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 181 毫秒
1.
We have reacted [Pt(dien)Cl]Cl, [Pt(en)(D2O)2]2+, and [Pt(Me4en)(D2O)2]2+ [Me4en = N,N,N′,N′-tetramethylethylenediamine] with selenomethionine (SeMet). When [Pt(dien)Cl]Cl is reacted with SeMet, [Pt(dien)(SeMet-Se)]2+ is formed; two Se-CH3 resonances are observed due to the different chiralities at the Se atom upon platination. In a reaction of [Pt(dien)Cl]Cl with an equimolar mixture of SeMet and Met, the SeMet product forms more quickly though a slow equilibrium with approximately equal amounts of both products is reached. [Pt(Me4en)(D2O)2]2+ reacts with SeMet to form [Pt(Me4en)(SeMet-Se)(D2O)]2+ initially but forms [Pt(Me4en)(SeMet-Se,N)]+ ultimately. One stereoisomer of the chelate, assigned to the R chirality at the Se atom, dominates within the first few minutes of reaction. [Pt(en)(D2O)2]2+ forms a variety of products depending on reaction stoichiometry; when one equivalent or less of SeMet is added, the dominant product is [Pt(en)(SeMet-Se,N)]+. In the presence of excess SeMet, [Pt(en)(SeMet-Se)2]2+ is the dominant initially, but displacement of the en ligand occurs leading to [Pt(SeMet-Se,N)2] as the eventual product. Displacement of the en ligand from [Pt(en)(SeMet-Se,N)]+ does not occur. In reactions of K2PtCl4 with two equivalents of SeMet, [Pt(SeMet-Se,N)2] is formed, and three sets of resonances are observed due to different chiralities at the Se atoms. Only the cis geometric isomers are observed by 1H and 195Pt NMR spectroscopy.  相似文献   

2.
S-100 protein absorbs to the calmodulin antagonist W-7 coupled to epoxy-activated Sepharose 6B in the presence of Ca2+ and is eluted by ethylene glycol bis(β-aminoethyl ether)-N,N′-tetraacetic acid buffer. S-100a and S-100b were separated and isolated by Ca2+-dependent affinity chromatography on W-7 Sepharose. The Ca2+-induced conformational changes of S-100a and S-100b were examined using circular dichroism, ultraviolet difference spectra, and a fluorescence probe. Differences in Ca2+-dependent conformational changes between S-100a and S-100b became apparent. Circular dichroism studies revealed that both S-100a and S-100b undergo a conformational change upon binding of Ca2+ in the aromatic and far-uv range. In the presence or absence of Ca2+, the aromatic CD spectrum of S-100a differed completely from that of S-100b, possibly due to the single tryptophan residue of S-100a. Far-uv studies indicate that α-helical contents of both S-100a and S-100b decreased with addition of Ca2+. Ca2+-induced conformational changes of S-100a and S-100b were also detected by uv difference spectra. The spectrum of S-100a also differed from that of S-100b. Fluorescence studies using 2-p-toluidinylnaphthalene-6-sulfonate (TNS), a hydrophobic probe for protein, revealed a slight difference in conformational changes of these two components. The interaction of TNS and S-100b was observed with concentrations above 3 μm Ca2+; on the other hand, S-100a required concentrations above 8 μm. This finding was supported by the difference in the binding affinities of S-100a and S-100b to the W-7 (N-(6-aminohexyl)-5-chloro-1-naphthalenesulfonamide)-Sepharose column; both S-100a and S-100b bound the column in the presence of Ca2+ but S-100a was eluted prior to S-100b. These results suggest that S-100a and S-100b differ in their dependence on Ca2+ and that the affinity-chromatographic separation of S-100a from S-100b on the W-7-Sepharose column makes feasible a rapid purification of these two components.  相似文献   

