首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Arginine deiminase (ADI, EC 3.5.3.6) is a potential antitumor drug for the treatment of arginine‐auxotrophic tumors such as hepatocellular carcinomas (HCCs) and melanomas, and studies on human lymphatic leukemia cell lines have confirmed that ADI has antiangiogenic activity. Recent studies showed that a combination of taxane and ADI‐PEG20, which induces caspase‐independent apoptosis, is more effective than taxane monotherapy for prostate cancer. The main limitation of ADI from Pseudomonas plecoglossicida (PpADI) and of many other ADI enzymes lies in their pH‐dependent activity profile. PpADI has a pH optimum at 6.5 and a pH shift from 6.5 to 7.5 results in an ~80 % activity drop (the pH of human plasma is 7.35 to 7.45). In 2010, we reported a proof of concept for ADI engineering by directed evolution that resulted in variant M2 (K5T/D44E/H404R). M2 has a pH optimum of pH 7.0, a fourfold higher kcat value than the wild‐type PpADI (pH 7.4, 0.5 M phosphate buffer), and an increased Km value for substrate arginine. In our latest work, variants M5 (K5T/D38H/D44E/A128T/H404R) and M6 (K5T/D38H/D44E/A128T/E296K/H404R) were generated by directed evolution by employing PBS buffer (pH 7.4), which mimics physiological conditions. The S0.5 value of parent M3 (K5T/D44E/A128T/H404R) decreased from 2.01 to 1.48 mM (M5) and 0.81 mM (M6). The S0.5 value of M6 (0.81 mM ) is lower than that of wild‐type PpADI (1.30 mM ); the kcat values improved from 0.18 s?1 (wild‐type PpADI) to 17.56 s?1 (M5, 97.6‐fold) and 11.64 s?1 (M6, 64.7‐fold).  相似文献   

2.
Arginine deiminase (ADI; EC 3.5.3.6) has been studied as a potential antitumor drug for the treatment of arginine‐auxotrophic tumors, such as hepatocellular carcinomas (HCCs) and melanomas. Studies with human lymphatic leukemia cell lines confirmed that ADI is an antiangiogenic agent for treating leukemia. The main limitation of ADI from Pseudomonas plecoglossicida (PpADI) lies in its pH‐dependent activity profile, its pH optimum is at 6.5. A pH shift from 6.5 to 7.5 results in an approximately 80 % drop in activity. (The pH of human plasma is 7.35 to 7.45.) In order to shift the PpADI pH optimum, a directed‐evolution protocol based on an adapted citrulline‐screening protocol in microtiter‐plate format was developed and validated. A proof of concept for ADI engineering resulted in a pH optimum of pH 7.0 and increased resistance under physiological and slightly alkaline conditions. At pH 7.4, variant M2 (K5T/D44E/H404R) is four times faster than the wild‐type PpADI and retains ~50 % of its activity relative to its pH optimum, compared to ~10 % in the case of the wild‐type PpADI.  相似文献   

3.
Glucose dehydrogenase (GDH) is a general tool for driving nicotinamide (NAD(P)H) regeneration in synthetic biochemistry. An increasing number of synthetic bioreactions are carried out in media containing high amounts of organic cosolvents or hydrophobic substrates/products, which often denature native enzymes, including those for cofactor regeneration. In this work, we attempted to improve the chemical stability of Bacillus megaterium GDH (BmGDHM0) in the presence of large amounts of 1-phenylethanol by directed evolution. Among the resulting mutants, BmGDHM6 (Q252L/E170K/S100P/K166R/V72I/K137R) exhibited a 9.2-fold increase in tolerance against 10 % (v/v) 1-phenylethanol. Moreover, BmGDHM6 was also more stable than BmGDHM0 when exposed to hydrophobic and enzyme-inactivating compounds such as acetophenone, ethyl 2-oxo-4-phenylbutyrate, and ethyl (R)-2-hydroxy-4-phenylbutyrate. Coupled with a Candida glabrata carbonyl reductase, BmGDHM6 was successfully used for the asymmetric reduction of deactivating ethyl 2-oxo-4-phenylbutyrate with total turnover number of 1800 for the nicotinamide cofactor, thus making it attractive for commercial application. Overall, the evolution of chemically robust GDH facilitates its wider use as a general tool for NAD(P)H regeneration in biocatalysis.  相似文献   

