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.
6-Mercaptopurine and 6-thioguanine strongly inhibited the zero-trans entry of hypoxanthine into Novikoff rat hepatoma cells which lacked hypoxanthine/guanine phosphoribosyltransferase, whereas 8-azaguanine had no significant effect. 6-Mercaptopurine was transported by the hypoxanthine carrier with about the same efficiency as its natural substrates (Michaelis-Menten constant = 372 ± 23 μM; maximum velocity = 30 ± 0.7 pmol/μl cell H2O per s). 8-Azaguanine entry into the cells, on the other hand, showed no sign of saturability and was not significantly affected by substrates of the hypoxanthine/guanine carrier. The rate of entry of 8-azaguanine at 10–100 μM amounted to only about 5% of that of hypoxanthine transport and was related to its lipid solubility in the same manner as observed for various substances whose permeation through the plasma membrane is believed to be non-mediated. Only the non-ionized form of 8-azaguanine (pKa = 6.6) permeated the cell membrane.Studies with wild type Novikoff cells showed that permeation into the cell was the main rate-determining step in the conversion of extracellular 8-azaguanine to intracellular aza-GTP and its incorporation into nucleic acids. In contrast, 6-mercaptopurine was rapidly transported into cells and phosphoribosylated; the main rate-determining step in its incorporation into nucleic acids was the further conversion of 6-mercaptopurine riboside 5'-monophosphate.  相似文献   

2.
Incorporation of thymidine into Novikoff rat hepatoma cells was analyzed with a rapid sampling technique which allowed collection of 12 time points in 20 sec. Transport was studied in the absence of metabolism by using either ATP-depleted cells or a thymidine kinase negative subline. Transport was a rapid, saturable, nonconcentrative process with a Km of about 85 μM. The intracellular thymidine pool was also rapidly labeled in cells which phosphorylated thymidine, so that a group translocation process involving thymidine kinase can be ruled out. Under all conditions examined, phosphorylation, not the transport, of thymidine was the rate-determining step in its incorporation into the acid-soluble pool. Estimation of transport rates from total incorporation into cells which phosphorylate the substrate is invalid in this cell system and must be questioned in all instances.  相似文献   

3.
Summary Phloridzin-insensitive, Na+-independentd-glucose uptake into isolated small intestinal epithelial cells was shown to be only partially inhibited by trypsin treatment (maximum 20%). In contrast, chymotrypsin almost completely abolished hexose transport. Basolateral membrane vesicles prepared from rat small intestine by a Percoll® gradient procedure showed almost identical susceptibility to treatment by these proteolytic enzymes, indicating that the vesicles are predominantly oriented outside-out. These vesicles with a known orientation were employed to investigate the kinetics of transport in both directions across the membrane. Uptake data (i.e. movement into the cell) showed aK t of 48mm and aV max of 1.14 nmol glucose/mg membrane protein/sec. Efflux data (exit from the cell) showed a lowerK t of 23mm and aV max of 0.20 nmol glucose/mg protein/sec.d-glucose uptake into these vesicles was found to be sodium independent and could be inhibited by cytochalasin B. TheK t for cytochalasin B as an inhibitor of glucose transport was 0.11 m and theK D for binding to the carrier was 0.08 m.d-glucose-sensitive binding of cytochalasin B to the membrane preparation was maximized withl- andd-glucose concentrations of 1.25m. Scatchard plots of the binding data indicated that these membranes have a binding site density of 8.3 pmol/mg membrane protein. These results indicate that the Na+-independent glucose transporter in the intestinal basolateral membrane is functionally and chemically asymmetric. There is an outward-facing chymotrypsin-sensitive site, and theK t for efflux from the cell is smaller than that for entry. These characteristics would tend to favor movement of glucose from the cell towards the bloodstream.  相似文献   

