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 814 毫秒
1.
The proton Bohr factor (phi H = alpha log PO2/alpha pH), the carbamate Bohr factor (phi C = alpha log PO2/alpha log PCO2), the total Bohr factor (phi HC = d log PO2/dpH[base excess) and the CO2 buffer factor (d log PCO2/dpH) were determined in the blood of 12 healthy donors over the whole O2 saturation (SO2) range. All three Bohr factors proved to be dependent on SO2, although to a lesser extent than reported in some of the recent literature. At SO2 = 50% and 37 degrees C, we found phi H = -0.428 +/- 0.010 (SE), phi C = 0.054 +/- 0.006, and phi HC = -0.488 +/- 0.007. The values obtained for phi H, phi C, and d log PCO2/dpH were used to calculate phi HC. Calculated and measured values of phi HC proved to be in good agreement. In an additional series of 12 specimens of human blood we determined the influence of PCO2 on phi H and the influence of pH on phi C. At SO2 = 50%, phi H varied from -0.49 +/- 0.009 at PCO2 = 15 Torr to -0.31 +/- 0.010 at PCO2 = 105 Torr and phi C from 0.157 +/- 0.015 at pH = 7.80 to 0.006 +/- 0.009 at pH = 7.00. When on the basis of these data a second-order term is taken into account, a still slightly better agreement between measured and calculated values of phi HC can be attained.  相似文献   

2.
Equations for proton equilibria of a single-phase binary buffer system have been applied to temperature-induced changes in pH and PCO2 of separated dog plasma at constant carbon dioxide content. Predicted behaviour, measured as deltapH/deltaT and deltalog PCO2 /deltaT, and pH and PCO2 as a function of temperature (range 8-45 degrees C), are in reasonable agreement with theory. Theory predicts and data confirm that deltapH/delta T and deltalog PCO2/deltaT functions of temperature; no single "temperature correction factor" is applicable. Comparison of whole blood with binary buffer equations also shows acceptable agreement between theory and experiment. Blood and separated plasma show similar responses in deltapH/deltaT and deltalog PCO2/delta T when compared over identical temperature intervals. For blood or plasma with initial pH (AT 37.5 DEGREES C) values in the range 7.53-7.45 deltapH/delta T (u/ degrees C) values are -0.0139 (37.5-27.5 degrees C) and -0.0192 (19-7 degrees C); comparable deltalog PCO2/deltaT values are 0.0195 (37.5-27.5 degrees C) and 0.0240 (19-7 degrees C). The charge state of protein components in this system remains nearly constant as temperature varies.  相似文献   

3.
The effects of temperature change (in vitro) on acid-base balance of skipjack tuna blood were investigated. By examining the relationship between blood pH and temperature (in vitro) under conditions of constant CO2 tension (open system), it was observed that dpH/dT = -0.013 U/degrees C. This value falls well within the range of in vivo values reported for other ectothermic vertebrates, and is only slightly different than results obtained in vitro under conditions of constant CO2 content (closed system; dpH/dT = -0.0165 U/degrees C). It is concluded that changes in pH following temperature changes can be accounted for solely by the passive, in vitro behaviour of the chemical buffer system found in the blood, so that active regulatory mechanisms of pH adjustment need not be postulated for skipjack tuna.  相似文献   

