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.
Two new genetic protein polymorphisms (CON 1 and CON 2) were identified in parotid saliva. Genetic polymorphisms of salivary CON 1 (concanavalin A) and CON 2 proteins are determined by autosomal inheritance of one expressed (dominant) and one unexpressed (recessive) allele for each gene. Autosomal inheritance is supported by studies in 26 families including 105 children for CON 1 and 23 families including 95 children for CON 2. Gene frequencies determined for randomly collected salivas from 134 whites, 79 Chinese, and 74 blacks are as follows: for whites, CON 1 += 0.396 and CON 1 = 0.604, CON 2 + = 0.034 and CON 2 = 0.966; for Chinese, CON 1 + = 0.580 and CON 1 = 0.420, CON 2 + = 0 and CON 2 = 1; for blacks, CON 1 + = 0.581 and CON 1 = 0.419, CON 2 + = 0.007 and CON 2 = 0.993. Both CON 1 and CON 2 proteins, transferred from SDS gels to nitrocellulose, react with concanavalin A. The CON 1 and CON 2 proteins react with antisera prepared to proline-rich proteins (PRP), and the CON 1 and CON 2 proteins have isoelectric points greater than pH 8.5. In randomly collected salivas, the CON 1 protein shows a strong association with Ps proteins, and the CON 2 protein shows a strong association with the PmF protein. On the basis of association data, PmS and CON 2 genes may be outside markers in a linear arrangement of the three genes, PmS, PmF, and CON 2. There is strong evidence for linkage of CON 1 and CON 2 to the SPC (salivary protein gene complex), CON 1 to Ps (15 families, lod score at = 0 is 6.77), CON 2 to PmF (7 families, lod score at = 0 is 5.93), and CON 2 to Gl (5 families, lod score at =0 is 3.91). In addition to immunologic reactions with the CON 1 and CON 2 proteins, antisera to PRP show extensive immunologic reactions with many other salivary proteins when tested by immunoblotting on nitrocellulose. Some of these proteins were previously identified PRP (proline-rich proteins) that are determined by different PRP loci.This study was supported by Grant (DEO 3658-18) from the National Institutes of Dental Research. Paper No. 2647 of the Laboratory of Genetics, University of Wisconsin, Madison, Wisconsin 53706.  相似文献   

2.
Synthesis of many T7 proteins is prevented in F′ episome-containing cells. In order to quantitate the degree of inhibition, we measured the activity of several T7 proteins in extracts prepared from T7-infected F and F′ cells and cells containing F factors mutant in phage inhibition [F′(PIF2A) and F′(PIF2A,2B)]. In addition, we were able to assign specific T7 proteins to the three translational units previously defined by polyacrylamide gel analysis of T7 proteins made in F and episome-containing cells. After T7 infection, the presence of the wild-type F′ (PIF+) episome led to greater than 90% inhibition of T7 DNA polymerase (product of gene 5), T7 lysozyme (gene 3.5), and gene 10 capsid protein synthesis. Nearly normal amounts of T7 RNA polymerase (gene 1) were made in these cells. T7 infection of cells containing the mutant F′ (PIF2A) episome led to normal synthesis of T7 RNA polymerase and T7 DNA polymerase; T7 lysozyme was synthesized at 30% of the maximal level in these cells; T7 gene 10 capsid protein synthesis was inhibited by 90%, and T7 DNA synthesis was arrested in these cells. T7 infection of cells containing the mutant F′ (PIF2A,2B) episome led to synthesis of normal levels of the enzymes assayed.  相似文献   

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

4.
The kinetics of the decomposition reactions of the CO(py)3(CO3)(H2O)+ ion have been investigated in aqueous perchloric acid solutions over a range of hydrogen ion concentrations (0.10 to 5.0 M) and at two ionic strengths (I = 1.0 and 5.0 M). At the lower ionic strength, plots of ln (AtA versus time show a nonlinearity that is consistent with that expected for consecutive first-order reactions. The rates of the faster reaction are similar to those reported for the spontaneous reduction of aquopyridine-cobalt(III) cations. At the higher ionic strength, the above noted curvature is not apparent and the decarboxylation kinetics of the title complex may be described by a pseudo-first-order rate constant: kobs = k[H3O+]. At 20°C, k = (1.75−+0.09) s−1 M−1 with activation parameters ofΔH = (97 −+ 4) kJ mol−1 and ΔS = −(54 −+ 32) J deg−1 mol−1. These kinetic parameters are compared with those previously reported for the similar complexes, Co(py)4CO3+ and Co(py)2(CO3)(H2O)2+.  相似文献   