4.
The magnesium chelate of the N(3)H tautomer of orotate, L3Mg, is the true substrate in the biosynthesis of orotidine 5′-monophosphate (OMP) catalyzed by yeast orotate phosphoribosyltransferase (OPRTase, E.C. 2.4.210) with a Michaelis constant KmL3Mg equal to 12(2) μM. It is postulated that Mg++ cations activate the transport of orotate to the active site by neutralizing the orotate charges; the ligand N(3)H is then exchanged between the incoming cation and the cation bound to the enzyme, thus ensuring the stabilization of the appropriate isomeric structure of orotate. This scheme, together with kinetic and thermodynamic data on orotate complexation by Mg++ and Ca++, accounts for the role of Ca++ cations that neither activate nor inhibit OMP synthesis.Cu++ and Ni++ inhibiting properties arise from the formation of inert complexes of orotate. Ni++ complexes have a poor affinity for the protein, whereas Cu++ complexes have a Michaelis constant similar to that of the L3Mg active species. The inertness of these complexes is tentatively understood in terms of low phosphoribosyl transfer rates as postulated from the kinetic study of the protonation of the complexes in water.  相似文献   

5.
The Role of MIP in Lens Fiber Cell Membrane Transport   总被引:1,自引:1,他引:0  
MIP has been hypothesized to be a gap junction protein, a membrane ion channel, a membrane water channel and a facilitator of glycerol transport and metabolism. These possible roles have been indirectly suggested by the localization of MIP in lens gap junctional plaques and the properties of MIP when reconstituted into artificial membranes or exogenously expressed in oocytes. We have examined lens fiber cells to see if these functions are present and whether they are affected by a mutation of MIP found in Cat Fr mouse lens. Of these five hypothesized functions, only one, the role of water channel, appears to be true of fiber cells in situ. Based on the rate of volume change of vesicles placed in a hypertonic solution, fiber cell membrane lipids have a low water permeability (p H2O ) on the order of 1 μm/sec whereas normal fiber cell membrane p H2O was 17 μm/sec frog, 32 μm/sec rabbit and 43 μm/sec mouse. Cat Fr mouse lens fiber cell p H2O was reduced by 13 μm/sec for heterozygous and 30 μm/sec for homozygous mutants when compared to wild type. Lastly, when expressed in oocytes, the p H2O conferred by MIP is not sensitive to Hg2+ whereas that of CHIP28 (AQP1) is blocked by Hg2+. The fiber cell membrane p H2O was also not sensitive to Hg2+ whereas lens epithelial cell p H2O (136 μm/sec in rabbit) was blocked by Hg2+. With regard to the other hypothesized roles, fiber cell membrane or lipid vesicles had a glycerol permeability on the order of 1 nm/sec, an order of magnitude less than that conferred by MIP when expressed in oocytes. Impedance studies were employed to determine gap junctional coupling and fiber cell membrane conductance in wild-type and heterozygous Cat Fr mouse lenses. There was no detectable difference in either coupling or conductance between the wild-type and the mutant lenses. Received: 17 February 1999/Revised: 16 April 1999  相似文献   

6.
Plant growth is achieved predominantly by cellular elongation, which is thought to be controlled on several levels by apoplastic auxin. Auxin export into the apoplast is achieved by plasma membrane efflux catalysts of the PIN‐FORMED (PIN) and ATP‐binding cassette protein subfamily B/phosphor‐glycoprotein (ABCB/PGP) classes; the latter were shown to depend on interaction with the FKBP42, TWISTED DWARF1 (TWD1). Here by using a transgenic approach in combination with phenotypical, biochemical and cell biological analyses we demonstrate the importance of a putative C‐terminal in‐plane membrane anchor of TWD1 in the regulation of ABCB‐mediated auxin transport. In contrast with dwarfed twd1 loss‐of‐function alleles, TWD1 gain‐of‐function lines that lack a putative in‐plane membrane anchor (HA–TWD1‐Ct) show hypermorphic plant architecture, characterized by enhanced stem length and leaf surface but reduced shoot branching. Greater hypocotyl length is the result of enhanced cell elongation that correlates with reduced polar auxin transport capacity for HA–TWD1‐Ct. As a consequence, HA–TWD1‐Ct displays higher hypocotyl auxin accumulation, which is shown to result in elevated auxin‐induced cell elongation rates. Our data highlight the importance of C‐terminal membrane anchoring for TWD1 action, which is required for specific regulation of ABCB‐mediated auxin transport. These data support a model in which TWD1 controls lateral ABCB1‐mediated export into the apoplast, which is required for auxin‐mediated cell elongation.  相似文献   

7.
Abstract

The interactions of octadecylamine and of positively charged liposomes (egg phosphatidylcholine/octadecylamine in molar ratio from 7:5 to 7:0.5) with human erythrocyte membrane have been studied by freeze-fracture and thin-section electron microscopy.

