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 62 毫秒
1.
Xiao X  Tan Y  Zhu L  Guo Y  Wen Z  Li M  Pu X  Tian A 《Journal of molecular modeling》2012,18(4):1389-1399
This work mainly studies the effects of the position (there are two possible hydrated sites) and the manner (i.e., whether water acts as a proton donor or acceptor) of hydration by various numbers of water molecules on the stability of 14 solvated N-methylacetamide structures, NMA-(H2O) n (n = 1–3), as well as the binding strength between the NMA and the water cluster, using molecular dynamics (MD) and B3LYP methods. Natural bond orbital (NBO) analysis is used to explore the origin of these effects. Some novel observations are obtained from the work. Our results show that monohydration at the carbonyl site favors stability and binding strength compared to monohydration at the amino site. Similarly, the preferred hydration at the carbonyl site is observed for dihydrated NMAs when the second water is added as a proton donor to the C=O group or the first water is H-bonded to the C=O group. However, unfavorable hydration at the C=O site occurs if the second water acts as a proton acceptor. Trihydration by a ring cluster of three water molecules at either the carbonyl site or the amino one yields relatively stable complexes, but significantly disfavors binding strength. The other trihydrated NMAs show similar behavior to dihydrated NMAs. In addition, our results show that the C=O and N–H frequencies can still be utilized to examine the H-bond effects of the water cluster.  相似文献   

2.
The microsolvation of taurine (TA) with one, two or three water molecules was investigated by a density functional theory (DFT) approach. Quantum theory of atoms in molecules (QTAIM) analyses were employed to elucidate the hydrogen bond (H-bond) interaction characteristics in TA-(H2O)n (n = 1–3) complexes. The results showed that the intramolecular H-bond formed between the hydroxyl and the N atom of TA are retained in most TA-(H2O)n (n = 1–3) complexes, and are strengthened via cooperative effects among multiple H-bonds from n = 1–3. A trend of proton transformation exists from the hydroxyl to the N atom, which finally results in the cleavage of the origin intramolecular H-bond and the formation of a new intramolecular H-bond between the amino and the O atom of TA. Therefore, the most stable TA-(H2O)3 complex becomes a zwitterionic complex rather than a neutral type. A many-body interaction analysis showed that the major contributors to the binding energies for complexes are the two-body energies, while three-body energies and relaxation energies make significant contributions to the binding energies for some complexes, whereas the four-body energies are too small to be significant.  相似文献   

3.
The microsolvation of aminoethanol (AE) with one, two, three or four water molecules was investigated using a density functional theory (DFT) approach. Quantum theory of atoms in molecules (QTAIM) analyses were employed to elucidate the hydrogen-bonding characteristics of AE–(H2O) n (n = 1–4) complexes. The results showed that AE tends to break its intramolecular OHAE···NAE hydrogen bond (H-bond) upon microsolvation and form intermolecular H-bonds with water molecules, while complexes that retain the intramolecular OHAE···NAE H-bond show reduced stabilities. The intermolecular H-bond that forms between the nitrogen atom of AE and the hydroxyl of a water molecule is the strongest one for the most stable AE–(H2O) n (n = 1–4) complexes, and as n increases from 1 to 4 they grow stronger. The partial covalent character of this H-bond was confirmed by QTAIM analyses. Many-body interaction analysis showed that the relaxation energies and two- and three-body energies make significant contributions to the binding energies of the complexes.  相似文献   

4.
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   

5.
The local meta-GGA exchange correlation density functional (TPSS) with a relativistic effective core potential was employed to systematically investigate the geometric structures, stabilities, and electronic properties of bimetallic Ca2Au n (n = 1–9) and pure gold Au n (n ≤ 11) clusters. The optimized geometries show that the most stable isomers for Ca2Au n clusters have 3D structure when n > 2, and that one Au atom capping the Ca2Au n−1 structure for different-sized Ca2Au n (n = 1–9) clusters is the dominant growth pattern. The average atomic binding energies and second-order difference in energies show that the Ca2Au4 isomer is the most stable among the Ca2Au n clusters. The same pronounced even–odd alternations are found in the HOMO–LUMO gaps, VIPs, and hardnesses. The polarizabilities of the Ca2Au n clusters show an obvious local minimum at n = 4. Moreover, the inverse corrections to the polarizabilities versus the ionization potential and hardness were found for the gold clusters.  相似文献   