5.
Genetic polymorphism of the Ps (parotid size variant) proteins found in saliva is determined by autosomal inheritance of two expressed and one unexpressed allele. This hypothesis is supported by studies in 43 families including 153 children. Gene frequencies determined for 150 randomly collected salivas from whites and 101 randomly collected salivas from blacks are as follows: for whites, Ps 1=0.598, Ps 2=0.101, Ps 0=0.301; for blacks, Ps 1=0.185, Ps 2=0.126, and Ps 0=0.689. The electrophoretic polymorphism is manifested by apparent differences in molecular weights between Ps proteins. The Ps proteins are glycosylated and have an approximate isoelectric point of pI 8.1 as determined by isoelectric focusing in gels. We have also found in saliva the presence of a protein (PmS) which shows strong positive correlations with the presence of the smaller sized Pm (PmF) salivary protein described by Ikemoto et al. (1977). This suggested that PmS is probably part of the Pm protein polymorphic system. For randomly collected salivas from whites, the gene frequencies are PmF+=0.15 (N=140) and PmS+=0.12 (N=150). For randomly collected salivas from blacks, the gene frequency is PmS+=0.24 (N=101). The gene frequency of PmF+ was not determined. Family studies support autosomal inheritance of PmF and PmS proteins. There is strong evidence for linkage of Pm to the proline-rich protein (PPP) region (11 families, lod score at =0.01 is 7.64) and of Ps to the PPP region (13 families, lod score at =0.01 is 11.50). Protein products of six linked loci (Pr, Pa, Db, Ps, Pm, and Gl), when tested against rabbit anti-Gl antiserum, show immunological reactions of partial or complete identity with each other by double diffusion analysis.This study was supported by a grant from the National Institutes of Dental Research (DEO 3658-14). Paper No. 2340 of the Laboratory of Genetics, University of Wisconsin, Madison, Wisconsin 53706.  相似文献   

6.
Irrigated olive is rapidly increasing in arid and semiarid areas, many of which may be negatively affected by soil salinity. We evaluated changes in trunk growth and leaf Cl, Na+ and K+ concentrations in young Arbequina olives (Olea europaea L.) grown in a saline-sodic field over a three-year period. The trunk diameter was measured at the beginning and the end of the 1999 (70 trees), 2000 (59 trees) and 2001 (42 trees) growing periods. Leaves, sampled in August of each year, were analyzed for Cl, Na+ and K+ concentrations. Soil salinity (apparent electrical conductivity, ECa) of each monitored tree was measured 14 times during the 1999–2001 experimental period with an electromagnetic sensor and converted to root zone electrical conductivity of the soil saturation extract (ECe) based on ECa–ECe calibration curves. Salinity tolerance was determined using the Maas and Hoffman threshold–slope response model. Based on salinity thresholds (ECethr), the tolerance of olive in terms of trunk growth was high in 1999 (ECethr = 6.7 dS m–1), but declined with age and time of exposure to salts by 30% in 2000 (ECethr = 4.7 dS m–1) and by 55% in 2001 (ECethr = 3.0 dS m–1). Based on the high absolute slopes obtained in all years (values between 16% and 23% dS–1 m), olive was classified as very sensitive to ECe values above the threshold. Trunk growth thresholds based on leaf ion concentrations varied, depending on years, between 2.6 and 4.0 mg g–1 (Clthr) and between 1.0 and 1.2 mg g–1 (Nathr), indicating that Arbequina olive was less sensitive to leaf Cl and much more sensitive to leaf Na+ than values reported as toxic in greenhouse studies. Leaf K+ slightly decreased with increasing salinity, whereas the K+/Na+ ratio sharply decreased with increasing salinity. We concluded that the initial salinity tolerance of olive was high, but declined sharply with time of exposure to salts and became quite sensitive due primarily to increasing toxic concentrations of Na+ in the leaves.  相似文献   

