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 46 毫秒
1.
2.
The binding site of glycyrrhizin (GLZ) on human serum albumin was detected through competitive displacement experiments with GLZ and ibuprofen (IBU) (diazepam site), warfarin (WAR), salicylate (SAL) (digitoxin site), or deoxycholic acid (DCA) by means of an ultrafiltration technique. The specific binding of GLZ was subject to competitive inhibition by IBU, WAR, SAL, or DCA (1 or 4 mM). The extent of displacement was in the order of: DCA greater than IBU greater than WAR greater than SAL. Conversely, the specific bindings of WAR and DCA and the low-affinity bindings of IBU and SAL were subject to competitive inhibition by GLZ (1 or 4 mM). The extent of inhibition by GLZ was in the order of DCA greater than IBU greater than WAR not equal to SAL. In addition, the low-affinity IBU binding and the specific DCA binding showed mutual competitive inhibition at 4 mM, with almost identical displacements. It was concluded that the specific GLZ binding site on human serum albumin may be located mostly within the low-affinity IBU binding site area (probably the same as the specific DCA binding site area) and partially within the specific WAR binding site area and the low-affinity SAL binding site area.  相似文献   

3.
Two classes of binding sites, a single high-affinity site with an association constant of 4·8×106 M−1 and two low-affinity sites with association constant of about 0·05×106 M−1 have been observed in the interaction of Naproxen with bovine serum albumin (BSA). Chemical modification of two tryptophan residues in BSA with 2-hydroxy-5-nitrobenzyl bromide has led to a reduction in the association constant of the high-affinity site by 89% and its number of binding sites by 66% suggesting the involvement of tryptophan residues in the high-affinity site. In contrast, the two low-affinity sites were not affected by the modification. Binding of Naproxen to the low-affinity sites of BSA induces microdisorganisation of the albumin structure leading to conformational changes as evident from fluorescence measurements with 1-anilino-8-naphthalenesulphonic acid as the probe.  相似文献   

4.
Human red blood cells (RBCs) were loaded with the Ca(2+)-sensitive fluorescent dye fura-2 to investigate the effects of media ionic strength and prostaglandin E2 (PGE2) on the intracellular free Ca2+ concentration ([Ca2+]i). [Ca2+]i of intact RBCs in a Ca(2+)-containing physiological (high) ionic strength (HIS) solution was 75.1 +/- 8.3 nM after 5 min incubation, increasing to 114.9 +/- 9.6 nM after 1 h. In Ca(2+)-containing low ionic strength (LIS) solutions, [Ca2+]i was significantly lower than in the Ca(2+)-containing HIS solution (p = 0.041 or 0.0385 for LIS solutions containing 200 or 250 mM sucrose, respectively), but, as in HIS solution, an increase of [Ca2+]i was seen after 1 h. In Ca(2+)-free (0 Ca2+ plus 15 microM EGTA) media, [Ca2+]i decreased (ranging from 15 to 21 nM), but were not significantly different in HIS or LIS, and did not change following 1 h incubation. The effect of the ionic strength and PGE2 on passive Ca2+ influx was investigated on ATP-depleted RBCs. Ca2+ influx was faster during the initial 10 min in comparison with the subsequent time period (10-45 min), both in HIS and LIS media, decreasing from 20.3 +/- 1.9 to 12.9 +/- 1.3 micromol/(lcells x h) in HIS, and from 36.7 +/- 5.3 to 8.6 +/- 1.2 micromol/(lcells x h) in LIS. Prostaglandin E2 (PGE2; 10(-7)-10(-11) M), dissolved in deionised water or in ethanol, did not affect [Ca2+]i in either normal or in ATP-depleted RBCs suspended in Ca(2+)-containing HIS medium. Finally, the addition of carbachol (100 microM) did not affect [Ca2+]i. The present findings suggest that stimulation of the Ca(2+)-activated K+ channel by PGE2, reported in [J. Biol. Chem. 271 (1996) 18651], cannot be mediated via increased [Ca2+]i.  相似文献   