6.
Nitric oxide (NO) is a short-lived radical that functions as a neurotransmitter in the central nervous system and plays a physiological role in the regulation of hypothalamic–pituitary–adrenal axis and vasopressinergic axis. In the present study, we aimed to investigate the interaction between the generation of NO and vasopressin (AVP) and corticosterone release after 3 days of water deprivation in rats. Animals were previously treated with intraperitoneal (i.p.) saline or l-nitro-arginine methyl ester (L-NAME) injection. l-NAME is a nonspecific inhibitor of nitric oxide synthases. In control rats given i.p. saline or l-NAME, hypothalamic, pituitary, and plasma AVP levels and plasma corticosterone did not change from baseline levels (p > 0.05). Three days of water deprivation increased significantly the corticosterone levels in plasma (p < 0.01) and AVP levels in hypothalamus and plasma (p < 0.01), but not in pituitary, which showed a significant decrease. These variations were concomitant with the elevation of nitrates/nitrates in plasma. l-NAME injection abolished significantly (p < 0.01) the elevation of plasma corticosterone and hypothalamic AVP levels induced by water deprivation. These findings showed that in water-deprived rats, nitric oxide synthase inhibition by l-NAME inhibits corticosterone and vasopressin release, suggesting a potent stimulatory role of NO.  相似文献   

7.
 The interaction of human carbonic anhydrase (hCA) isozymes I and II with cyanamide, a linear molecule isoelectronic with the main physiological substrate of the enzyme, CO2, was investigated through spectroscopic, kinetic, and X-ray crystallographic studies. We show here that cyanamide is hydrated to urea in the presence of CAs, and that it also acts as a weak non-competitive inhibitor (K I=61±3 mM and 238±9 mM for hCA II and hCA I, respectively) towards the esterasic activity of these enzymes, as tested with 4-nitrophenyl acetate. Changes in the spectrum of the Co(II)-hCA II derivative observed in the presence of cyanamide suggest that it likely binds the metal ion within the CA active site, adding to the coordination sphere, not substituting the metal-bound solvent molecule. It thereafter undergoes a nucleophilic attack from the metal-bound hydroxide ion, forming urea which remains bound to the metal, as observed in the X-ray crystal structure of hCA II soaked in cyanamide solutions for several hours. The urea molecule is directly coordinated to the active site Zn(II) ion through a protonated nitrogen atom. Several hydrogen bonds involving active site residues Thr199 and Thr200 as well as three water molecules (Wat99, Wat122, and Wat123) further stabilize the urea-hCA II adduct. Kinetic studies in solution further proved that urea acts as a tight binding inhibitor of the two isozymes hCA I and hCA II, with very slow binding kinetics (k on=2.5×10–5 s–1 M–1). A mechanism to explain the hydration process of cyanamide by CAs, as well as the tight binding of urea in the active site, is also proposed based on the hypothesis that urea is deprotonated when bound to the enzyme. Cyanamide is thus the first true suicide substrate of this enzyme for which binding has been documented by means of X-ray crystallographic and spectroscopic studies. Received: 26 February 1999 / Accepted: 25 May 1999  相似文献   

8.
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   

9.
Galdieria sulphuraria (Galdieri) Merola can grow heterotrophically on at least ten different polyols. We investigated their metabolic path to glycolysis/gluconeogenesis and identified two NAD-dependent polyol dehydrogenases. Activity of other enzymes metabolizing mannitol or sorbitol could not be detected. The two dehydrogenases had a broad substrate specificity and were termed xylitol dehydrogenase (EC 1.1.1.14; substrate specificity: xylitol > d-sorbitol > d-mannitol > l-arabitol) and d-arabitol dehydrogenase (EC 1.1.1.11; substrate specificity: d-arabitol > l-fucitol > d-mannitol > d-threitol) according to the substrate with the lowest K m value. The xylitol dehydrogenase was stable during purification. In contrast, the d-arabitol dehydrogenase was thermolabile and depended on divalent ions for stability and activity, preferentially Mn2+ and Ni2+. The molecular mass of the xylitol dehydrogenase was estimated to be 295 kDa by size-exclusion chromatography and 220 kDa by rate-sedimentation centrifugation. The d-arabitol dehydrogenase had a molecular mass of 105 kDa as determined by rate-sedimentation centrifugation. The specific activity of both enzymes increased about fourfold when cells were transferred from autotrophic to heterotrophic conditions regardless of whether sugars or polyols were supplied as substrates. The significance of polyol metabolism in Galdieria sulphuraria with regard to the natural habitat of the alga is discussed. Received: 15 January 1997 / Accepted: 12 February 1997  相似文献   

