinsic base-flipping rate constant for DNA and RNA.
- Author
-
Spies MA and Schowen RL
- Subjects
- Chemical Phenomena, Chemistry, Physical, Kinetics, Spectrophotometry, Ultraviolet, Cyclodextrins chemistry, DNA chemistry, RNA chemistry, alpha-Cyclodextrins, beta-Cyclodextrins, gamma-Cyclodextrins
- Abstract
Beta-cyclodextrin, which forms stable host-guest complexes with purine bases, induces the melting of RNA and DNA duplexes below their normal melting temperatures. Alpha-cyclodextrin, which does not form stable complexes, has no effect on either RNA or DNA. Gamma-cyclodextrin, which forms weaker complexes, has no effect on RNA and a smaller effect than beta-cyclodextrin on DNA. The rate of melting is kinetically first-order in duplex and, above about 20 mM beta-cyclodextrin, is independent of the beta-cyclodextrin concentration with a first-order rate constant, common to both RNA and DNA, of (3.5 +/- 0.5) x 10(-3) s(-1) at 61 degrees C (DNA) and at 50 degrees C (RNA). This is taken to be the rate constant for spontaneous "flipping out" of a base from within the duplex structure of the nucleic acids, the exposed base being rapidly trapped by beta-cyclodextrin. Like beta-cyclodextrin, nucleic acid methyltransferases bind the target base for methylation in a site that requires it to have flipped out of its normal position in the duplex. The spontaneous flip-out rate constant of around 10(-3) s(-1) is near the value of k(cat) for the methyltransferases (ca. 10(-3) to 10(-1) s(-1)). In principle, the enzymes, therefore, need effect little or no catalysis of the flipping-out reaction. Nevertheless, the flip-out rate in enzyme/DNA complexes is much faster. This observation suggests that the in vivo circumstances may differ from in vitro models or that factors other than a simple drive toward higher catalytic power have been influential in the evolution of these enzymes.
- Published
- 2002
- Full Text
- View/download PDF
30. The elicitation of carboxylesterase activity in antibodies by reactive immunization with labile organophosphorus antigens: a role for flexibility.
- Author
-
Schowen RL
- Subjects
- Binding Sites, Catalysis, Cross Reactions, Immunization, Antibodies, Catalytic metabolism, Carboxylic Ester Hydrolases metabolism, Organophosphorus Compounds immunology
- Abstract
For the creation of powerfully catalytic antibodies, the technique of reactive immunization solves the problem inherent in immunization with transition-state analogs (TSAs), namely, that many interesting target reactions are multistep reactions, with multiple transition states, and thus in general, no single analog can adequately simulate all the transition states along the reaction path. In contrast, immunization with chemically reactive antigens such as phosphonylating agents, which phosphonylate B-cells during the immune response, produces antibodies that have been "trained" to recognize, bind, and stabilize all the actual transition states involved in the phosphonylation reaction. Therefore, catalytic antibodies have been selected by the immune system on the basis of their capacity to stabilize any number of transition states that occur during the target reaction. Somewhat surprisingly, phosphonolysis catalysts generated in this way commonly also catalyze esterolysis reactions. Esterolysis reactions should pass through transition states with a roughly tetrahedral disposition of ligands about a central carbon atom, while phosphonolysis reactions should pass through transition states with a roughly trigonal-bipyramidal disposition of ligands about a central phosphorus atom. These two divergent transition-state geometries suggest that the same active site should have difficulty recognizing and stabilizing both kinds of transition state. The observations thus indicate a puzzling form of "cross-reactivity" toward transition states. A possible explanation arises from evidence that at least some nucleophilic displacements at phosphorus do not pass through a trigonal-bipyramidal adduct, with a bond-formation transition state preceding it and a bond-fission transition state succeeding it. Instead a single transition state is traversed in which both bond-formation and bond-fission occur simultaneously. Such a concerted-reaction transition state should have two weak, partial bonds to phosphorus, one for formation of the nucleophile-P bond and one for fission of the P-leaving group bond. In a stepwise reaction through an intermediate, only one bond is partial and weak in each of the two transition states. The concerted-reaction transition state, with two weak bonds to phosphorus, may be more easily compressed, expanded, and otherwise distorted because of the lower force constants associated with partial bonds; particularly distortions of angles involving the two partial bonds should require relatively low energies. This may lend a high level of flexibility to phosphonolysis transition states, allowing them to be accommodated within an active site (or a range of active sites) with strong catalytic stabilization. Included among these active sites may be a majority that can also stabilize esterolysis transition states. Indeed many of the target esterolysis reactions studied to date may occur through a single concerted-reaction transition state rather than through separate transition states before and after a tetrahedral intermediate. Thus, the esterolysis transition states may also be highly flexible. Finally, flexibility present in germline antibodies may be specifically preserved in reactive immunization. The high flexibility of both kinds of ligands and of the antibody combining site may then account for the catalytic "cross-reactivity" of these antibodies.