4.
Docking simulations were used to predict the most favorable interaction between the T315I mutated form of Abl (invariably associated with resistance to the tyrosine kinase inhibitor imatinib mesylate, IM) and C6‐unsubstituted and substituted pyrazolo[3,4‐d]pyrimidines previously found to be dual Src/Abl inhibitors. Two C6‐unsubstituted ( 1 and 2 ) and eight C6‐substituted compounds ( 3 – 10 ) were selected and assayed for their effects on the Ba/F3 cell line transducing the wild‐type p210Bcr–Abl construct, which is IM‐sensitive, or three of the most common mutations associated with IM resistance in vivo (T315I, Y253F, and E255K), and driven to drug resistance by saturating doses of IL‐3 or by the expression of the Bcr–Abl construct coding for the p185 protein of acute lymphoblastic leukemia. Compounds 1 and 2 were active against all cell lines assayed (LD50 range: 0.7–4.3 μM ), whereas C6‐substituted compounds exhibited lower activity (LD50~8 μM for compound 3 toward the T315I mutant). Notably, 1 and 2 were also effective toward the T315I mutation, which is insensitive to dual Src/Abl inhibitors. The cytotoxic effects of 1 and 2 on IM‐sensitive and IM‐resistant Ba/F3 cells were attributable, at least in part, to their pro‐apoptotic activity. Taken together, such findings suggest that C6‐unsubstituted pyrazolo[3,4‐d]pyrimidines may represent useful inhibitors to target IM‐resistant chronic myeloid leukemia.  相似文献   

5.
A novel L ‐pantolactone hydrolase, Lph, from Agrobacterium tumefaciens Lu681 was characterized, which stereospecifically hydrolyses L ‐pantolactone to L ‐pantoic acid yielding D ‐pantolactone with > 95% enantiomeric excess. The enzyme was found to be a 30 kDa‐Zn2+‐hydrolase with a Km for L ‐pantolactone of 7 mM and a Vmax of 30 U/mg. The corresponding lph gene was identified as an 807 bp ORF and cloned into E. coli. It was overexpressed under control of Ptac and Prha yielding enzyme activities of up to 600 U/g dry weight. Resolution of d,l ‐pantolactone in repeated batches with isolated Lph and enzyme recovery by membrane filtration gave D ‐pantolactone with 50% yield and 90–95% ee over 6 days. Covalent immobilization to EupergitC led to a stable biocatalyst easy to handle in a repeated batch production of D ‐pantolactone. Further improvements in the activity of Lph were achieved by directed evolution of the enzyme. Activities of mutants F62S, K197D and F100L were increased 2.3, 1.7, and 1.5 fold, respectively.  相似文献   

6.
The four stereoisomers of azetidine‐2,3‐dicaroxylic acid (L ‐trans‐ADC, L ‐cis‐ADC, D ‐trans‐ADC, and D ‐cis‐ADC) were synthesized in a stereocontrolled fashion following two distinct strategies: one providing the two cis‐ADC enantiomers and one giving access to the two trans‐ADC enantiomers. The four azetidinic amino acids were characterized in a radioligand binding assay ([3H]CGP39653) at native NMDA receptors: L ‐trans‐ADC showed the highest affinity (Ki=10 μM ) followed by the D ‐cis‐ADC stereoisomer (21 μM ). In contrast, the two analogues L ‐cis‐ADC and D ‐trans‐ADC were low‐affinity ligands (>100 and 90 μM , respectively). Electrophysiological characterization of the ADC compounds at the four NMDA receptor subtypes NR1/NR2A, NR1/NR2B, NR1/NR2C, and NR1/NR2D expressed in Xenopus oocytes showed that L ‐trans‐ADC displayed the highest agonist potency at NR1/NR2D (EC50=50 μM ), which was 9.4‐, 3.4‐, and 1.9‐fold higher than the respective potencies at NR1/NR2A–C. D ‐cis‐ADC was shown to be a partial agonist at NR1/NR2C and NR1/NR2D with medium‐range micromolar potencies (EC50=720 and 230 μM , respectively). A subsequent in silico ligand–protein docking study suggested an unusual binding mode for these amino acids in the agonist binding site.  相似文献   