7.
Summary The function of the caecal bulb, and its adaptation to chronic high- or low-Na+ intake, was investigated by in vivo perfusion of anaesthetised birds. Effects of acute aldosterone injection (125 g·kg–1 body mass) were also measured.Evidence was found for primary active net absorption of Na+, inducing parallel Na-linked absorption of water and Cl and secretion of K+. Around 20–35% of total Cl absorption and K+ secretion were independent of Na+ fluxes, and these components appear to be driven by passive processes with apparent conductances of 6.3×10–3 (G Cl) and 1.1×10–3 (G K) S·cm–2.Acetate (40mM) stimulated Na+ fluxes (8.5–9.9 Eq·cm–2·h–1) and Na-linked water fluxes (27–44 l·cm–2·h–1). Increased coupling ratios (2.9–4.6 l·Eq–1) and other data indicate that these effects may be due to increased osmotic permeabilities of barriers involved in the Na-linked water transfer pathway.Low-Na+ maintenance enhanced EPD (49–69 mV, serosa positive) and all net fluxes:J Na (6.8–11.6);J K (–3.2––4.3);J Cl (4.3–5.6 Eq·cm serosal area–2·h–1);J v (28–43 l·cm–2·h–1) (mucosal-serosal fluxes positive).Acute aldosterone enhancedJ Na (10.8–14.0 Eq·cm–2·h–1) and EPD (54–66 mV) by 3 h after injection, but had no effect on the Na-linked components ofJ K orJ Cl.Abbreviations ECPD, EPD Electrochemical or electrical potential difference - G Cl ,G K ionic conductances (Cl, K+) - J v ,J ion net volume or ion flux rate, mucosa-serosa positive;P d (Cl) diffusive permeability coefficient (of Cl) - SEDM standard error of difference between means  相似文献   

8.
Clostridium botulinum produce the antigenically distinct 150 kD neurotoxin serotypes (e.g., A, B, C1, and E) and simultaneously proteins, A Hn+, B Hn+, C Hn+, and E Hn, that have high, low, and no hemagglutinating activity. A Hn+ and B Hn+ are serologically cross-reactive. A Hn+, B Hn+, and C Hn+ found as large aggregates (900–220 kD) can be dissociated on SDS-PAGE into multiple subunits, the smallest for A Hn+, B Hn+ is 17 kD and 27 kD for C Hn+. The 116 kD E Hn does not aggregate. We determined the sequences of 10–33 amino terminal residues of the 17, 21.5, 35, and 57 kD subunits of A Hn+ and B Hn+. Each of these subunits have unique sequences, indicating that the larger units studied are not homomers or heteromers of smaller units. The subunits of A Hn+ and B Hn+ of comparable size have striking sequence identity (e.g., 21.5 kD subunits from the two are identical and 57 kD subunits have 80% identity).In vitro proteolysis of 116 kD E Hn with different proteases did not impart hemagglutinating activity to the fragments. The 116 kD E Hn and one of its proteolytic fragments (87 kD) were partially sequenced. Sixty-two base pairs downstream from the termination codon of the cloned 33 kD subunit of C Hn+, there is an initiation codon followed by an open reading frame for at least 34 amino acid residues (Tsuzukiet al., 1990). The derived amino acid sequence of this open reading frame, we found, has 73–84% sequence identity with those of the 17 kD subunits of A Hn+ and B Hn+ and significant identity with the N-terminal of E Hn. These highly conserved sequences show existence of genetic linkage among the Hn+ and Hn proteins.  相似文献   