4.
Oxygenation studies with the whole blood of Phrynops hilarii show a P50 of 38 torr at extracellular pH (pHe) of 7.4 which corresponds to an intracellular pH (pHi) of 7.05 at 25 degrees C. The blood CO2 Bohr effect was -0.56 when related to pHi. pHi is related to pHe by the following equation: pHi = 0.75.pHe + 1.54 (r = 0.99); pHi = 0.72. pHe + 1.72 (r = 0.96) at 10 and 25 degrees C respectively. Blood pHe, for 25 degrees C, was 7.519 +/- 0.254 (n = 6). Blood gas partial pressures were: pCO2 = 25.8 +/- 3.8 torr (n = 6); pO2 = 61.7 +/- 21.2 torr (n = 6). The major red cell phosphates, in mmole/l erythrocytes, n = 6, were: ATP (3.66 +/- 0.86); GTP (0.53 +/- 0.28); 2.3-DPG (0.32 +/- 0.12) and inorganic phosphates (2.00 +/- 0.35). The plasma inorganic ion composition, n = 6, was, in mEq/l: K+ (3.04 +/- 0.40); Na+ (148.4 +/- 12.6); Ca2+ (4.75 +/- 1.32); Cl- (106.6 +/- 5.0). Additional blood parameters of interest (n = 6) were: lactate (2.07 +/- 1.72 mM in plasma); erythrocytes/mm3 (416 X 10(3) +/- 4.6 X 10(3)); leucocytes/mm3 (44636 +/- 2618); haematocrit (%) (14.5 +/- 3.6); haemoglobin, g/dl (3.2 +/- 0.5); plasma protein g/dl (4.4 +/- 0.4); osmolarity (293 +/- 10 mOsm/l). The non-bicarbonate buffer value was -22.6 mmol/kg H2O/pH. For a constant CO2 content, delta pHe/delta t = 0.0141 +/- 0.002 (n = 18) and delta pHi/delta t = 0.0157 +/- 0.003 (n = 18).  相似文献   

5.
Theory predicts that the rate of rise of the oscillation in arterial CO2 partial pressure (PaCO2) is linearly dependent on CO2 flux from venous blood to alveolar gas. We have measured, in the anesthetized cat, CO2 output (VCO2) and oscillations in arterial pH. The pH signal was differentiated to give the maximum rate of fall of pH on the downstroke of the oscillation (dpH/dt decreases max). Since oscillations in pH are due to oscillations in arterial PCO2, dpH/dt decreases max was considered to be equivalent to the maximum rate of rise of the PCO2 oscillation. VCO2 was increased by ventilating the intestines with CO2 and by the intra-arterial infusion of 2,4-dinitrophenol. VCO2 was decreased by filling the intestines with isotonic tris(hydroxymethyl)methylamine buffer. The maximum range of VCO2 covered was 7.8-51 ml/min, and the mean range was from 13.6 +/- 1.3 to 29.7 +/- 1.6 (SE) ml/min. Although CO2 loading produced a small rise and CO2 unloading a small fall in mean PaCO2, the changes were not statistically significant, so that overall the response was close to isocapnia. Over the limited range of VCO2 studied there was a highly significant linear association between dpH/dt decreases max and VCO2 which supports the contention that the slope of the upstroke of the PaCO2 oscillation is determined by the CO2 flux from mixed venous blood to alveolar gas. As such this slope is a potential chemical signal linking ventilation to CO2 production.  相似文献   

6.
We assessed the effects of cold and submergence on blood oxygen transport in common map turtles (Graptemys geographica). Winter animals were acclimated for 6-7 wk to one of three conditions at 3 degrees C: air breathing (AB-3 degrees C), normoxic submergence (NS-3 degrees C), and hypoxic (PO2=49 Torr) submergence (HS-3 degrees C). NS-3 degrees C turtles exhibited a respiratory alkalosis (pH 8.07; PCO2=7.9 Torr; [lactate]=2.2 mM) relative to AB-3 degrees C animals (pH 7.89; PCO2=13.4 Torr; [lactate]=1.1 mM). HS-3 degrees C animals experienced a profound metabolic acidosis (pH 7.30; PCO2=7.9 Torr; [lactate]=81 mM). NS-3 degrees C turtles exhibited an increased blood O2 capacity; however, isoelectric focusing revealed no seasonal changes in the isohemoglobin (isoHb) profile. Blood O2 affinity was significantly increased by cold acclimation; half-saturation pressures (P50's) for air-breathing turtles at 3 degrees and 22 degrees C were 6.5 and 18.8 Torr, respectively. P50's for winter animals submerged in normoxic and hypoxic water were 5.2 and 6.5 Torr, respectively. CO2 Bohr slopes (Delta logP50/Delta pH) were -0.15, -0.16, and -0.07 for AB-3 degrees C, NS-3 degrees C, and HS-3 degrees C turtles, respectively; the corresponding value for AB-22 degrees C was -0.37. The O2 equilibrium curve (O2EC) shape was similar for AB-3 degrees C and NS-3 degrees C turtles; Hill plot n coefficients ranged from 1.8 to 2.0. The O2EC shape for HS-3 degrees C turtles was anomalous, exhibiting high O2 affinity below P50 and a right-shifted segment above half-saturation. We suggest that increases in Hb-O2 affinity and O2 capacity enhance extrapulmonary O2 uptake by turtles overwintering in normoxic water. The anomalous O2EC shape and reduced CO2 Bohr effect of HS-3 degrees C turtles may also promote some aerobic metabolism in hypoxic water.  相似文献   

