首页 | 官方网站   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.
The inhibition by cupric ions of the hydration of CO2 catalyzed by carbonic anhydrase II is interesting because of the results of Tuet al. obtained at chemical equilibrium, indicating that Cu2+ inhibits specifically a proton transfer in the catalytic pathway. We have measured this inhibition at steady state, using stopped-flow methods. The inhibition by Cu2+ of the hydration of CO2 catalyzed by carbonic anhydrase II had aK I near 1×10?6 M atpH 7.0 and gave inhibition that is noncompetitive atpH 6.0 and mixed, but close to uncompetitive, atpH 6.8. ThepH dependence of this binding is consistent with a binding site for Cu2+ on the enzyme with apK a near 7. The binding interaction between Cu2+ and the fluorescent inhibitor 5-dimethylaminonaphthalene-l-sulfonamide on carbonic anhydrase II was noncompetitive, indicating that the binding site for Cu2+ is distinct from the coordination sphere of zinc in which the actual interconversion of CO2 and HCO 3 ? and the binding of sulfonamides takes place.  相似文献   

2.
The inhibition by cupric ions of the hydration of CO2 catalyzed by carbonic anhydrase II is interesting because of the results of Tuet al. obtained at chemical equilibrium, indicating that Cu2+ inhibits specifically a proton transfer in the catalytic pathway. We have measured this inhibition at steady state, using stopped-flow methods. The inhibition by Cu2+ of the hydration of CO2 catalyzed by carbonic anhydrase II had aK I near 1×10–6 M atpH 7.0 and gave inhibition that is noncompetitive atpH 6.0 and mixed, but close to uncompetitive, atpH 6.8. ThepH dependence of this binding is consistent with a binding site for Cu2+ on the enzyme with apK a near 7. The binding interaction between Cu2+ and the fluorescent inhibitor 5-dimethylaminonaphthalene-l-sulfonamide on carbonic anhydrase II was noncompetitive, indicating that the binding site for Cu2+ is distinct from the coordination sphere of zinc in which the actual interconversion of CO2 and HCO 3 and the binding of sulfonamides takes place.  相似文献   

3.
Carbonic anhydrases (CAs, EC 4.2.1.1) are ubiquitous metalloenzymes, grouped into seven different classes, which catalyze the reaction of CO2 hydration to bicarbonate and protons. All of the fifteen human isoforms reported to date belong to the α-class and contain zinc as a cofactor. The structure of human Zn,Cu-CA II has been solved which contains a copper ion bound at its N-terminal, coordinated to His4 and His64. In the active site a dioxygen molecule is coordinated to the zinc ion. Since dioxygen is a rather unexpected CA ligand, molecular dynamics (MD) simulations were performed which suggested a superoxide character of the zinc bound O2.  相似文献   

4.
A one-dimensional photochemical model was used to examine the effect of bolide impacts on the oxidation state of Earth's primitive atmosphere. The impact rate should have been high prior to 3.8 Ga before present, based on evidence derived from the Moon. Impacts of comets or carbonaceous asteroids should have enhanced the atmospheric CO/CO2 ratio by bringing in CO ice and/or organic carbon that can be oxidized to CO in the impact plume. Ordinary chondritic impactors would contain elemental iron that could have reacted with ambient CO2 to give CO. Nitric oxide (NO) should also have been produced by reaction between ambient CO2 and N2 in the hot impact plumes. High NO concentrations increase the atmospheric CO/CO2 ratio by increasing the rainout rate of oxidized gases. According to the model, atmospheric CO/CO2 ratios of unity or greater are possible during the first several hundred million years of Earth's history, provided that dissolved CO was not rapidly oxidized to bicarbonate in the ocean. Specifically, high atmospheric CO/CO2 ratios are possible if either: (1) the climate was cool (like today's climate), so that hydration of dissolved CO to formate was slow, or (2) the formate formed from CO was efficiently converted into volatile, reduced carbon compounds, such as methane. A high atmospheric CO/CO2 ratio may have helped to facilitate prebiotic synthesis by enhancing the production rates of hydrogen cyanide and formaldehyde. Formaldehyde may have been produced even more efficiently by photochemical reduction of bicarbonate and formate in Fe++-rich surface waters.  相似文献   