7.
An analysis of 503 available triosephosphate isomerase sequences revealed nine fully conserved residues. Of these, four residues—K12, H95, E97 and E165—are capable of proton transfer and are all arrayed around the dihydroxyacetone phosphate substrate in the three‐dimensional structure. Specific roles have been assigned to the residues K12, H95 and E165, but the nature of the involvement of E97 has not been established. Kinetic and structural characterization is reported for the E97Q and E97D mutants of Plasmodium falciparum triosephosphate isomerase (Pf TIM). A 4000‐fold reduction in kcat is observed for E97Q, whereas the E97D mutant shows a 100‐fold reduction. The control mutant, E165A, which lacks the key catalytic base, shows an approximately 9000‐fold drop in activity. The integrity of the overall fold and stability of the dimeric structure have been demonstrated by biophysical studies. Crystal structures of E97Q and E97D mutants have been determined at 2.0 Å resolution. In the case of the isosteric replacement of glutamic acid by glutamine in the E97Q mutant a large conformational change for the critical K12 side chain is observed, corresponding to a trans‐to‐gauche transition about the Cγ? Cδ (χ3) bond. In the E97D mutant, the K12 side chain maintains the wild‐type orientation, but the hydrogen bond between K12 and D97 is lost. The results are interpreted as a direct role for E97 in the catalytic proton transfer cycle. The proposed mechanism eliminates the need to invoke the formation of the energetically unfavourable imidazolate anion at H95, a key feature of the classical mechanism.  相似文献   

8.
CYP106A2 is known as a 15β‐hydroxylase, but also shows minor 11α‐hydroxylase activity for progesterone. 11α‐Hydroxyprogesterone is an important pharmaceutical compound with anti‐androgenic and blood‐pressure‐regulating activity. This work therefore focused on directing the regioselectivity of the enzyme towards hydroxylation at position 11 in the C ring of the steroid through a combination of saturation mutagenesis and rational site‐directed mutagenesis. With the aid of data from a homology model of CYP106A2 containing docked progesterone, together with site‐directed mutagenesis of active‐site residues (Lisurek et al. ChemBioChem 2008 , 9, 1439–1449), a saturation mutagenesis library at positions A395 and G397 was created. Screening of the library identified the mutants A395I and A395W/G397K as having 11α‐hydroxylase activities 8.9 and 11.5 times higher than that of the wild type (WT). In the next step, additional mutations were integrated by a rational site‐directed mutagenesis approach to increase the catalytic efficiency. Of the 40 candidates analyzed, the mutants A106T/A395I, A106T/A395I/R409L, and T89N/A395I turned out to display increased 11α‐hydroxylase selectivities and activities relative to the WT (14.3‐, 12.6‐, and 11.8‐fold increases in selectivity and 39.3‐, 108‐, and 24.4‐ in kcat/Km). In the last step of the study, the best mutants were applied in a whole‐cell biotransformation. In these experiments the production (percentage) of 15β‐hydroxyprogesterone decreased from 50.4 % (wild type) to 4.8 % (mutant T89N/A395I), whereas that of 11α‐hydroxyprogesterone increased from 27.7 to 80.9 %, thus demonstrating an impressive regioselectivity.  相似文献   