9.
Summary The reaction of abdominal skins of the frog speciesRana temporaria on mucosal K+-containing solutions was studied in an Ussing-type chamber by recording transepithelial potential difference (PD), short-circuit current (SCC) and conductance (G). With Na-Ringer's as serosal medium, a linear correlation between PD and the logarithm of the mucosal K+-concentration ([K] o ) was obtained. The K+-dependent SCC saturated with increasing [K] o , and could quickly and reversibly be depressed by addition of Rb+, Cs+, and H+, Li+, Na+, and NH 4 + did not influence K+ current. A large scatter was obtained for kinetic parameters like the slope of the PD-log [K] o -line (18–36.5 mV/decade), the apparent Michaelis constant (13–200mm), and the maximal current of the saturable SCC (6–50 A·cm–2), as well as for the degree of inhibition by Cs+ ions. This seemed to be caused by a time-dependent change during long time exposure to high [K] o (more than 30 sec), thereby inducing a selectivity loss of K+-transporting structures, together with an increase in SCC andG and a decrease in PD. Short time exposure to K+-containing solutions showed a competitive inhibition of K+ current by Cs+ ions, and a Michaelis constant of 6.6mm for the inhibitory action of Cs+. Proton titration resulted in a decrease of K+ current at pH<3. An acidic membrane component (apparent dissociation constant 2.5×10–3 m) is virtually controlling K+ transfer. Reducing the transepithelial K+-concentration gradient by raising the serosal potassium concentration was accompanied by the disappearance of SCC and PD.  相似文献   

10.
Summary Antibodies which were raised against highly purified membrane-bound (Na+–K+)-ATPase from the outer medulla of rat kidneys inhibit the (Na+–K+)-ATPase activity up to 95%. The antibody inhibition is reversible. The time course of enzyme inhibition and reactivation is biphasic in semilogarithmic plots.In the purified membrane-bound (Na+–K+)-ATPase negative cooperativity was observed (a) for the ATP dependence of the (Na+–K+)-ATPase activity (n=0.86), (b) for the ATP binding to the enzyme (n=0.58), and (c) for the ouabain inhibition of the (Na+–K+)-ATPase activity (n=0.77). By measuring the Na+ dependence of the (Na+–K+-ATPase reaction, a positive homotropic cooperativity (n=1.67) was found.As reactivation of the antibody-inhibited enzyme proceeds very slowly (t 0.5=5.2hr), it was possible to measure characteristics of the antibody-(Na+–K+)-ATPase complex: The antibodies exerted similar effects on the ATP dependence of the (Na+–K+)-ATPase reaction and on the ATP binding of the enzyme.V max of the (Na+–K+)-ATPase reaction and the number of ATP binding sites were reduced whileK 0.5 ATP for the (Na+–K+)-ATPase activity and for the ATP binding were increased by the antibodies. The Hill coefficients for the ATP binding and for the ATP dependence of the enzyme activity were not significantly altered by the antibodies. The antibodies increased theK 0.5 value for the Na+ stimulation of the (Na+–K+)-ATPase activity, but they did not alter the homotropic interactions between the Na+-binding sites. The negative cooperativity which was observed for the ouabain inhibition of the (Na+–K+)-ATPase activity was abolished by the antibodies.The data are tentatively explained by the following model: The antibodies bind to the (Na+–K+)-ATPase from the inner membrane side, reduce the ATP binding symmetrically at the ATP binding sites and reduce thereby also the (Na+–K+)-ATPase activity of the enzyme. The antibodies may inhibit the ATP binding by a direct interaction or by means of a conformational change at the ATP binding sites. This may possibly also lead to the alteration of the Na+ dependence of the (Na+–K+)-ATPase activity and to the observed alteration of the dose response to the ouabain inhibition.  相似文献   