- Published
- 2002
- Full Text
- View/download PDF
31. Contributions of active site residues to the partial and overall catalytic activities of human S-adenosylhomocysteine hydrolase.
- Author
-
Elrod P, Zhang J, Yang X, Yin D, Hu Y, Borchardt RT, and Schowen RL
- Subjects
- Adenosylhomocysteinase, Binding Sites genetics, Catalysis, Humans, Hydrolases chemistry, Hydrolases genetics, Models, Chemical, Mutagenesis, Site-Directed, Oxidation-Reduction, Point Mutation, Protein Structure, Quaternary genetics, Protein Structure, Secondary genetics, Recombinant Proteins chemistry, Recombinant Proteins isolation & purification, Recombinant Proteins metabolism, Thermodynamics, Hydrolases metabolism
- Abstract
Residues glutamate 156 (E156), aspartate 190 (D190), asparagine 181 (N181), lysine 186 (K186), and asparagine 191 (N191) in the active site of S-adenosylhomocysteine (AdoHcy) hydrolase have been mutated to alanine (A). AdoHcy hydrolase achieves catalysis of AdoHcy hydrolysis to adenosine (Ado) and homocysteine (Hcy) by means of a redox partial reaction (3'-oxidation of AdoHcy at the beginning and 3'-reduction of Ado at the end of the catalytic cycle) spanning an elimination/addition partial reaction (elimination of Hcy from the oxidized substrate and addition of water to generate the oxidized product), with the enzyme in an open NAD(+) form in the ligand-free state and in a closed NADH form during the elimination/addition partial reaction. Mutation K186A reduces the rate of a model enzymatic reaction for the redox partial reaction by a factor of 280000 and the rate of a model reaction for the elimination/addition partial reaction by a factor of 24000, consistent with a primary catalytic role in both partial reactions as a proton donor/acceptor at the 3'-OH/3'-keto center. Secondary roles for N181 and N191 in localizing the flexible side chain of K186 in a catalytically effective position are supported by rate reduction factors for N181A of 2500 (redox) and 240 (elimination/addition) and for N191A of 730 (redox) and 340 (elimination/addition). A role of D190 in orienting the substrate for effective transition-state stabilization is consistent with rate reduction factors of 1300 (redox) and 30 (elimination/addition) for D190A. Residue E156 may act to maintain K186 in the desired protonation state: rate deduction factors are 1100 (redox) and 70 (elimination/addition). The mutational increases in free energy barriers for k(cat)/K(M) are described by a linear combination of the effects for the partial reactions with the coefficients equal to the fractional degree that each partial reaction determines the rate for k(cat)/K(M). A similar linear equation for k(cat) overestimates the barrier increase by a uniform 5 kJ/mol, probably reflecting reactant-state stabilization by the wild-type enzyme that is abolished by the mutations.
- Published
- 2002
- Full Text
- View/download PDF
32. Computational characterization of substrate binding and catalysis in S-adenosylhomocysteine hydrolase.
- Author
-
Hu Y, Yang X, Yin DH, Mahadevan J, Kuczera K, Schowen RL, and Borchardt RT
- Subjects
- Adenosine metabolism, Adenosylhomocysteinase, Animals, Catalysis, Catalytic Domain, Computer Simulation, Humans, In Vitro Techniques, Kinetics, Models, Molecular, Protein Conformation, Protein Structure, Tertiary, Rats, S-Adenosylhomocysteine metabolism, Software, Substrate Specificity, Thermodynamics, Hydrolases chemistry, Hydrolases metabolism
- Abstract
S-Adenosylhomocysteine (AdoHcy) hydrolase catalyzes the reversible hydrolysis of AdoHcy to adenosine (Ado) and homocysteine (Hcy), playing an essential role in modulating the cellular Hcy levels and regulating activities of a host of methyltransferases in eukaryotic cells. This enzyme exists in an open conformation (active site unoccupied) and a closed conformation (active site occupied with substrate or inhibitor) [Turner, M. A., Yang, X., Yin, D., Kuczera, K., Borchardt, R. T., and Howell, P. L. (2000) Cell Biochem. Biophys. 33, 101-125]. To investigate the binding of natural substrates during catalysis, the computational docking program AutoDock (with confirming calculations using CHARMM) was used to predict the binding modes of various substrates or inhibitors with the closed and open forms of AdoHcy hydrolase. The results have revealed that the interaction between a substrate and the open form of the enzyme is nonspecific, whereas the binding of the substrate in the closed form is highly specific with the adenine moiety of a substrate as the main recognition factor. Residues Thr57, Glu59, Glu156, Gln181, Lys186, Asp190, Met351, and His35 are involved in substrate binding, which is consistent with the crystal structure. His55 in the docked model appears to participate in the elimination of water from Ado through the interaction with the 5'-OH group of Ado. In the same reaction, Asp131 removes a proton from the 4' position of the substrate after the oxidation-reduction reaction in the enzyme. To identify the residues that bind the Hcy moiety, AdoHcy was docked to the closed form of AdoHcy hydrolase. The Hcy tail is predicted to interact with His55, Cys79, Asn80, Asp131, Asp134, and Leu344 in a strained conformation, which may lower the reaction barrier and enhance the catalysis rate.