7.
1. Arterial blood was sampled at 15 min-intervals in European hamsters Cricetus cricetus fitted with indwelling catheters, from deep hibernation to full arousal. Temperature-corrected pH and PCO2, respectively pH* and P*CO2, were directly measured at 37 degrees C. 2. Deep hibernation corresponded to a respiratory acidosis: pH* = 7.01 +/- 0.01 (mean +/- SE), P*CO2 = 160 +/- 4 Torr (n = 9 animals). 3. Three periods could be distinguished in the arousal: (i) a period of hyperventilation (28 +/- 5 min), in which P*CO2 was reduced to 79 +/- 4 Torr, while cheek pouch temperature increased only by 0.9 +/- 0.2 degrees C; (ii) a period of metabolic acidification by lactate accumulation (84 +/- 6 min), corresponding to the period of peak thermogenesis; (iii) a progressive return to euthermic conditions (104 +/- 10 min), by simultaneous respiratory and metabolic alkalinization. 4. Over 60% of the blood CO2 stores accumulated at the beginning of the hibernation bout were released by hyperventilation during the first period, prior to the full development of thermogenesis. This is in agreement with the hypothesis of an inhibitory role of the respiratory acidosis in hibernation.  相似文献   

8.
After 30-min or longer incubation of rabbit blood under 42 degrees C the blood O2 affinity was increased. During governed hyperthermia of conscious rabbits PO2, pH and the oxygen content in the mixed venous blood were decreased. Similar events were observed after 1 hour from the ceasing of hyperthermia. In the real blood circulation the oxygen affinity of haemoglobin was decreased because of the higher temperature. But after the recalculation of P50 to the standard conditions (t = 37 degrees C, pH 7.4, PCO2 = 40 Torr) it's value was below the initial one by 4.0 +/- 0.68 Torr. The mechanism of the increase of the oxygen affinity of haemoglobin is discussed.  相似文献   

9.
We studied salt and water absorption in isolated rabbit superficial proximal straight tubules perfused and bathed with solutions providing oppositely directed transepithelial anion gradients similar to those which might obtain in vivo. The perfusing solution contained 138.6 mM Cl- 3.8 mM HCO-3 (pH 6.6) while the bathing solution contained 113.6 mM Cl- and 25 mM HCO-3 (pH 7.4); the system was bubbled with 95% O2-5% CO2. At 37 degrees C, net volume absorption (Jv nl min-1 mm-1) was 0.32 +/- 0.03 (SEM); Ve, the transepithelial voltage (millivolts; lumen to bath), was +3.1 +/- 0.2. At 21 degrees C, Ve rose to +3.7 +/- 0.1 and Jv fell to 0.13 +/- 0.01 (significantly different from zero at P less than 0.001); in the presence of 10(-4)M ouabain at 37 degrees C, Ve rose to +3.8 +/- 0.1 and Jv fell to 0.16 +/- 0.01 (P less than 0.001 with respect to zero). In paired experiments, the ouabain- and temperature-insensitive moieties of Jv and Ve became zero when transepithelial anion concentration gradients were abolished. Titrametric determinations net chloride flux at 21 degrees C or at 37 degrees C with 10(-4) M ouabain showed that chloride was the sole anion in an isotonic absorbate. And, combined electrical and tracer flux data indicated that the tubular epithelium was approximately 18 times more permeable to Cl- than to HCO-3. We interpret these results to indicate that, in these tubules, NaCl absorption depends in part on transepithelial anion concentration gradients similar to those generated in vivo and in vitro by active Na+ absorption associated with absorption to anions other than chloride. A quantitative analysis of passive solute and solvent flows in lateral intercellular spaces indicated that fluid absorption occurred across junctional complexes when the osmolality of the lateral intercellular spaces was equal to or slightly less than that of the perfusing and bathing solutions; the driving force for volume flow under these conditions depended on the fact that sigmaHCO3 exceeded sigmaCl.  相似文献   

