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 31 毫秒
1.
It is shown that the glass electrode may be used without appreciable error to measure pH of ammonia or ammonium chloride buffers, but that corrections must be applied above pH 8.6 if sodium ions are present in the unknown solution. Corrections are given for values from pH 8.6 to 9.4. A slight further modification of form of glass electrode used by previous workers is described.  相似文献   

2.
微生物培养基质量控制技术和标准   总被引:1,自引:0,他引:1  
微生物培养基的酸碱度、凝胶强度和选择性等直接影响到培养基的质量,在理化试验方法中采用连接可渗透陶器型液体接头的电极和平头电极或者连接微型探头的电极可分别测定液体和固体培养基的pH值,而采用Gelometer和the LFRA Texture Analyser可测定固体培养基的凝胶强度。在微生物学方法中固体培养基采用倾注平板法、涂布法、划线法(半定量法)、改良的Miles-Misra法等测定生长情况,液体培养基采用稀释法测定生长率,用目标菌和杂菌的混合菌株评价选择性增菌培养基的选择性,利用OD值评价液体培养基生长率等。ICFMH(国际食品微生物学和卫生学委员会培养基工作组)、ISO、FDA以及我国卫生部等相继制定了培养基质量控制的标准,但目前还没有一个系统的适合我国国情的培养基质量控制国家标准,以致各相关单位采用的标准不一致,所以制定培养基质量控制国家标准非常关键。  相似文献   

3.
Min KR  Zimmer MN  Rickard AH 《Biofouling》2010,26(8):931-940
The aim of this study was to explore the physicochemical parameters that influence coaggregation between the freshwater bacteria Sphingomonas natatoria 2.1 and Micrococcus luteus 2.13. Using visual coaggregation assays, the effect of different buffers, solutions of differing ionic strength, pH, temperature, and viscosity on the degree of coaggregation was assessed. Coaggregation occurred maximally in distilled water but was inhibited when coaggregates were suspended in a commonly-used oral bacterial coaggregation buffer, saline solutions, and Tris-Cl buffers. Coaggregation was weakly expressed in standard laboratory buffers. The ionic strength of inorganic salt solutions required to inhibit coaggregation depended upon the inorganic salt being tested. Coaggregation occurred at a pH of 3-10, between 5 and 80°C and was inhibited in solutions with a viscosity of 22.5 centipoises at 20°C. Inhibition of coaggregation with NaCl impaired biofilm development. When developing buffers to test for coaggregation, the natural liquid environment should be considered. Coaggregation between S. natatoria 2.1 and M. luteus 2.13 is only affected by physicochemical conditions beyond those typically found in natural freshwater ecosystems. Such a robust ability to coaggregate may enhance the ability of S. natatoria 2.1 and M. luteus 2.13 to develop a niche in freshwater biofilms.  相似文献   

4.
Summary Buffer solutions and incubating media for enzyme histochemistry are discussed in terms of pH, ionic strength and buffering capacity. A specially written program is presented. This program enables (i) buffers and media of known pH and ionic strength to be formulated; (ii) the ionic strength of a buffer solution of known molarity, and (iii) the thermodynamic acid dissociation constant of a buffer substance, to be calculated.  相似文献   

5.
Discrepancies of one pH unit and more have been observed after a few days, between continuous on-line in situ pH measurements and instant off-line pH measurements during anaerobic digestion of an agroindustrial wastewater. Concomitantly, the electrical resistance across the porous diaphragm of the on-line electrode increased, and a black clogging developed on its diaphragm. Measurements of the relative liquid junction potential in KCl or Na(2)S solutions excluded that high concentrations of ions such as, K(+), Na(+), Cl(-), HS(-), or S(2-) were the major cause of the drifts in pH values. It has been possible to limit the rapid increase of the liquid junction potential and the black clogging formation on the porous diaphragm either by acidification and/or by overpressurization of the electrode-filling liquid. Continuous on-line in situ pH values consistent with instant off-line pH values over long periods of time have been obtained with a newly designed pH sensor in which a special jellied electrode filling replaced the porous diaphragm.  相似文献   