3.
N.m.r. and c.d. spectroscopy have been used to study the interactions of cyclic hexapeptide cyclo(Pro-Sar-Sar)2 with metal ions and ammonium ions. Cyclo(Pro-Sar-Sar)2 was found to form complexes with Li+, K?, Ba2+ and Cu2+, accompanying the conformational change into a single conformer, and the conformation of cyclo(Pro-Sar-Sar)2 in the Li+-complex was different from that in the Cu2+-complex. These findings indicate conformational flexibility of cyclo(Pro-Sar-Sar)2. The equilibrium constant for the complexation with Li+ was 2.3 × 102l mol?1, and cyclo(Pro-Sar-Sar)2 adopted an asymmetric conformation in the complex. The addition of α-amino acid ester hydrochloride also caused the conformational change of cyclo(Pro-Sar-Sar)2), but in this case it did not converge into a single conformation. This type of interaction was strengthened with aromatic α-amino acid ester hydrochloride due to the aromatic-amide interactions. Finally, the rates of exchange between unbound α-amino acid ester hydrochlorides and those complexed with cyclo(Pro-Sar-Sar)2 were found to be different, according to the nature of α-amino acid.  相似文献   

4.
NH 4 + excretion was undetectable in N2-fixing cultures ofRhodospirillum rubrum (S-1) and nitrogenase activity in these cultures was repressed by the addition of 10 mM NH 4 + to the medium. The glutamate analog,l-methionine-dl-sulfoximine (MSX), derepressed N2 fixation even in the presence of 10 mM extracellular NH 4 + . When 10 mg MSX/ml was added to cultures just prior to nitrogenase induction they developed nitrogenase activity (20% of the control activities) and excreted most of their fixed N2 as NH 4 + . Nitrogenase activities and NH 4 + production from fixed N2 were increased considerably when a combined nitrogen source, NH 4 + (>40 moles NH 4 + /mg cell protein in 6 days) orl-glutamate (>60 moles NH 4 + /mg cell protein in 6 days) was added to the cultures together with MSX.Biochemical analysis revealed thatR. rubrum produced glutamine synthetase and glutamate synthase (NADP-dependent) but no detectable NADP-dependent glutamate dehydrogenase. The specific activity of glutamine synthetase was observed to be maximal when nitrogenase activity was also maximal. Nitrogenase and glutamine synthetase activities were repressed by NH 4 + as well as by glutamate.The results demonstrate that utilization of solar energy to photoproduce large quantities of NH 4 + from N2 is possible with photosynthetic bacteria by interfering with their regulatory control of N2 fixation.  相似文献   

5.
The syntheses and structural characterization of four cobalt(II)-salicylate complexes, [(TPA)CoII(HSA)](ClO4) (1), [(isoBPMEN)CoII(HSA)](BPh4) (2), [(TPzA)CoII(HSA)](ClO4) (3) and [(6Me3TPA)CoII(HSA)](BPh4) (4) [TPA = tris(2-pyridylmethyl)amine, isoBPMEN = N1,N1-dimethyl-N2,N2-bis(2-pyridylmethyl)ethane-1,2-diamine, TPzA = tris((3,5-dimethyl-1H-pyrazole-1-yl)methyl)amine and 6Me3TPA = tris(6-methyl-2-pyridylmethyl)amine] are described. While 2, 3 and 4 are unreactive towards dioxygen, 1 reacts slowly with molecular oxygen to a cobalt(III)-salicylate complex, [(TPA)CoIII(SA)](ClO4) (1a). Two different crystalline forms, 1a and 1a·4H2O were isolated depending upon the condition of oxidation and crystallization. The solid-state structures of cobalt(III)-salicylate unit in both 1a and 1a·4H2O show a six-coordinate distorted octahedral coordination geometry at the cobalt(III) center ligated by the tetradentate ligand (TPA) where the dianionic salicylate (SA) binds in a bidentate fashion through one carboxylate and one phenolate oxygen. The hydrated form 1a·4H2O reveals a hexameric water cluster formation in the inorganic lattice host. The complex cation and the perchlorate counterion are involved in stabilizing the (H2O)6 cluster in a rare ‘pentamer planar+1’ conformation. A one-dimensional water tape consisting of edge-shared water hexamers is observed. The water tape represents a subunit of ice structure.  相似文献   