For the first time a very fast adsorption of liposomes to the cell membranes (less than 1 sec) is shown, and their intensive incorporation into plasma membrane (probably within the first 2-5 sec) without visible changes in cell morphology.

The prolonged incubation of the cells with liposomes results in certain morphologic changes: the transition of diskocytes to stomatocytes (30-80 sec) accompanied by the formation of isolated membrane vesicles in cell matrix (80-120 sec); the formation of pentalaminar contacts between plasma membrane of spherocytes and the membrane of isolated matrix vesicles (2.5-3 min); and the incorporation of matrix vesicle membranes into the spherocyte membranes (5 min+).

Possible molecular mechanisms underlying the observed structural changes are discussed briefly.  相似文献   

8.
The protein fraction released from human erythrocyte membranes with 90% acetic acid enhanced the transport of several sugar species when enclosed in erythrocyte ghosts. Both the influx and the efflux of d-glucose were increased so that permeation rather than sugar binding was involved. The permeation increase was selective, being found with d-glucose, d-galactose and d-xylose but not with l-glucose or lactose. The protein-dependent sugar transport was saturable and the incorporation of proteins into the ghost membrane brought Jmax to the level corresponding to intact erythrocytes, leaving Km unchanged.  相似文献   

9.
Transporters and ion channels are conventionally categorised into distinct classes of membrane proteins. However, some membrane proteins have a split personality and can function as both transporters and ion channels. The excitatory amino acid transporters (EAATs) in particular, function as both glutamate transporters and chloride (Cl?) channels. The EAATs couple the transport of glutamate to the co-transport of three Na+ ions and one H+ ion into the cell, and the counter-transport of one K+ ion out of the cell. The EAAT Cl? channel is activated by the binding of glutamate and Na+, but is thermodynamically uncoupled from glutamate transport and involves molecular determinants distinct from those responsible for glutamate transport. Several crystal structures of an EAAT archaeal homologue, GltPh, at different stages of the transport cycle, alongside numerous functional studies and molecular dynamics simulations, have provided extensive insights into the mechanism of substrate transport via these transporters. However, the molecular determinants involved in Cl? permeation, and the mechanism by which this channel is activated are not entirely understood. Here we will discuss what is currently known about the molecular determinants involved in EAAT-mediated Cl? permeation and the mechanisms that underlie their split personality.  相似文献   

10.
A binding component with a high affinity for 5-methyltetrahydrofolate (KD = 0.11μm) is present on the external surface of L1210 cells. The amount of binder (1 pmol/mg protein) corresponds to 8 × 104 sites per cell. The participation of this component in the high-affinity 5-methyltetrahydrofolate/methotrexate transport system is supported by similarities in the KD values for 5-methyltetrahydrofolate and methotrexate binding and the Kt values of these compounds for transport. Relative affinities for other folate substrates (aminopterin, 5-formyltetrahydrofolate, and folate) and various competitive inhibitors (thiamine pyrophosphate, ADP, AMP, arsenate, and phosphate) are also similar for both the binding component and the transport system. The measured binding activity does not represent low-temperature transport of substrate into cells, since it is readily saturable with time and is eliminated by either washing the cells with buffer or by the addition of excess unlabeled substrate.  相似文献   