5.
Recent ab initio studies reported in the literature have challenged the mechanistic assignments made on the basis of volume of activation data [1,2]. In addition to that ab initio molecular orbital calculations on hydrated zinc(II)-ions were used to elucidate the general role of this ion in metalloproteins [3]. Due to our interest in both inorganic reaction mechanisms and enzymatic catalysis we started a systematic investigation of solvent exchange processes on divalent zinc-ion using density functional calculations. Our investigations cover aqua complexes of the general form [Zn(H2O)n]2+·mH20 with n=3-6 and m=0-2, where n and m represent the number of water molecules in the coordination and solvation sphere, respectively.The complexes [Zn(H2O)5]2+·2H2O and [Zn(H2O)4]2+·2H2O turnend out to be the most stable zinc complexes with seven and six water molecules, respectively. This implies that a heptacoordinated zinc(II) complex, where all water molecules are located in the co-ordination sphere, should be energetically highly unfavorable and that [Zn(H2O)6]2+ can quite readily push two coordinated water molecules into the solvation sphere. For the pentaqua complex [Zn(H2O)5]2+ only one water molecule is easily lost to the solvation sphere, which makes the [Zn(H2O)4]2+·H2O complex the most favorable in order to consider the limiting dissociative and associative water exchange process of hexacoordinated zinc(II). The dehydration and hydration energies using the most stable zinc(II) complexes [Zn(H2O)4]2+·2H2O, [Zn(H2O)5]2+·2H2O and [Zn(H2O)4]2+·H2O were calculated to be 24.1 and -21.0 kcal/mol, respectively.  相似文献   

6.
The singnificance of the zinc hydroxide–Thr-199–Glu-106 hydrogen-bond network in the active site of human carbonic anhydrase II has been examined by X-ray crystallographic analyses of site-specific mutants. Mutants with Ala-199 and Ala-106 or Gln-106 have low catalytic activities, while a mutant with Asp-106 has almost full CO2 hydration activity. The structures of these four mutants, as well as that of the bicarbonate complex of the mutant with Ala-199, have been determined at 1.7 to 2.2 Å resolution. Removal of the γ atoms of residue 199 leads to distorted tetrahedral geometry at the zine ion, and a catalytically important zinc-bound water molecule has moved towards Glu-106. In the bicarbonate complex of the mutant with Ala-199 one oxygen atom from bicarbonate binds to zinc without displacing this water molecule. Tetrahedral coordination geometries are retained in the mutants at position 106. The mutants with Ala-106 and Gln-106 have a zinc-bound sulfate ion, whereas this sulfate site is only partially occupied in the mutant with Asp-106. The hydrogen-bond network seems to be “reversed” in the mutants with Ala-106 and Gln-106. The network is preserved as in native enzyme in the mutant with Asp-106 but the side chain of Asp-106 is more extended than that of Glu-106 in the native enzyme. These results illustrate the importance of Glu-106 and Thr-199 for controlling the precise coordination geometry of the zinc ion and its ligand preferences with results in an optimal orientation of a zine-bound hydroxide ion for an attack on the CO2 substrate. © 1993 Wiley-Liss, Inc.  相似文献   

7.
A new catalytic mechanism is proposed for the hydration of CO2 by the zinc metalloenzyme carbonic anhydrase. This mechanism identifies the group controlling catalytic activity as an active site histidine, in which the protonated imidazole ring coordinates to zinc, losing a proton. Geometric constraints on the histidine unit make the metal-ligand bond a strained and, therefore, labile one. In the hydration reaction, the metal-bound neutral histidine moiety serves as a proton acceptor for the transient ionization of metal-bound water. Zinc-bound hydroxide attacks the carbon of the substrate to generate metal-bound bicarbonate, and the system regenerates itself by losing the elements of carbonic acid.  相似文献   