9.
Syntheses and biodegradation of statistical copolymers of D ,L ‐lactide (D ,L ‐LA) with trimethylene carbonate (TMC), rac‐1‐methyltrimethylene carbonate (1‐MTMC) and 2,2‐dimethyltrimethylene carbonate (2,2‐DTMC) were investigated at various monomer ratios using SmMe(C5Me5)2THF as an initiator at 80 °C for 24 h in toluene. Biodegradations of poly(D ,L ‐LA‐co‐racemo‐1‐MTMC) (95/5) and poly(D ,L ‐LA‐co‐2,2‐DTMC) (98/2) with a compost at 60 °C proceed rapidly. Enzymatic degradations of these polymers were also performed using cholesterol esterase, lipoprotein lipase and proteinase K. Only poly(D ,L ‐LA‐co‐TMC) was biodegraded with cholesterol esterase, while poly(TMC), poly(1‐MTMC), poly(2,2‐DTMC) and poly(D ,L ‐LA) were barely degraded with these enzymes. Biodegradations of poly(D ,L ‐LA‐co‐TMC) (87/13) and poly(D ,L ‐LA‐co‐racemo‐1‐MTMC) (95/5) are rapid using proteinase K. Physical properties of these copolymers were also described. © 2003 Society of Chemical Industry  相似文献   

10.
The effects of the molecular weight of poly(D ‐lactic acid) (PDLA), which forms stereocomplex (SC) crystallites with poly(L ‐lactic acid) (PLLA), and those of processing temperature Tp on the acceleration (or nucleation) of PLLA homocrystallization were investigated using PLLA films containing 10 wt% PDLA with number‐average molecular weight (Mn) values of 5.47 × 105, 9.67 × 104 and 3.67 × 104 g mol–1 (PDLA‐H, PDLA‐M and PDLA‐L, respectively). For the PLLA/PDLA‐H and PLLA/PDLA‐M films, the SC crystallites that were ‘non’‐melted and those that were ‘completely’ melted at Tp values just above their endset melting temperature and recrystallized during cooling were found to act as effective accelerating (or nucleation) agents for PLLA homocrystallization. In contrast, SC crystallites formed from PDLA‐L, having the lowest Mn, were effective accelerating agents without any restrictions on Tp. In this case, the accelerating effects can be attributed to the plasticizer effect of PDLA‐L with the lowest Mn. The accelerating effects of SC crystallites in the PLLA/PDLA‐H and PLLA/PDLA‐M films was dependent on crystalline thickness for Tp values below the melting peak temperature of SC crystallites, whereas for Tp values above the melting peak temperature the accelerating effects are suggested to be affected by the interaction between the SC crystalline regions and PLLA amorphous regions.  相似文献   

11.
A conductive poly(aniline codoped with dodecyl benzene sulfonic acid and hydrochloric acid) [PANI‐D/H, yield: 32.2%, intrinsic viscosity ([η]): 1.39 dL/g, electrical conductivity: 7.3 S/cm] was synthesized by chemical oxidative polymerization from aniline‐dodecylbenzene sulfonic acid salt (A‐DS)/aniline‐hydrochloric acid salt (A‐HS) (6/4M ratio) in an aqueous system. Waterborne polyurethane (WBPU) dispersion obtained from isophorone diisocyanate/poly(tetramethylene oxide)glycol/dimethylol propionic acid/ethylene diamine/triethylene amine/water was used as a matrix polymer. The blend films of WBPU/PANI‐D/H with various weight ratios (99.9/0.1–25/75) were prepared by solution blending/casting. Effect of PANI‐D/H content on the mechanical property, dynamic mechanical property, hardness, electrical conductivity, and antistaticity of WBPU/PANI‐D/H blend films was investigated. The dynamic storage modulus and initial tensile modulus increased with increasing PANI‐D/H content up to 1 wt %, and then it was significantly decreased about the content. With increasing PANI‐D/H content, the glass transition temperature of soft segment (Tgs) and hard segment (Tgh) of WBPU/PANI‐D/H blend films were shifted a bit to lower the temperature. The tensile strength and hardness of WBPU/PANI‐D/H blend films increased a little with increasing PANI‐D/H content up to 0.5 wt %, and then it was dramatically decreased over the content. The elongation at break of WBPU/PANI‐D/H decreased with an increase in PANI‐D/H content. From these results, it was concluded that 0.5–1 wt % of PANI‐D/H was the critical concentration to reinforce those various properties of WBPU/PANI‐D/H blend films prepared in this study. The electrical conductivity of WBPU/ultrasonic treated PANI‐D/H (particle size: 0.7 μm) blend films prepared here increased from 4.0 × 10?7 to 0.33 S/cm with increasing PANI‐D/H content from 0.1 to 75 wt %. The antistatic half‐life time (τ1/2) of pure WBPU film was about 110 s. However, those of WBPU/ultrasonic treated PANI‐D/H blend films (τ1/2: 8.2–0.1 s, and almost 0 s) were found to decrease exponentially with increasing PANI‐D/H content (0.1–9 wt %, and above 9 wt %). © 2004 Wiley Periodicals, Inc. J Appl Polym Sci 93: 700–710, 2004  相似文献   