6.
The impact of pH and ionic strength on the mobility (individual and co-transport) and deposition kinetics of TiO2 and ZnO NPs in porous media was systematically investigated in this study. Packed column experiments were performed over a series of environmentally relevant ionic strengths with both NaCl (0.1−10 mM) and CaCl2 (0.01–0.1mM) solutions and at pH 5, 7, and 9. The transport of TiO2 NPs at pH 5 was not significantly affected by ZnO NPs in solution. At pH 7, a decrease in TiO2 NP transport was noted with co-existence of ZnO NPs, while at pH 9 an increase in the transport was observed. At pH 5 and 7, the transport of ZnO NPs was decreased when TiO2 NPs was present in the solution, and at pH 9, an increase was noted. The breakthrough curves (BTC) were noted to be sensitive to the solution chemistries; the decrease in the breakthrough plateau with increasing ionic strength was observed under all examined pH (5, 7, and 9). The retention profiles were the inverse of the plateaus of BTCs, as expected from mass balance considerations. Overall, the results from this study suggest that solution chemistries (ionic strength and pH) are likely the key factors that govern the individual and co-transport behavior of TiO2 and ZnO NPs in sand.  相似文献   

7.
K. R. Min  M. N. Zimmer 《Biofouling》2013,29(8):931-940
The aim of this study was to explore the physicochemical parameters that influence coaggregation between the freshwater bacteria Sphingomonas natatoria 2.1 and Micrococcus luteus 2.13. Using visual coaggregation assays, the effect of different buffers, solutions of differing ionic strength, pH, temperature, and viscosity on the degree of coaggregation was assessed. Coaggregation occurred maximally in distilled water but was inhibited when coaggregates were suspended in a commonly-used oral bacterial coaggregation buffer, saline solutions, and Tris-Cl buffers. Coaggregation was weakly expressed in standard laboratory buffers. The ionic strength of inorganic salt solutions required to inhibit coaggregation depended upon the inorganic salt being tested. Coaggregation occurred at a pH of 3–10, between 5 and 80°C and was inhibited in solutions with a viscosity of 22.5 centipoises at 20°C. Inhibition of coaggregation with NaCl impaired biofilm development. When developing buffers to test for coaggregation, the natural liquid environment should be considered. Coaggregation between S. natatoria 2.1 and M. luteus 2.13 is only affected by physicochemical conditions beyond those typically found in natural freshwater ecosystems. Such a robust ability to coaggregate may enhance the ability of S. natatoria 2.1 and M. luteus 2.13 to develop a niche in freshwater biofilms.  相似文献   

8.
Electrophoresis through agarose and polyacrylamide-type gels is the standard method to separate, identify, and purify nucleic acids. Properties of electrophoresis buffers such as pH, ionic strength, and composition affect performance. The buffers in use contain a weak acid or weak base buffered by a compound with a dissimilar pK. Herein, three pK-matched buffers were developed, each containing two effective buffering components: one weak base and one weak acid which have similar pKa at 25 degrees C (within 0.3 pK units): (i) Ethanolamine/Capso, pH 9.6; (ii) triethanolamine/Tricine, pH 7.9; and (iii) Bis-Tris/Aces, pH 6.7. On agarose gels, the buffers in various concentrations were tested for separation of double-stranded DNA fragments with various DNA markers, agarose gel concentrations, and field strengths. Mobility was inversely proportional to the logarithm of molecular weight. The buffers provided high resolution without smearing at more dilute concentration than is possible with standard TAE (Tris/acetate, pH 8.0) or TBE (Tris/borate, pH 8.3) buffers. The buffers were also tested in 7 M urea denaturing LongRanger sequencing gels and in nondenaturing polyacrylamide SSCP gels. The pK-matched buffers provide good separation and high resolution, at a broad range of potential pH values. In comparison to TAE and TBE, pK-matched buffers provide higher voltage and current stability, lower working concentration, more concentrated stock solutions (up to 200x), and lower current per unit voltage, resulting in less heat generation.  相似文献   