10.
We have examined the interaction of the Escherichia coli trp aporepressor with its ligand, L-tryptophan, using both equilibrium dialysis and flow dialysis methods. Results obtained by the two procedures were equivalent and indicate that the trp aporepressor binds L-tryptophan with an equilibrium dissociation constant (Kd) of 40 microM at 25 degrees C under standard binding assay conditions (10 mM potassium phosphate, pH 7.4, 0.2 M potassium chloride, 0.1 mM EDTA, 5% glycerol). Molecular sizing of the purified trp aporepressor shows that in the absence of ligand the regulatory protein exists as a dimeric species with greater than 99% purity and an apparent molecular weight of 30,000. Under the storage and assay conditions used, the dimer appears quite stable, and essentially no monomer or higher multimeric species are detected. Analysis of binding data by Scatchard and direct linear plot methods shows two identical and independent ligand-binding sites/native trp aporepressor dimer. When examined as a function of temperature, L-tryptophan binding by trp aporepressor varied over 7-fold (Kd = 28 microM at 6.5 degrees C to Kd = 217 microM at 40 degrees C). At the optimal growth temperature for E. coli (37 degrees C), the dissociation constant was 160 microM for the ligand, L-tryptophan. From the relationship between temperature and L-tryptophan binding by trp aporepressor, the apparent enthalpy change delta H = -10.6 +/- 0.6 kcal mol-1 and the apparent entropy change delta S = -17 +/- 2 cal degree-1 mol-1 were determined.  相似文献   

11.
1. The hydrated electron reacts with ferrocytochrome c to form an unstable intermediate. This intermediate decays in a first-order manner to give, in the first instance, a product which has a similar absorption spectrum in the range 400-610 nm as normal ferricytochrome c. 2. At 21 degrees C the rate constant for the reaction of hydrated electrons with ferrocytochrome c at pH 7.4 (2 mM phosphate buffer) is (3.0 +/- 0.3) = 10(10) M-1 - S-1. As the pH is increased above pH 8.0 the rate constant steadily decreases. The dependence of the rate constant on pH can be explained if ferrocytochrome c has a pK of around 9.2. 3. At 21 degrees C and pH 7.4, the rate constant for the decay of the intermediate is (1.40 +/- 0.15) - 10(5) S-1. This reaction shows no pH dependence in the range 6-2-11.0. 4. A mechanism is proposed whereby the central metal atom of the ferrocytochrome c is oxidased and a thioether bond is reduced. The resulting ferricytochrome c species then slowly develops an absorbance at 606 nm due to the attack of the sulfhydryl group on the haem.  相似文献   

12.
The thermodynamics of zinc hematoporphyrin (ZnHP) dimerization and ZnHP-membrane binding were studied. The dimerization equilibrium was determined over the temperature range 19-40 degrees C, using fluorometric techniques. The dimerization constant obtained at 37 degrees C (neutral pH in phosphate-buffered saline) is 4.6 (+/- 0.6) X 10(4) M-1. The dimerization was found to decrease with temperature over the range 19-36 degrees C, the data allowing the extraction of the following thermodynamic parameters for the temperature range 19-31 degrees C: delta G0 = -9.3 kcal/mol, delta H0 = -7.4 kcal/mol, delta S0 = -6.4 eu. For temperatures above 36 degrees C the dimerization was found to be temperature independent, giving the following parameters: delta G0 = -6.6 kcal/mol, delta H0 = 0 kcal/mol, delta S0 = 21.2 eu. On the basis of the data the case is made for the existence of two types of ZnHP dimers, differing in the location of the fifth Zn2+ ligand and in the nature of the contribution of the solvent to the dimerization. For the membrane binding, large unilamellar liposomes served to model biological membranes. The binding of ZnHP to the liposomes was found to be similar, quantitatively, to the corresponding metal-free molecule, namely, fitting a case of one type of site and giving a binding constant of 1600 +/- 160 M (neutral pH and 37 degrees C) which is independent of the length of the porphyrin-liposome.  相似文献   