8.
The carboxylase activities of crude carboxysome preparations obtained from the wild-type Synechococcus elongatus strain PCC 7942 strain and the mutant defective in the carboxysomal carbonic anhydrase (CA) were compared. The carboxylation reaction required high concentrations of bicarbonate and was not even saturated at 50 mM bicarbonate. With the initial concentrations of 50 mM and 25 mM for bicarbonate and ribulose-1,5-bisphosphate (RuBP), respectively, the initial rate of RuBP carboxylation by the mutant carboxysome (0.22 μmol mg?1 protein min?1) was only 30 % of that observed for the wild-type carboxysomes (0.71 μmol mg?1 protein min?1), indicating the importance of the presence of CA in efficient catalysis by ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco). While the mutant defective in the ccmLMNO genes, which lacks the carboxysome structure, could grow under aeration with 2 % (v/v) CO2 in air, the mutant defective in ccaA as well as ccmLMNO required 5 % (v/v) CO2 for growth, indicating that the cytoplasmically localized CcaA helped utilization of CO2 by the cytoplasmically localized Rubisco by counteracting the action of the CO2 hydration mechanism. The results predict that overexpression of Rubisco would hardly enhance CO2 fixation by the cyanobacterium at CO2 levels lower than 5 %, unless Rubisco is properly organized into carboxysomes.  相似文献   

9.
《BBA》2022,1863(1):148503
The uptake of inorganic carbon in cyanobacteria is facilitated by an energetically intensive CO2-concentrating mechanism (CCM). This includes specialized Type-1 NDH complexes that function to couple photosynthetic redox energy to CO2 hydration forming the bicarbonate that accumulates to high cytoplasmic concentrations during the operation of the CCM, required for effective carbon fixation. Here we used a Synechococcus PCC7942 expression system to investigate the role of conserved histidine and cysteine residues in the CupB (also designated, ChpX) protein, which has been hypothesized to participate in a vectoral CO2 hydration reaction near the interface between CupB protein and the proton-pumping subunits of the NDH-1 complex. A homology model has been constructed and most of the targeted conserved residues are in the vicinity of a Zn ion modeled to form the catalytic site of deprotonation and CO2 hydration. Growth and CO2 uptake assays show that the most severe defects in activity among the targeted residues are due to a substitution of the predicted Zn ligand, CupB-His86. Mutations at other sites produced intermediate effects. Proteomic analysis revealed that some amino acid substitution mutations of CupB caused the induction of bicarbonate uptake proteins to a greater extent than complete deletion of CupB, despite growth under CO2-enriched conditions. The results are discussed in terms of hypotheses on the catalytic function of this unusual enzyme.  相似文献   

10.
The formation constants of equimolar and bis-chelate copper(II) and zinc(II) complexes with three aliphatic and four aromatic-substituted β-aminophosphonates have been determined in water solution by potentiometric studies. Spectroscopic parameters clearly indicate involvement of {NH3, PO3 2−} in both metal ions coordination. The comparison of the stability constants reveals slightly higher coordination power of the aliphatic-substituted β-aminophosphonic acids, which may be due to the higher basicity of their amino groups. All studied ligands are more effective in Cu2+ and Zn2+ coordination than phosphonic analogue of simple β-amino acid.  相似文献   

11.
Summary The effect of acidosis on the myocardial Ca2+ distribution was examined at 15°C in ventricular strips of the flounder (Platichthys flesus) and at 30°C in atrial strips of the rat (Rattus norvegicus).Lowering the Ringer pH from 7.6 to 6.9 by increasing its CO2 (flounder 2% to 12%, rat 4% to 14%), resulted in an elevated Ca2+ efflux in resting strips as well as in strips stimulated (12/min) to contraction. A decrease in pH of the Ringer used for the flounder myocardium by a lowering of bicarbonate (30 mM to 5 mM) also resulted in an elevation of the Ca2+ efflux, but the effect was smaller than that produced by an increased CO2.With 11 mM Ca2+ and 10 mM EGTA added to the Ringer to reduce the amount of45Ca2+ bound to extracellular sites, an increased CO2 with a concomitant drop in Ringer pH resulted in an increased Ca2+ efflux in both myocardia. The Ca2+ efflux was only marginally elevated in the flounder myocardium and unchanged in that of rat when the same drop in Ringer pH was produced with a lowering in bicarbonate.In a nominally Ca2+-free Ringer with 0.1 mM EGTA the45Ca2+ efflux was stimulated for both myocardia by an increase in CO2.The flounder myocardium was exposed to high CO2 in a nominally Na+, Ca2+-free Ringer and again the45Ca2+ efflux increased. After a return to Na, Ca and low CO2 in the Ringer, a higher efflux persisted in the strips being subjected to a high CO2 than in the controls.The Ca2+ uptake rate was about the same at high and low CO2 for both myocardia.Based on these results the measured increase in Ca efflux following an increase in CO2 or a decrease in bicarbonate probably results from an elevated cytoplasmatic Ca2+ activity. It seems unlikely that an increased uptake rate of Ca2+ or a direct stimulation of Ca2+ transporting mechanisms in the cell membrane are responsible for the change.  相似文献   