9.
We investigated the effects of pH and ionic strength of solutions used for antigen retrieval to elucidate the mechanism of heat-induced antigen retrieval (HIAR) in immunohistochemistry. The immunostaining intensity of nuclear, cytoplasmic, cell membrane, and extracellular matrix antigens with 17 different antibodies was evaluated in formaldehyde-fixed and paraffin-embedded mouse and human tissues. Deparaffinized sections were autoclaved for 10 min in buffers with different pH values ranging from 3.0 to 10.5. To test the influence of ionic strength on immunoreactions, the sections were autoclaved for 10 min in 20 mM Tris-HCl buffers (TB) at pH 9.0 and 10.5 with or without 25, 50, and 100 mM NaCl. There were two immunostaining patterns for pH dependency of HIAR. First, the majority of antibodies recovered their antigenicity when heated in the buffers with both acidic pH (pH 3.0) and basic pH (pH 9.0 and 10.5). Second, some antibodies showed strong immunostaining only at basic pH values (pH 9.0 and 10.5). When the sections were autoclaved in TB at pH 9.0, immunostaining of all eight antibodies examined decreased as the NaCl concentration increased. On the other hand, when the sections were treated with TB at pH 10.5, all antibodies yielded stronger reactions in the buffer containing NaCl than in the buffer without NaCl; five antibodies exhibited the strongest immunoreaction at concentrations from 25 to 50 mM. These results suggest that the extended polypeptides by heating are charged negatively or positively at basic or acidic pH, and that an electrostatic repulsion force acts to prevent random entangling of polypeptides caused by hydrophobic attractive force and to expose antigenic determinants, during cooling process of HIAR solution.  相似文献   

10.
The free solution mobility of a 20-bp double-stranded DNA oligomer has been measured in diethylmalonate (DM) and Tris-acetate buffers, with and without added NaCl or TrisCl. DM buffers have the advantage that the buffering ion is anionic, so the cation composition in the solution can be varied at will. The results indicate that the free solution mobility of DNA decreases linearly with the logarithm of ionic strength when the ionic strength is increased by increasing the buffer concentration. The mobility also decreases linearly with the logarithm of ionic strength when NaCl is added to NaDM buffer or TrisCl is added to TrisDM buffer. Nonlinear effects are observed if the counterion in the added salt differs from the counterion in the buffer. The dependence of the mobility on ionic strength cannot be predicted using the Henry, Debye-Hückel-Onsager, or Pitts equations for electrophoresis. However, the mobilities observed in all buffer and buffer/salt solutions can be predicted within approximately 20% by the Manning equation for electrophoresis, using no adjustable parameters. The results suggest that the electrostatic shielding of DNA is determined not only by the relative concentrations of the various ions in the solution, but also by their equivalent conductivities.  相似文献   

11.
Many organic anions bind free Ca2+, the total concentration of which must be adjusted in experimental solutions. Because published values for the apparent dissociation constant (Kapp) describing the Ca2+ affinity of short chain fatty acids (SCFAs) and gluconate are highly variable, Ca2+ electrodes coupled to either a 3 M KCl or a Na+ selective electrode were used to redetermine Kapp. All solutions contained 130 mM Na+, whereas the concentration of the studied anion was varied from 15 to 120 mM, replacing Cl that was decreased concomitantly to maintain osmolarity. This induces changes in the liquid junction potential (LJP) at the 3 M KCl reference electrode, leading to a systematic underestimation of Kapp if left uncorrected. Because the Na+ concentration in all solutions was constant, a Na+ electrode was used to directly measure the changes in the LJP at the 3 M KCl reference, which were under 5 mV but twice those predicted by the Henderson equation. Determination of Kapp either after correction for these LJP changes or via direct reference to a Na+ electrode showed that SCFAs do not bind Ca2+ and that the Kapp for the binding of Ca2+ to gluconate at pH 7.4, ionic strength 0.15 M, and 23 °C was 52.7 mM.  相似文献   