6.
The effects of the solvents deuterated water (2H2O) and dimethyl sulfoxide (Me2SO) on [3H]ouabain binding to (Na+,K+)-ATPase under different ligand conditions were examined. These solvents inhibited the type I ouabain binding to the enzyme (i.e., in the presence of Mg2++ATP+Na+). In contrast, both solvents stimulated type II (i.e., Mg2++Pi-, or Mn2+-dependent) binding of the drug. The solvent effects were not due to pH changes in the reaction. However, pH did influence ouabain binding in a differential manner, depending on the ligands present. For example, changes in pH from 7.05 to 7.86 caused a drop in the rate of binding by about 15% in the presence of Mg2++Na++ATP, 75% in the Mg2++Pi system, and in the presence of Mn2+ an increase by 24% under similar conditions. Inhibitory or stimulatory effects of solvents were modified as various ligands, and their order of addition, were altered. Thus, 2H2O inhibition of type I ouabain binding was dependent on Na+ concentration in the reaction and was reduced as Na+ was elevated. Contact of the enzyme with Me2SO, prior to ligands for type I binding, resulted in a greater inhibition of ouabain binding than that when enzyme was exposed to Na++ATP first and then to Me2SO. Likewise, the stimulation of type II binding was greater when appropriate ligands acted on enzyme prior to addition of the solvent. Since Me2SO and 2H2O inhibit type I ouabain binding, it is proposed that this reaction is favored under conditions which promote loss of H2O, and E1 enzyme conformation; the stimulation of type II ouabain binding in the presence of the solvents suggests that this type of binding is favored under conditions which promote the presence of H2O at the active enzyme center and E2 enzyme conformation. This postulation of a role of H2O in modulating enzyme conformations and ouabain interaction with them is in concordance with previous observations.  相似文献   

7.
This paper concerns the Circular Dichroism (CD) and Nuclear Magnetic Resonance (NMR) structural studies of the quadruple helix arrangements adopted by three tailored oligodeoxyribonucleotide analogues, namely d(TGMeGGT), d(TGGMeGT) and d(TGGGMeT), where dGMe represents a 8-methyl-2′-deoxyguanosine residue. The results of this study clearly demonstrate that the effects of the incorporation of dGMe instead of a dG residue are strongly dependant upon the positioning of a single base replacement along the sequence. As such, d(TGMeGGT), d(TGGMeGT) have been found to form 4-fold symmetric quadruplexes with all strands parallel and equivalent to each other, each more stable than their natural counterpart. NMR experiments clearly indicate that [d(TGMeGGT)]4 possesses a GMe-tetrad with all dGMe residues in a syn-glycosidic conformation while an anti-arrangement is apparent for the four dGMe of [d(TGGMeGT)]4. As the two complexes show a quite different CD behaviour, a possible relationship between the presence of residues adopting syn-glycosidic conformations and CD profiles is briefly discussed. As far as d(TGGGMeT) is concerned, NMR data indicate that at 25°C it exists primarily as a single-strand conformation in equilibrium with minor amounts of a quadruplex structure.  相似文献   

8.
4′,5′-Unsaturated nucleosides are obtained by the action of 1,5-diazabicyclo-[5.4.0]undec-5-ene on N1- and N3-(methyl 2,3,4-tri-O-acetyl-β-d-glucopyranosyluronate)-5-fluorouracil. The 2H1 conformation of N1- and N3-(methyl 4-deoxy-α-l-threo-hex-4-enopyranosyluronate)-5-fluorouracil has been established by 1H-n.m.r. and c.d. methods. Interaction of the heterocyclic base and the double bond of the sugar moiety is demonstrated.  相似文献   