- Published
- 2001
- Full Text
- View/download PDF
33. In vitro metabolism studies of the prodrug, 2',3',5'-triacetyl-6-azauridine, utilizing an automated analytical system.
- Author
-
Zhou J, Riley CM, and Schowen RL
- Subjects
- Algorithms, Animals, Buffers, Esterases metabolism, Humans, Hydrogen-Ion Concentration, In Vitro Techniques, Kinetics, Liver enzymology, Microdialysis, Temperature, Antifungal Agents metabolism, Azauridine analogs & derivatives, Azauridine metabolism, Prodrugs metabolism
- Abstract
The purpose was to study in vitro metabolism of 2',3',5'-triacetyl-6-azauridine (1) by porcine liver esterase (PLE) and in human plasma using an automated analytical system developed previously. A gradient-LC method was developed to study the concentration-time course of 1 and its metabolites. A fast-LC assay was used to study the temperature effect on the metabolism of 1 by the PLE. 1 and all of its proposed possible metabolites were separated by the gradient-LC method in less than 10 min. Two simplified kinetic schemes were developed to describe the time course of 1, the intermediates and final metabolites with only five rate constants for the metabolisms of 1 by PLE and four rate constants in human plasma. Both enthalpy and entropy of activation in the in vitro metabolism of 1 by PLE were obtained.
- Published
- 2001
- Full Text
- View/download PDF
34. Formaldehyde production by Tris buffer in peptide formulations at elevated temperature.
- Author
-
Song Y, Schowen RL, Borchardt RT, and Topp EM
- Subjects
- Formaldehyde chemical synthesis, Hot Temperature, Peptides chemistry, Tromethamine chemistry
- Abstract
This technical note provides evidence for the degradation of Tris buffer in a peptide formulation stored at elevated temperature (70 degrees C). The buffer degrades to liberate formaldehyde, which is shown to react with the peptide tyrosine residue. Those involved in peptide/protein formulation should be aware of the possible instability in this common biological buffer., (Copyright 2001 Wiley-Liss, Inc.)
- Published
- 2001
- Full Text
- View/download PDF
35. Effect of 'pH' on the rate of asparagine deamidation in polymeric formulations: 'pH'-rate profile.
- Author
-
Song Y, Schowen RL, Borchardt RT, and Topp EM
- Subjects
- Amino Acid Sequence, Buffers, Catalysis, Kinetics, Amides chemistry, Asparagine chemistry, Hydrogen-Ion Concentration, Polymers chemistry
- Abstract
The rate of Asn deamidation of a model hexapeptide (L-Val-L-Tyr-L-Pro-L-Asn-Gly-L-Ala) was measured as a function of effective pH ('pH') in glassy and rubbery polymeric solids containing poly(vinyl pyrrolidone) (PVP) and in solution controls at 70 degrees C. The reaction exhibited pseudo-first-order kinetics in all samples over a wide 'pH' range (0.5 < 'pH' < 12); the formation of similar products suggests that the reaction mechanism is unaffected by matrix type. Rates of deamidation were comparable for the polymeric and solution samples in the acidic range ('pH' < 4). Solution-state rates were faster than those in polymeric solids at neutral 'pH' (6 < 'pH' < 8), increasing to a > 10,000-fold difference in the basic range ('pH' > 8). Specific base catalysis was observed in solution and in the polymeric solids under neutral conditions (6 < 'pH' < 8). In solution, the reaction exhibited general base catalysis for 'pH' > 8, whereas the reaction was 'pH'-independent in the polymeric solids in this range. The 'pH'-rate profile and supporting buffer catalysis data are consistent with a change in the rate-determining step in the basic range from 'pH'-dependent attack of the deprotonated backbone amide nitrogen on the Asn side chain in solution to 'pH'-independent ammonia expulsion in the polymeric solids. The results suggest that polymer matrix incorporation not only affects the magnitude of the deamidation rate constant but also the 'pH' dependency of the reaction and the rate-determining step in the basic 'pH' range.