12.
Soot is an unwanted by-product of combustion or pyrolysis of organic compounds. At least four morphologically distinct forms of particulate carbon can be recognized in various types of soot:
1. (1) aciniform carbon (AC), composed of spheroidal carbon particles fused together in aggregates of colloidal dimensions
2. (2) carbonaceous microgel (CM), in which spheroidal carbon particles are embedded in carbon or carbonaceous material
3. (3) carbon cenospheres
4. (4) coke and char fragments.
We have examined 18 soots from various sources. Domestic chimney soots have very little AC but consist mainly of char fragments, CM, ash, and soluble organic fraction (SOF); whereas diesel soots consist mainly of AC and SOF. Carbon black is a manufactured product consisting almost entirely of AC. Carbon blacks can be distinguished from many soots by their morphology, and from all soots investigated in the present study by their purity. Thus, carbon blacks have less than 0.3% SOF removable by toluene, and less than 1% ash, whereas soots exceed at least one of these limits by several fold. Because of differences in origin, function, and composition, common nomenclature carbon black with soot is inappropriate.

References

1.U. Colombo and M.W. Thring, Combust. Sci. Technol. 5 (1972), p. 189. Full Text via CrossRef2.Webster's New International Dictionary (2nd Edn.), G. and C. Merriam Co, Springfield (1961).3.A.I. Medalia and F.A. Heckman, J. Colloid Interf. Sci. 36 (1971), p. 173. Abstract | Article | PDF (14240 K) | View Record in Scopus | Cited By in Scopus (6)4.A.I. Medalia, Rubber Chem. Technol. 47 (1974), p. 411. View Record in Scopus | Cited By in Scopus (46)5.D.J. Lloyd (2nd Edn.) In: J. Alexander, Editor, Colloid Chemistry Vol. 1, Chemical Catalog Co, New York (1962), pp. 767–782.6.D.H. Everett Preparer (2nd Edn.), IUPAC Division of Physical Chemistry, Manual of Symbols and Terminology for Physicochemical Quantities and Units, Butterworths, London (1971) Part I..7.W.O. Baker, Ind. Eng. Chem. 41 (1949), p. 511. Full Text via CrossRef8.A.I. Medalia, J. Polymer Sci. 6 (1951), p. 423. Full Text via CrossRef9.W.C. McCrone and J.G. Delly, The Particle Atlas (2nd Edn.), Ann Arbor Science Publishers, Ann Arbor (1972).10.E.M. Dannenberg, Carbon black (3rd Edn.), Kirk-Othmer Encyclopedia of Chemical Technology Vol. 4, Wiley, New York (1978), p. 631.11.S.S. Butcher and E.M. Sorenson, J. Air Poll. Contr. Assoc. 29 (1979), p. 724. View Record in Scopus | Cited By in Scopus (10)12.J.L. Muhlbaier, paper presented at Int. Conf. on Residential Solid Fuels Portland, OR, June, General Motors Research Publication GMR-3730 (1981).13.B. Michel, M. Neufelder and D. Pruggmayer, Staub Reinh. Luft in English 34 (1974), p. 129.14.W.H. Lipkea, J.H. Johnson and C.T. Vuk, SAE paper 780108 (1978).15.L.E. Frisch, J.H. Johnson and D.G. Leddy, SAE paper 790417 (1979).16.C.T. Hare, K.J. Springer and R.L. Bradow, SAE paper 760130 (1976).17.G. Hunter, J. Scholl, F. Hibbler, S. Bagley, D. Leddy, D. Abata and J.H. Johnson, SAE Technical Publications 810263 (1981).18.C.T. Vuk, M.A. Jones and J.H. Johnson, SAE paper 760131 (1976).19.K. Carpenter and J.H. Johnson, SAE paper 790815 (1979).20.D.B. Kittelson, D.F. Dolan, R.B. Diver and E. Aufderkeide, SAE paper 780787 (1978).21.F. A. Heckman, private communication.22.E.F. Funkenbusch, D.G. Leddy and J.H. Johnson, SAE paper 790418 (1979).23.A.I. Medalia, D. Rivin and D.R. Sanders, Paper No. 7 to Meeting of ACS Rubber Division Cleveland (Oct. 13, 1981).24.ASTM Procedure D3037-76 (Method C).25.H.E. Rose, J. Appl. Chem. 7 (1957), p. 244.26.L.G. Nabors and M.L. Studebaker, Rubber Chem. Technol. 40 (1967), p. 1323.  相似文献   