10.
Structure of nematogenic p-n-Alkoxy cinnamic acids (nOCAC) with various alkyl chain carbon atoms (n = 2, 4, 6, 8) has been optimized using density functional B3LYP with 6-31+G (d) basis set using crystallographic geometry as input. Using the optimized geometry, electronic structure of the molecules has been evaluated using the semiempirical methods and DFT calculations. Molecular charge distribution and phase stability of these systems have been analyzed based on Mulliken and L?wdin population analysis. The electronic absorption spectra of nOCAC molecules have been simulated by employing DFT method, semiempirical CNDO/S and INDO/S parameterizations. Two types of calculations have been performed for model systems containing single and double molecules of nOCAC. UV-Visible spectra have been calculated for all single molecules. The UV stability of the molecules has been discussed in light of the electronic transition oscillator strength (f). The dimer complexes of higher homologues (n = 6, 8) have also been reported to enable the comparison between single and double molecules.  相似文献   

11.
The origin, age, and dynamics of carbonate sediments in Kailua Bay on Oahu, Hawaii, are described. The shoreface (from shoreline to 4 km offshore) consists of a broad (5 km2) fringing coral reef ecosystem bisected by a sinuous, shore-normal, sand-filled paleostream channel 200–300 m wide. The median grain diameter of surface sands is finest on the beach face (<0.3 mm) and increases offshore along the channel axis. Kailua sands are >90% biogenic carbonate, dominated by skeletal fragments of coralline algae (e.g. Porolithon, up to 50%) followed by the calcareous green alga Halimeda (up to 32%), coral fragments (1–24%), mollusc fragments (6–21%), and benthic foraminifera (1–10%). Sand composition and age across the shoreface are correlated to carbonate production. Corals and coralline algae, principal builders of the reef framework, are younger and more abundant in sands along the channel axis and in offshore reefal areas, while Halimeda, molluscs, and foraminifera are younger and more dominant in nearshore waters shoreward of the main region of framework building. Shoreface sediments are relatively old. Of 20 calibrated radiocarbon dates on skeletal constituents of sand, only three are younger than 500 years b.p.; six are 500–1000 years b.p.; six are 1000–2000 years b.p.; and five are 2000–5000 years b.p. Dated fine sands are older than medium to coarse sands and hence may constitute a reservoir of fossil carbonate that is distributed over the entire shoreface. Dominance of fossiliferous sand indicates long storage times for carbonate grains, which tend to decrease in size with age, such that the entire period of relative sea-level inundation (∼5000 years) is represented in the sediment. Despite an apparently healthy modern coral ecosystem, the surficial sand pool of Kailua Bay is dominated by sand reflecting an antecedent system, possibly one that existed under a +1–2 m sea-level high stand during the mid- to late Holocene. Accepted: 20 December 1999  相似文献   

12.
When Phanerochaete chrysosporium was grown with laminarin (a β-1,3/1,6-glucan) as the sole carbon source, a β-1,3-glucanase with a molecular mass of 36 kDa was produced as a major extracellular protein. The cDNA encoding this enzyme was cloned, and the deduced amino acid sequence revealed that this enzyme belongs to glycoside hydrolase family 16; it was named Lam16A. Recombinant Lam16A, expressed in the methylotrophic yeast Pichia pastoris, randomly hydrolyzes linear β-1,3-glucan, branched β-1,3/1,6-glucan, and β-1,3-1,4-glucan, suggesting that the enzyme is a typical endo-1,3(4)-β-glucanase (EC 3.2.1.6) with broad substrate specificity for β-1,3-glucans. When laminarin and lichenan were used as substrates, Lam16A produced 6-O-glucosyl-laminaritriose (β-d-Glcp-(1–>6)-β-d-Glcp-(1–>3)-β-d-Glcp-(1–>3)-d-Glc) and 4-O-glucosyl-laminaribiose (β-d-Glcp-(1–>4)-β-d-Glcp-(1–>3)-d-Glc), respectively, as one of the major products. These results suggested that the enzyme strictly recognizes β-d-Glcp-(1–>3)-d-Glcp at subsites −2 and −1, whereas it permits 6-O-glucosyl substitution at subsite +1 and a β-1,4-glucosidic linkage at the catalytic site. Consequently, Lam16A generates non-branched oligosaccharide from branched β-1,3/1,6-glucan and, thus, may contribute to the effective degradation of such molecules in combination with other extracellular β-1,3-glucanases.  相似文献   