- Published
- 2001
- Full Text
- View/download PDF
36. A functional assay for quantitation of the apparent affinities of ligands of P-glycoprotein in Caco-2 cells.
- Author
-
Gao J, Murase O, Schowen RL, Aubé J, and Borchardt RT
- Subjects
- Biological Transport, Biological Transport, Active, Caco-2 Cells, Drug Interactions, Humans, Ligands, Models, Theoretical, Tritium, ATP Binding Cassette Transporter, Subfamily B, Member 1 metabolism, Antineoplastic Agents, Phytogenic metabolism, Paclitaxel metabolism
- Abstract
Purpose: To develop a facile functional assay for quantitative determination of the apparent affinities of compounds that interact with the taxol binding site of P-glcoprotein (P-gp) in Caco-2 cell monolayers., Methods: A transport inhibition approach was taken to determine the inhibitory effects of compounds on the active transport of [3H]-taxol, a known substrate of P-gp. The apparent affinities (K(I) values) of the compounds were quantitatively determined based on the inhibitory effects of the compounds on the active transport of [3H]-taxol. Intact Caco-2 cell monolayers were utilized for transport inhibition studies. Samples were analyzed by liquid scintillation counting., Results: [3H]-Taxol (0.04 microM) showed polarized transport with the basolateral (BL) to apical (AP) flux rate being about 10-20 times faster than the flux rate in the AP-to-BL direction. This difference in [3H]-taxol flux could be totally abolished by inclusion of (+/-)-verapamil (0.2 mM), a known inhibitor of P-gp, in the incubation medium. However, inclusion of probenecid (1.0 mM), a known inhibitor for the multidrug resistance associated protein (MRP), did not significantly affect the transport of [3H]-taxol under the same conditions. These results suggest that P-gp, not MRP, was involved in taxol transport. Quinidine, daunorubicin, verapamil, taxol, doxorubicin, vinblastine, etoposide, and celiprolol were examined as inhibitors of the BL-to-AP transport of [3H]-taxol with resulting K(I) values of 1.5+/-0.8, 2.5+/-1.0, 3.0+/-0.3, 7.3+/-0.7, 8.5+/-2.8, 36.5+/-1.5, 276+/-69, and 313+/-112 microM, respectively. With the exception of that of quinidine, these K(I) values were comparable with literature values., Conclusions: This assay allows a facile quantitation of the apparent affinities of compounds to the taxol-binding site in P-gp, however, this assay does not permit the differentiation of substrates and inhibitors. The potential of drug-drug interactions involving the taxol binding site of P-gp can be conveniently estimated using the protocol described in this paper.
- Published
- 2001
- Full Text
- View/download PDF
37. Effects of solution polarity and viscosity on peptide deamidation.
- Author
-
Li R, D'Souza AJ, Laird BB, Schowen RL, Borchardt RT, and Topp EM
- Subjects
- Computer Simulation, Glycerol chemistry, Glycerol metabolism, Kinetics, Povidone chemistry, Povidone metabolism, Refractometry, Regression Analysis, Static Electricity, Viscosity, Amides metabolism, Oligopeptides chemistry, Oligopeptides metabolism, Solutions chemistry, Solutions metabolism
- Abstract
Deamidation kinetics were measured for a model hexapeptide (L-Val-L-Tyr-L-Pro-L-Asn-Gly-L-Ala, 0.02 mg/mL) in aqueous solutions containing glycerol (0-50% w/w) and poly(vinyl pyrrolidone) (PVP, 0-20% w/w) at 37 degrees C and pH 10 to determine the effects of solution polarity and viscosity on reactivity. The observed pseudo-first order deamidation rate constants, k(obs), decreased markedly when the viscosity increased from 0.7 to 13 cp, but showed no significant change at viscosities >13 cp. Values of k(obs) also increased with increasing dielectric constant and decreasing refractive index. Molecular dynamics simulations indicated that the free energy associated with Asn side-chain motion is insensitive to changes in dielectric constant, suggesting that the observed dielectric constant dependence is instead related primarily to the height of the transition state energy barrier. An empirical model was proposed to describe the effects of the viscosity, refractive index and dielectric constant on k(obs). Analysis of the regression coefficients suggested that both permanent and induced dipoles of the medium affect the deamidation rate constant, but that solution viscosity is relatively unimportant in the range studied.
- Published
- 2000
- Full Text
- View/download PDF
38. Reactivity toward deamidation of asparagine residues in beta-turn structures.
- Author
-
Xie M, Aubé J, Borchardt RT, Morton M, Topp EM, Vander Velde D, and Schowen RL
- Subjects
- Amino Acid Sequence, Asparagine metabolism, Circular Dichroism, Deamination, Kinetics, Models, Chemical, Molecular Sequence Data, Peptides, Cyclic, Structure-Activity Relationship, Asparagine chemistry, Protein Structure, Secondary
- Abstract
Mimetics of beta-turn structures in proteins have been used to calibrate the relative reactivities toward deamidation of asparagine residues in the two central positions of a beta-turn and in a random coil. N-Acetyl-Asn-Gly-6-aminocaproic acid, an acyclic analog of a beta-turn mimic undergoes deamidation of the asparaginyl residue through a succinimide intermediate to generate N-acetyl-Asp-N-Gly-6-aminocaproic acid (6-aminocaproic acid, hereafter Aca) and N-acetyl-L-iso-aspartyl (isoAsp)-Gly-Aca (pH 8.8, 37 degrees C) approximately 3-fold faster than does the cyclic beta-turn mimic cyclo-[L-Asn-Gly-Aca] with asparagine at position 2 of the beta-turn. The latter compound, in turn, undergoes deamidation approximately 30-fold faster than its positional isomer cyclo-[Gly-Asn-Aca] with asparagine at position 3 of the beta-turn. Both cyclic peptides assume predominantly beta-turn structures in solution, as demonstrated by NMR and circular dichroism characterization. The open-chain compound and its isomer N-acetyl-Gly-Asn-Aca assume predominantly random coil structures. The latter isomer undergoes deamidation 2-fold slower than the former. Thus the order of reactivity toward deamidation is: asparagine in a random coil approximately 3x(asparagine) in position 2 of a beta-turn approximately 30x (asparagine) in position 3 of a beta-turn.