11.
3-Quinuclidinyl benzilate (QNB), a potent antagonist of muscarinic acetylcholine receptors, has been demonstrated to inhibit specifically the zona pellucida (ZP)-inducud acrosome reaction (AR) in mouse sperm (Florman and Storey, 1982; Dev Biol 91:121–130). In this study we describe the solubilization and partial purification of the mouse sperm QNB binding activity which may represent a component of the putative receptor complex for ZP on the sperm plasma membrane. Sperm membranes were isolated from cell homogenates of washed, capacitated, epididymal mouse sperm. Scatchard plots of QNB binding to these membranes indicated a single class of binding sites with KD = 7.2 nM and Bmax = 8700 sites/cell. These binding characteristics are similar to those seen with QNB binding to whole cells (Florman and Storey, 1982, J Androl 3:157–164). Sperm membranes were solubilized using 1% digitonin/0.2% cholate, and the resultant detergent-soluble fraction possessed QNB binding activity similar to that of intact membranes. The detergent-soluble fraction maintained intact ZP receptor(s)–G protein coupling in that treatment of this fraction with either ZP or mastoparan resulted in a 35% or 65% increase in specific GTPγS binding, respectively. The solubilized membrane preparation was fractionated by gel permeation HPLC. A majority of specific QNB binding activity was confined to one HPLC fraction. Analysis of this fraction by SDS–PAGE revealed a complex of approximately 5 proteins unique to this fraction. The most prominent protein had a Mr of 72 kDa, which is within the Mr range for muscarinic receptors. A protein with Mr = 41 kDa was also present within this fraction. Subsequent pertussis toxin (PTX)-catalyzed ADP-ribosylation of this fraction revealed this protein to be the α subunit of the Gi class of G proteins. Although the QNB binding activity could not be positively identified, we propose that it is contained in one or more of the proteins unique to this fraction and that these proteins, including Gi, may act as part of a sperm receptor complex for the ZP. © 1994 Wiley-Liss, Inc.  相似文献   

12.
The effects of culture age, muscle activity, and cell fusion on the metabolism of acetylcholine receptors in the mouse muscle cell line, C2, were determined. Receptor degradation followed complex kinetics and was dependent on culture age. One or two day old myotubes degraded receptors rapidly (t50 = 7–8 h) in a nearly single exponential process. Four or five day old myotubes, however, degraded receptors more slowly (t50 = 12–16 h) in a process that deviated substantially from single exponential kinetics. A similar complex pattern of receptor degradation was seen with the L6 cell line, but receptor degradation followed single exponential kinetics and was independent of culture age in primary rat myotubes and the BC3H-1 cell line. Acetylcholine receptors on C2 myotubes were immunologically similar to the extrajunctional receptors of denervated mouse muscle. Clustered receptors were degraded at approximately the same rate as the total receptor population and receptor turnover was not changed when spontaneous contractions of the C2 myotubes were inhibited. Newly synthesized receptors were more rapidly degraded than older receptors. Finally, receptors on fusion-arrested C2 myoblasts were degraded at the same rate (t50 = 16 h) regardless of culture age.  相似文献   

13.
A mathematical model has been analysed describing uridine uptake in mammalian cells as a tandem process that involves membrane transport and uridine phosphorylation within the cell. The measurement of kinetic parametres of uridine uptake in 3T6 cells showed that the transport system possesses a low affinity to uridine (Kt = 145 microM) and a high velocity (Vt = 10 microM/sec), whereas the phosphorylation system possesses a high affinity for uridine (Ke = 10 microM) and a low velocity (Ve = 0.17 microM/sec). A method of construction of "ideal" curves was proposed, describing the time dependence of uridine uptake which helps to verify values of kinetic parameters obtained. On the basis of the theoretical analysis and generalization of experimental data it was concluded that the optimum conditions of uridine transport parameters measuring at 25 degrees C involve the uridine concentration in the medium equal to 20-200 microM, and the time of cell incubation, 2-20 sec, while the optimum conditions of uridine phosphorilation parameters measuring being its concentration in the medium 5-20 microM and the cell incubation longer than 1 minute.  相似文献   