12.
Ribulose diphosphate carboxylase was found to exist in two distinct kinetic forms in spinach leaf extracts. One form displayed an apparent Km for CO2 in excess of 200 μm and is likely to be the form purified and studied by many previous workers. However, if leaf extracts were prepared in the presence of Mg2+ and atmospheric levels of CO2, the recently described high-affinity form was obtained. It had a Km for CO2 of about 20 μm, was quite stable even at 25 °C, and its properties were consistent with it being the form which operates in photosynthesis in vivo. Mg2+ was also able to convert the high-Km (CO2) form to the low-Km (CO2) form when it was added to an extract which had been prepared in its absence. Mg2+ was more effective in causing this conversion if bicarbonate was added as well. This activating effect of bicarbonate is a probable cause of previously reported apparent homotropic effects of bicarbonate on ribulose diphosphate carboxylase activity. It is possible that the apparently high-Km (CO2) form is not intrinsically active and appears to have activity only by virtue of the low-Km (CO2) form produced by contact with Mg2+ and bicarbonate (or CO2) during the course of the assay. Extracts prepared with ribose 5-phosphate in the absence of Mg2+ also showed low-Km (CO2) carboxylase activity initially, but the presence of this sugar phosphate was deleterious during storage at 25 °C, where it promoted conversion to the apparently high-Km (CO2) form.Effects on the affinity of ribulose diphosphate carboxylase for CO2 were paralleled by effects on the activity of the associated ribulose diphosphate oxygenase. Treatments which produced the low-Km (CO2) form of the carboxylase also resulted in high oxygenase activity, and it is possible that the apparently high-Km (CO2) form of the carboxylase has little, if any, oxygenase activity associated with it.The carboxylase and oxygenase activities of the low-Km (CO2) form showed broad and quite similar responses to pH variation, and the oxygenase had a Km for O2 of 0.22 mm.The stability of the low-Km (CO2) form in the presence of Mg2+ and bicarbonate was quite sufficient for it to be partially purified by Sepharose chromatography. The significance of the low-Km (CO2) form is discussed with respect to activation of photosynthesis by Mg2+.  相似文献   

13.
The approach of CO2 to a series of active site model complexes of human carbonic anhydrase II (HCAII) and its catalytic hydration to bicarbonate anion have been investigated using semiempirical MO theory (AM1). The results show that direct nucleophilic attack of zinc-bound hydroxide to the substrate carbon occurs in each model system. Further rearrangement of the bicarbonate complex thus formed via a rotation-like movement of the bicarbonate ligand can only be found in active site model systems that include at least one additional water molecule. Further refinement of the model complex by adding a methanol molecule to mimic Thr-199 makes this process almost activationless. The formation of the final bicarbonate complex by an internal (intramolecular) proton transfer is only possible in the simplest of all model systems, namely {[Im3Zn(OH)]+·CO2}. The energy of activation for this process, however, is 36.8 kcal·mol–1 and thus too high for enzymatic catalysis. Therefore, we conclude that within the limitations of the model systems presented and the level of theory employed, the overall mechanism for the formation of the bicarbonate complex comprises an initial direct nucleophilic attack of zinc-bound hydroxide to carbon dioxide followed by a rotation-like rearrangement of the bicarbonate ligand via a penta-coordinate Zn2+ transition state structure, including the participation of an extra active site water molecule.Electronic Supplementary Material available.  相似文献   

14.
15.
West J  Hill R 《Plant physiology》1967,42(6):819-826
Pea chloroplasts isolated in salt media show decreased rates of 2:6 dichlorophenolindophenol (DCPIP) and ferricyanide reduction when depleted of CO2 at pH values below 7.5. The greatest effect of CO2 was on uncoupled systems. The incorporation of 10−2, 2 × 10−2 and 4 × 10−2 m sodium acetate into the reaction mixtures progressively increased the bicarbonate concentration required for half maximal rates of reduction of DCPIP. The reaction was saturated by bicarbonate concentrations of 1 to 4 × 10−2 m. With both DCPIP and ferricyanide, the addition of bicarbonate to illuminated chloroplast systems depleted of CO2 gave very rapid increases in the rates of reduction. Bicarbonate also stimulated oxygen uptake by the illuminated chloroplasts when added hydrogen acceptors had been reduced. There was no effect of bicarbonate on ferricyanide reduction at low light intensities, but with DCPIP reduction, the apparent magnitude of the effect was independent of light intensity. This suggests that DCPIP reacts with the chloroplast electron transport chain at a site nearer to a photochemical stage than does ferricyanide. It also suggests that CO2 has at least 2 sites of action.  相似文献   