- Published
- 2000
- Full Text
- View/download PDF
39. Substrate binding stabilizes S-adenosylhomocysteine hydrolase in a closed conformation.
- Author
-
Yin D, Yang X, Hu Y, Kuczera K, Schowen RL, Borchardt RT, and Squier TC
- Subjects
- Adenosylhomocysteinase, Catalysis, Fluorescence Polarization, Fluorescent Dyes, Half-Life, Maleimides, Models, Chemical, Motion, Oxidation-Reduction, Protein Conformation, Spectrometry, Fluorescence, Sulfhydryl Reagents, Hydrolases metabolism, NAD metabolism, S-Adenosylhomocysteine metabolism
- Abstract
Comparison of crystal structures of S-adenosylhomocysteine (AdoHcy) hydrolase in the substrate-free, NAD(+) form [Hu, Y., Komoto, J., Huang, Y., Gomi, T., Ogawa, H., Takata, Y., Fujioka, M., and Takusagawa, F. (1999) Biochemistry 38, 8323-8333] and a substrate-bound, NADH form [Turner, M. A., Yuan, C.-S., Borchardt, R. T., Hershfield, M. S., Smith, G. D., and Howell, P. L. (1998) Nat. Struct. Biol. 5, 369-376] indicates large differences in the spatial arrangement of the catalytic and NAD(+) binding domains. The substrate-free, NAD(+) form exists in an "open" form with respect to catalytic and NAD(+) binding domains, whereas the substrate-bound, NADH form exists in a closed form with respect to those domains. To address whether domain closure is induced by substrate binding or its subsequent oxidation, we have measured the rotational dynamics of spectroscopic probes covalently bound to Cys(113) and Cys(421) within the catalytic and carboxyl-terminal domains. An independent domain motion is associated with the catalytic domain prior to substrate binding, suggesting the presence of a flexible hinge element between the catalytic and NAD(+) binding domains. Following binding of substrates (i.e., adenosine or neplanocin A) or a nonsubstrate (i.e., 3'-deoxyadenosine), the independent domain motion associated with the catalytic domain is essentially abolished. Likewise, there is a substantial decrease in the average hydrodynamic volume of the protein that is consistent with a reduction in the overall dimensions of the homotetrameric enzyme following substrate binding and oxidation observed in earlier crystallographic studies. Thus, the catalytic and NAD(+) binding domains are stabilized to form a closed active site through interactions with the substrate prior to substrate oxidation.
- Published
- 2000
- Full Text
- View/download PDF
40. Hydrogen bonds and proton transfer in general-catalytic transition-state stabilization in enzyme catalysis.
- Author
-
Schowen KB, Limbach HH, Denisov GS, and Schowen RL
- Subjects
- Crystallography, Isotopes, Magnetic Resonance Spectroscopy, Models, Chemical, Models, Theoretical, Proton-Motive Force, Structure-Activity Relationship, Substrate Specificity, Catalysis, Enzymes chemistry, Hydrogen Bonding, Protons
- Abstract
The question of the nature of the proton bridge involved in general acid-base catalysis in both enzymic and non-enzymic systems is considered in the light of long-known but insufficiently appreciated work of Jencks and his coworkers and of more recent results from neutron-diffraction crystallography and NMR spectroscopic studies, as well as results from isotope-effect investigations. These lines of inquiry lead toward the view that the bridging proton, when between electronegative atoms, is in a stable potential at the transition state, not participating strongly in the reaction-coordinate motion. Furthermore they suggest that bond order is well-conserved at unity for bridging protons, and give rough estimates of the degree to which the proton will respond to structural changes in its bonding partners. Thus if a center involved in general-catalytic bridging becomes more basic, the proton is expected to move toward it while maintaining a unit total bond order. For a unit increase in the pK of a bridging partner, the other partner is expected to acquire about 0.06 units of negative charge. The implications are considered for charge distribution in enzymic transition states as the basicity of catalytic residues changes in the course of molecular evolution or during progress along a catalytic pathway.
