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 15 毫秒
1.
To date, most spectroscopic studies on mammalian purple acid phosphatases (PAPs) have been performed at a single pH, typically pH 5. The catalytic activity of these enzymes is, however, pH dependent, with optimal pH values of 5.5–6.2 (depending on the form). For example, the pH optimum of PAPs isolated as single polypeptides is around pH 5.5, which is substantially lower that of proteolytically cleaved PAPs (ca. pH 6.2). In addition, the catalytic activity of single polypeptide PAPs at their optimal pH values is four to fivefold lower than that of the proteolytically cleaved enzymes. In order to elucidate the chemical basis for the pH dependence of these enzymes, the spectroscopic properties of both the single polypeptide and proteolytically cleaved forms of recombinant human PAP (recHPAP) and their complexes with inhibitory anions have been examined over the pH range 4 to 8. The EPR spectra of both forms of recHPAP are pH dependent and show the presence of three species: an inactive low pH form (pH<pK a,1), an active form (pK a,1<pH<pK a,2), and an inactive high pH form (pH>pK a,2). The pK a,1 values observed by EPR for the single polypeptide and proteolytically cleaved forms are similar to those previously observed in kinetics studies. The spectroscopic properties of the enzyme–phosphate complex (which should mimic the enzyme–substrate complex), the enzyme–fluoride complex, and the enzyme–fluoride–phosphate complex (which should mimic the ternary enzyme–substrate–hydroxide complex) were also examined. EPR spectra show that phosphate binds to the diiron center of the proteolytically cleaved form of the enzyme, but not to that of the single polypeptide form. EPR spectra also show that fluoride binds only to the low pH form of the enzymes, in which it presumably replaces a coordinated water molecule. The binding of fluoride and phosphate to form a ternary complex appears to be cooperative.Electronic Supplementary Material Supplementary material is available for this article at  相似文献   

2.
A simple method of hollow fiber-liquid phase microextraction (HF-LPME) combined with gas chromatography (GC) was developed for the analysis of four phenothiazine drugs (promethazine, promazine, chlorpromazine and trifluoperazine) in human urine samples. All variables affecting the extraction of target analytes including organic solvent type, stirring rate, extraction time, extraction temperature, pH of sample solution and ionic strength were carefully studied and optimized. Under the optimal conditions, the analytical performance of HF-LPME-GC-flame photometric detector (FPD) and HF-LPME-GC-flame ionization detector (FID) were evaluated and compared. The results showed that the HF-LPME-GC-FID was more sensitive than HF-LPME-GC-FPD for the determination of four target phenothiazine drugs, while the signal peak shape and resolution obtained by HF-LPME-GC-FPD was better than that obtained by HF-LPME-GC-FID. HF-LPME-GC-FPD/FID was successfully applied for the assay of the interested phenothiazine drugs in urine sample, and the excretion of the drugs was also investigated by monitoring the variation of the concentration of chlorpromazine in urine of a psychopath within 8 h after drug-taking. The proposed method provided an effective and fast way for the therapeutic drug monitoring (TDM) of phenothiazine.  相似文献   

3.
The pre-steady states of Pseudomonas species lipase inhibitions by p-nitrophenyl-N-substituted carbamates (1–6) are composed of two steps: (1) formation of the non-covalent enzyme–inhibitor complex (E:I) from the inhibitor and the enzyme and (2) formation of the tetrahedral enzyme–inhibitor adduct (E–I) from the E:I complex. From a stopped-flow apparatus, the dissociation constant for the E:I complex, KS, and the rate constant for formation of the tetrahedral E–I adduct from the E:I complex, k2 are obtained from the non-linear least-squares of curve fittings of first-order rate constant (kobs) versus inhibition concentration ([I]) plot against kobs=k2+k2[I]/(KS+[I]). Values of pKS, and log k2 are linearly correlated with the σ* values with the ρ* values of −2.0 and 0.36, respectively. Therefore, the E:I complexes are more positive charges than the inhibitors due to the ρ* value of −2.0. The tetrahedral E–I adducts on the other hand are more negative charges than the E:I complexes due to the ρ* value of 0.36. Formation of the E:I complex from the inhibitor and the enzyme are further divided into two steps: (1) the pre-equilibrium protonation of the inhibitor and (2) formation of the E:I complex from the protonated inhibitor and the enzyme.  相似文献   