11.
pH i recovery in acid-loaded Ehrlich ascites tumor cells and pH i maintenance at steady-state were studied using the fluorescent probe BCECF.Both in nominally HCO 3 -free media and at 25 mm HCO 3 , the measured pH i (7.26 and 7.82, respectively) was significantly more alkaline than the pH i . value calculated assuming the transmembrane HCO 3 gradient to be equal to the Cl gradient. Thus, pH i in these cells is not determined by the Cl gradient and by Cl/HCO 3 exchange.pH i recovery following acid loading by propionate exposure, NH 4 + withdrawal, or CO2 exposure is mediated by amiloride-sensitive Na+/H+ exchange in HCO3 free media, and in the presence of HCO 3 (25 mm) by DIDS-sensitive, Na+-dependent Cl/HCO 3 exchange. A significant residual pH i recovery in the presence of both amiloride and DIDS suggests an additional role for a primary active H+ pump in pH i regulation. pH i maintenance at steady-state involves both Na+/H+ exchange and Na+-dependent Cl/HCO 3 exchange.Acute removal of external Cl induces a DIDS-sensitive, Na+-dependent alkalinization, taken to represent HCO 3 influx in exchange for cellular Cl. Measurements of 36Cl efflux into Cl-free gluconate media with and without Na+ and/or HCO 3 (10 mm) directly demonstrate a DIDS-sensitive, Na+ dependent Cl/HCO 3 exchange operating at slightly acidic pH i (pHo 6.8), and a DIDS-sensitive, Na+-independent Cl/HCO 3 exchange operating at alkaline pH i (pH o 8.2).The excellent technical assistance of Marianne Schiødt and Birgit B. Jørgensen is gratefully acknowledged. The work was supported by the Carlsberg Foundation (B.K.) and by a grant from the Danish Natural Science Foundation (E.K.H. and L.O.S.).  相似文献   

12.
C. I. Ullrich  J. Guern 《Planta》1990,180(3):390-399
During the reduction of extracellular [Fe(CN)6]3– at the plasmalemma of intact, K+-starvedLemna gibba L. fronds, the external medium was acidified and K+ released, in the absence of inhibitors with rates of 10 e/8.5 H+/1.5 K+ (mol·(g FW)–1·–1). In K+ plants the larger K+ efflux caused a lag phase in extracellular acidification and a change in rates to 10 e/6 H+/4 K+ and in the presence of CN+salicylhydroxamic acid at pH 5 to 5.2 e/0 H+/6.6 K+. The e transfer was accompanied by a membrane depolarization of up to 100 mV and a cytosolic acidification of about 0.6 pH units, but only in K+ plants, where the extracellular acidification was smaller. These results indicate that a stimulation of the plasmalemma H+-ATPase may be triggered either by a cytosolic acidification or by a strong membrane depolarization. It is concluded that the redox system catalyses only uncoupled e transfer without H+ transfer across the plasmalemma. The obligatory, but secondary charge compensation is partially achieved by the rapid K+ release upon membrane depolarization and partially by the activity of the plasma membrane H+-ATPase, but not by an e/anion exchange. The extracellular acidification during [Fe(CN)6]3– reduction is generated by the conversion of a strong trivalent into a strong tetravalent anion. This acidification is caused by changes in the concentration ratio of strong cations to strong anions. Efflux of K+ and not the production of organic acids or NAD(P)H oxidation is the chemical cause of the measurable cytosolic acidification. Extracellular acidification was inversely correlated with intracellular acidification. Similarly, fusicoccin-induced pH changes were correlated with changes in the strong-ion concentration difference. Extracellular ± FC-dependent acidification and intracellular alkalinization of up to 0.6 pH units were strongly dependent on K+ fluxes. The ferricyanide-triggered trans-plasmalemma electron-transfer system is an example of how measurable pH changes are the consequence and not the cause of charge-transfer-induced changes in strong-ion fluxes.Abbreviations DCCD dicyclohexylcarbodiimide - Em electrical membrane potential difference - FC fusicoccin - pHc cytosolic pH - FW fresh weight - PM plasmalemma - SHAM salicylhydroxamic acid - SID strong-ion concentration difference This work was supported by the Deutsche Forschungsgemeinschaft. We gratefully acknowledge the Alexander von Humboldt award donated to J.G. We thank Professor Ulrich Lüttge (TH Darmstadt, FRG) for his kind support and Annett Ehrhardt and Dr. Karl Fischer (TH Darmstadt, FRG) for their valuable help with Cl and CO2 experiments. Special thanks are due to Professor Erasmo Marrè (Università di Milano, Italy) for continuous discussions and also to Professor Alessandro Ballio (Università di Roma, Italy) for their kind gifts of fusicoccin.  相似文献   