- Published
- 2000
- Full Text
- View/download PDF
41. Deamidation of a model hexapeptide in poly(vinyl alcohol) hydrogels and xerogels.
- Author
-
Lai MC, Schowen RL, Borchardt RT, and Topp EM
- Subjects
- Asparagine chemistry, Hydrogels chemistry, Amides chemistry, Delayed-Action Preparations chemistry, Gels chemistry, Oligopeptides chemistry, Polyvinyl Alcohol chemistry
- Abstract
Polymeric controlled release systems have been proposed to prolong the half-lives of protein and peptide drugs in vivo and to deliver active drug at a controlled rate. These systems are ineffective, however, if the drug is not stable during storage and release. This study addresses the effect of poly(vinyl alcohol) on the stability and release of an incorporated hexapeptide, VYPNGA, which undergoes deamidation. Two types of peptide-loaded poly(vinyl alcohol) matrices were formed, a semisolid hydrogel and a lower water content 'xerogel', and stored at 50 degrees C for up to 122 days. The hexapeptide was less stable in both poly(vinyl alcohol) matrices than in aqueous buffer or lyophilized polymer-free powders. The type of poly(vinyl alcohol) matrix appeared to influence the degradation mechanism, since the product distributions differ in the hydrogel and the xerogel. The results suggest that, rather than stabilizing this peptide, incorporation in poly(vinyl alcohol) matrices reduces stability relative to solution and lyophilized controls.
- Published
- 2000
- Full Text
- View/download PDF
42. Catalytic antibodies for complex reactions: hapten design and the importance of screening for catalysis in the generation of catalytic antibodies for the NDA/CN reaction.
- Author
-
DeSilva BS, Orosz G, Egodage KL, Carlson RG, Schowen RL, and Wilson GS
- Subjects
- Animals, Antibodies, Monoclonal metabolism, Antibody Affinity, Catalysis, Cyanides metabolism, Fluorescent Dyes metabolism, Haptens chemistry, Hybridomas immunology, In Vitro Techniques, Mice, Naphthalenes immunology, Naphthalenes metabolism, Antibodies, Catalytic metabolism
- Abstract
Success in generating catalytic antibodies as enzyme mimics lies in the strategic design of the transition-state analog (TSA) for the reaction of interest, and careful development of screening processes for the selection of antibodies that are catalysts. Typically, the choice of TSA structure is straightforward, and the criterion for selection in screening is often binding of the TSA to the antibody in a microtiter-plate assay. This article emphasizes the problems of TSA design in complex reactions and the importance of selecting antibodies on the basis of catalysis as well as binding to the TSA. The target reaction is the derivatization of primary amines with naphthalene-2,3-dicarboxaldehyde (NDA) in the presence of cyanide ion. The desired outcome is selective catalysis of formation of the fluorescent derivative in preference to nonfluorescent side-products. In the study, TSA design was directed toward the reaction branch leading to the fluorescent product. Here, we describe a microtiter plate-based assay that is capable of detecting antibodies showing catalytic activity at an early stage. Of the antibodies selected, 36% showed no appreciable binding to any of the substrates tested, but did show catalytic activity in derivatizing one or more of the amino acids screened. In contrast, only two out of 77 clones that showed binding did not show catalysis. Thus, in this complex system, observation of binding is a good predictor of the presence of catalytic activity, and failure to observe binding is a poor predictor of the absence of catalytic activity.
- Published
- 2000
- Full Text
- View/download PDF
43. Catalyst design for reactions with multiple transition states.
- Author
-
DeSilva BS, Wilson GS, and Schowen RL
- Subjects
- Aldehydes, Animals, Catalysis, Cations, Drug Design, Kinetics, Naphthalenes, Thermodynamics, Antibodies, Catalytic chemistry, Antibodies, Catalytic metabolism
- Published
- 2000
- Full Text
- View/download PDF
44. Chemical stability of peptides in polymers. 2. Discriminating between solvent and plasticizing effects of water on peptide deamidation in poly(vinylpyrrolidone).
- Author
-
Lai MC, Hageman MJ, Schowen RL, Borchardt RT, Laird BB, and Topp EM
- Subjects
- Adsorption, Algorithms, Amides chemistry, Drug Stability, Freeze Drying, Glycerol chemistry, Kinetics, Nonlinear Dynamics, Oligopeptides analysis, Solvents, Temperature, Thermodynamics, Water, Oligopeptides chemistry, Pharmaceutic Aids chemistry, Povidone chemistry
- Abstract
The mechanistic role of water in the deamidation of a model asparagine-containing hexapeptide (Val-Tyr-Pro-Asn-Gly-Ala) in lyophilized formulations containing poly(vinylpyrrolidone) (PVP) and glycerol was investigated. Glycerol was used as a plasticizer to vary formulation glass transition temperature (T(g)) without significantly changing water content or activity. Increases in moisture and glycerol contents increased the rate of peptide deamidation. This increase was strongly correlated with T(g) at constant water content and activity, suggesting that increased matrix mobility facilitates deamidation. In rubbery systems (T > T(g)), deamidation rates appeared to be independent of water content and activity in formulations with similar T(g)s. However, in glassy formulations with similar T(g)s, deamidation increased with water content, suggesting a solvent/medium effect of water on reactivity in this regime. An increase in water content also affected the degradation product distribution; less of the cyclic imide intermediate and more of the hydrolytic products, isoAsp- and Asp-hexapeptides, were observed as water content increased. Thus, residual water appears to facilitate deamidation in these solid PVP formulations both by enhancing molecular mobility and by solvent/medium effects, and also participates as a chemical reactant in the subsequent breakdown of the cyclic imide.
