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.
采用辣根过氧化物酶顺、逆行标记方法对鸣禽鸟蜡嘴雀控制发声的神经核团、脑干听觉核团及神经通路,从外周至中枢逐级进行了追踪研究。结果表明:1.控制发声的神经核团及通路,前脑古纹状体腹内侧粗核是大脑控制发声的重要核团之一,它发出枕中脑后束经端脑前联合呈双侧支配延脑中间核,中间核又发出舌下神经经气管鸣管分支支配鸣肌,中间核同时也接受中脑背内侧核的支配,2.脑干听觉中枢及通路,中脑背外侧核是脑干较高级听觉中枢、初级中枢耳蜗核由角核和前庭外侧核组成,NA发出以对侧为主的纤维经外侧丘系可直接传入中脑背外侧核形成脑干听觉直接通路。  相似文献   

2.
A hydrogen gas (H(2)) biosensor was developed in which hydrogenase (H(2)ase) was immobilized and sandwiched between two layers of a montmorillonite clay and poly(butylviologen) (PBV) mixture on a glass carbon electrode. The immobilized PBV efficiently enhanced the electron transfer among the electrode, H(2)ase, and methyl viologen in solution. Both PBV and methyl viologen acted as the electron carrier in the clay-PBV-H(2)ase modified electrode. The clay-PBV-H(2)ase electrode catalyzed the oxidation of H(2) to protons (H(+)) with the electrons being transferred by viologen groups to the electrode. The activation energy of this process was 38+/-2 kJ/mol at pH 7. The catalytic current of the clay-PBV-H(2)ase electrode increased linearly when exposed to increasing concentrations of H(2) gas. In contrast, this electrode showed no activity when exposed to three combustible compounds, namely, carbon monoxide, methane and methanol. The optimum pH range for the oxidation of H(2) by the clay-PBV-H(2)ase electrode was from 7 to 10. Electron transfer process in the clay-PBV-H(2)ase electrode is discussed.  相似文献   

3.
A novel biosensor for superoxide radical (O(2)(*-)) detection based on Pseudomonas aeruginosa azurin immobilized on gold electrode was designed. The rate constant of azurin reduction by O(2)(*-) was found to be 10(5)M(-1)s(-1) in solution and five times lower, i.e., 0.2 x 10(5)M(-1)s(-1), for azurin coupled to gold by 3,3'-dithiobis(sulfosuccinimidylpropionate) (DTSSP). The electron transfer rate between the protein and the electrode ranged from 2 to 6s(-1). The sensitivity of this biosensor to O(2)(*-) was 6.8 x 10(2)Am(-2)M(-1). The response to the interference substances, such as uric acid, H(2)O(2), and dimethylsulfoxide was negligible below 10 microM. The electrode was applied in three O(2)(*-) generating systems: (i) xanthine oxidase (XOD), (ii) potassium superoxide (KO(2)), and (iii) stimulated neutrophil granulocytes. The latter was compared with luminol-amplified chemiluminescence. The biosensor responded to O(2)(*-) in all three environments, and the signals were antagonized by superoxide dismutase.  相似文献   

4.
The susceptibility of Trichoplusia ni larvae to several Bacillus thuringiensis insecticidal crystal proteins (ICPs) was tested. Neonatal larvae proved to be susceptible to solubilized trypsin-treated CryIA(a), CryIA(b), and CryIA(c) (50% lethal concentrations [LC(50)s], 570, 480, and 320 ng/cm, respectively) but showed little susceptibility to CryIB and CryID (LC(50)s, 5,640 and 2,530 ng/cm, respectively). The toxicity of ICPs was correlated to binding to the epithelial brush border of the midgut, as revealed by immunocytochemical staining with monoclonal antibodies. In vitro binding experiments with iodinated ICPs and brush border membrane vesicles indicated that CryIA(b) and CryIA(c) share the same high-affinity binding site, whereas CryIA(a) binds to a different one. The affinities of CryIA(b) and CryIA(c) for the binding site were similar (K(d) = 3.6 and 4.7 nM, respectively), and the mean binding-site concentration was 0.71 pmol/mg of vesicle protein. Selection of a population with increasing concentrations of CryIA(b) produced 31-fold resistance in seven generations. The realized heritability (h) was 0.19. The increase of homozygosity (for resistance factors) as selection proceeded was reflected in the increase in the slopes of the dose-mortality curves. Resistance was specific for CryIA(b) and did not extend to CryIA(a) or even to CryIA(c). This result was not predicted by the binding-site model, in which CryIA(b) and CryIA(c) bind to the same high-affinity binding site. This result may suggest a more complicated relationship between in vitro binding of ICPs to specific sites in the epithelial membrane of the midgut and the in vivo toxic effect.  相似文献   