13.
The solubility of polystyrene with molecular weight of 100 000 and 260 000 g mol?1 was measured at temperatures from 313 to 333 K and at pressures from 5.0 to 13.0 MPa in subcritical 1,1,1,2‐tetrafluoroethane (R134a). The effects of pressure, temperature and molecular weight on the solubility of polystyrene were investigated. Meanwhile, the solubility of polystyrene was correlated using six density‐based semi‐empirical models (Chrastil, A‐L, K‐J, S‐S, M‐S‐T and Bartle). The M‐S‐T model was used to verify the self‐consistency of the experimental data, and the enthalpy values of polystyrene, including ΔHtotal, ΔHsub and ΔHsol, were estimated through the Chrastil and Bartle models. In addition, a modified M‐S‐T model was proposed through a detailed study of semi‐empirical model formulas and verified by the solubility of polystyrene in subcritical R134a and 38 other solid solutes in supercritical carbon dioxide and subcritical R134a. © 2018 Society of Chemical Industry  相似文献   

14.
Mixed‐substituent fluoroalkoxyphosphazene polymers bearing ~15% 1H,1H,2H,2H‐perfluorooctan‐1‐oxy or 1H,1H,2H,2H‐perfluorodecan‐1‐oxy side groups together with trifluoroethoxy cosubstituent groups were synthesized. The low reactivity of the long‐chain fluoroalkoxides and their limited solubility in organic solvents prevented higher levels of substitution. Moreover, the sodium alkoxides with two methylene residues adjacent to the oxygen proved to be unstable in solution due to elimination of NaF and precipitation of side products, and this limited the time available for chlorine replacement reactions. The resulting cosubstituent polymers were characterized by proton nuclear magnetic resonance (1H‐NMR), 31P‐NMR, 19F‐NMR, gel‐permeation chromatography, and differential scanning calorimetry. Unlike homo‐ or mixed‐substituent fluoroalkoxyphosphazene polymers, such as [NP(OCH2CF3)2]n (a microcrystalline thermoplastic, Tg ~ ?63°C, Tm ~ 242°C) or [NP(OCH2CF3)(OCH2(CF2)xCF2H)]n (PN‐F, a rubbery elastomer, Tg ~ ?60°C, but no detectable Tm), the new polymers are gums (Tg ~ ?50°C, but no detectable Tm) with molecular weights in the 105 g/mol rather than the 106 g/mol range. POLYM. ENG. SCI., 54:1827–1832, 2014. © 2013 Society of Plastics Engineers  相似文献   