5.
ISU (eukaryotes) and IscU (prokaryotes) are a homologous family of proteins that appear to provide a platform for assembly of [2Fe-2S] centers prior to delivery to an apo target protein. The intermediate [2Fe-2S] ISU-bound cluster is formed by delivery of iron and sulfur to the apo ISU, with the latter delivered through an IscS-mediated reaction. The identity of the iron donor has thus far not been established. In this paper we demonstrate human frataxin to bind from six to seven iron ions. Iron binding to frataxin has been quantitated by iron-dependent fluorescence measurements [K(D)(Fe(3+)) approximately 11.7 microM; (K(D)(Fe(2+)) approximately 55.0 microM] and isothermal titration calorimetry (ITC) [K(D)(Fe(3+)) approximately 10.2 microM]. Enthalpies and entropies for ferric ion binding were determined from calorimetric measurements. Both fluorescence (K(D) 0.45 microM) and ITC measurements (K(D) 0.15 microM) demonstrate holo frataxin to form a complex with ISU with sub-micromolar binding affinities. Significantly, apo frataxin does not bind to ISU, suggesting an important role for iron in cross-linking the two proteins and/or stabilizing the structure of frataxin that is recognized by ISU. Holo frataxin is also shown to mediate the transfer of iron from holo frataxin to nucleation sites for [2Fe-2S] cluster formation on ISU. We have demonstrated elsewhere [J. Am. Chem. Soc. 2002, 124, 8774-8775] that this iron-bound form of ISU is viable for assembly of holo ISU, either by subsequent addition of sulfide or by NifS-mediated sulfur delivery. Provision of holo frataxin and inorganic sulfide is sufficient for cluster assembly in up to 70% yield. With NifS as a sulfur donor, yields in excess of 70% of holo ISU were obtained. Both UV-vis and CD spectroscopic characteristics were found to be consistent with those of previously characterized ISU proteins. The time course for cluster assembly was monitored from the 456 nm absorbance of holo ISU formed during the [2Fe-2S] cluster assembly reaction. A kinetic rate constant k(obs) approximately 0.075 min(-)(1) was determined with 100 microM ISU, 2.4 mM Na(2)S, and 40 microM holo frataxin in 50 mM Tris-HCl (pH 7.5) with 4.3 mM DTT. Similar rates were obtained for NifS-mediated sulfur delivery, consistent with iron release from frataxin as a rate-limiting step in the cluster assembly reaction.  相似文献   

6.
NikR represses expression of a nickel transporter in response to elevated levels of Ni(II). Recent results suggest that repression is elicited by binding of nickel to a high-affinity site, but a low-affinity binding pocket may also play a role.  相似文献   

7.
人血清白蛋白与季铵盐双子表面活性剂的相互作用   总被引:2,自引:0,他引:2  
在298.15 K下, 应用等温滴定量热法研究了人血清白蛋白(HSA)与两种季铵盐双子表面活性剂[(CnN)2Cl2, n=12, 14]在缓冲溶液(pH=7.0)中相互作用的热力学性质. 实验结果表明, HSA对这两种表面活性剂有两类结合位点, 分别为结合时需要吸收热量的强结合位点和可放出热量的弱结合位点. 两种表面活性剂对应的第一类结合——强结合为熵驱动过程, 且该结合位点对应的结合位点数、结合常数和热力学参数差别不大. 至于第二类结合——弱结合位点, 由于 (C14N)2Cl2 疏水链过长, 只有部分进入HSA的疏水空腔内, 因此相应的的结合位点数和放热量减小, 而熵变增加, 为焓和熵共同驱动的反应. 圆二色研究表明(CnN)2Cl2的加入使HSA的二级结构发生变化, 这说明(CnN)2Cl2与HSA的相互作用既包含结合反应也包含(CnN)2Cl2诱导该蛋白部分结构改变的过程.  相似文献   

8.
Microcalorimetric titrations allow to recognize and investigate high-affinity ligand binding to Na,K-ATPase. Titrations with the cardiac glycoside Ouabain, which acts as a specific inhibitor of the enzyme, have provided not only the thermodynamic parameters of high-affinity binding with a stoichiometric coefficient of about 0.6 but also evidence for low-affinity binding to the lipid. The marked enthalpic contribution of -95 kJ mol-1 at 298.2 K is partially compensated by a large negative entropy change, attributed to an increased interaction between water and the protein. The calorimetric ADP and ATP titrations at 298.2 K are indicative of high-affinity nucleotide binding either in 3 mM NaCl, 3 mM MgCl2 or at high ionic strength such as 120 mM choline chloride. However, no binding is detected in the buffer solution alone at low ionic strength. The affinities for ADP and ATP are similar, around 106 M-1 and the stoichiometric coefficients are close to that of Ouabain binding. The exothermic binding of ADP is characterized by a ΔH and ΔS value of -65 kJ mol-1 and -100 J mol-1 K-1, respectively. TheΔH value for ATP binding is larger than for ADP and is compensated by a larger, unfavorable ΔS value. This leads to an enthalpy/entropy compensation, which could express that H-bond formation represents the major type of interaction. As for Ouabain, the negative ΔS values that are also characteristic of nucleotide binding can indicate an increase of solvate interaction with the protein due to a conformational transition occurring subsequent to the binding process. The resulting binding constants are discussed with regard to the results of other studies employing different techniques. A molecular interaction model for nucleotide binding is suggested. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