12.
Ox muscle troponin was shown by equilibrium- and velocity-sedimentation studies to undergo concentration-dependent dissociation into its constituent subunits as well as self-association in imidazole buffers, pH 6.9. The extent of troponin association was found to be strongly dependent on ionic strength and also to exhibit a dependence on pH and temperature; under conditions physiological in regard to pH, temperature and ionic strength the extent of polymerization of troponin is considerable in 2 mg/ml solutions. The ability of polymeric troponin to bind to tropomyosin has been inferred from studies of mixtures containing actin-tropomyosin and an excess of troponin over the amount required for the normal 7:1:1 actin-tropomyosin-troponin complex. These findings should be relevant to studies of reconstituted actin-tropomyosin-troponin preparations, since they signify possible chemical as well as physical differences between the gel, paracrystalline and filamentous states of the complex that result from adoption of different preparative procedures for analogues of the native thin filament.  相似文献   

13.
Information is given concerning two standard buffer solutions suitable as pH references in 30, 40, and 50 mass% dimethyl sulfoxide (DMSO)/H2O mixed solvents at subzero temperatures from -20 to 0 degrees C, with the intention of establishing a pH (designated pH*) scale. The two buffers selected were the ampholytes N,N-bis(2-hydroxyethyl)-2-aminoethane sulfonic acid ("bes") and N-tris(hydroxymethyl)methylglycine ("tricine"), and the reference standard consisted of equal molal quantities of the buffer and its respective sodium salt. The assignment of pH* values was based on measurements of the emf of cells without liquid junction of the type: Pt;H2(g,l atm) /Bes, Na Besate, NaCl / AgCl;Ag and Pt;H2(g,l atm) /Tricine, Na Tricinate, NaCl /AgCl;Ag and the pH* was derived from a determination of K2, the equilibrium constant for the dissociation process (Buffer)+/- in equilibrium with(Buffer)- + H+.  相似文献   

14.
A computer program that allows the preparation of buffers containing known concentrations of metal-ligand complexes at defined pH values and temperatures is described. Ligands are defined as compounds that bind metals and may include AMP, ADP, ATP, GMP, GDP, GTP, EGTA, EDTA, BAPTA, phosphate, sulfate, chloride, monocarboxylic acids, dicarboxylic acids, organophosphates, and/or citric acid. Metals may include sodium, potassium, magnesium, calcium, and/or manganese. The program uses association constants corrected for temperature and ionic strength so that solutions between 0 and 40 degrees C and between pH values of 4 and 10 can be defined. The program can perform the following: (i) calculate the concentration of all metal-ligand complexes when total metal and total ligand concentrations are known, (ii) calculate the concentration of metal ion required to make a solution of known free metal ion concentration when total ligand concentrations are known, (iii) calculate the concentration of ligand required to make a solution of known free metal ion concentration when total metal concentrations are known, and (iv) calculate the total concentrations of metal and ligand required to make a buffer of known metal-ligand concentration. Options i-iii are useful for making buffers of defined free metal ion concentrations; option iv is useful for making buffers of defined metal-nucleotide concentrations.  相似文献   

15.
A potentiometric procedure for cysteine thiol group concentration monitoring in media generating free radicals was developed using a thiol specific silver-mercury electrode. Electrolytic deposition of mercury on a silver wire and treatment with 20 mM cysteine in 0.5 M NaOH were used to produce the electrode. A silver-chloride electrode in saturated KCl was the reference. A glass capillary with 1 M KNO3 in 1% agarose gel was the liquid junction. The electrode responded to cysteine concentration in the range from 0.01 to 20 mM yielding a perfect linear relationship for the dependence of log [cysteine] versus electrode potential [mV], with b0 (constant) = -373.43 [mV], b1 (slope) = -53.82 and correlation coefficient r2 = 0.97. The electrode potential change per decade of cysteine concentration was 57 mV. The minimal measurable signal response was at a cysteine concentration of 0.01 mM. The signal CV amounted to 4-6% for cysteine concentrations of 0.01 to 0.05 mM and to less than 1% for cysteine concentrations of 0.5 to 20 mM. The response time ranged from about 100 s for cysteine concentrations of 0.01 to 0.1 mM to 30 s at higher cysteine concentrations. The standard curve reproducibility was the best at cysteine concentrations from 0.1 to 20 mM. In a reaction medium containing cysteine and copper(II)-histidine complex ([His-Cu]2+) solution in 55 mM phosphate buffer pH 7.4 the electrode adequately responded to changes in cysteine concentration. Beside cysteine, the silver-mercury electrode responded also to thiol groups of homocysteine and glutathione, however, the Nernst equation slope was about half of that for cysteine.  相似文献   