15.
Directed evolution of Cp*RhIII-linked nitrobindin (NB), a biohybrid catalyst, was performed based on an in vitro screening approach. A key aspect of this effort was the establishment of a high-throughput screening (HTS) platform that involves an affinity purification step employing a starch-agarose resin for a maltose binding protein (MBP) tag. The HTS platform enables efficient preparation of the purified MBP-tagged biohybrid catalysts in a 96-well format and eliminates background influence of the host E. coli cells. Three rounds of directed evolution and screening of more than 4000 clones yielded a Cp*RhIII-linked NB(T98H/L100K/K127E) variant with a 4.9-fold enhanced activity for the cycloaddition of acetophenone oximes with alkynes. It is confirmed that this HTS platform for directed evolution provides an efficient strategy for generating highly active biohybrid catalysts incorporating a synthetic metal cofactor.  相似文献   

16.
Proteases have niche applications in diagnostic kits that use cell lysis and thereby require high resistance towards chaotropic salts and detergents, such as guanidinium chloride (GdmCl) and sodium dodecylsulfate (SDS). Subtilisin E, a well‐studied serine protease, was selected to be re‐engineered by directed evolution into a “chaophilic” protease that would be resistance to GdmCl and SDS, for application in diagnostic kits. In three iterative rounds of directed evolution, variant SeSaM1–5 (S62I/A153V/G166S/I205V) was generated, with improved activity (330 %) and increased half life in 1 M GdmCl (<2 min to 4.7 h) or in 0.5 % SDS (<2 min to 2.7 h). Saturation mutagenesis at each site in the wild‐type subtilisin E revealed that positions 62 and 166 were mainly responsible for increased activity and stability. A double mutant, M2 (S62I/G166M), generated by combination of the best single mutations showed significantly improved kinetic constants; in 2 M GdmCl the Km value decreased (29‐fold) from 7.31 to 0.25 mM , and the kcat values increased (fourfold) from 15 to 61 s?1. The catalytic efficiency, kcat/Km, improved dramatically (GdmCl: 247 mM ?1 s?1 (118‐fold); SDS, 179 mM ?1 s?1 (13‐fold)). In addition, the SeSaM1–5 variant showed higher stability in 2.0 % SDS when compared to the wild‐type (t1/2 54.8 min (>27‐fold)). Finally, molecular dynamics simulations of the wild‐type subtilisin E showed that Gdm+ ions could directly interact with active site residues, thereby probably limiting access of the substrate to the catalytic centre.  相似文献   

17.
(9‐Ethyl‐carbazol‐6‐yl) methyl methacrylate/methyl acrylate (E/A) copolymers of different compositions were prepared by solution polymerization by varying the molar infeed ratio, using AIBN as initiator at 60°C. The reactivity ratios calculated by Kelen–Tudos (KT) method were found to be rE = 1.16 ± 0.02 and rA = 0.69 ± 0.01 whereas those calculated from RREVM method were found to be rE = 1.18 and rA = 0.68. The molecular weights (Mw) and polydispersity index (PDI, Mw/Mn) were determined using gel permeation chromatography (GPC). Glass transition temperatures (Tg) for different compositions of E/A copolymers were determined using differential scanning calorimetry (DSC). Copolymer molar outfeed ratio (FE) was calculated from 1H NMR spectra. The α‐methyl, methine, backbone methylene, and quaternary carbon resonance signals of the copolymers were distinguished using 13C{1H}, DEPT‐45, ‐90, and ‐135 NMR techniques. The α‐methyl and β‐methylene showed compositional and configurational sensitivity up to pentad and tetrad level, respectively, whereas methine showed only compositional sensitivity up to pentad level. Unambiguous assignments for 1H and 13C{1H} NMR spectra were done by correlating 1D (1H, 13C{1H}, DEPT) and 2D (HSQC, TOCSY) NMR data. The spectral assignments for carbonyl region were done by studying higher bond order couplings by heteronuclear multibond correlation (HMBC) spectra. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 102: 5595–5606, 2006  相似文献   