4.
Minimal inhibitory concentrations (MICs) of 14 first and second-line antituberculous drugs against drug-susceptible and drug-resistant clinical isolates of Mycobacterium tuberculosis (including the multiple drug-resistant or MDR-TB isolates), as well as the type strain H37Rv, were determined radiometrically by the Bactec 460-TB methodols. MICs (μg/ml) of all the fourteen drugs were within an extremely narrow range in case of susceptible strains; isoniazid (0.02–0.04), rifampin (0.2–0.4), ethambutol and streptomycin (0.5–2.0), ethionamide (0.25–0.5), D-cycloserine (25–75), capreomycin (1–2), kanamycin (2–4), amikacin (0.5–1.0), clofazimine (0.1–0.4), ofloxacin (0.5–1.0), ciprofloxacin (0.25–1.0), and sparfloxacin (0.1–0.4). The activity of second-line drugs remained unaltered against MDR-TB isolates resistant to routine first-line drugs. With peak serum level concentrations (Cmax), the intracellular killing of the virulent H37Rv strain was studied in detail in cultured human macrophages. Based on an decreasing order of bactericidal activity, our results showed the following spectrum of intracellular drug action: among the first-line drugs, rifampin > ethionamide = isoniazid > ethambutol > streptomycin > D-cycloserine; among second-line drugs, clofazimine = amikacin > kanamycin = capreomycin; among fluoroquinolones, sparfloxacin > ofloxacin > ciprofloxacin. On the other hand, contrary to atypical mycobacteria, the macrolide drug clarithromycin was inactive against both extracellular and intracellular M. tuberculosis. Received: 23 January 1996 / Accepted: 5 April 1996  相似文献   

5.
The dissociation constants (pKms) of the phenothiazine drugs promazine, chlorpromazine, and triflupromazine, incorporated in the phosphatidylcholine (PC) bilayer of small unilamellar vesicles (SUV), were investigated by a 13C nuclear magnetic resonance (NMR) titration method employing their N-13CH3 (ionizable group) labelled derivatives. Use of the labelled drugs enabled direct observations of the ionization equilibrium of the N-dimethyl group. A second derivative spectrophotometric study proved that 95-98% of the phenothiazine species in the sample solutions (200 μM phenothiazine in the presence of 27 mM PC SUV) were incorporated into the PC bilayer, which simplified the calculation of pKm values by allowing that the phenothiazines in the aqueous phase could be neglected. The pKm values were calculated from the chemical shift dependence of the N-dimethyl 13C NMR signal on the pH value of sample solutions. The pKm values obtained were smaller than those measured in aqueous solutions by about one unit. The existence of cholesterol (30 mol%) in the PC bilayer showed little effect on the pKm values, suggesting that cholesterol in the bilayer does not largely affect the interfacial region where the N-dimethyl group of the incorporated phenothiazines is located. The results offered clear evidence for the pKm decrease and provided their precise values.  相似文献   

6.
The antipsychotic phenothiazines may have other therapeutic applications because of their ability to kill bacteria, plasmids and tumor cells. They are also known to undergo a peroxidase-catalysed oxidation to form cation radicals that are stable at acid pH, but are not detected at a neutral pH. The objective of this project was to determine whether phenothiazine cation radical metabolites could cause oxidative stress at a neutral pH resulting in cytotoxicity. At a neutral pH, catalytic amounts of phenothiazines were found to be oxidised by a peroxidase/H2O2 system and also caused ascorbate, GSH and NADH cooxidation. NADH and GSH co-oxidation was accompanied by oxygen uptake and was increased by the addition of catalytic amounts of superoxide dismutase, indicating that the superoxide radical was formed. The phenothazines were different from other peroxidase substrates in that the NADH, ascorbate or GSH cooxidation was faster at pH 6.0 than pH 7.4, thereby partly reflecting the cation radical stability. The order of catalytic effectiveness found was promazine > chlorpromazine > trifluoperazine. Peroxidase/H2O2 also markedly increased phenothiazine cytotoxicity towards isolated rat hepatocytes at nontoxic phenothiazine concentrations. At both pH 6.0 and 7.4, the same order of phenothiazine catalytic effectiveness was observed as seen in the co-oxidation experiments. Cytotoxicity to hepatocytes could be attributed to oxidative stress as most hepatocyte glutathione oxidation and lipid peroxidation preceded phenothiazine induced cytotoxicity and that cytotoxicity was prevented by the antioxidant butylated hydroxyanisole. This hepatocyte/peroxidase/H2O2 system could be a useful model for studying drug induced idiosyncratic hepatic injury enhanced by inflammation.  相似文献   