13.
A putative N-acyl-d-glucosamine 2-epimerase from Caldicellulosiruptor saccharolyticus was cloned and expressed in Escherichia coli. The recombinant enzyme was identified as a cellobiose 2-epimerase by the analysis of the activity for substrates, acid-hydrolyzed products, and amino acid sequence. The cellobiose 2-epimerase was purified with a specific activity of 35 nmol min–1 mg–1 for d-glucose with a 47-kDa monomer. The epimerization activity for d-glucose was maximal at pH 7.5 and 75°C. The half-lives of the enzyme at 60°C, 65°C, 70°C, 75°C, and 80°C were 142, 71, 35, 18, and 4.6 h, respectively. The enzyme catalyzed the epimerization reactions of the aldoses harboring hydroxyl groups oriented in the right-hand configuration at the C2 position and the left-hand configuration at the C3 position, such as d-glucose, d-xylose, l-altrose, l-idose, and l-arabinose, to their C2 epimers, such as d-mannose, d-lyxose, l-allose, l-gulose, and l-ribose, respectively. The enzyme catalyzed also the isomerization reactions. The enzyme exhibited the highest activity for mannose among monosaccharides. Thus, mannose at 75 g l–1 and fructose at 47.5 g l–1 were produced from 500 g l–1 glucose at pH 7.5 and 75°C over 3 h by the enzyme.  相似文献   

14.
Reactions of lithium halide (LiX, X = F, Cl, Br and I) and methyl halide (CH3X, X = F, Cl, Br and I) have been investigated at the B3LYP/6-31G(d) level of theory using the microhydration model. Beginning with hydrated lithium ion, four or two water molecules have been conveniently introduced to these aqueous-phase halogen-exchange SN2 reactions. These water molecules coordinated with the center metal lithium ion, and also interacted with entering and leaving halogen anion via hydrogen bond in complexes and transition state, which to some extent compensated hydration of halogen anion. At 298 K the reaction profiles all involve central barriers ΔE cent which are found to decrease in the order F > Cl > Br > I. The same trend is also found for the overall barriers (ΔE ovr ) of the title reaction. In the SN2 reaction of sodium iodide and methyl iodide, the activation energy agrees well with the aqueous conductometric investigation.  相似文献   

15.
The modulation of the calmodulin-induced inhibition of the calcium release channel (ryanodine receptor) by two sulfhydryl oxidizing compounds, 4-(chloromercuri)phenyl–sulfonic acid (4-CMPS) and 4,4′-dithiodipyridine (4,4′-DTDP) was determined by single channel current recordings with the purified and reconstituted calcium release channel from rabbit skeletal muscle sarcoplasmic reticulum (HSR) and [3H]ryanodine binding to HSR vesicles. 0.1 μm CaM reduced the open probability (P o ) of the calcium release channel at maximally activating calcium concentrations (50–100 μm) from 0.502 ± 0.02 to 0.137 ± 0.022 (n= 28), with no effect on unitary conductance. 4-CMPS (10–40 μm) and 4,4′-DTDP (0.1–0.3 mm) induced a concentration dependent increase in P o (> 0.9) and caused the appearance of longer open states. CaM shifted the activation of the calcium release channel by 4-CMPS or 4,4′-DTDP to higher concentrations in single channel recordings and [3H]ryanodine binding. 40 μm 4-CMPS induced a near maximal (P o > 0.9) and 0.3 mm 4,4′-DTDP a submaximal (P o = 0.74) channel opening in the presence of CaM, which was reversed by the specific sulfhydryl reducing agent DTT. Neither 4-CMPS nor 4,4′-DTDP affected Ca-[125I]calmodulin binding to HSR. 1 mm MgCl2 reduced P o from 0.53 to 0.075 and 20–40 μm 4-CMPS induced a near maximal channel activation (P o > 0.9). These results demonstrate that the inhibitory effect of CaM or magnesium in a physiological concentration is diminished or abolished at high concentrations of 4-CMPS or 4,4′-DTDP through oxidation of activating sulfhydryls on cysteine residues of the calcium release channel. Received: 22 July 1999/Revised: 15 November 1999  相似文献   