5.
This study investigated the effect of rice straw char (RSC) on the immobilization of Cr(VI) in soils. The Cr(VI) sorption experiments on the RSC and RSC-amended soils were conducted using the batch method. RSC exhibited Cr(VI) reduction capacity due to its black carbon content. The addition of RSC to the soils enhanced the overall Cr(VI) immobilization of the soils, which is primarily attributed to the Cr(VI) reduction capacity of RSC. The effects of RSC amendment on the Cr(VI) sorption of the soils increased with increasing RSC content in the soils and decreased with increasing pH or anion contents in the soil solutions. After Cr(VI) was sorbed by the soils, a portion of the Cr(VI) was converted to Cr(III) and the remainder was sorbed onto the soils. The presence of RSC in the soils decreased the portion of sorbed Cr(VI) in the soils and therefore lowered the potential remobilization of Cr(VI) from the soils. The results suggested that RSC amendment can be applied to develop a cost-effective method for immobilizing Cr(VI) in polluted soils, thus lowering the environmental risk from Cr(VI) toxicity.  相似文献   

6.
Hay S  Westerlund K  Tommos C 《Biochemistry》2005,44(35):11891-11902
De novo protein design and electrochemistry were used to measure changes in the potential and pK(A) of a phenol when its OH group is moved from a solvent-exposed to a sequestered protein position. A "phenol rotation strategy" was adopted in which phenols, containing a SH in position 4, 3, or 2 relative to the OH group, were bound to a buried protein site. The alpha(3)C protein used here is a tryptophan to cysteine variant of the structurally defined alpha(3)W protein (Dai et al. (2002) J. Am. Chem. Soc. 124, 10952-10953). The protein characteristics of alpha(3)C and the three mercaptophenol-alpha(3)C (MP-alpha(3)C) proteins are shown to be close to those of alpha(3)W. Moreover, the phenol OH group is fully solvent exposed in 4MP-alpha(3)C and more sequestered in 3MP-alpha(3)C and 2MP-alpha(3)C. Here we compare the redox properties of the three mercaptophenols when bound to alpha(3)C and to cysteine free in water. The pK(A) and E(peak) values are essential identical when 4MP is ligated to alpha(3)C relative to when it is free in solution. In contrast, these values are increased in 3MP-alpha(3)C and 2MP-alpha(3)C relative to the solvated compounds. The E(peak) vs pH plots all display a approximately 59 mV/pH unit dependence. We conclude that interactions with the OH group dominate the phenol redox characteristics. In 3MP-alpha(3)C and 2MP-alpha(3)C, hydrogen bonds between the protein and the bound phenols appear to either stabilize the reduced phenol or destabilize the radical, relative to the aqueous buffer, raising the potential by 0.11 and 0.12 V, respectively.  相似文献   

7.
8.
The Km and Vmax values for primers d(pA)n, d(pT)n, r(pA)n, r(pU)n where n = 1-16, were compared. The Km values for minimal primers dTMP, dAMP, rUMP, rAMP were found to be 48, 71, 602 and 602 microM, respectively. The Vmax value for any NMP made up approximately 7% of that for (pN)10. The lengthening of any primer per one mononucleotide unit for n from 1 to 10 resulted in the decrease of the Km value 1.8-fold and the increase of the Vmax value 1.35-fold. The ratios of the Km values for primers r(pA)n-d(pA)n and r(pU)n-d(pT)n were 7.5 and 12.5, respectively, for any n. The Km value for [d[pT)8]r(pU) primer was the same as for r(pU)9, but not for d(pT)9. Decanucleotide [d(Tp)9]ddT interacted with the polymerase competitively to the template, but not to the primer. The primer's 3'-OH group was supposed to form the hydrogen bond with the enzyme. The absence of 3'-hydroxygroup in [d(Tp)9]ddT resulted in its inability to compete effectively with the primer. The difference of the affinity of ribo- and deoxyriboprimers is due, apparently, to the existence of the different conformation of the furanose rings in the ribose and deoxyribose.  相似文献   