9.
Whether La3+ can enter human peripheral blood lymphocytes by the Na+/Ca2+ exchanger or not and the effect of La3+on the Na+/Ca2+ exchanger activity are examined by fura-2 technique. And that whether La3+ is sequestered by intracellular organelles (mainly endoplasmic reticulum and mitochondria) is studied by this method. La3+uptake is obviously stimulated by pre-treating the cells with ouabain and by removing extracellular Na+, and intracellular La3+concentration ([La3+]i) is directly proportional to its extracellular concentration ([La3+]o). But when [La3+]o exceeds 0.4 mmol/L, the 340/380 nm ratio of fluorescence is no longer varied and the maximum [La3+], is 1.5×10-12 mol · L-1. The higher concentration of La3+ (0.1 mmol/L) increases Na+/Ca2+ exchange-mediated calcium influx, but lower concentration (10 μmol/L) appears to block calcium influx. The results also suggest that cytosolic La3+ is transported by the ATP-dependent Ca2+ pump. Intracellular Ca2+ stores are depleted by ionomycin, and then ion  相似文献   

10.
The surface properties of ferrihydrite were studied by combining wet chemical data, Cd(K) EXAFS data, and a surface structure and protonation model of the ferrihydrite surface. Acid-base titration experiments and Cd(II)-ferrihydrite sorption experiments were performed within 3<-log[H(+)]<10.5 and 0.5<[Cd(t)]<12 mM in 0.3 M NaClO(4) at 25 degrees C, where [Cd(t)] refers to total Cd concentration. Measurements at -5.5triple bond Fe-OH(-1/2),logk((int))=-8.29, assuming the existence of a unique intrinsic microscopic constant, logk((int)), and consequently the existence of a single significant type of acid-base reactive functional groups. The surface structure model indicates that these groups are terminal water groups. The Cd(II) data were modeled assuming the existence of a single reactive site. The model fits the data set at low Cd(II) concentration and up to 50% surface coverage. At high coverage more Cd(II) ions than predicted are adsorbed, which is indicative of the existence of a second type of site of lower affinity. This agrees with the surface structure and protonation model developed, which indicates comparable concentrations of high- and low-affinity sites. The model further shows that for each class of low- and high-affinity sites there exists a variety of corresponding Cd surface complex structure, depending on the model crystal faces on which the complexes develop. Generally, high-affinity surface structures have surface coordinations of 3 and 4, as compared to 1 and 2 for low-affinity surface structures.  相似文献   

11.
Thioflavin-T is one of the most important amyloid specific dyes and has been used for more than 50 years; however, the molecular mechanism of staining is still not understood. Chemically synthesized short polyglutamine peptides (Q(n), n = 5-10) were subjected to the thioflavin-T (ThT) staining assay. It was found that the minimum Q(n) peptide that stained positive to ThT was Q(6). Two types of ThT-binding sites, a high-affinity site (k(d1) = 0.1-0.17 μM) and a low-affinity site (k(d2) = 5.7-7.4 μM), were observed in short polyQs (n = 6-9). (13)C{(2)H}REDOR NMR experiments were carried out to extract the local structure of ThT binding sites in Q(8) peptide aggregates by observing the intermolecular dipolar coupling between [3-Me-d(3)]ThT and natural abundance Q(8) or residue-specific [1,2-(13)C(2)] labeled Q(8)s. (13)C{(2)H}REDOR difference spectra of the [3-Me-d(3)]ThT/natural abundance Q(8) (1/9) complex indicated that all of the five carbons of the glutamine residue participated in the formation of ThT-binding sites. (13)C{(2)H}DQF-REDOR experiments of [3-Me-d(3)]ThT/residue-specific [1,2-(13)C(2)] labeled Q(8) (1/50) complexes demonstrated that the N-terminal glutamine residue had direct contact with the ThT molecule at the high-affinity ThT-binding sites.  相似文献   