9.
The kinetics of binding of Cu (II), Tb (III) and Fe(III) to ovotransferrin have been investigated using the stopped-flow technique. Rate constants for the second-order reaction, k +, were determined by monitoring the absorbance change upon formation of the metal-transferrin complex in time range of milliseconds to seconds. The N and C sites appeared to bind a particular metal ion with the same rate; thus, average formation rate constants k + (average) were 2.4 × 104 M–1 s–1 and 8.3 × 104 M–1 S –1 for Cu (II) and Tb (III) respectively. Site preference (N site for Cu (II) and C site for Tb (III)) is then mainly due to the difference in dissociation rate constant for the metals. Fe (III) binding from Fe-nitrilotriacetate complex to apo-ovotransferrin was found to be more rapid, giving an average formation rate constant k + (average) of 5 × 105 M–1 s–1, which was followed by a slow increase in absorbance at 465 nm. This slow process has an apparent rate constant in the range 3 s–1 to 0.5 s–1, depending upon the degree of Fe (III) saturation. The variation in the rate of the second phase is thought to reflect the difference in the rate of a conformational change for monoferric and diferric ovotransferrins. Monoferric ovotransferrin changes its conformation more rapidly (3.4s–1) than diferric ovotransferrin (0.52 s–1). A further absorbance decrease was observed over a period of several minutes; this could be assigned to release of NTA from the complex, as suggested by Honda et al. (1980).Abbreviations Tf ovotransferrin - NTA nitrilotriacetate Jichi Medical School, School of Nursing, Yakushiji 3311-159, Minamikawachi, Tochigi, 329-04 Japan  相似文献   

10.
1. Modification of the Class II sulphydryl groups on the (Na+ + K+)-ATPase from rectal glands of Squalus acanthias with N-ethylmaleimide has been used to detect conformational changes in the protein. The rates of inactivation of the enzyme and the incorporation of N-ethylmaleimide depend on the ligands present in the incubation medium. With 150 mM K+ the rate of inactivation is largest (k1 = 1.73 mM?1 · min?1) and four SH groups per α-subunit are modified. The rate of inactivation in the presence of 150 mM Na+ is smaller (k1 = 1.08 mM?1 · min-1) but the incorporation of N-ethylmaleimide is the same as with K+. 2. ATP in micromolar concentrations protects the Class II groups in the presence of Na+ (k1 = 0.08 mM?1 · min?1 at saturating ATP) and the incorporation id drastically reduced. ATP in millimolar concentrations protects the Class II groups partially in the presence of K+ (k1 = 1.08 mM?1 · min?1) and three SH groups are labelled per α subunit. 3. The K+ -dependent phosphatase is inhibited in parallel to the (Na+ + K+)-ATPase under all conditions, and the ligand-dependent incorporation of N-ethylmaleimide was on the α-subunit only. 4. It is shown that the difference between the Na+ and K+ conformations sensed with N-ethylmaleimide depends on the pH of the incubation medium. At pH 6 there is a very small difference between the rates of inactivation in the presence of Na+ and K+, but at higher pH the difference increases. It is also shown that the rate of inactivation has a minimum at pH 6.9, which suggests that the conformation of the enzyme changes with pH. 5. Modification of the Class III groups with N-ethylmaleimide-whereby the enzyme activity is reduced from about 16% to zero-shows that these groups are also sensitive to conformational changes. As with the Class II groups, ATP in micromolar concentrations protects in the presence of Na+ relative to Na+ or K+ alone. ATP in millimolar concentrations with K+ present increases the rate of inactivation relative to K+ alone, in contrast to the effect on the Class II groups. 6. Modification of the Class II groups with a maleimide spin label shows a difference between Class II groups labelled in the presence of Na+ (or K+) and Class II groups labelled in the presence of K + ATP, in agreement with the difference in incorporation of N-ethylmaleimide. The spectra suggest that the SH group protected by ATP in the presence of K+ is buried in the protein. 7. The results suggest that at least four different conformations of the (Na+ + K+)-ATPase can be sensed with N-ethylmaleimide: (i) a Na+ form of the enzyme with ATP bound to a high-affinity site (E1-Na-ATP); (ii) a Na+ form without ATP bound (E1-Na); (iii) a K+ form without ATP bound (E2-K); and (iv) an enzyme form with ATP bound to a low-affinity site in the presence of K+, probably and E1-K-ATP form.  相似文献   