7.
Enzyme access, kinetic behavior, and protein–protein interactions are critical for explaining reaction of the metabolites contained within the myriad compartments of biological systems. To explore these relationships, the reaction kinetics of oil bodies versus oil emulsions as substrates for lipolytic reactions were measured. The initial rate of hydrolysis for the oil body system was comparatively very low due to a brief latency period. However, the complete activation of the lipase at the interface resulted in an enzyme–membrane complex that was catalytically enhanced 3–15-fold over the emulsion system for substrate concentrations in the measured range of approximately 1–5.5 mM. This disparity is explained by the availability of substrate to the enzyme active site (defined as the availability parameter “A”) which varies between the two substrates by 40-fold. A simple hyperbolic kinetic mechanism is proposed with K m replaced by the parameter, A, to account for this phenomenon, leading to a maximum rate of approximately 1450 IU/mg protein. The interaction is verified through separation of the enzyme–membrane complex which shows nearly double the activity towards an emulsified soybean oil substrate (activity ratio of 5:3) when compared to the native enzyme.  相似文献   

8.
Stereoselectivity of reversible inhibition of butyrylcholinesterase (BChE; EC 3.1.1.8) by optically pure ethopropazine [10-(2-diethylaminopropyl)phenothiazine hydrochloride] enantiomers and racemate was studied with acetylthiocholine (0.002–250 mM) as substrate. Molecular modelling resulted in the reaction between BChE and ethopropazine starting with the binding of ethopropazine to the enzyme peripheral anionic site. In the next step ethopropazine ‘slides down’ the enzyme gorge, resulting in interaction of the three rings of ethopropazine through π–π interactions with W82 in BChE. Inhibition mechanism was interpreted according to three kinetic models: A, B and C. The models differ in the type and number of enzyme–substrate, enzyme–inhibitor and enzyme–substrate–inhibitor complexes, i.e., presence of the Michaelis complex and/or acetylated BChE. Although, all three models reproduced well the BChE activity in absence of ethopropazine, model A was poor in describing inhibition with ethopropazine, while models B and C were better, especially for substrate concentrations above 0.2 mM. However model C was singled out because it approaches fulfilment of the one step-one event criteria, and confirms the inhibition mechanism derived from molecular modelling. Model C resulted in dissociation constants for the complex between BChE and ethopropazine: 61, 140 and 88 nM for R-enantiomer, S-enantiomer and racemate, respectively. The respective dissociation constants for the complexes between acetylated BChE and ethopropazine were 268, 730 and 365 nM. Butyrylcholinesterase had higher affinity for R-ethopropazine.  相似文献   

9.
Commercial available lipases viz. Lipozyme™, Novozyme-735 and Candida antartica lipase-B (CAL-B) were immobilized on seven different supports by simple adsorption process. The importance of suitable enzyme–support combination in esterification of lauric acid and iso-propanol was validated experimentally. Effect of long chain fatty acids (C4–C18) and small chain monohydric alcohols (C1–C6) on specific activities of different immobilized lipases were evaluated. Lauric acid (C12) was found to be the most preferred fatty acid and t-amyl alcohol (C5) being the best alcohol. CAL-B adsorbed on Lewatit was the most efficient immobilized enzyme for esterification reaction. Selectivity constant for lauric acid (3.4) was the highest among all fatty acids tested, whereas there was not much difference in selectivity between different alcohols. Furthermore, increase in fatty acid unsaturation leads to decrease catalytic efficiency of immobilized CAL-B. The optimum conditions for t-amyllaurate synthesis were as follows: lauric acid—0.5 M, t-amyl alcohol—0.3 M and amount of immobilized enzyme—150 mg. Finally, CAL-B adsorbed on Lewatit was reused for three consecutive cycles.  相似文献   