13.
Summary We have measured the intracellular potassium activity, [K+]i and the mechanisms of transcellular K+ transport in reabsorptive sweat duct (RSD) using intracellular ion-sensitive microelectrodes (ISMEs). The mean value of [K+]i in RSD is 79.8±4.1mm (n=39). Under conditions of microperfusion, the [K+]i is above equilibrium across both the basolateral membrane, BLM (5.5 times) and the apical membrane, APM (7.8 times). The Na+/K+ pump inhibitor ouabain reduced [K+]i towards passive distribution across the BLM. However, the [K+]i is insensitive to the Na+/K+/2 Cl cotransport inhibitor bumetanide in the bath. Cl substitution in the lumen had no effect on [K+]i. In contrast, Cl substitution in the bath (basolateral side) depolarized BLM from –26.0±2.6 mV to –4.7*±2.4 mV (n=3;* indicates significant difference) and decreased [K+]i from 76.0±15.2mm to 57.7* ±12.7mm (n=3). Removal of K+ in the bath decreased [K+]i from 76.3±15.0mm to 32.3*±7.6mm (n=4) while depolarizing the BLM from –32.5±4.1 mV to –28.3*±3.0 mV (n=4). Raising the [K+] in the bath by 10-fold increased [K+]i from 81.7±9.0mm to 95.0*±13.5mm and depolarized the BLM from –25.7±2.4 mV to –21.3*±2.9 mV (n=4). The K+ conductance inhibitor, Ba2+, in the bath also increased [K+]i from 85.8±6.7mm to 107.0*±11.5mm (n=4) and depolarized BLM from –25.8±2.2 mV to –17.0*±3.1 mV (n=4). Amiloride at 10–6 m increased [K+]i from 77.5±18.8mm to 98.8*±21.6mm (n=4) and hyperpolarized both the BLM (from –35.5±2.6 mV to –47.8*±4.3 mV) and the APM (from –27.5±1.4 mV to –46.0* ±3.5 mV,n=4). However, amiloride at 10–4 m decreased [K+]i from 64.5±0.9mm to 36.0*±9.9mm and hyperpolarized both the BLM (from –24.7±1.4 mV to –43.5*±4.2 mV) and APM (from –18.3±0.9 mV to –43.5*±4.2 mV,n=6). In contrast to the observations at the BLM, substitution of K+ or application of Ba2+ in the lumen had no effect on the [K+]i or the electrical properties of RSD, indicating the absence of a K+ conductance in the APM. Our results indicate that (i) [K+]i is above equilibrium due to the Na+/K+ pump; (ii) only the BLM has a K+ conductance; (iii) [K+]i is subject to modulation by transport status; (iv) K+ is probably not involved in carrier-mediated ion transport across the cell membranes; and (v) the RSD does not secrete K+ into the lumen.  相似文献   

14.
Summary Ouabain-resistant (OR), volume-or N-ethylmaleimide (NEM)-stimulated K+(Rb+)Cl fluxes were measured in low-K+ sheep red cells and found to be functionally separate but immunologically similar. In anisosmotic solutions both K+ effluxes and Rb+ influxes of NEM-treated and control cells were additive. In contrast to the NEM-stimulated K+Cl flux, metabolic depletion did not reduce K+Cl flux of normal or swollen cells. The anion preference of OR K+ efflux in NEM-treated cells was Br>Cl>HCO 3 =FI=NO 3 =CNS, and hence consistent with a reported Br>Cl>NO 3 sequence of the volume-dependent K+Cl transport. Alloimmune anti-Ll antibodies known to decrease passive K+ fluxes in low K+ cells reduced by 51% both volume-and NEM-stimulated, furosemidesensitive Rb+Cl fluxes suggesting their immunologic identity, a conclusion also supported by anti-L1 absorption studies. Since pretreatment with anti-L1 prevented the activation of Rb+ influx by NEM, and the impermeant glutathionmaleimide-I did not stimulate Rb+Cl influx, the NEM reactive SH groups must be located apart from the L1 antigen either within the membrane or on its cytoplasmic face. A model is proposed consisting of a K+Cl transport path(s) regulated by a protein with two functional subunits or domains; a chemically (C s) and a volume (V s)-stimulated domain, both interfacing with the L1 surface antigen. Attachment of alloanti-L1 from the outside reduces K+Cl transport stimulated throughC s by NEM orV s by cell swelling.  相似文献   