12.
Glycan-targeting antibodies and pseudo-antibodies have been extensively studied for their stoichiometry, avidity, and their interactions with the rapidly modifying glycan shield of influenza A. Broadly neutralizing antiviral agents bind in the same order when they neutralize enveloped viruses regardless of the location of epitopes to the host receptor binding site. Herein, we investigated the binding of cyanovirin-N (CV–N) to surface-expressed glycoproteins such as those of human immunodeficiency virus (HIV) gp120, hemagglutinin (HA), and Ebola (GP)1,2 and compared their binding affinities with the binding response to the trimer-folded gp140 using surface plasmon resonance (SPR). Binding-site knockout variants of an engineered dimeric CV–N molecule (CVN2) revealed a binding affinity that correlated with the number of (high-) affinity binding sites. Binding curves were specific for the interaction with N-linked glycans upon binding with two low-affinity carbohydrate binding sites. This biologically active assembly of a domain-swapped CVN2, or monomeric CV–N, bound to HA with a maximum KD of 2.7 nM. All three envelope spike proteins were recognized at a nanomolar KD, whereas binding to HIV neutralizing 2G12 by targeting HA and Ebola GP1,2 was measured in the µM range and specific for the bivalent binding scheme in SPR. In conclusion, invariant structural protein patterns provide a substrate for affinity maturation in the membrane-anchored HA regions, as well as the glycan shield on the membrane-distal HA top part. They can also induce high-affinity binding in antiviral CV–N to HA at two sites, and CVN2 binding is achieved at low-affinity binding sites.  相似文献   

13.
In this paper, we present a computational study of the dynamics of the potent anti-HIV virucidal protein cyanovirin in complex with mannose disaccharides. Recently, it has been experimentally demonstrated that cyanovirin binds mannose oligomers on the surface of glycoprotein gp120. gp120, a protein on the surface of the HIV virus, is key in the process of viral docking and transfer of genetic material into human cells. Cyanovirin prevents the transfer of viral RNA into human cells. In this study, we found that, among all residues that show nuclear Overhauser effects in the solution NMR experiments, residues Glu41 and Arg76 appear to interact with the sugar at the high-affinity binding site through stronger Coulombic interactions. In particular, Arg76 participates in a dynamical mechanism that caps and locks the sugar once it is bound to the protein. We also studied the distribution of glycosidic torsional angles of mannose disaccharides in solution and compared it with those when bound at the high- and low-affinity sites of the protein. Throughout our 20 ns simulations, we find that the sugar bound to the high-affinity site preserves the most favorable conformation in solution while the sugar bound at the low-affinity site does not. The sugar at the low-affinity site can adopt both conformations, but we find it most predominantly on the one that is least probable for the free sugar in solution. We also carried out a detailed study of the interactions between the disaccharides and different amino acids as well as between the disaccharide and the solvent at both binding locations.  相似文献   

14.
A negatively charged region on the surface of photosystem II (PSII) near Q(A) has been identified as a docking site for cationic exogenous electron acceptors. Oxygen evolution activity, which is inhibited in the presence of the herbicide 3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU), is recovered by adding Co(III) complexes. Thus, a new electron-transfer pathway is created with Co(III) as the new terminal electron acceptor from Q(A)(-). This binding site is saturated at ~2.5 mM [Co(III)], which is consistent with the existence of low-affinity interactions with a solvent-exposed surface. This is the first example of a higher plant PSII in which the electron-transfer pathway has been redirected from the normal membrane-associated quinone electron acceptors to water-soluble electron acceptors. The proposed Co(III) binding site may enable efficient collection of electrons generated from photochemical water oxidation by PSII immobilized on an electrode surface.  相似文献   