11.
《Inorganica chimica acta》2006,359(6):1855-1869
A series of discrete, mononuclear palladium(II)–methyl complexes, together with several palladium(II)–chloro analogues, of pyridine-functionalised bis-NHC ligands have been prepared via ligand transmetallation from the silver(I)-NHC complexes. The reported complexes comprise examples with both the methylene-bridged 2,6-bis[(3-R-imidazolin-2-yliden-1-yl)methyl]pyridine (RCNC; R = Mes, dipp, tBu) and planar 2,6-bis(3-R-imidazolin-2-yliden-1-yl)pyridine (RCNC; R = Mes, dipp) ligands and, when combined with the previously reported MeCNC/MeCNC examples, cover a broad spectrum of ligand substituent steric and electronic properties, including the bulky Mes and dipp groups frequently used in catalytic applications. The palladium(II) complexes have been characterised by a variety of methods, including single crystal X-ray crystallography, with the shielding of the Pd–Me groups in the proton NMR spectra of some of the N-aryl substituted examples correlated with the proximity of the aryl rings to the methyl group in the solid state structures. The [PdMe(RCNC/RCNC)]+ complexes undergo thermal degradation via reductive methyl-NHC coupling to give 2-methyl-3-R-imidazolium-1-yl species with relative stabilities in the order of [PdMe(MesCNC)]BF4 > [PdMe(MeCNC)]BF4  [PdMe(MesCNC)]BF4 > [PdMe(MeCNC)]BF4 > [PdMe(tBuCNC)]BF4  [PdMe(tBuCNC)]BF4 (not isolable). A comparison of the activity of the complexes as precatalysts in a model Heck coupling reaction shows greatest activity in those species bearing bulkier N-substituents, with complexes bearing RCNC ligands generally more efficient precatalysts than those bearing RCNC ligands.  相似文献   

12.
[Pt(Me2pipdt)2](BF4)2 salts [Me2pipdt = N,N-dimethyl-piperazine-2,3-dithione] bearing cationic dithiolene complexes react with (Bu4N)2[Pt(X)4] (X = SCN, CN) to form [Pt(Me2pipdt)2][Pt(SCN)4 ] (1) and [Pt(Me2pipdt)2][Pt(CN)4] (2) salts by metathesis. Black crystals of 1 have been structurally characterized showing that the two metals lie on inversion centers and exhibit a square planar coordination. The Pt-S bond distances in the anion complex (2.324(2) Å) are longer than in the cation complex (2.280(2) Å) whereas the C-S bond distances are shorter in SCN (average 1.669 Å) than in Me2pipdt (average 1.694 Å). The chelating Me2Pipdt ligand is found disordered in the λ/δ conformations with site occupancies of 50/50, respectively. The cation and anion complexes run parallel to a.  相似文献   

13.
This paper deals with measurements of the heat of the helix-coil transition by microcalorimetry for kappa carrageenans; data are given in the absence of gel in various solvents (H2O, Me2SO, and formamide) with two counterions (K+, Rb+). In water, the influence of gel formation on ΔH is pointed out and the influence of the time of ageing is demonstrated. The ΔH values measured in the absence of gel are interpreted in terms of the electrostatic model proposed by Manning; the agreement is quite good if a double helix is formed in water, but a monochain, ordered conformation is suggested for solutions in Me2SO and formamide.  相似文献   