15.
Ion and acid–base balance were examined in the freshwater-adapted mummichog (Fundulus heteroclitus) using a series of treatments designed to perturb the coupling mechanisms. Unidirectional Cl uptake (JClin) was extremely low whereas JNain was substantial (three- to sixfold higher); comparable differences occurred in unidirectional efflux rates (JClout, JNaout). JClin was refractory to all treatments, suggesting that Cl/base exchange was unimportant or absent. Indeed, no base excretion or modulation of ion fluxes occurred for acid–base balance for up to 8 h after NaHCO3 loading (injections of 1000 or 3000 nequiv.·g−1). Acute environmental low pH (4.5) and amiloride (10−4 M) treatments caused concurrent inhibition of JNain and net H+ excretion (JH+net), indicating the presence of Na+/H+ exchange. JNain was elevated and JH+net restored during recovery from both treatments, but this exchange did not appear to be dynamically adjusted for acid–base homeostasis. High external ammonia exposure (1 mmol·l−1) initially blocked ammonia excretion (JAmmnet) but had no effect on JNain, whereas high pH (9.4) reduced both JAmmnet and JNain. Inhibition of JNain by the low pH and amiloride treatments had no effect on JAmmnet. These results indicate that ammonia excretion is entirely diffusive and independent of both Na+uptake and the protons that are transported via the Na+/H+ coupling. In addition, ureagenesis served as a compensatory mechanism during high external ammonia exposure, as a marked elevation in urea excretion partially replaced the inhibited JAmmnet. In all treatments, changes in the Na+–Cl net flux differential were consistent with changes in JH+net measured by traditional water titration techniques, indicating that the former can be used as an estimate of the acid–base status of the fish. Overall, the results demonstrate that the freshwater-adapted F. heteroclitus does not conform to the ion/acid–base relationships described in the standard model based on commonly studied species such as trout, goldfish, and catfish.  相似文献   

16.
Summary Transbasal electrical potential (V b) and intraepithelial potassium chemical activity ((K+) i ) were measured in isolated midgut epithelium of tobacco hornworm (Manduca sexta) using double-barrelled glass microelectrodes. Values ofV b ranging from +8 to –48 mV (relative to blood side) were recorded. For all sites, (K+) i is within a few millivolts of electrochemical equilibrium with the blood side bathing solution. Sites more negative than –20 mV show relatively high sensitivity ofV b to changes in blood side K+ concentration: 43% of these sites can be marked successfully with iontophoresed Lucifer yellow CH dye and shown to represent epithelial cells of all three types present in the midgut. In about half of successful marks, dye-coupling of several adjacent cells is seen. Low potential sites — those withV b less negative than –20 mV —typically do not show high sensitivity ofVb to changes of external K+, but rather (K+) i rapidly approaches the K+ activity of blood side bathing solution. These sites can seldom be marked with Lucifer yellow (4% success). The mean (K+) i of the high potential sites is 95±29 (sd)mm under standard conditions, a value which is in accord with published values for the whole tissue.  相似文献   