16.
We have measured the exchange of 18O between CO2 and H2O in stirred suspensions of Chlorella vulgaris (UTEX 263) using a membrane inlet to a mass spectrometer. The depletion of 18O from CO2 in the fluid outside the cells provides a method to study CO2 and HCO3 kinetics in suspensions of algae that contain carbonic anhydrase since 18O loss to H2O is catalyzed inside the cells but not in the external fluid. Low-CO2 cells of Chlorella vulgaris (grown with air) were added to a solution containing 18O enriched CO2 and HCO3 with 2 to 15 millimolar total inorganic carbon. The observed depletion of 18O from CO2 was biphasic and the resulting 18C content of CO2 was much less than the 18O content of HCO3 in the external solution. Analysis of the slopes showed that the Fick's law rate constant for entry of HCO3 into the cell was experimentally indistinguishable from zero (bicarbonate impermeable) with an upper limit of 3 × 10−4 s−1 due to our experimental errors. The Fick's law rate constant for entry of CO2 to the sites of intracellular carbonic anhydrase was large, 0.013 per second, but not as great as calculated for no membrane barrier to CO2 flux (6 per second). The experimental value may be explained by a nonhomogeneous distribution of carbonic anhydrase in the cell (such as membrane-bound enzyme) or by a membrane barrier to CO2 entry into the cell or both. The CO2 hydration activity inside the cells was 160 times the uncatalyzed CO2 hydration rate.  相似文献   

17.
Stemler A 《Plant physiology》1980,65(6):1160-1165
High concentrations of both bicarbonate and formate inhibit photosynthetic O2 evolution at pH 8.0. At this pH, only 2.4% of the total dissolved carbon dioxide exists as CO2. At pH 7.3, where 11% of the total dissolved carbon dioxide exists as CO2, HCO3 no longer inhibits. While formate still inhibits O2 evolution at pH 7.3, its effect can be partially overcome if CO2 is also present. The rate of binding of added 14C-labeled inorganic carbon is nearly 10-fold more rapid when the internal pH of thylakoid membranes is at 6.0 than when it is at 7.8. These observations suggest that CO2, not HCO3, is initially bound to the photosystem II reaction center and that the location of the binding site is on the inside thylakoid surface. However, additional data presented here suggest that, after binding, CO2 is hydrated to HCO3 + H+ in a pH-dependent reaction. Two possible explanations of the “bicarbonate effect” are presented.  相似文献   

18.
Summary Although it is generally believed thatChara and some fresh-water angiosperms transport bicarbonate ions inwards across their plasma membranes, there has been no direct demonstration of such transport in these plants. The (indirect) arguments for their transporting HCO 3 are arguments against the inward diffusion of CO2 at the observed rates. They rest on calculations of the equilibrium concentration of CO2 or of the maximum rate at which CO2 might be produced from HCO 3 at the pH of the medium outside the cells. SinceChara acidifies the medium over about half the cell surface during C assimilation, these calculations have been based on questionable premises.We propose a model forChara in which the acidification is attributed to active efflux of H+, and we calculate that both the equilibrium concentration of CO2 and its rate of production outside the cell can be high enough to support the observed rates of C assimilation, without postulating transport of the species HCO 3 or H2CO3.Calculations are presented also for alternative models in which there is membrane transport of HCO 3 . The first includes symport of H+ with HCO 3 , again dependent on active H+ efflux. In the second, there is active electrogenic transport of HCO 3 . In this case the low pH in the medium outside the cell is caused by the dissociation of H2CO3 produced by hydration of CO2 which leaks from the cell cytoplasm.All three models are consistent with the observations to date, but the first is more economical of postulates. It can also explain the apparent transport of HCO 3 by fresh-water angiosperms such asEgeria.  相似文献   