15.
Adduct formation between cationic reagents and targets on DNA are facilitated by the ability of DNA to attract cations to its surface. The electrostatic interactions likely provide the basis for the documented preference exhibited by cisplatin and related compounds for nuclear DNA over other cellular constituents. As an extension of a previous communication, we here present an investigation illustrating how the rate of adduct formation with the naturally occurring base guanine (G-N7) can be modulated by i) bulk solvent conditions, ii) local nature and size of the surrounding DNA and, iii) increasing DNA concentration. A series of single-stranded DNA oligomers of the type d(TnGTm); n= 0, 2, 4, 6, 8, 10, 12, 14, 16 and m= 16 -n or n=m= 4, 6, 8, 12, 16, 24 were allowed to react with the active metabolite of a potential orally active platinumIV drug, cis-[PtCl(NH3))(c-C6H11NH2)(OH2)]+ in the presence of three different bulk cations; Na+, Mg2+, and Mn2+. For all positions along the oligomers, a change from monovalent bulk cations to divalent ones results in a decrease in reactivity, with Mn2+ as the more potent inhibitor as exemplified by the rate constants determined for interaction with d(T8GT8): 10(3) x k obs/s(-1)= 6.5 +/- 0.1 (Na+), 1.8 +/- 0.1 (Mg2+), 1.0 +/- 0.1 (Mn2+) at pH 4.2 and 25 degrees C. Further, the adduct formation rate was found to vary with the exact location of the binding site in the presence of both Na+ and Mg2+, giving rise to reactivity maxima at the middle position. Increasing the size of the DNA-fragments was found to increase the reactivity only up to a total length of ca. 20 bases. The influence from addition of further bases to the reacting DNA was found to be salt dependent. At [Na+]= 0.5 mM a retardation in reactivity was observed whereas [Na+] < or = 4.5 mM give rise to length independent kinetics. Finally, for the first time we have here been able to evaluate the influence from an increasing concentration of non-reactive DNA bases on the adduct formation process. The latter data were successfully fitted to an inhibition model suggesting that non-productive association of the platinum complex with sites distant from G-N7 competes with productive ones in the vicinity of the G-N7 target. Taken together, the kinetics support a reaction mechanism in which access to suitable association sites in the direct vicinity of the target site controls the rate of platination.  相似文献   

16.
Binding of human beta-endorphin (beta-EP) to rat renal basolateral membranes was characterized using [125I]Tyr27-beta-EP ([125I]beta-EP) as a primary ligand. Ten millimolar of ethylenediaminetetra acetic acid (EDTA) completely inhibited the degradation of [125I]beta-EP in the incubation mixture at 4 degrees C, thus making it possible to quantitatively examine the [125I]beta-EP binding. The specific binding of [125I]beta-EP to the basolateral membranes was reversible and saturable, and a nonlinear least-squares regression analysis of a saturation isotherm revealed two different classes of specific binding sites. One class had an apparent dissociation constant (Kd) of 0.68 nM and a lower number of binding sites (33 fmol/mg protein), whereas the other class had a lower affinity (apparent Kd of 210 nM) and a higher number of binding sites (7.3 pmol/mg protein). Inhibition of the [125I]beta-EP binding by naloxone (10 microM) was approximately only 20%, and that by D-Ala2-D-Leu5-enkephalin (10 microM) was null, suggesting the major role of a non-opioid binding component in specific [125I]beta-EP binding to basolateral membranes. Moreover, a 50% inhibition by 10 microM of dynorphin(1-13) suggests that a certain region of the primary structure of beta-EP, excluding at least the NH2-terminal enkephalin sequence, is of particular importance for the [125I]beta-EP binding. These lines of evidence suggest the existence of two different classes of specific binding sites for beta-EP on the renal basolateral membranes, and the high-and low-affinity bindings may be attributed to opioid and non-opioid receptors, respectively, as judged by known characteristics of opioid and non-opioid receptors in other peripheral tissues.  相似文献   

17.
The speciation in the phosphitomolybdate system, H+-MoO4(2-)-(HP)O(3)2-, has been determined from combined potentiometric and 31P NMR measurements in 0.600 M Na(Cl) medium at 298(1) K. Potentiometric titration data were collected in the ranges 2.5<-log[H+]<6.2, 40.0相似文献   