9.
Eliot AC  Kirsch JF 《Biochemistry》2002,41(11):3836-3842
The active sites of the homologous pyridoxal phosphate- (PLP-) dependent enzymes 1-aminocyclopropane-1-carboxylate (ACC) synthase and aspartate aminotransferase (AATase) are almost entirely conserved, yet the pK(a)'s of the two internal aldimines are 9.3 and 7.0, respectively, to complement the substrate pK(a)'s (S-adenosylmethionine pK(a) = 7.8 and aspartate pK(a) = 9.9). This complementation is required for maximum enzymatic activity in the physiological pH range. The most prominent structural difference in the active site is that Ile232 of ACC synthase is replaced by alanine in AATase. The I232A mutation was introduced into ACC synthase with a resulting 1.1 unit decrease (from 9.3 to 8.2) in the aldimine pK(a), thus identifying Ile232 as a major determinant of the high pK(a) of ACC synthase. The mutation also resulted in reduced k(cat) (0.5 vs 11 s(-1)) and k(cat)/K(m) values (5.0 x 10(4) vs 1.2 x 10(6) M(-1) s(-1)). The effect of the mutation is interpreted as the result of shortening of the Tyr233-PLP hydrogen bond. Addition of the Y233F mutation to the I232A ACC synthase to generate the double mutant I232A/Y233F raised the pK(a) from 8.2 to 8.8, because the Y233F mutation eliminates the hydrogen bond between that residue and PLP. The introduction of the retro mutation A224I into AATase raised the aldimine pK(a) of that enzyme from 6.96 to 7.16 and resulted in a decrease in single-turnover k(max) (108 vs 900 s(-1) for aspartate) and k(max)/K(m)(app) (7.5 x 10(4) vs 3.8 x 10(5) M(-1) s(-1)) values. The distance from the pyridine nitrogen of the cofactor to a conserved aspartate residue is 2.6 A in AATase and 3.8 A in ACC synthase. The D230E mutation introduced into ACC synthase to close this distance increases the aldimine pK(a) from 9.3 to 10.0, as would be predicted from a shortened hydrogen bond.  相似文献   

10.
The decomposition of lipid hydroperoxides (LOOH) into peroxyl radicals is a potential source of singlet molecular oxygen ((1)O(2)) in biological systems. Recently, we have clearly demonstrated the generation of (1)O(2) in the reaction of lipid hydroperoxides with biologically important oxidants such as metal ions, peroxynitrite and hypochlorous acid. The approach used to unequivocally demonstrate the generation of (1)O(2) in these reactions was the use of an isotopic labeled hydroperoxide, the (18)O-labeled linoleic acid hydroperoxide, the detection of labeled compounds by HPLC coupled to tandem mass spectrometry (HPLC-MS/MS) and the direct spectroscopic detection and characterization of (1)O(2) light emission. Using this approach we have observed the formation of (18)O-labeled (1)O(2) by chemical trapping of (1)O(2) with anthracene derivatives and detection of the corresponding labeled endoperoxide by HPLC-MS/MS. The generation of (1)O(2) was also demonstrated by direct spectral characterization of (1)O(2) monomol light emission in the near-infrared region (lambda = 1270 nm). In summary, our studies demonstrated that LOOH can originate (1)O(2). The experimental evidences indicate that (1)O(2) is generated at a yield close to 10% by the Russell mechanism, where a linear tetraoxide intermediate is formed in the combination of two peroxyl radicals. In addition to LOOH, other biological hydroperoxides, including hydroperoxides formed in proteins and nucleic acids, may also participate in reactions leading to the generation (1)O(2). This hypothesis is currently being investigated in our laboratory.  相似文献   