10.
This study shows the preparation and application of enzyme–nanofiber composites for long-term stable operation. The enzyme–nanofiber composite was prepared by coating an enzyme aggregate, the esterase from Rhizopus oryzae, on the surface of the nanofibers. After immobilization on the nanofiber, the apparent K m for the immobilized esterase was 1.48-fold higher than that of the free esterase, with values of 0.98 and 1.35 mM for the free and immobilized enzymes, respectively. It was found that enzyme–nanofiber was very stable, even when the fibers were shaken in glass vials, preserving 80% of the initial activity for 100 days. In addition, the enzyme–nanofiber composite was used repeatedly in 30 cycles of substrate hydrolysis and still remained active. Consequently, the esterase–nanofiber composite was employed within a continuous reactor system to evaluate its use in a long-term and stable continuous substrate hydrolysis reaction. It was found that the production of p-nitrophenol was stable for at least 400 h. This study demonstrates that the enzyme–nanofiber composite can be used in both repeated-batch mode and a continuous mode for a long-term stable operation.  相似文献   

11.
We report geometries, stabilization energies, symmetry adapted perturbation theory (SAPT) and quantum theory of atoms in molecules (QTAIM) analyses of a series of carbene–BX3 complexes, where X = H, OH, NH2, CH3, CN, NC, F, Cl, and Br. The stabilization energies were calculated at HF, B3LYP, MP2, MP4 and CCSD(T)/aug-cc-pVDZ levels of theory using optimized geometries of all the complexes obtained from B3LYP/aug-cc-pVTZ. Quantitatively, all the complexes indicate the presence of B–Ccarbene interaction due to the short B–Ccarbene distances. Inspection of stabilization energies reveals that the interaction energies increase in the order NH2 > OH > CH3 > F > H > Cl > Br > NC > CN, which is the opposite trend shown in the binding distances. Considering the SAPT results, it is found that electrostatic effects account for about 50% of the overall attraction of the studied complexes. By comparison, the induction components of these interactions represent about 40% of the total attractive forces. Despite falling in a region of charge depletion with ∇2 ρ BCP >0, the B–Ccarbene bond critical points (BCPs) are characterized by a reasonably large value of the electron density (ρ BCP) and HBCP <0, indicating that the potential energy overcomes the kinetic energy density at BCP and the B–Ccarbene bond is a polar covalent bond.  相似文献   

12.
The antibacterial effect of Zn(II), tetraphenyl porphyrin (TPP), propdioxyl bridged tetraphenyl bisporphyrin 1, and its metallobisporphyrin complexes (ZnMnbisporphyrin 2 and ZnZnbisporphyrin 3) towards Staphylococcus aureus growth was investigated by microcalorimetry at 37°C. Differences in their capacities to inhibit the growth metabolism of S. aureus were observed. By analyzing the power–time curves, crucial parameters such as the rate constant of bacterial growth (k), inhibitory ratio (I), and generation time (t G) were determined. The growth rate constant (k) of S. aureus (in the log phase) in the presence of the drugs decreased linearly with increasing concentrations of the complexes. The sequence of the antibacterial activities of these compounds tested was 3 > 2 > 1 > Zn(II) > TPP. ZnZnbisporphyrin 3 is proposed to benefit from the synergetic effects of Zn(II) and 1.  相似文献   

13.
Phenothiazine cation radicals (PTZ + •) irreversibly inactivated Trypanosoma cruzi dihydrolipoamide dehydrogenase (LADH). These radicals were obtained by phenothiazine (PTZ) peroxidation with myeloperoxidase (MPO) or horseradish peroxidase (HRP/H 2 O 2 ) systems. LADH inactivation depended on PTZ structure and incubation time. After 10 min incubation of LADH with the MPO-dependent systems, promazine, trimeprazine and thioridazine were the most effective; after 30 min incubation, chlorpromazine, prochlorperazine and promethazine were similarly effective. HRP-dependent systems were equally or more effective than the corresponding MPO-dependent ones. Chloro, trifluoro, propionyl and nitrile groups at position 2 of the PTZ ring significantly decreased molecular activity, specially with the MPO/H 2 O 2 systems. Comparison of inactivation values for LADH and T. cruzi trypanothione reductase demonstrated a greater sensitivity of LADH to chlorpromazine and perphenazine and a 10-fold lower sensitivity to promazine, thioridazine and trimeprazine. Alkyl-amino, alkyl-piperidinyl or alkyl-piperazinyl groups at position 10 modulated PTZ activity to a limited degree. Production of PTZ + • radicals was demonstrated by optical and ESR spectroscopy methods. PTZ + • radicals stability depended on their structure as demonstrated by promazine and thioridazine radicals. Thiol compounds such as GSH and N -acetylcysteine, l -tyrosine, l -tryptophan, the corresponding peptides, ascorbate and Trolox, prevented LADH inactivation by the MPO/H 2 O 2 /thioridazine system, in close agreement with their action as PTZ + • scavengers. NADH (not NAD + ) produced transient protection of LADH against thioridazine and promazine radicals, the protection kinetics being affected by the relatively fast rate of NADH oxidation by these radicals. The role of the observed effects of PTZ radicals for PTZ cytotoxicity is discussed.  相似文献   