18.
Binding of muscimol-conjugated quantum dots to GABAC receptors   总被引:1,自引:0,他引:1  
Functionalization of highly fluorescent CdSe/ZnS core-shell nanocrystals (quantum dots, qdots) is an emerging technology for labeling cell surface proteins. We have synthesized a conjugate consisting of approximately 150-200 muscimols (a GABA receptor agonist) covalently joined to the qdot via a poly(ethylene glycol) (PEG) linker (approximately 78 ethylene glycol units) and investigated the binding of this muscimol-PEG-qdot conjugate to homomeric rho1 GABAC receptors expressed in Xenopus oocytes. GABAC receptors mediate inhibitory synaptic signaling at multiple locations in the central nervous system (CNS). Binding of the conjugate was analyzed quantitatively by determining the fluorescence intensity of the oocyte surface membrane in relation to that of the surrounding incubation medium. Upon 5- to 10-min incubation with muscimol-PEG-qdots (34 nM in qdot concentration), GABAC-expressing oocytes exhibited a fluorescent halo at the surface membrane that significantly exceeded the fluorescence of the incubation medium. This halo was absent following muscimol-PEG-qdot treatment of oocytes lacking GABAC receptors. Incubation of the oocyte with free muscimol (100 microM-5 mM), PEG-muscimol (500 microM), or GABA (100 microM - 5 mM) substantially reduced or eliminated the fluorescence halo produced by muscimol-PEG-qdots, and the removal of GABA or free muscimol led to a recovery of muscimol-PEG-qdot binding. Unconjugated qdots and PEG-qdots that lacked conjugated muscimol neither exhibited significant binding activity nor diminished the subsequent binding of muscimol-PEG-qdots. The results indicate that muscimol joined to qdots via a long-chain PEG linker exhibits specific binding activity at the ligand-binding pocket of expressed GABAC receptors, despite the presence of both the long PEG linker and the sterically bulky qdot.  相似文献   

19.
Calcium green I, a ratiometric probe based on fluorescence lifetime measurements, was used to monitor intracellular calcium activity ([Ca2+]i) in RINm5F cells using a time-resolved fluorescence confocal microscope. The probe affinity constant has been recalibrated in single cells using ionomycin as a calcium ionophore and ethylenebis(oxyethylenenitrilo)tetraacetic acid as a calcium buffer; Kd was found to equal 150 nmol/L. The kinetics of ionomycin equilibration showed that the calcium release from calcium stores occurs before equilibration with extracellular calcium. The response to the muscarinic agonist carbachol, measured on 17 cells receiving three consecutive applications was characterized both by a [Ca2+]i peak lasting 50 s without any trailing plateau and by desensitization with a 30% decrease in the response. The dose-dependent response was obtained for a carbachol concentration from 5 mumol/L to 0.5 mmol/L. The ability of our set-up to obtain a value every 10 ms enabled us to record asynchronous spikes of [Ca2+]i in the RINm5F cells. The spikes, lasting less than 1 s, are significantly bigger than the noise, and they are not observed in the colonic HT29 cells.  相似文献   

20.
Hu W  Haddad PR  Hasebe K  Tanaka K 《The Analyst》2001,126(5):555-558
An ion chromatographic (IC) method based on the use of titrant (strong acid) as the stationary phase was developed for simultaneous determination of total alkalinity (TA) and monovalent cations. The titrant used in this study was obtained by initially loading lithium dodecylsulfate (Li-DS) onto a reversed-phase material and then conditioning the column with a slightly acidified aqueous LiCl solution (a mixture of 50.0 mM LiCl and 0.1 mM H2SO4). When a small amount of a basic sample was injected onto a column prepared in this way, the basic species (Bn-) reacted predominantly with H+ on the stationary phase and the reaction with the eluent phase was negligible due to the very low concentration of eluent H+ (in the eluent, a molar ratio of [Li+]/[H+] = 250:1 applied). The stationary phase H+ consumed in the acid-base reaction was then re-supplied by H+ from the eluent. By monitoring the conductance of the eluent using conductivity, an induced peak resulting from the basic species was observed. Calibration graphs of peak areas vs. molar concentration of the basic species for OH-, HCO3- and H2PO4- were found to be identical. CO3(2-), HPO4(2-), and B4O7(2-) also gave identical calibration curves but their slope values were twice those for HCO3-. The detection limit for HCO3- was less than 3.2 microM and the calibration curve was linear up to 12.3 mM (injection volume, 100 microL). Seawater was directly analyzed and its total alkalinity was found to be 2.87 mM (RSD 0.53%, n = 5), which was in good agreement with the result of 2.88 mM (RSD 3.2%, n = 5) obtained using auto-potentiometric titration. Na+ and K+ were determined simultaneously and the concentrations were 481.6 and 10.6 mM, respectively.  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