19.
Cellular metabolite analyses by 13C-NMR showed that C. reinhardtii cells assimilate acetate at a faster rate in heterotrophy than in mixotrophy. While heterotrophic cells produced bicarbonate and CO2 aq, mixotrophy cells produced bicarbonate alone as predominant metabolite. Experiments with singly 13C-labelled acetate (13CH3-COOH or CH3-13COOH) supported that both the 13C nuclei give rise to bicarbonate and CO2 aq. The observed metabolite(s) upon further incubation led to the production of starch and triacylglycerol (TAG) in mixotrophy, whereas in heterotrophy the TAG production was minimal with substantial accumulation of glycerol and starch. Prolonged incubation up to eight days, without the addition of fresh acetate, led to an increased TAG production at the expense of bicarbonate, akin to that of nitrogen-starvation. However, such TAG production was substantially high in mixotrophy as compared to that in heterotrophy. Addition of mitochondrial un-coupler blocked the formation of bicarbonate and CO2 aq in heterotrophic cells, even though acetate uptake ensued. Addition of PSII-inhibitor to mixotrophic cells resulted in partial conversion of bicarbonate into CO2 aq, which were found to be in equilibrium. In an independent experiment, we have monitored assimilation of bicarbonate via photoautotrophy and found that the cells indeed produce starch and TAG at a much faster rate as compared to that in mixotrophy and heterotrophy. Further, we noticed that the accumulation of starch is relatively more as compared to TAG. Based on these observations, we suggest that acetate assimilation in C. reinhardtii does not directly lead to TAG formation but via bicarbonate/CO2 aq pathways. Photoautotrophic mode is found to be the best growth condition for the production of starch and TAG and starch in C. reinhardtii.  相似文献   

20.
Summary The effects of carbonic anhydrase inhibitors on secretion by macropodine parotid and mandibular glands were investigated using anaesthetized red kangaroos. In the parotid gland, acetazolamide (500 mol·l-1) reduced a stable acetylcholine-evoked, half-maximal flow rate of 2.02±0.034 to 0.27±0.023 ml·min-1 (87% reduction). Concurrently, salivary bicarbonate concentration and secretion fell (129.4±1.46 to 80.9±1.63 mmol·l-1 and 264.8±7.96 to 22.3±2.30 mol·min-1, respectively), phosphate and chloride concentrations rose (14.0±0.79 to 27.6±0.85 mmol·l-1 and 5.6±0.25 to 27.5±1.32 mmol·l-1, respectively), sodium concentration and osmolality were unaltered, and potassium concentration fell (8.8±0.33 to 6.4±0.29 mmol·l-1). High-rate cholinergic stimulation during acetazolamide blockade was unable to increase salivary flow beyond 11±0.9% of that for equivalent unblocked control stimulation. However, superimposition of isoprenaline infusion on the acetylcholine stimulation caused a three-fold increase in the blocked flow rate. These treatments were accompanied by small increases in salivary phosphate and chloride concentrations but not bicarbonate concentration. Methazolamide infusion caused similar changes in parotid secretion. In the mandibular gland, acetazolamide infusion had no effect on salivary flow rate during either low- or high-level acetylcholine stimulation. Acetazolamide caused no alterrations in salivary electrolyte secretion at low flow rates, but curtailed the rise in bicarbonate concentration associated with high-level acetylcholine stimulation. Acetazolamide administration did not affect the increase in salivary flow rate associated with isoprenaline infusion, but did block the concomitant increase in bicarbonate concentration and secretion substantially. It was concluded that neither cholinergic nor adrenergic stimulation of mandibular fluid secretion depends on secretion of bicarbonate derived from catalysed hydration of CO2, but a substantial proportion of the increase in bicarbonate secretion during isoprenaline administration, which is probably ductal in origin, is so dependent. In contrast to other salivary glands, including the ovine parotid, fluid secretion by the kangaroo parotid gland during cholinergic stimulation is largely dependent (about 90%) on secretion of bicarbonate derived from hydration of CO2 catalysed by glandular carbonic anhydrase. Fluid secretion during adrenergic stimulation is not bicarbonate dependent.Abbreviations b.w. body weight - PAH p-aminohippurate - PCO2 partial pressure carbon dioxide - PCO2 partial pressure of oxygen  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-23

京公网安备 11010802026262号