14.
Heteropolyoxometalate complexes have been widely applied in many fields. In this paper, the interaction between a series of novel rare earth molybdotungstosilicate heteropolyoxometalates, K10H3[Ln(SiMo6W5O39)2]·xH2O (abbr. LnW5, Ln = Pr (x = 30), Gd (x = 29), Dy (x = 28), and Yb (x = 31)), and bovine serum albumin (BSA) was investigated by spectroscopic approach under the physiological conditions. In the mechanism discussion, it was proved that the fluorescence quenching of BSA by LnW5 is a result of the formation of LnW5–BSA complex. Fluorescence quenching constants were determined using the Stern–Volmer equation to provide a measure of the binding affinity between LnW5 and BSA. The binding affinity ranked in the order GdW5 > DyW5 > PrW5 > YbW5. The results of thermodynamic parameters ΔG, ΔH, and ΔS at different temperatures indicate that van der Waals interactions and hydrogen bonds play a major role for LnW5–BSA association. Furthermore, the distance r between donor (BSA) and acceptor (LnW5) was obtained according to fluorescence resonance energy transfer.  相似文献   

15.
Type I Baeyer–Villiger monooxygenases (BVMOs) strongly prefer NADPH over NADH as an electron donor. In order to elucidate the molecular basis for this coenzyme specificity, we have performed a site-directed mutagenesis study on phenylacetone monooxygenase (PAMO) from Thermobifida fusca. Using sequence alignments of type I BVMOs and crystal structures of PAMO and cyclohexanone monooxygenase in complex with NADP+, we identified four residues that could interact with the 2′-phosphate moiety of NADPH in PAMO. The mutagenesis study revealed that the conserved R217 is essential for binding the adenine moiety of the nicotinamide coenzyme while it also contributes to the recognition of the 2′-phosphate moiety of NADPH. The substitution of T218 did not have a strong effect on the coenzyme specificity. The H220N and H220Q mutants exhibited a ~3-fold improvement in the catalytic efficiency with NADH while the catalytic efficiency with NADPH was hardly affected. Mutating K336 did not increase the activity of PAMO with NADH, but it had a significant and beneficial effect on the enantioselectivity of Baeyer–Villiger oxidations and sulfoxidations. In conclusion, our results indicate that the function of NADPH in catalysis cannot be easily replaced by NADH. This finding is in line with the complex catalytic mechanism and the vital role of the coenzyme in BVMOs.  相似文献   

16.
The intracellular localization of two neurotropic drugs, flunitrazepam (benzodiazepine) and triflupromazine (phenothiazine), was studied by secondary ion mass spectrometry microscopy (SIMS) in three varieties of cells. The images of the intracellular distributions of the two drugs are easily obtained by selecting the fluorine atom of the molecules. These images show that the drug from the benzodiazepine group is mainly located in the nuclei, whereas the phenothiazine is exclusively located inside the cytoplasm.  相似文献   

17.
Selectivity of nuclear probes is controlled by competitive accumulation of the probe by cellular organelles as well as the high affinity for nucleic acids. Physicochemical features of probes which favor nucleic acid binding include cationic character and a planar aromatic system above a minimum size. Features of probes which permit entry into cells are low protein and lipid binding. Features which reduce accumulation in non-nuclear sites include high base strength and hydrophilicity of the cation. The overall quantitative structure–activity (QSAR) model specifying nuclear accumulation may be expressed as follows: CBN<40; 8>log P neutral species>0; AI<8; Z>0; -5<log P cation<0; pK a >10; LCF>17; LCF/CBN>0.70 (where CBN is the conjugated bond number, log P x the logarithm of the water–octanol partition coefficient of species x, AI the amphilicity index, Z the electric charge, pK a the negative logarithm of the equilibrium constant for the free base–protonated base reaction, and LCF the largest conjugated fragment). Preliminary applications of the QSAR model—to the selection of anticancer drugs, minimization of dye and drug toxicity and the designed synthesis of fluorescent probes—are outlined.  相似文献   