- Published
- 1999
- Full Text
- View/download PDF
45. Chemical stability of peptides in polymers. 1. Effect of water on peptide deamidation in poly(vinyl alcohol) and poly(vinyl pyrrolidone) matrixes.
- Author
-
Lai MC, Hageman MJ, Schowen RL, Borchardt RT, and Topp EM
- Subjects
- Adsorption, Algorithms, Amides chemistry, Drug Stability, Kinetics, Oligopeptides analysis, Polymers, Temperature, Thermodynamics, Water, Oligopeptides chemistry, Pharmaceutic Aids chemistry, Polyvinyl Alcohol chemistry, Povidone chemistry
- Abstract
This paper examines the effect of water content, water activity, and glass transition temperature (T(g)) on the deamidation of an asparagine-containing hexapeptide (VYPNGA; Asn-hexapeptide) in lyophilized poly(vinyl alcohol) (PVA) and poly(vinyl pyrrolidone) (PVP) at 50 degrees C. The rate of Asn-hexapeptide deamidation increases with increasing water content or water activity and, hence, decreasing T(g). The rate of deamidation is more sensitive to changes in these parameters in PVA than in PVP. Deamidation is clearly evident in the glassy state in both formulations. In the glassy state, the peptide is more stable in PVA than in PVP formulations but is less stable in the rubbery state. No single variable (water content, water activity, or T(g)) could account for the variation in deamidation rates in PVA and PVP formulations. Deamidation rates were correlated with the degree of plasticization by water (distance of T(g) from the dry intrinsic glass transition temperature); coincident curves for the two polymers were obtained with this correlation. Deamidation in PVA and PVP was closely correlated with the extent of water-induced plasticization experienced by the formulation relative to its glass transition at 50 degrees C, suggesting that the physical state of formulations could be used to predict chemical stability.
- Published
- 1999
- Full Text
- View/download PDF
46. The linkage of catalysis and regulation in enzyme action: oxidative diversion in the hysteretically regulated yeast pyruvate decarboxylase.
- Author
-
Hajipour G, Schowen KB, and Schowen RL
- Subjects
- 2,6-Dichloroindophenol pharmacology, Allosteric Regulation, Binding Sites, Dose-Response Relationship, Drug, Kinetics, Models, Chemical, Saccharomyces cerevisiae enzymology, Catalysis, Oxygen metabolism, Pyruvate Decarboxylase metabolism
- Abstract
The reaction catalyzed by the thiamin-diphosphate-dependent yeast pyruvate decarboxylase, which is hysteretically regulated by pyruvate, undergoes paracatalytic oxidative diversion by 2,6-dichlorophenolindophenol, which traps a carbanionic intermediate and diverts the product from acetaldehyde to acetate (Christen, P. Meth. Enzymol. 1977, 46, 48). This reaction is now shown to exhibit an oxidant on-rate constant somewhat faster than that for pyruvate in the normal catalytic cycle and a product off-rate constant about 60-fold smaller than that for acetaldehyde. Both on-rates and off-rates exhibit an inverse solvent isotope effect of 1.5-2, observed in normal catalysis as a signal of sulfhydryl addition to the keto group of pyruvate at the allosteric regulatory site. The findings are consistent with a model for regulation in which the sulfhydryl-addition process mediates access to a fully catalytically competent active site, the oxidative-diversion reaction being forced to make use of the normal entry exit machinery.
- Published
- 1999
- Full Text
- View/download PDF
47. Secondary structure and protein deamidation.
- Author
-
Xie M and Schowen RL
- Subjects
- Animals, Asparagine chemistry, Asparagine metabolism, Humans, Proteins chemistry, Amides metabolism, Protein Structure, Secondary, Proteins metabolism
- Abstract
The deamidation reactions of asparagine residues in alpha-helical and beta-turn secondary structural environments of peptides and proteins are reviewed. Both kinds of secondary structure tend to stabilize asparagine residues against deamidation, although the effects are not large. The effect of beta-sheet structures on asparagine stability is unclear, although simple considerations suggest a stabilization in this environment also.
- Published
- 1999
- Full Text
- View/download PDF
48. Transition-state theoretical interpretation of the catalytic power of pyruvate decarboxylases: the roles of static and dynamical considerations.