14.
Soluble gangliosides in cultured neurotumor cells   总被引:3,自引:3,他引:0  
Abstract: The biosynthesis and degradation of glycosphingolipids were studied in cytosolic and membrane fractions obtained from rat glioma C6 cells. Both pools had a similar composition of neutral glycosphingolipids but the soluble pool contained only a few percent of the total. The major ganglioside in C6 cells was GM3, of which only 2% was soluble. Whereas the bulk of the membrane GM3 was accessible to surface labeling procedures, the soluble GM3 was not. Mouse neuroblastoma N18 cells also contained small amounts of cytoplasmic gangliosides corresponding to GM3, GM2, GM1, and GDla. When C6 cells were incubated with medium containing [3H]galactose at 37°C, the specific activity of soluble GM3 initially increased more rapidly than that of membrane GM3; by 4 h, the specific activities in both pools became equal. Total incorporation into the membrane pool, however, was always several-fold greater even at the shortest incubation times examined. The labeling pattern of neutral glycosphingolipids in both soluble and membrane fractions indicated the existence of a precursor-product relationship between glucosylceramide and other glycosphingolipids. When labeled cells were transferred to nonradioactive medium, glucosylceramide disappeared the most rapidly, with a 50% loss within <6 h. The turnover rates of other glycosphingolipids were much slower. Although cytosolic GM3 was degraded more rapidly (t1/2= 26 h) than membrane-bound GM3 (t1/2= 44 h), its turnover rate was much slower than the time required for transport of GM3 to the cell surface (20–30 min). Our results are consistent with the existence of a small intracellular pool of soluble gangliosides and neutral glycosphingolipids that is stable and independent of the main membrane-bound pool. Although the role of these cytosolic glycolipids is unknown, they do not appear to represent a transport pool between the site of synthesis and the plasma membrane.  相似文献   

15.
Time courses of [3H]uridine uptake as a function of uridine concentration were determined at 25° in untreated and ATP-depleted wild-type and uridine kinase-deficient Novikoff cells and in mouse L and P388 cells, Chinese hamster ovary cells and human HeLa cells. Short term uptake was measured by a rapid sampling technique which allows sampling of cell suspensions in intervals as short as one and one-half seconds. The initial segments of the time courses were the same in untreated, wild-type cells in which uridine is rapidly phosphorylated and in cells in which uridine phosphorylation was prevented due to lack of ATP or uridine kinase. The initial rates of uptake, therefore, reflected the rate of uridine transport. Uridine uptake, however, was approximately linear for only five to ten seconds at uridine concentrations from 20–160 μM and somewhat longer at higher concentrations. In phosphorylating cells the rate of uridine uptake (at 80 μM) then decreased to about 20–30% of the initial rate and this rate was largely determined by the rate of phosphorylation rather than transport. At uridine concentrations below 1 μM, however, the rate of intracellular phosphorylation in Novikoff cells approached the transport rate. The apparent substrate saturation of phosphorylation suggests the presence of a low Km uridine phosphorylation system in these cells. The “zero-trans” (zt) Km for the facilitated transport of uridine as estimated from initial uptake rates fell between 50 and 240 μM for all cell lines examined. The zero-trans Vmax values were also similar for all the lines (4–15 pmoles/μ1 cell H2O.sec). The time courses of uridine uptake by CHO cells and the kinetic constants for transport were about the same whether the cells were propagated (and analyzed for uridine uptake) in suspension or monolayer culture. When Novikoff cells were preloaded with 10 μM uridine the apparent Km and Vmax values (infinite-trans) were two to three times higher than the corresponding zero-trans values. Uridine transport was inhibited in a simple competitive manner by several other ribo- and deoxyribonucleosides. All nucleosides seem to be transported by the same system, but with different efficiencies. Uridine transport was also inhibited by hypoxanthine, adenine, thymine, Persantin, papaverin, and o-nitrobenzylthioinosine, and by pretreatment of the cells with p-chloromercuri-benzoate, but not by high concentrations of cytosine, D-ribose or acronycin. The inhibition of uridine transport by Persantin involved changes in both V and K. Because of the rapidity of transport, some loss of intracellular uridine occurred when cells were rinsed in buffer solution to remove extracellular substrate, even at 0°. This loss was prevented by the presence of a transport inhibitor, Persantin, in the rinse fluid or by separating suspended cells from the medium by centrifugation through oil. Metabolic conversion of intracellular uridine were also found to continue during the rinse period. The extent of artifacts due to efflux and metabolism during rinsing increased with duration of the rinse.  相似文献   