13.
The mangrove cockle Anadara grandis (Broderip and Sowerby, 1829) is a potential candidate for aquaculture and for bioremediation of aquaculture effluents in the tropical and subtropical coastal areas of the eastern Pacific Ocean. Laboratory-produced spat are available, but there is no information on their responses to the range of environmental conditions to which they might be subject during the growth cycle. The aim of this study was to evaluate the filtration and clearance rates ofA. grandis spat (shell length 9.50+/-0.37 mm) with a food concentration (7.5 mgxl(-1)) at four different temperatures (22, 25, 28 and 31 degrees C, with pH=7.5+/-0.2 and O2 concentration of 6.4+/-0.5 mgxl(-1); experiment one); and with a temperature (25 degrees C) and five concentrations of suspended matter (from 7.5 to 29 mgxl(-1) and pH and O2 values of 7.9+/-0.2 and 6.8+/-0.4 mgxl(-1); experiment two). Filtration and clearance rates were highest at 25 degrees C and significantly different (p<.05) from those obtained at 22, 28 and 31 degrees C; the clearance rates had the same tendency but the differences were not significant (p>.05). In the second experiment filtration increased according to the amount of food available, but there were no significant differences (p>.05) between 7.5 and 11 mgxl(-1) and from 22.4 to 29 mgxl(-1). The trend was similar for clearance, and in this case significant differences were found (p<.05) between 7.5, 22.4 and 29 mgxl(-1). Filtration at 31 degrees C was close to 80% at the optimum temperature of 25 degrees C, which indicates that A. grandis is a good candidate for tropical aquaculture. Clearance increased with high concentrations of suspended solids, but the production of biodeposits could be a source of environmental concern. Therefore, the possibility of using this species for bioremediation of aquaculture effluents should be studied with larger specimens and at higher seston concentrations.  相似文献   

14.
We studied unidirectional [14C]HCO3- efflux from human resealed red cell ghosts with 1 mM acetazolamide under self-exchange conditions at pH = pH(i = o) 7.4-9.0 and 0-38 degrees C by means of the Millipore- Swinnex and continuous flow tube filtering techniques. 14CO2 loss from cells to efflux medium and further to the atmosphere was insignificant. [14C]HCO3- efflux was determined at pH 7.8, 38 degrees C under symmetric variation of the HCO3- concentrations (C(i = o)), and asymmetric conditions: C(i) varied, C(o) constant, or C(o) varied, C(i) constant. MM-fit, Jeff = Jmaxeff x C x (C + K1/2)-1, used to describe the concentration dependence of Jeff,o when only C(o) varied, yields at C(i) = 50 mM: K1/2o = 3.8 mMJ, Jmaxeff.o = 20 nmol cm-2 s-1; at C(i) = 165 mM: K1/2o = 10 mM, Jmaxeff.o = 32 nmol cm-2 s-1. When C(i) varied, noncompetitive self inhibition by HCO3- binding (inhibitor constant K1) to an intracellular site was included (MS-fit). Under conditions of (a) symmetry: C(i = o) = 9-600 mM, K1/2s = 173 mM, K1 = 172 mM, and Jmaxeff,s = 120 nmol cm-2 s-1, (b) asymmetry: C(o) = 50 mM, K1/2i = 116 mM, K1 = 136 mM, and Jmaxeff,i = 92 nmol cm-2 s-1. All flux parameters accord with the ping-pong model for anion exchange. The data for C(i) < 200 mM also fit well to the MM equation, but K1/2 and Jmaxeff are different from the MS-fit and are inconsistent with the ping-pong model. Thus, self-inhibition (MS-fit) must be included even at low concentrations. As at 0 degree C, the system is asymmetric: 8-10 times more unloaded transport sites face inward than outward when C(i = o). Jeff,s was not mono-exponentially dependent on temperature at 0-38 degrees C, indicating that the transmembrane anion transport is controlled by several rate constants with different temperature dependencies. Jeff,s was not significantly affected by increasing pH(i = o) from 7.4 to 7.8, but it decreased by 50% when pH was raised to 9.0.  相似文献   