- Author
-
Hong J, Sun S, Derrick T, Larive C, Schowen KB, and Schowen RL
- Subjects
- Bacteria enzymology, Catalysis, Models, Biological, Thermodynamics, Thiamine Pyrophosphate metabolism, Yeasts enzymology, Pyruvate Decarboxylase metabolism
- Abstract
The catalytic power of two thiamin diphosphate (ThDP)-dependent enzymes, yeast pyruvate decarboxylase (the hysteretically regulated enzyme from Saccharomyces cerevisiae, SCPDC) and bacterial pyruvate decarboxylase (the unregulated enzyme from Zymomonas mobilis, ZMPDC), are analyzed by thorough-going application of transition-state theory, i.e. by a static approach that emphasizes the state-function character of the free energy of activation and takes no explicit account of dynamical considerations. The overall catalytic reaction is resolved into manifolds for addition (conversion of free enzyme and substrate to the complex of enzyme with the pyruvate:ThDP adduct), decarboxylation, and elimination (conversion of the complex of enzyme with the acetaldehyde:ThDP adduct formed by decarboxylation into free product and free enzyme). For SCPDC, the addition manifold is most strongly catalyzed (3x1012-fold, corresponding to net transition-state stabilization of 72 kJ/mol, transition-state stabilization of 83 kJ/mol diminished by reactant-state stabilization of 11 kJ/mol), the decarboxylation manifold is least strongly catalyzed (5x107-fold, corresponding to net transition-state stabilization of 41 kJ/mol, transition-state stabilization of 68 kJ/mol diminished by reactant-state stabilization of 27 kJ/mol), and the elimination manifold is catalyzed to an intermediate degree (2x1010-fold, corresponding to net transition-state stabilization of 59 kJ/mol, transition-state stabilization of 76 kJ/mol diminished by reactant-state stabilization of 17 kJ/mol). A similar situation holds for ZMPDC. There is no need to make an explicit analysis of dynamical factors in order to describe the catalytic mechanism and catalytic power of these complex enzymes.
- Published
- 1998
- Full Text
- View/download PDF
49. Ion channel properties of a protein complex with characteristics of a glutamate/N-methyl-D-aspartate receptor.
- Author
-
Aistrup GL, Szentirmay M, Kumar KN, Babcock KK, Schowen RL, and Michaelis EK
- Subjects
- 2-Amino-5-phosphonovalerate pharmacology, Animals, Aspartic Acid pharmacology, Binding Sites, Brain metabolism, Calcium metabolism, Excitatory Amino Acid Antagonists pharmacology, Glutamic Acid pharmacology, Glycine pharmacology, Ion Channels antagonists & inhibitors, Ion Channels chemistry, Ion Channels isolation & purification, Lipid Bilayers chemistry, Liposomes chemistry, Patch-Clamp Techniques, Rats, Ion Channels metabolism, Receptors, Glutamate metabolism, Receptors, N-Methyl-D-Aspartate metabolism, Synaptic Membranes chemistry
- Abstract
The functional reconstitution of glutamate receptor proteins purified from mammalian brain has been difficult to accomplish. However, channels activated by L-glutamate (L-Glu) and N-methyl-D-aspartate (NMDA) were detected in planar lipid bilayer membranes (PLMs) following the reconstitution of a complex of proteins with binding sites for NMDA receptor (NMDAR) ligands. The presence of glycine was necessary for optimal activation. A linear current-voltage relationship was observed with the reversal potential being zero. Channels activated by L-Glu had conductances of 23, 47 and 65 pS, and were suppressed partially by competitive and fully by noncompetitive inhibitors of NMDARs. Magnesium had little effect on the reconstituted channels.
- Published
- 1996
- Full Text
- View/download PDF
50. Facilitated Intramolecular Conjugate Addition of Amides of 3-(3',6'-Dioxo-2',4'-dimethyl-1',4'-cyclohexadienyl)-3,3-dimethylpropionic Acid. 2. Kinetics of Degradation.
- Author
-
Nicolaou MG, Wolfe JL, Schowen RL, and Borchardt RT
- Abstract
The chemical stability studies of amides of 3-(3',6'-dioxo-2',4'-dimethyl-1',4'-cyclohexadienyl)-3,3-dimethylpropionic acid (Qa) [Qop(a-j)] were conducted in order to determine the utility of this redox-sensitive system as a potential prodrug promoiety or redox-sensitive protecting group in organic synthesis. This study showed that quinone propionic amides of aniline derivatives [Qop(a-d)] underwent rapid degradation in mildly acidic conditions (pH 4.5-6.0) to yield degradation products resulting from the intramolecular 1,2- or 1,4- conjugate addition of the amide nitrogen to the quinone ring. This conjugate addition was found to be specific base-catalyzed and independent of the para substituent on the aromatic ring of the amine. The predominant route of degradation yielded a five-membered ring spirolactam. By altering the nature of the amine component of the amide, these degradation reactions were prevented. Amides of Qa other than those of the aniline type [Qop(e-j)] were found to be substantially more stable and were thus proposed as the more suitable candidates for this potential redox-sensitive prodrug system and redox-sensitive protecting group for amines and alcohols in organic synthesis.
- Published
- 1996
- Full Text
- View/download PDF
Catalog
Discovery Service for Jio Institute Digital Library
For full access to our library's resources, please sign in.