11.
Structure of human oxyhaemoglobin at 2.1 A resolution   总被引:16,自引:0,他引:16  
The structure of human oxyhaemoglobin was determined by single crystal X-ray analysis at 2.1 A resolution. Data were collected on an Arndt-Wonacott camera at -2 degrees C. The structure was refined to an R factor of 0.223 by the Jack-Levitt method, starting from Baldwin's model of human carbon monoxide haemoglobin. The active sites in the alpha and beta subunit are distinct. The iron atoms are 0.16(8) A and 0.00(8) A from the mean plane of the porphyrin carbons and nitrogens (0.12(8) A and -0.11(8) A from the mean plane of the porphyrin nitrogens) in the alpha and beta subunit, respectively, in correlation with the orientation of HisF8 relative to the porphyrin nitrogens. The haem group appears to be nearly planar in the alpha subunit but ruffled in the beta subunit. The Fe-O(1)-O(2) angles are 153(7) degrees and 159(12) degrees in the alpha and beta subunit, respectively. The oxygen molecule forms a hydrogen bond to N epsilon of HisE7 in the alpha, but either none or a weak one in the beta subunit. The following bond lengths were found: Fe-N epsilon (HisF8) = 1.94(9) A (alpha) and 2.07(9) A (beta); Fe-O(1) = 1.66(8) A (alpha) and 1.87(13) A (beta); Fe-Nporph (mean = 1.99(5) A (alpha) and 1.96(6) A (beta). These dimensions agree with the values obtained in oxymyoglobin and model compounds. The C-terminal residues, ArgHC3(141 alpha) and HisHC3(146 beta), are relatively delocalized, and their positions do not enable them to form the intersubunit salt bridges in which they are involved in deoxyhaemoglobin. The penultimate tyrosine residues, TyrHC2 140 alpha and 145 beta, are relatively localized and maintain the hydrogen bonds to the carbonyl oxygens of ValFG5 (93 alpha and 98 beta), with only minor variations compared to their geometry in deoxyhaemoglobin. TyrHC2(145 beta), however, alternates between a major and a minor site, in conjunction with CysF9(93 beta), both sharing the internal pocket between the F and H helices while in the major conformation. This suggests that the role of the penultimate tyrosines in the allosteric mechanism may differ from that previously proposed by Perutz. The overall quaternary structure of oxyhaemoglobin is identical, within experimental error, to that of carbon monoxide haemoglobin, and thus confirms the applicability of the allosteric mechanisms proposed by Perutz and Baldwin & Chothia to the process of oxygen binding.  相似文献   

12.
The brain uptake and brain to blood efflux transport of (14)C-GABA were studied in spontaneously hypertensive rats (SHR) and normotensive Wistar Kyoto (WKY) rats using 20 min bilateral in situ brain perfusion in rats anesthetized using urethane. The volume of distribution (Vd) of (14)C-GABA into cerebrospinal fluid (CSF) and brain regions (cortex, diencephalon, cerebellum, and brain stem) was significantly greater in SHR than in the corresponding regions in WKY rats (p<0.05). The estimated Vd value of (14)C-GABA in CSF of SHR was 3.4 fold greater than that in WKY. Also compared to WKY, the Vd of (14)C-GABA into cerebellum and cortex of SHR was 15.3 fold and 19.4 fold greater, respectively. Although the study of blood-brain barrier (BBB) integrity using (3)H-mannitol revealed increased paracellular permeability at the brain capillaries of SHR when compared to WKY rats, this was found to be only partially responsible for the increased (14)C-GABA uptake. The study of brain to blood efflux transport of (14)C-GABA (after loading of brain with (14)C-GABA by vascular perfusion) revealed that the half-time of elimination was significantly shorter in SHR (5.35+/-0.66 min) than in WKY rats (14.83+/-1.94 min), (p<0.001). HPLC analysis revealed that GABA concentrations in brain extracts and CSF of SHR were similar to those in WKY rats (p>0.05). The faster efflux in SHR might be, at least partially, responsible to compensate for increased uptake of this neurotransmitter and to preserve the protective function of BBB towards GABA. The protective function of the BCSFB towards GABA appears to be also preserved, since systemic infusion of GABA within a wide range of administered doses (0.004-5.00 mg/kg) produced an increase in GABA CSF concentration from around 0.5 microM to only 11 microM, and the obtained pattern of CSF GABA concentrations under these conditions did not differ between SHR and WKY rats, as revealed by HPLC.  相似文献   