15.
W R Harris 《Biochemistry》1986,25(4):803-808
Equilibrium constants for the successive binding of 2 equiv of Ga3+ to human lactoferrin have been measured by difference ultraviolet spectroscopy in 0.1 M 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid containing 5 mM bicarbonate at pH 7.4 and 25 degrees C. Ethylenediamine-N,N'-diacetic acid was used as the competing chelating agent. Values of the effective binding constants for the stated experimental conditions are log K1 = 21.43 +/- 0.18 and log K2 = 20.57 +/- 0.16. Comparison of these results with literature values for the gallium-transferrin binding constants indicates that lactoferrin binds gallium more strongly by a factor of approximately 90. The ratios of successive binding constants for the two proteins are essentially identical. A linear free energy relationship (LFER) for the complexation of gallium(III) and iron(III) has been prepared and used to estimate an iron(III)-lactoferrin binding constant for pH 7.4. The LFER prediction is compared with thermodynamic data on iron binding at pH 6.4 and gallium binding at pH 7.4. The results indicate that the ratio of iron binding constants for lactoferrin and transferrin is likely in the range of 50-90.  相似文献   

16.
The strong ion approach provides a quantitative physicochemical method for describing the mechanism for an acid-base disturbance. The approach requires species-specific values for the total concentration of plasma nonvolatile buffers (A(tot)) and the effective dissociation constant for plasma nonvolatile buffers (K(a)), but these values have not been determined for human plasma. Accordingly, the purpose of this study was to calculate accurate A(tot) and K(a) values using data obtained from in vitro strong ion titration and CO(2) tonometry. The calculated values for A(tot) (24.1 mmol/l) and K(a) (1.05 x 10(-7)) were significantly (P < 0.05) different from the experimentally determined values for horse plasma and differed from the empirically assumed values for human plasma (A(tot) = 19.0 meq/l and K(a) = 3.0 x 10(-7)). The derivatives of pH with respect to the three independent variables [strong ion difference (SID), PCO(2), and A(tot)] of the strong ion approach were calculated as follows: dpH/dSID(+) = [1 + 10(pK(a)-pH)](2)/(2.303 x [SPCO(2)10(pH-pK'(1)[1 + 10(pK(a)-pH](2) + A(tot)10(pK(a)-PH]]; dpH/dPCO(2) = S10(-pK'(1)/[2.303[A(tot)10(pH)(10(pH + 10(pK(a))(-2) - SID(+)10(-pH)]], dpH/dA(tot) = -1/[2.303[SPCO(2)10(pH-pK'(1) + SID(+)10(pK(a)-pH)]], where S is solubility of CO(2) in plasma. The derivatives provide a useful method for calculating the effect of independent changes in SID(+), PCO(2), and A(tot) on plasma pH. The calculated values for A(tot) and K(a) should facilitate application of the strong ion approach to acid-base disturbances in humans.  相似文献   