16.
We measured the self-diffusion coefficients of water in a Nafion membrane and two sulfonated polyethersulfone (SPES) membranes with varying ion-exchange capacities (IEC) in terms of relative humidity using the pulse field gradient NMR (PFG-NMR) technique. The self-diffusion coefficients were plotted against the number of water molecules per sulfonic acid group, λ, and compare these values with the results of molecular dynamics (MD) simulations. Classical MD simulations for all membranes were carried out using a consistent force field at λ = 3, 6, 9, 12, and 15. The dynamic properties of water (H2O) and hydronium (H3O+) on a molecular level were estimated as self-diffusion coefficients and residence times around a sulfonate group ( \textSO3- {\text{SO}}_3^{-} ). The diffusion coefficients of H2O and H3O+ followed the order, Nafion > SPES with IEC = 1.4 > SPES with IEC = 1.0 > SPES with IEC = 0.75, which agreed with the experimental data. The residence time distribution of H2O around \textSO3- {\text{SO}}_3^{-} in Nafion was in the range of 1–6 ps, whereas H2O in the SPES exhibited a residence time of greater than 20 ps.  相似文献   

17.
 The stability constants of the 1 : 1 complexes formed between Pb2+ and several simple phosphate monoesters (4-nitrophenyl phosphate, phenyl phosphate, d-ribose 5-monophosphate, n-butyl phosphate) or phosphonate ligands (methylphosphonate, ethylphosphonate) (R-PO2– 3) were determined by potentiometric pH titrations in aqueous solution (25  °C;I=0.1 M, NaNO3). The construction of a log K P P b b(R-PO3) versus pK H H(R-PO3) plot for the mentioned ligand systems results in a straight line on which the data pairs (the corresponding equilibrium constants were also measured) for uridine 5′-monophosphate (UMP2–) and thymidine 5′-monophosphate (dTMP2–) also fall; this result shows that in the Pb2+ complexes of UMP2– and dTMP2– the nucleobase residues do not interfere, in neither a positive nor a negative way, with the binding of Pb2+ and that the stability of all these complexes is determined by the basicity of the phosph(on)ate group. The mentioned straight-line correlation (as defined by the least-squares procedure) allowed us to demonstrate (via constants determined now) that the stability of the Pb2+ complex of cytidine 5′-monophosphate (CMP2–) is also solely determined by the basicity of its phosphate group. A similar evaluation, based on literature data, for the Pb(HPO4) complex reveals that its stability corresponds closely to the expectations based on the Pb(R-PO3) data, though there is a slight hint that Pb(HPO4) may be somewhat more stable [which would be in agreement with previous observations of other M(HPO4) complexes]; clearly, more such comparisons are possible with the reference line given now. Based on the stability constants of the monoprotonated Pb(H;CMP)+ complex and the Pb(cytidine)2+ species (which was also measured now), it is concluded that in Pb(H;CMP)+ the proton is located at the phosphate group and Pb2+ mainly at the N3/(C2)O site of the cytosine residue. Regarding nucleic acids in solution, it is further concluded that the affinity of Pb2+ towards the negatively mono-charged phosphate unit, —O—P(O)2 —O—, of a nucleic acid backbone is comparable to that of the cytosine moiety, the affinity towards other nucleobase residues being smaller. This information may prove helpful regarding the properties of lead ribozymes. Received: 16 April 1999 / Accepted: 2 June 1999  相似文献   