13.
A sensitive chemiluminescence method for vitamin B(12) using a charge-coupled device (CCD) photodetector combined with on-line UV-persulfate oxidation in a simple continuous flow system has been developed. The principle for the determination of vitamin B(12) is based on the enhancive effect of cobalt (II) on the chemiluminescence reaction between luminol and percarbonate in alkaline medium. In addition, percarbonate has been investigated and proposed as a powerful source of hydrogen peroxide as oxidant agent in this chemiluminescence reaction. The digestion of vitamin B(12) to release the cobalt (II) is reached by UV irradiation treatment in a persulfate medium. The CCD detector, directly connected to the flow cell, is used with the continuous flow manifold to obtain the full spectral characteristics of cobalt (II) catalyzed luminol-percarbonate reaction. The vitamin B(12) oxidation process and chemical conditions for the chemiluminescence reaction were investigated and optimized. The increment of the emission intensity was proportional to the concentration of vitamin B(12) , giving a second-order calibration graph over the cobalt (II) concentration range from 10 to 5000 μg L(-1)(r(2) = 0.9985) with a detection limit of 9.3 μg L(-1). The proposed method was applied to the determination of vitamin B(12) in different kinds of pharmaceuticals.  相似文献   

14.
When the arginyl-transfer ribonucleic acid (tRNA) species isolated from unshaken and from shaken cultures of Neurospora were compared by co-chromatography, a marked change in the relative abundance of the two main tRNA(arg) species was found. The two arginine tRNA species had different codon responses in ribosome binding assays. The tRNA(arg) eluting first (prevalent in shaken cultures) bound strongly to polyadenylic-guanylic acid [poly(A,G)] and to a lesser extent to polycytidylic-guanylic-adenylic acid [poly(C,G,A)]. The second tRNA(arg) species (prevalent in unshaken cultures) bound to poly(C,G,A) but not to poly(A, G). The possible significance of these observations is briefly discussed. Several modifications that improve the yield of tRNA from Neurospora were introduced in a standard isolation procedure.  相似文献   

15.
16.
The pK(A) values of (4RS)-[4-carboxy-5,8,11-tris(carboxymethyl)-1-phenyl-2-oxa-5,8,11-triazatridecan-13-oic acid] (BOPTA), a polyprotic molecule whose gadolinium complex is an important magnetic resonance imaging contrast agent for clinical use, have been determined in water, in physiologic solution (PS), in serum (S), and in cerebrospinal fluid (CSF), by means of 13C nuclear magnetic resonance spectroscopy data processed by a dedicated software package called DISCO. The aim of this study was to supply the BOPTA pK(A) values in media very similar to the in vivo environment and, consequently, to get a picture of the in vivo behavior of its Gd complex, whose thermodynamic stability is directly linked to the pK(A) values. The pK(A) values appeared to be almost equal both in D(2)O and in PS, while pK(1) and pK(5) values in CSF differ a little. In S, only pK(2) and pK(3) were calculated due to the narrow pH range used for data collection. However, these pK(A) values were found equal to those in the other media. These results represent the first direct spectroscopic evidence of a substantial invariability of BOPTA behavior in different media and they justify the extrapolation to biological fluids of the data obtained in water. The values also confirmed the high-quality performance of DISCO in calculating pK(A) values of polyprotic molecules in complex media.  相似文献   

17.
We report the results of NMR studies and computer simulations of potent antagonists reflective of the alpha(IIb)beta(3) receptor-bound conformations. The peptides c[Mpa-(15)N-Arg(1)-(15)N-Gly(2)-(15)N-Asp(3)-(15)N-Phe(4)-(15)N-Arg(5)-Cys]-NH(2) (Phe-Arg analog) (Mpa: 3-mercaptopropionic acid) and c[Mpa-(15)N-Arg(1)-(15)N-Gly(2)-(15)N-Asp(3)-(15)N-Asp(4)-(15)N-Val(5)-Cys]-NH(2) (Asp-Val analog) were subjected to (15)N-edited NMR experiments to study the conformations of these peptides in the absence and in the presence of alpha(IIb)beta(3) receptor. The NMR studies of the Phe-Arg analog, a selective alpha(IIb)beta(3) antagonist, resulted in distinctly different experimental data in the presence and absence of the receptor. The computer simulations for this peptide resulted in one large family of structures consistent with the experimental data. This conformation suggests a type I beta-turn spanning residues Arg(1) and Gly(2) when bound to the receptor and we were able to establish a model for the three dimensional arrangement of the pharmacophores. The studies on the Asp-Val analog, an alpha(v)beta(3) antagonist that binds to the alpha(IIb)beta(3) with moderate affinity, resulted in conformations that are not as well defined as those for the Phe-Arg analog but are consistent with the model established for this analog. These results are important for the design of novel alpha(IIb)beta(3) antagonists.  相似文献   