18.
Drew DP  Lunde C  Lahnstein J  Fincher GB 《Planta》2007,225(4):945-954
Monodehydroascorbate reductase (MDHAR; EC 1.6.5.4) catalyses the reduction of the monodehydroascorbate (MDHA) radical to ascorbate, using NADH or NADPH as an electron donor, and is believed to be involved in maintaining the reactive oxygen scavenging capability of plant cells. This key enzyme in the ascorbate-glutathione cycle has been studied here in the moss Physcomitrella patens, which is tolerant to a range of abiotic stresses and is increasingly used as a model plant. In the present study, three cDNAs encoding different MDHAR isoforms of 47 kDa were identified in P. patens, and found to exhibit enzymic characteristics similar to MDHARs in vascular plants despite low-sequence identity and a distant evolutionary relationship between the species. The three cDNAs for the P. patens MDHAR enzymes were expressed in Escherichia coli and the active enzymes were purified and characterized. Each recombinant protein displayed an absorbance spectrum typical of flavoenzymes and contained a single non-covalently bound FAD coenzyme molecule. The K m and k cat values for the heterologously expressed PpMDHAR enzymes ranged from 8 to 18 μM and 120–130 s−1, respectively, using NADH as the electron donor. The K m values were at least an order of magnitude higher for NADPH. The K m values for the MDHA radical were ∼0.5–1.0 μM for each of the purified enzymes, and further kinetic analyses indicated that PpMDHARs follow a ‘ping–pong’ kinetic mechanism. In contrast to previously published data, site-directed mutagenesis indicated that the conserved cysteine residue is not directly involved in the reduction of MDHA.  相似文献   

19.
Jin  Xin  Wang  Xueye  Ren  Cuihuan  Miao  Yuan  Yi  Ling 《Journal of molecular modeling》2011,17(4):913-920
The inclusion complex of β-cyclodextrin (β-CD) and 4-methylbenzyl-4′-[trans-4- (guanidinomethyl)cylohexylcarbonyloxy]-biphenyl-4-carboxlylate monohydrochloride (TG44) had been investigated by using densify functional theory (DFT) and PM3 semiempirical method. The results indicate that the β-CD includes predominantly the biphenyl moiety of TG44, and the inclusion complex formed by TG44 entering into the cavity of β-CD from its narrow side (the primary hydroxyl group side) is more stable than that formed by TG44 entering into the cavity of β-CD from its wide side (the secondary hydroxyl group side). The negative enthalpy changes calculated from the statistical thermodynamic calculations at 1 atm and 298.15 K suggest that the inclusion complexes are favored enthalpy-driven processes. The molecular modeling results are in good agreement with the experiment for 2D 1H–13C H HETCOR spectroscopic and H-NMR spectroscopic observations.  相似文献   

20.
A simple and rapid method for the determination of seven phenothiazines derivatives (chlorpromazine, promethazine, levomepromazine, prochlorperazine, trifluoperazine, fluphenazine and thioridazine) in human urine samples is presented. The analytes are extracted from the sample in 50 μL of the ionic liquid 1-butyl-3-methyl-imidazolium hexafluorophosphate working in an automatic flow system under dynamic conditions. The chemical affinity between the extractant and the analytes allows a good isolation of the drugs from the sample matrix achieving at the same time their preconcentration. The separation and detection of the extracted compounds is accomplished by liquid chromatography and UV detection. The proposed method is a valuable alternative for the analysis of these drugs in urine within the concentration range 0.07–10 μg mL?1. Limits of detection were in the range from 21 ng mL?1 (thioridazine) to 60 ng mL?1 (levomepromazine). The repeatability of the proposed method expressed as RSD (n = 5) varied between 2.2% (levomepromazine) and 3.9% (chlorpromazine).  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