14.
This report describes synthesis and evaluation of cationic complexes, [99mTc(CO)3(L)]+ (L = N-methoxyethyl-N,N-bis[2-(bis(3-ethoxypropyl)phosphino)ethyl]amine (L1), N-[(15-crown-5)-2-yl]-N,N-bis[2-(bis(3-ethoxypropyl)phosphino)ethyl]amine (L2) and N-[(18-crown-6)-2-yl]-N,N-bis[2-(bis(3-ethoxypropyl)phosphino)ethyl]amine (L3)) as potential radiotracers for heart imaging. Preliminary results from biodistribution studies in female adult BALB-c mice indicated that the cationic 99mTc(I)-tricarbonyl complex, [99mTc(CO)3(L2)]+, has a significant localization in the heart at 60 min post-injection. To understand the coordination chemistry of these bisphosphine ligands with the 99mTc(I)-tricarbonyl core, we prepared [Re(CO)3(L4)]Br (L4: N,N-bis[(2-diphenylphosphino)ethyl]methoxyethylamine) as a model compound. [Re(CO)3(L4)]Br has been characterized by elemental analysis, IR, ESI-MS, NMR (1H, 13C, 1H-1H COSY, and 1H-13C HMQC) methods, and X-ray crystallography. In solid state, [Re(CO)3(L4)]+ has a distorted octahedron coordination geometry with PNP occupying one facial plane. The chelator backbone adopts a “chair” conformation with phosphine-P atoms at equatorial positions and the amine-N at the apical site. In solution, [Re(CO)3(L4)]+ is able to maintain its cationic nature with no dissociation of carbonyl ligands or any of the three PNP donors.  相似文献   

15.
Poly(rI) stabilized by either Na+ or K+ was investigated using uv resonance Raman (UVRR) spectroscopy. Raman excitation profiles of inosine 5′-monophosphate demonstrated the 250 nm excitation selectively enhances base stacking interactions, while ribose and carbonyl stretching vibrations are preferentially enhanced with 210 nm excitation. These wavelengths were used to examine the structure of poly(rI) in the presence of either K+ or Na+ as a function of temperature. UVRR studies revealed that the K+ stabilized form is more thermally stable, yielding a Tm of ∼ 47°C compared to a Tm of ∼ 30°C for the Na+ stabilized form. We observed that both the ribosyl conformation and the coordination of the carbonyl groups depend on the nature of the cation. The C6O stretching frequency indicates that Na+ coordinates much more strongly to the carbonyl groups than K+ (1672 cm−1 Na+ vs 1684 cm−1 K+ at 4°C). Conformationally sensitive modes of the phosphate backbone and the ribosyl ring indicate that Na+ stabilized poly(rI) predominantly exists in a C3′-endo ribose conformation, whereas K+ stabilized poly(rI) adopts a C2′-endo conformation possibly as a consequence of the larger ionic radius of the K+ ion. © 1998 John Wiley & Sons, Inc. Biopoly 46: 475–487, 1998  相似文献   

16.
Fractionation of Acacia senegal gum has been carried out on Sephacryl gels S-400 and S-500. The shape and relative height of the chromatograms are sample dependent.Physicochemical data including circular dichroism and the weight-average molecular weights distribution show that about 70% of the material is composed of homogeneous polysaccharide chains with a very low nitrogen content. The remaining material is a combination of polysaccharide and nitrogen moieties (probably an arabinogalactan-protein complex).Our data are consistent with the new and simplified structural models proposed recently for this polysaccharide.  相似文献   

17.
The effect of different co-anions on the formation and aggregation of the ordered structure of the anionic polysaccharide kappa-carrageenan has been investigated by optical rotation, differential scanning calorimetry, and halide-n.m.r. spectroscopy. The mid-point temperature (Tm) of the disorder—order transition increases systematically with the Hofmeister number for the anion through the lyotropic series SO42? < F? < Cl? < Br? < NO3? < I? < SCN? when salt concentration and cation (Me4N+ or K+) are held constant. A corresponding increase is observed in transition enthalpy (ΔHcal) and entropy (ΔScal). Helix—helix aggregation (as indicated by turbidity, gel formation, and hysteresis between heating and cooling scans) also shows a systematic dependence on the Hofmeister number for the anion, but in the opposite sense. Thus, with tetramethylammonium as the sole counterion present, clear solutions with no thermal hysteresis in the order—disorder transition are observed at all temperatures with I?, Br?, NO3?, or Cl? as co-anion, whereas weak, turbid gels with significant thermal hysteresis between melting and setting are formed in the presence of SO42?, and to a lesser degree F?. With K+ as counterion, a similar regular progression is observed through the anion lyotropic series from rapid formation of very turbid gels in the presence of F?, to very slow development of clear gels with I? or SCN?. In agreement with previous studies, an increase in 127I-n.m.r. linewidth was observed on conformational ordering of kappa-carrageenan (Me4N+ salt form) in the presence of Me4NI. However, closely similar behaviour was observed for 35Cl and 81Br, indicating a simple charge-cloud interaction rather than the specific site-binding of I? which has previously been suggested.  相似文献   