17.
We have measured fluid secretion rate in Rhodnius prolixus upper Malpighian tubules (UMT) stimulated to secrete with 5-OH-tryptamine. We used double perfusions in order to have access separately to the basolateral and to the apical cell membranes. Thirteen pharmacological agents were applied: ouabain, Bafilomycin A1, furosemide, bumetanide, DIOA, Probenecid, SITS, acetazolamide, amiloride, DPC, BaCl2, pCMBS and DTT. These agents are known to block different ion transport functions, namely ATPases, co- and/or counter-transporters and ion and water channels. The basic assumption is that water movement changes reflect changes in ion transport mechanisms, which we localize as follows: (i) At the basolateral cell membrane, fundamental are a Na+-K+-2Cl cotransporter and a Cl-HCO3 exchanger; of intermediate importance are the Na+-K+-ATPase, Cl channels and Rp-MIP water channels; K+ channels play a lesser role: (ii) At the apical cell membrane, most important are a K+-Cl cotransport that is being located for the first time, a V-H+-ATPase; and a Na+-H+ exchanger; a urate-anion exchanger and K+ channels are less important, while Cl channels are not important at all. A tentative model for the function of the UMT cell is presented.Symbols and abbreviations:ACTZ, acetazolamide; cAMP, cyclic adenosine-mono-phosphate; DIOA, [(dihydroindenyl)oxy] alkanoic acid; DPC, diphenylamine-2-carboxylate; DTT, dithiothreitol; 5-HT, 5-hydroxy-tryptamine; IR, Insects Ringer; Jv, secretion rate [nl/cm2.s]; pCMBS, parachloro-mercuri-benzene-sulphonate; Rp-MIP, Rhodnius prolixus water channels; SITS, 4-acetamido-4-isothiocyanatostilbene -2,2-disulfonic Acid; UMT, upper malpighian tubules.  相似文献   

18.
The nitrogen relations of an inbred line of white clover (Trifolium repens L.) thought to exhibit an abnormally low capacity for NO3 uptake (line LNU) were compared with a line regarded as normal with respect to NO3 uptake (line NNU). Growth, nodulation, N2 fixation and NO3 uptake were measured over 7 weeks in flowing solution culture (Experiment 1) by plants dependent for N acquisition on either (i) NO3 uptake, (ii) NO3 uptake +N2 fixation, or (iii) N2 fixation only. Effects of plant N status on the short-term uptake and translocation of 15NH4 + and 15NO3 were also investigated (Experiment 2). Nitrate uptake per plant by –fix/+NO3 line LNU was 50% of uptake by line NNU over 35 days, and there were significant differences in specific uptake rates of NO3 between the lines over the first 24 days. The `low NO3 uptake' phenotype was indistinct under +fix/+NO3 treatment. Nitrate lowered specific rates of nitrogen fixation by line NNU but had no effect on line LNU. Only low N status line LNU plants had lower short-term rates of NH4 + and NO3 uptake than line NNU. It is concluded that the `low NO3 uptake' phenotype of line LNU is inconsistently expressed. Circumstantial evidence points to increased NO3 efflux and decreased xylem translocation of NO3 as possible explanations for the lower NO3 uptake by line LNU.  相似文献   

19.
Summary Sodium- and potassium-dependent adenosine triphosphatase (Na+–K+-ATPase) is demonstrated in the branchial heart of Sepia officinalis L. by biochemical, cytochemical and autoradiographical methods. The biochemical data indicate the presence of Na+–K+-ATPase, shown by potassium and magnesium dependency and inhibition by ouabain. Cytochemically and autoradiographically, the enzyme is localized in the sarcolemma of the muscle cells. The positive reaction of the transparent cells (type I cells) is due to activity of alkaline phosphatases. The dark cells (type II cells) react negatively. In addition to the Na+–K+-ATPase, a magnesium-activated adenosine triphosphatase (Mg2+-ATPase) and a bicarbonate-stimulated ATPase (HCO 3 - -ATPase) are localized in the mitochondria.This study was supported by the Deutsche Forschungsgemeinschaft and is part of the doctoral dissertation  相似文献   

20.
Summary In the freshwater snailLymnaea stagnalis the influxes of Na+ and Cl were studied at different external concentrations of these ions. The characteristies of the Na+- and Cl-influxes are similar with respect to saturation kinetics,K m (0.1 mM) and activation by low-salt adaptation. In short-term experiments the Na+- and Cl-influxes are independent. Because of the counter-ions (H+ and HCO 3 ) involved, this indicates a potential acid-base regulatory capacity. Low-salt adaptation, due to either Na+-or Cl-depletion, activates both the Na+- and the Cl-influx. It is suggested that under both conditions the number of active integumental pumps, involved in Na+- as well as in Cl-uptake, is increased.  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