18.
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   

19.
 A novel heptacoordinating ligand consisting of a thirteen-membered tetraazamacrocycle containing the pyridine ring and bearing three methylenephosphonate groups (PCTP-[13]) has been synthesized. Its Gd(III) complex displays a remarkably high longitudinal water proton relaxivity (7.7 mM–1 s–1 at 25  °C, 20 MHz and pH 7.5) which has been accounted for in terms of contributions arising from (1) one water molecule bound to the metal ion, (2) hydrogen-bonded water molecules in the second coordination sphere, or (3) water molecules diffusing near the paramagnetic chelate. Variable-temperature 17O-NMR transverse relaxation data indicate that the residence lifetime of the metal-bound water molecule is very short (8.0 ns at 25  °C) with respect to the Gd(III) complexes currently considered as contrast agents for magnetic resonance imaging. Furthermore, GdPCTP-[13] interacts with human serum albumin (HSA), likely through electrostatic forces. By comparing water proton relaxivity data for the GdPCTP-[13]-HSA adduct, measured as a function of temperature and magnetic field strength, with those for the analogous adduct with GdDOTP (a twelve-membered tetraaza macrocyclic tetramethylenephosphonate complex lacking a metal-bound water molecule), it has been possible to propose a general picture accounting for the main determinants of the relaxation enhancement observed when a paramagnetic Gd(III) complex is bound to HSA. Basically, the relaxation enhancement in these systems arises from (1) water molecules in the hydration shell of the macromolecule and protein exchangeable protons which lie close to the interaction site of the paramagnetic complex and (2) the metal bound water molecule(s). As far as the latter contribution is concerned, the interaction with the protein causes an elongation of the residence lifetime of the metal-bound water molecule, which limits, to some extent, the potential relaxivity enhancement expected upon the binding of the paramagnetic complex to HSA. Received: 27 January 1997 / Accepted: 12 May 1997  相似文献   

20.
 The tumour-suppressor gene p53 is pivotal in the regulation of apoptosis, and point mutations within p53 are the commonest genetic alterations in human cancers. Cytotoxic T lymphocytes (CTL) recognise peptide-MHC complexes on the surface of tumour cells and bring about lysis. Therefore, p53-derived peptides are potential candidates for immunisation strategies designed to induce antitumour CTL in patients. Conformational changes in the p53 protein, generated as a result of point mutations, frequently expose the 240 epitope, RHSVV (amino acids 212–217), which may be processed differently from the wild-type protein resulting in an altered MHC-associated peptide repertoire recognised by tumour-specific CTL. In this study 42 peptides (37 overlapping nonameric peptides, from amino acids 193–237 and peptides 186–194, 187–197, 188–197, 263–272, 264–272, possessing binding motifs for HLA-A2) derived from the wild-type p53 protein sequence were assayed for their ability to stabilise HLA-A2 molecules in MHC class I stabilisation assays. Of the peptides tested, 24 stabilised HLA-A2 molecules with high affinity (fluorescence ratio>1.5) at 26 °C, and five (187–197, 193–200, 217–224, 263–272 and 264–272) also stabilised the complexes at 37 °C. Peptides 188–197, 196–203 and 217–225 have not previously been identified as binders of HLA-A2 molecules and, of these, peptide 217–225 stabilised HLA-A2 molecules with the highest fluorescence ratio. Peptide 217–225 was chosen to generate HLA-A2-restricted CTL in vitro; peptide 264–272 was used as a positive control. The two primary CTL thus generated (CTL-217 using peptide 217–225; and CTL-264 using peptide 264–272) were capable of specifically killing peptide-pulsed T2 or JY cells. In order to determine whether these peptides were endogenously processed and to test the hypothesis that mutants expressing different protein conformations would generate an alternative peptide repertoire at the cell surface, a panel of target cells was generated. HLA-A2+ SaOs-2 cells were transfected with p53 cDNA containing point mutations at either position 175 (R → H) or 273 (R → H) (SaOs-2/175 and SaOs-2/273). Two HLA-A2-negative cell lines, A431 and SKBr3, naturally expressing p53 mutations at positions 273 and 175 respectively, were transfected with a cDNA encoding HLA-A2. The results showed that primary CTL generated in response to both peptides were capable of killing SaOs-2/175 and SKBr3-A2 cells, which possess the same mutation, but not SaOs-2/273, A431-A2 or SKBr3 cells transfected with control vector. This suggests that these peptides are presented on the surface of SaOs-2/175 and SKBr3-A2 cells in a conformation-dependent manner and represent potentially useful target peptides for immunotherapy. Received: 23 March 2000 / Accepted: 22 June 2000  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