18.
Synthesis and characterisation of the new macrocyclic ligand 1,7-dimethyl-4,10-di(methylcarbamoylmethy)-1,4,7,10-tetraazacyclododecane (L) are reported. The ligand, based on cyclen (1,4,7,10-tetraazacyclododecane), has been functionalised by the insertion of two methyl groups and two amidic pendant arms linked to the amine nitrogens. The interaction of L with H+, Na(I), Ca(II), Cu(II), Zn(II), Pb(II), and Gd(III) ions has been studied by potentiometric titrations, microcalorimetric and 1H NMR measurements in 0.1 mol dm−3 Me4NCl aqueous solution at 298.1±0.1 K. The thermodynamic data suggest that the N4 moiety is the binding site for Cu(II) and Zn(II), while in the case of Pb(II) also the pendant arms are coordinated to the metal ion. The crystal structure of [PbL](ClO4)2 (space group P21/a, a=12.883(2) Å, b=12.259(3) Å, c=17.275(5) Å, β=108.65(2)°, V=2585.0(11) Å3, Z=4, R=0.0660, RW 2=0.1467) shows the metal ion hexa-coordinated by the four nitrogen atoms of the cyclic tetra-amine and by the two amidic oxygens of the pendant arms.  相似文献   

19.
Abstract

The conformational properties of the tetrapeptide Ser1-Pro2-Phe3-Arg4, the C-terminal fragment of the nonapeptide hormone bradykinin, have been studied by circular dichroism and two-dimensional NMR techniques. Measurements of coupling constants, NH temperature dependence rates and nuclear Overhauser effects (performed with rotating frame nuclear Overhauser spectroscopy, ROESY) in H2O and CD3OH/D2O (80/20, v/v) reveal different conformations in the corresponding solvent. In aqueous solution the molecule exists in a random conformation or as an average of several conformations in rapid exchange. In CD3OH/D2O, however, the conformation is well-defined. The backbone of the peptide is extended, and the side-chains of Phe3 and Arg4 exhibit unusual rigidity for a peptide of this size. Evidently, the secondary structure is stabilized by a charge interaction between the guanidino group of Arg4 and the terminal carboxyl group, since experiments at various pH's show clearly that the definition of conformation decreases strongly upon protonation of the carboxyl function. A NH3 +(Ser1)-COO?(Arg4) salt bridge, as well as any form of turn stabilized by hydrogen bonds can be ruled out with certainty.  相似文献   

20.
To prepare novel hydrogels for use in water technologies, guar gum was subjected to acid hydrolysis. The depolymerized guar gum obtained there from and the native guar gum were oxidized to their respective polycarboxylic forms using NOx as oxidant. All these polymers were crosslinked with N,N-methylenebisacrylamide, and were used as Cu2+ sorbents. The candidate hydrogel exhibiting the highest uptake was used further to investigate the effect of external stimuli on sorption. The sorption on hydrogels was fast as the highest sorption was observed after 2 h at 40 °C and 20 ppm of Cu2+ ions. The hydrogel prepared from the oxidized guar gum afforded the maximum sorption capacity of 125.893 mg g−1. Langmuir and Freundlich isotherms, and pseudo second order kinetics matches the experimental data. The evidence of sorption was obtained by characterizing Cu2+-loaded hydrogels by FTIR spectroscopy.  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