16.
Previous studies of the coat protein of tobacco mosaic virus (TMVP) have shown that TMVP presumably exists as linear stacks of two-ring cylindrical disks in the 0.7 M ionic strength buffer used for crystallizing the disks for X-ray diffraction studies [Raghavendra, K., Adams, M.L., & Schuster, T.M. (1985) Biochemistry 24, 3298-3304]. The spectroscopic and sedimentation studies of solutions of TMVP under these crystallizing conditions have demonstrated a long-term metastability of these disk aggregates when they are placed in 0.1 M ionic strength buffers, as are used for reconstituting tobacco mosaic virus from TMVP and viral RNA. The present work describes an electron microscopic study of TMVP disk aggregates under the same solution conditions employed in the previous spectroscopic and sedimentation studies. The results show that in the pH 8.0 0.7 M ionic strength crystallization buffer TMVP exists as stacks of disks which range in size from about 6 to 24 layers, corresponding to 3-12 2-layer disk aggregates having 17 subunits per layer. These TMVP aggregates persist in a metastable form in 0.1 M ionic strength virus reconstitution buffer with no apparent changes in structure of the stacked disks. The results are consistent with the conclusions of the solution physical-chemical studies which suggest that the disk structure may not be related to the 20S TMVP aggregate that is the nucleation species in virus  相似文献   

17.
Diffusion and partitioning of proteins in charged agarose gels.   总被引:4,自引:2,他引:2       下载免费PDF全文
The effects of electrostatic interactions on the diffusion and equilibrium partitioning of fluorescein-labeled proteins in charged gels were examined using fluorescence recovery after photobleaching and gel chromatography, respectively. Measurements were made with BSA, ovalbumin, and lactalbumin in SP-Sepharose (6% sulfated agarose), in phosphate buffers at pH 7 and ionic strengths ranging from 0.01 to 1.0 M. Diffusivities in individual gel beads (D) and in the adjacent bulk solution (D infinity) were determined from the spatial Fourier transform of the digitized two-dimensional fluorescence recovery images. Equilibrium partition coefficients (phi) were measured by recirculating protein solutions through a gel chromatography column until equilibrium was reached, and using a mass balance. Diffusion in the gel beads was hindered noticeably, with D/D infinity = 0.4-0.5 in each case. There were no effects of ionic strength on BSA diffusivities, but with the smaller proteins (ovalbumin and lactalbumin) D infinity increased slightly and D decreased at the lowest ionic strength. In contrast to the modest changes in diffusivity, there were marked effects of ionic strength on the partition coefficients of these proteins. We conclude that for diffusion of globular proteins through gel membranes of like charge, electrostatic effects on the effective diffusivity (Deff = phi D) are likely to result primarily from variations in phi with only small contributions from the intramembrane diffusivity.  相似文献   

18.
The diffusion coefficient and viscosity of lysozyme solutions were measured at 25°C in various buffers with and without sodium chloride. Measurements were made over the entire concentration range in each case and were extended into the supersaturated region. The results show that diffusion coefficient behavior depends strongly on the buffer used and the ionic strength of the solution, which means the amount of sodium chloride used in buffer solution. Viscosity measurements indicate a small degree of time dependence, with the viscosity increasing with solution age.  相似文献   