16.
Currents entering through single channels with conductivity 10 pS were produced on the membrane of an isolated neuron of the fresh-water molluskPlanorbarius corneus in the presence of suberyldicholine (5 µM) by the patch-clamp technique (cell-attached configuration). The times of stay of the channels in the open and closed states, as well as the durations of pulse bursts and clusters, were measured. The distributions of the time intervals obtained experimentally were approximated for open states by one exponential function: to=27±3 msec (n=21), and for closed states by a sum of three exponentials: tc1=9.5±1.0 msec (n=21); tc2=171±33 msec (n=19); tc3=5.2±1.0 sec (n=21). The burst durations are characterized by the presence of two exponential functions in the distribution: tb2=20±14 msec (n=10), tb2=203±23 msec (n=10), and the clusters by three exponential functions: tk1=33±11 msec (n=8), tk2=274±84 msec (n=8), and tk3=1.5±0.5 sec (n=9). Thus, for work of a chemoactivated channel associated with nicotinic-type cholinoreceptors in a mollusk neuron we can suggest a kinetic scheme with one open and three nonconducting states: C O D1A2 D2A2. The two "long-lived" closed states of the channel may be associated with desensitization of the integral response of the neurons to the application of suberyldicholine. Values were obtained for the rate constants of these proposed reactions. It is suggested that this model may be useful in analyzing the action of cholinomimetics and blockers on the molluskan neuronal membrane.I. M. Sechenov Institute of Evolutionary Physiology and Biochemistry, Academy of Sciences of the USSR, Leningrad. Translated from Neirofiziologiya, Vol. 23, No. 5, pp. 588–595, September–October, 1991.  相似文献   

17.
The binding of cis(c)- and trans(t)-Pt(NH3)2Cl2 to DNA at platinum/DNA-nucleotide ratios (Ri) of 0.1 or less has been studied by means of radioactive 195mPt-labeled compounds. Kinetic data are consistent with the following scheme:
At 25°C and pH 5–6 in 5 mM NaClO4, the values for the rate constants in the above scheme for the c-isomer are k2 = 2.2 × 10?5 sec?1, k7 = 0.32 (sec M)?1, and k8 = 143 (sec M)?1; for the t-isomer the values are k2 < 0.5 × 10?5 sec?1 and k7 = 0.95 (sec M)?1. Platinum-DNA adducts do not undergo detectable exchange after 3 days at 37°C, indicating the absence of a dynamic equillibrium. For both isomers the rate of binding is the same for single- and double-stranded DNA. The conclusions derived from Ag+ and H+ titration studies are consistent with binding at guanine N(7) for Ri < 0.1. The reaction rate is competitively inhibited by various salts and buffers and is suppressed by raising the pH (50% inhibition of initial rates at pH 7.3). At 37°C and pH 7 in 0.15 M NaCl, 6–8% of both the c- and t-isomers bind to DNA in 24 h, suggesting that both compounds should bind to DNA under biological conditions.  相似文献   

18.
Dihydroorotate dehydrogenase from rat liver is found to be located on the outer surface of the inner membrane of mitochondria. Dihydroorotate can diffuse freely from the cytosol into the mitochondria. Orotate can also diffuse freely from the mitochondria into the cytosol for futher conversion to UMP. Therefore, no active transport of either dihydroorotate or orotate is required in pyrimidine biosynthesis. The Km for l-dihydroorotate is 5.2 ± 0.6 μm. pd-Dihydroorotate is not a substrate for the enzyme but is a competitive inhibitor with a Ki of 1.4 mm. Of the compounds tested as analogs for dihydroorotate or metabolites related to pyrimidine biosynthesis, orotate is the strongest inhibitor, with a Ki of 8.4 μm. The Ki values for 2,4-dinitrophenol and barbiturate are 180 and 56 μm, respectively.  相似文献   