18.
Four integrins, namely alpha(1)beta(1), alpha(2)beta(1), alpha(10)beta(1), and alpha(11)beta(1), form a special subclass of cell adhesion receptors. They are all collagen receptors, and they recognize their ligands with an inserted domain (I domain) in their alpha subunit. We have produced the human integrin alpha(10)I domain as a recombinant protein to reveal its ligand binding specificity. In general, alpha(10)I did recognize collagen types I-VI and laminin-1 in a Mg(2+)-dependent manner, whereas its binding to tenascin was only slightly better than to albumin. When alpha(10)I was tested together with the alpha(1)I and alpha(2)I domains, all three I domains seemed to have their own collagen binding preferences. The integrin alpha(2)I domain bound much better to fibrillar collagens (I-III) than to basement membrane type IV collagen or to beaded filament-forming type VI collagen. Integrin alpha(1)I had the opposite binding pattern. The integrin alpha(10)I domain was similar to the alpha(1)I domain in that it bound very well to collagen types IV and VI. Based on the previously published atomic structures of the alpha(1)I and alpha(2)I domains, we modeled the structure of the alpha(10)I domain. The comparison of the three I domains revealed similarities and differences that could potentially explain their functional differences. Mutations were introduced into the alphaI domains, and their binding to types I, IV, and VI collagen was tested. In the alpha(2)I domain, Asp-219 is one of the amino acids previously suggested to interact directly with type I collagen. The corresponding amino acid in both the alpha(1)I and alpha(10)I domains is oppositely charged (Arg-218). The mutation D219R in the alpha(2)I domain changed the ligand binding pattern to resemble that of the alpha(1)I and alpha(10)I domains and, vice versa, the R218D mutation in the alpha(1)I and alpha(10)I domains created an alpha(2)I domain-like ligand binding pattern. Thus, all three collagen receptors appear to differ in their ability to recognize distinct collagen subtypes. The relatively small structural differences on their collagen binding surfaces may explain the functional specifics.  相似文献   

19.
Adenosine A(3) receptors are reported to couple negatively to adenylyl cyclase (AC) but their mediation of anti-inflammatory effects in human eosinophils prompted us to investigate their coupling to AC. The A(3)-selective agonists IB-MECA and Cl-IB-MECA evoked a concentration-dependent generation of cAMP (EC(50), 3.2 and 1.8 microM, respectively) and were more potent than the A(2A) agonist CGS 21680 (EC(50)=15.4 microM) and adenosine (EC(50)=19.2 microM). The cAMP response was additive to that produced by forskolin (10 microM). The effect of IB-MECA was insensitive to A(1) and A(2A) receptor antagonists, but was antagonized by the A(3)-selective antagonist MRS 1220 (0.1-2.5 microM) in a competitive manner. The estimated K(B) of 190 nM was, however, atypical. The cyclo-oxygenase inhibitor, indomethacin, had no effect on the cAMP response. A general inverse relationship between cAMP generation and inhibition of degranulation was seen. We conclude that in human eosinophils, an atypical form of A(3) receptors positively coupled to AC may exist. The resulting cAMP generation may underlie the anti-inflammatory actions of A(3) agonists in eosinophils.  相似文献   

20.
N-Cbz-Ser(OMe)-(Pm----O)-d(TpT) was synthesized as a diastereomeric mixture of approx. (1:1) by reacting d(TpT) with 2,4,6-triisopropylbenzenesulfonyl chloride and N-Cbz-Ser-OMe. The phosphoesteric bond of N-Cbz-Ser (OMe)-(Pm----O)-d(TpT) was found to be stable in acid (1 N HCl, 1 h, 37 degrees C), but labile in alkaline solution (0.1-1 N NaOH, 1 h, 37 degrees C). The products of alkaline hydrolysis were determined to be d(TpT) and the amino acid derivatives. Furthermore, the phosphotriester N-Cbz-Ser(OMe)-(Pm----O)-d(TpT) was more labile than the diesteric analogue N-Cbz-Ser (OMe)-pdT. The internucleotide phosphotriester linkage of N-Cbz-Ser (OMe)-(Pm----O)-d(TpT) was also found to be resistant to enzymatic digestion with spleen and snake venom phosphodiesterases.  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