17.
The effects of arterial alphastat regulation on brain intracellular pH (pHi) and several phosphate metabolites were assessed in anesthetized rats during hypothermia (28.6 +/- 0.2 degrees C) and normothermia (36.2 +/- 0.2 degrees C) by using 31P high-field (8.5 T) nuclear magnetic resonance (NMR). There were significant differences in pHi and metabolite ratios at the two temperatures under conditions of equal minute ventilation. During hypothermia, the brain pHi was 0.09 U higher, the phosphocreatine-to-inorganic phosphate (PCR/Pi) ratio 49% larger, and Pi-to-ATP 20% lower than at normothermia. These changes were fully reversible on warming the animal. The change in brain pHi/temperature was -0.011U/degrees C (95% confidence interval -0.007 to -0.016). The brain's ability to regulate its pHi and phosphate metabolism during hypercapnic acid-base stress was studied by using 10% CO2 ventilation. Hypothermic rats showed a larger fall in brain pHi (0.145 +/- 0.01 U, 7.15-7.01) with 10% CO2 than normothermic rats (0.10 +/- 0.02 U, 7.06-6.96). Similarly ventilated rats had a larger fall in arterial pH with 10% CO2 at hypothermia (0.36 +/- 0.04 U) than normothermia (0.24 +/- 0.01 U), so the delta brain pH/delta arterial pH was the same at both temperatures. The brain PCr-to-Pi ratio decreased approximately 20% during 10% CO2 breathing in both hypothermic and normothermic animals. Brain pHi and metabolite ratios returned to base line 30-50 min after CO2 washout in both groups. In summary, lowering body temperature while maintaining constant ventilation leads to changes in brain pHi and metabolites.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
The thermodynamics of the conversion of aqueous L-aspartic acid to fumaric acid and ammonia have been investigated using both heat conduction microcalorimetry and high-pressure liquid chromatography. The reaction was carried out in aqueous phosphate buffer over the pH range 7.25-7.43, the temperature range 13-43 degrees C, and at ionic strengths varying from 0.066 to 0.366 mol kg(-1). The following values have been found for the conversion of aqueous L-aspartateH- to fumarate2- and NH4+ at 25 degrees C and at zero ionic strength: K = (1.48 +/- 0.10) x 10(-3), DeltaG degrees = 16.15 +/- 0.16 kJ mol(-1), DeltaH degrees = 24.5 +/- 1.0 kJ mol(-1), and DeltaC(p) degrees = -147 +/- 100 J mol(-1) K(-1). Calculations have also been performed which give values of the apparent equilibrium constant for the conversion of L-aspartic acid to fumaric acid and ammonia as a function of temperature, pH and ionic strength.  相似文献   

19.
The thermodynamics of the conversion of aqueous D-psicose to D-allose has been investigated using high-pressure liquid chromatography. The reaction was carried out in phosphate buffer at pH 7.4 over the temperature range 317.25-349.25 K. The following results are obtained for the conversion process at 298.15 K: DeltaG degrees = - 1.41 +/- 0.09 kJ mol(-1), DeltaH degrees = 7.42 +/- 1.7 kJ mol(-1), and DeltaC(p) degrees = 67 +/- 50 J mol(-1) K(-1). An approximate equilibrium constant of 0.30 is obtained at 333.15 K for the conversion of aqueous D-psicose to D-altrose. Available thermodynamic data for isomerization reactions involving aldohexoses and aldopentoses are summarized.  相似文献   

20.
Oxygen utilization (VO2) and lactate production by an isolated perfused canine hindlimb was evaluated at various hydrogen ion concentrations. A membrane lung perfusion system was established such that blood flow and temperature could be fixed at normal levels. Oxygen, nitrogen, and carbon dioxide (CO2) gas flows to the membrane lung were independently regulated to provide a fixed arterial oxygen content (CaO2). By changing CO2 flow, the pH of the arterial blood was varied between 6.9 and 7.6 at 10-min intervals. The mean O2 delivery (CaO2 X blood flow) was between 16.3 ML O2/min and 20.5 ml O2/min. Standard error of the mean in each dog, however, was less than 0.4 ml O2/min. VO2 was linearly related to the pH of the perfusing blood: VO2% = 100.1 pH - 643 (r = 0.866). Oxygen consumption was inversely related to PCO2: VO2% = -0.62 PCO2 + 124, but the correlation was less good (r = 0.729). Lactate production was linearly related to the pH of the perfusing blood (above a pH of 7.4): lactate produced = 22.5 pH - 162.5 (r = 0.75). At a pH below 7.4, lactate was not produced. Oxygen consumption of skeletal muscle appears critically dependent on extracellular fluid pH. A change in pH of 0.1 alters VO2 almost exactly 10%. Alkalosis is a potent stimulus to lactic acid production by skeletal muscle.  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