19.
The zero-trans influx of 500 μM uridine by CHO, P388, L1210 and L929 cells was inhibited by nitrobenzylthioinosine (NBTI) in a biphasic manner; 60–70% of total uridine influx by CHO cells and about 90% of that in P388, L1210 and L929 cells was inhibited by nmolar concentrations of NBTI (ID50 = 3?10 nM) and is designated NBTI-sensitive transport. The residual transport activity, designated NBTI-resistant transport, was inhibited by NBTI only at concentrations above 1 μM (ID50 = 10?50 μM). S49 cells exhibited only NBTI-sensitive uridine transport, whereas Novikoff cells exhibited only NBTI-resistant uridine transport. In all instances NBTI-sensitive transport correlated with the presence of between 7·104 and 7·105 high-affinity NBTI binding sites/cell (Kd = 0.3?1 nM). Novikoff cells lacked such sites. The two types of nucleoside transport, NBTI-resistant and NBTI-sensitive, were indistinguishable in substrate affinity, temperature dependence, substrate specificity, inhibition by structurally unrelated substances, such as dipyridamole or papaverine, and inhibition by sulfhydryl reagents or hypoxanthine. We suggest, therefore, that a single nucleoside transporter can exist in an NBTI-sensitive and an NBTI-resistant form depending on its disposition in the plasma membrane. The sensitive form expresses a high-affinity NBTI binding site(s) which is probably made up of the substrate binding site plus a hydrophobic region which interacts with the lipophilic nitrobenzyl group of NBTI. The latter site seems to be unavailable in NBTI-resistant transporters. The proportion of NBTI-resistant and sensitive uridine transport was constant during proportion of NBTI-resistant and sensitive uridine transport was constant during progression of P388 cells through the cell cycle and independent of the growth stage of the cells in culture. There were additional differences in uridine transport between cell lines which, however, did not correlate with NBTI sensitivity and might be related to the species origin of the cells. Uridine transport in Novikoff cells was more sensitive to inhibition by dipyridamole and papaverine than that in all other cell lines tested, whereas uridine transport in CHO cells was the most sensitive to inactivation by sulfhydryl reagents.  相似文献   

20.
Abstract

The structural stability and transport properties of the cyclic peptide nanotube (CPN) 8?×?[Cys–Gly–Met–Gly]2 in different phospholipid bilayers such as POPA (1-palmitoyl-2-oleoyl-sn-glycero-3-phosphatidic acid), POPE (1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoethanolamine), POPC (1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine), POPG (1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoglycerol) and POPS (1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoserine) with water have been investigated using molecular dynamics (MD) simulation. The hydrogen bonds and non-bonded interaction energies were calculated to study the stability in different bilayers. One µs MD simulation in POPA lipid membrane reveals the stability of the cyclic peptide nanotube, and the simulations at various temperatures manifest the higher stability of 8?×?[Cys–Gly–Met–Gly]2. We demonstrated that the presence of sulphur-containing amino acids in CPN enhances the stability through disulphide bonds between the adjacent rings. Further, the water permeation coefficient of the CPN is calculated and compared with human aquaporin-2 (AQP2) channel protein. It is found that the coefficients are highly comparable to the AQP2 channel though the mechanism of water transport is not similar to AQP 2; the flow of water in the CPN is taking place as a two-line 1–2–1–2 file fashion. In addition to that, the transport behavior of Na+ and K+ ions, single water molecule, urea and anti-cancer drug fluorouracil were investigated using pulling simulation and potential of mean force calculation. The above transport behavior shows that Na+ is trapped in CPN for a longer time than other molecules. Also, the interactions of the ions and molecules in Cα and mid-Cα plane were studied to understand the transport behavior of the CPN. Abbreviations AQP2 Aquaporin-2

CPN Cyclic peptide nanotube

MD Molecular dynamics

POPA 1-Palmitoyl-2-oleoyl-sn-glycero-3-phosphatidic acid

POPE 1-Palmitoyl-2-oleoyl-sn-glycero-3-phosphoethanolamine

POPG 1-Palmitoyl-2-oleoyl-sn-glycero-3-phosphoglycerol

POPS 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoserine

Communicated by Ramaswamy H. Sarma  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