19.
R A Alberty  R N Goldberg 《Biochemistry》1992,31(43):10610-10615
The criterion for chemical equilibrium at specified temperature, pressure, pH, concentration of free magnesium ion, and ionic strength is the transformed Gibbs energy, which can be calculated from the Gibbs energy. The apparent equilibrium constant (written in terms of the total concentrations of reactants like adenosine 5'-triphosphate, rather than in terms of species) yields the standard transformed Gibbs energy of reaction, and the effect of temperature on the apparent equilibrium constant at specified pressure, pH, concentration of free magnesium ion, and ionic strength yields the standard transformed enthalpy of reaction. From the apparent equilibrium constants and standard transformed enthalpies of reaction that have been measured in the adenosine 5'-triphosphate series and the dissociation constants of the weak acids and magnesium complexes involved, it is possible to calculate standard Gibbs energies of formation and standard enthalpies of formation of the species involved at zero ionic strength. This requires the convention that the standard Gibbs energy of formation and standard enthalpy of formation for adenosine in dilute aqueous solutions be set equal to zero. On the basis of this convention, standard transformed Gibbs energies of formation and standard transformed enthalpies of formation of adenosine 5'-trisphosphate, adenosine 5'-diphosphate, adenosine 5'-monophosphate, and adenosine at 298.15 K, 1 bar, pH = 7, a concentration of free magnesium ions of 10(-3) M, and an ionic strength of 0.25 M have been calculated.  相似文献   

20.
We have used a stopped flow rapid reaction pH apparatus to investigate the carbamate equilibrium in glycylglycine solutions and in three biological tissues, human plasma, sheep muscle, and sheep brain, as well as to investigate the kinetics of carbamate formation in glyclyglycine solution and in human plasma. The rapid reaction apparatus was equipped with a pH sensitive glass electrode in order to follow the time course of pH from 0.005 to 100 s after rapid mixing of a solution of amine or protein and CO2. Two phases of the pH curve were observed: a fast phase representing carbamate formation, and a slow phase due to the hydration of CO2 which was uncatalyzed since a carbonic anhydrase inhibitor was added to the biological solutions. From the time course of pH change during the fast phase K2, the R-NH2 ionization constant, and Kc, the carbamate equilibrium constant as well as the velocity constant for the formation of carbamate, ka could be calculated from data at different pH and pCO2. The carbamate formed in glycylglycine solutions over a wide range of pH and pCO2 was found consistent with the theory of carbamate formation and with published data. At ionic strength 0.16 and 37 degrees pK is 7.67. pKc 4.58. The heat of the carbamate reaction (deltaH) was calculated to be -3.2 kcal/mol between 20 degrees and 37 degrees. Kt of glycylglycine depends quantitatively on ionic strength as predicted by the Debye-Huckel theory. With ionic strength 0.16 ku was found to be 2,500 M1 S1 at 37 degrees. The activation energy of carbamate formation is 6.7 kcal/mol. Carbamate measurements in human plasma at pCO2 from 38 to 359 Torr. pH from 6.9 to 8.3, temperature 37 degrees, and ionic strength 0.15 provided evidence that two kinds of amino groups participate in carbamate formation. From the equilibrium constants computed for the two species they could be identified as alpha- and epsilon-amino groups. On the basis of a protein molecular weight of 69.000. 0.6 alpha-amino groups/molecule with pKz=7.0 and pKc=4.2, and 5.9 epsilon-amino groups/molecule with pKz=9.0 and pKc=4.3 contribute to carbamate formation. The velocity constant ka was estimated to be 4,950 M1 S1 for the alpha-amino groups and 13,800 M1 S1 for the epsilon-amino groups. Under physiological conditions (pCO2=40 Torr. pH=7.4). The concentration of carbamate in plasma is 0.6 mM and the half-time of carbamate formation is 0.05 s. In extracts prepared from sheep brain at 37 degrees pH=7 and pCO2=35 Torr. the carbamate formation was estimated to be 0.8 mM. With pCO2=70 Torr and the same pH and temperature the carbamate concentration in muscle approximates 0.3 mM and increases to 7 mM as pH rises to 8. It is concluded that, as in plasma, a considerable number of epsilon-amino groups appear to be available for carbamate formation in these tissues.  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