18.
A significant improvement in the treatment of trypanosomiases has been achieved with the recent development of nifurtimox–eflornithine combination therapy (NECT). As an alternative to drug combinations and as a means to overcome most of the antitrypanosomatid drug discovery challenges, a multitarget drug design strategy has been envisaged. To begin testing this hypothesis, we designed and developed a series of quinone–coumarin hybrids against glyceraldehyde‐3‐phosphate dehydrogenase/trypanothione reductase (GAPDH/TR). These enzymes belong to metabolic pathways that are vital to Trypanosoma brucei and Trypanosoma cruzi, and have thus been considered promising drug targets. The synthesized molecules were characterized for their dual‐target antitrypanosomal profile, both in enzyme assays and in in vitro parasite cultures. The merged derivative 2‐{[3‐(3‐dimethylaminopropoxy)‐2‐oxo‐2H‐chromen‐7‐yl]oxy}anthracene‐1,4‐dione ( 10 ) showed an IC50 value of 5.4 μM against TbGAPDH and a concomitant Ki value of 2.32 μM against TcTR. Notably, 2‐{4‐[6‐(2‐dimethylaminoethoxy)‐2‐oxo‐2H‐chromen‐3‐yl]phenoxy}anthracene‐1,4‐dione (compound 6 ) displayed a remarkable EC50 value for T. brucei parasites (0.026 μM ) combined with a very low cytotoxicity toward mammalian L6 cells (7.95 μM ). This promising low toxicity of compound 6 might be at least partially due to the fact that it does not interfere with human glutathione reductase.  相似文献   

19.
D ,L ‐lactide (LA) was first successfully ring‐opening polymerized in melt by Schiff base complexes K[ML]nH2O [M = Cu(II), Zn(II), Co(II), Ni(II); n = 2, 2, 3, 3.5; H3L = L‐aspartic acid‐salicylidene Schiff base], which were prepared by Schiff base ligand derived from salicylidene and L‐aspartic acid and corresponding acetates. The effects of various complexes, the molar ratio of K[ML]nH2O/LA, the polymerization temperature, and time were studied in detail. The results show that all complexes studied have the ability to initiate the ring‐opening polymerization of D ,L ‐lactide in melt. More than 90% high polymerization conversion and narrow molecular weight distribution (MWD) can be obtained very easily. However, the Ni(II) complex shows better catalytic property than other complexes on the polymerization and the molecular weight (MW) of poly(D ,L ‐lactide) (PLA) produced. With a rise in temperature and a prolongation of time, the MW of PLA decreased remarkably. The MW of PLA prepared by all complexes is not very high, which might be related to the crystalline water of complexes. X‐ray study indicated that PLA produced by Ni(II) complex is an amorphous polymer. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 86: 3312–3315, 2002  相似文献   

20.
Enzyme efficiency results from the cooperation of functional groups in the catalytic site. In order to mimic a natural enzyme, a definite 3D scaffold must be carefully designed so that the functional groups can work cooperatively. During the HIV‐1 fusion process, the gp41 N‐ and C‐terminal heptad repeat regions form a coiled‐coil six‐helical bundle (6HB) that brings the viral and target cell membranes into close proximity for fusion. We used 6HB as the molecular model for a novel scaffold for the design of an artificial enzyme, in which the modified C34 and N36 peptides formed a unique 6HB structure through specific molecular recognition, and the position and orientation of the side‐chain groups on this scaffold were predictable. The histidine modified 6HB C34H13/20/N36H15/22 showed enzyme‐like hydrolytic activity towards p‐nitrophenyl acetate (PNPA; kcat/KM=3.66 M ?1 s?1) through the cooperation of several inter‐ or intrahelical imidazole groups. Since the catalytic activity of 6HB depends on the C‐ and N‐peptide assembly, either HIV fusion inhibitors that can compete with the formation of catalytic 6HB or denaturants that can destroy the ordered structure were able to modulate its activity. Further engineering of the solvent‐exposing face with Glu?‐Lys+ salt bridges enhanced the helicity and the stability of 6HB. As a result, the population and stability of cooperative catalytic units increased. In addition, the Glu?‐Lys+‐stabilized 6HB SC35H13/20/N36H15/22 had increased catalytic efficiency (kcat/KM=6.30 M ?1 s?1). A unique 6HB system was specifically assembled and provided a scaffold sufficiently stable to mimic the function of enzymes or other biomolecules.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号