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 625 毫秒
1.
Cell survival during freezing applications in biomedicine is highly correlated to the temperature history and its dependent cellular biophysical events of dehydration and intracellular ice formation (IIF). Although cell membranes are known to play a significant role in cell injury, a clear correlation between the membrane state and the surrounding intracellular and extracellular water is still lacking. We previously showed that lipid hydration in LNCaP tumor cells is related to cellular dehydration. The goal of this study is to build upon this work by correlating both the phase state of the membrane and the surrounding water to cellular biophysical events in three different mammalian cell types: human prostate tumor cells (LNCaP), human dermal fibroblasts (HDF), and porcine smooth muscle cells (SMC) using Fourier Transform Infrared spectroscopy (FTIR). Variable cooling rates were achieved by controlling the degree of supercooling prior to ice nucleation (− 3 °C and − 10 °C) while the sample was cooled at a set rate of 2 °C/min. Membranes displayed a highly cooperative phase transition under dehydrating conditions (i.e. NT = − 3 °C), which was not observed under IIF conditions (NT = − 10 °C). Spectral analysis showed a consistently greater amount of ice formation during dehydrating vs. IIF conditions in all cell types. This is hypothesized to be due to the extreme loss of membrane hydration in dehydrating cells that is manifested as excess water available for phase change. Interestingly, changes in residual membrane conformational disorder correlate strongly with cellular volumetric decreases as assessed by cryomicroscopy. A strong correlation was also found between the activation energies for freezing induced lyotropic membrane phase change determined using FTIR and the water transport measured by cryomicroscopy. Reduced lipid hydration under dehydration freezing conditions is suggested as one of the likely causes of what has been termed as “solution effects” injury in cryobiology.  相似文献   

2.
To survive freezing, cells must not undergo internal ice formation during cooling. One vital factor is the cooling rate. The faster cells are cooled, the more their contents supercool, and at some subzero temperature that supercooled cytoplasm will freeze. The question is at what temperature? The relation between cooling rate and cell supercooling can be computed. Two important parameters are the water permeability (Lp) and its temperature dependence. To avoid intracellular ice formation (IIF), the supercooling must be eliminated by dehydration before the cell cools to its ice nucleation temperature. With an observed nucleation temperature of −25 °C, the modeling predicts that IIF should not occur in yeast cooled at <20 °C/min and it should occur with near certainty in cells cooled at ?30 °C/min. Experiments with differential scanning calorimetry (DSC) confirmed these predictions closely. The premise with the DSC is that if there is no IIF, one should see only a single exotherm representing the freezing of the external water. If IIF occurs, one should see a second, lower temperature exotherm. A further test of whether this second exotherm is IIF is whether it disappears on repeated freezing. IIF disrupts the plasma membrane; consequently, in a subsequent freeze cycle, the cell can no longer supercool and will not exhibit a second exotherm. This proved to be the case at cooling rates >20 °C/min.  相似文献   

3.
The current study presents a new and novel analysis of heat release signatures measured by a differential scanning calorimeter (DSC) associated with water transport (WT), intracellular ice formation (IIF) and extracellular ice formation (EIF). Correlative cryomicroscopy experiments were also performed to validate the DSC data. The DSC and cryomicroscopy experiments were performed on human dermal fibroblast cells (HDFs) at various cytocrit values (0–0.8) at various cooling rates (0.5–250 °C/min). A comparison of the cryomicroscopy experiments with the DSC analysis show reasonable agreement in the water transport (cellular dehydration) and IIF characteristics between both the techniques with the caveat that IIF measured by DSC lagged that measured by cryomicroscopy. This was ascribed to differences in the techniques (i.e. cell vs. bulk measurement) and the possibility that not all IIF is associated with visual darkening. High and low rates of 0.5 °C/min and 250 °C/min were chosen as HDFs did not exhibit significant IIF or WT at each of these extremes respectively. Analysis of post-thaw viability data suggested that 10 °C/min was the presumptive optimal cooling rate for HDFs and was independent of the cytocrit value. The ratio of measured heat values associated with IIF (qIIF) to the total heat released from both IIF and water transport or from the total cell water content in the sample (qCW) was also found to increase as the cooling rate was increased from 10 to 250 °C/min and was independent of the sample cytocrit value. Taken together, these observations suggest that the proposed analysis is capable of deconvolving water transport and IIF data from the measured DSC latent heat thermograms in cell suspensions during freezing.  相似文献   

4.
In the present study, we report the effects of cooling ejaculated and epididymal rhesus monkey (Macacamulatta) sperm with and without the presence of a cryoprotective agent, glycerol. Water transport data during freezing of ejaculated and epididymal sperm cell suspensions were obtained at a cooling rate of 20 °C/min in the absence of any cryoprotective agents and in the presence of 0.7 M of glycerol, as well. Using previously published values, the macaque sperm cell was modeled as a cylinder of length 73.83 μm with a radius of 0.40 μm and an osmotically inactive cell volume, Vb, of 0.772Vo, where Vo is the isotonic cell volume. This translated to a surface area, SA to initial water volume, WV ratio of ∼22 μm−1. By fitting a model of water transport to the experimentally determined volumetric shrinkage data, the best-fit membrane permeability parameters (reference membrane permeability to water at 0 °C, Lpg or Lpg[cpa] and the activation energy, ELp or ELp[cpa]) were found to range from: Lpg or Lpg[cpa] = 0.0020-0.0029 μm/min-atm; ELp or ELp[cpa]) = 10.6-18.3 kcal/mole. By incorporating these membrane permeability parameters in a recently developed equation (optimal cooling rate, ; where the units of Bopt are °C/min, ELp or ELp[cpa] are kcal/mole, Lpg or Lpg[cpa] are μm/min-atm and SA/WV are μm−1), we determined the optimal rates of freezing macaque sperm to be ∼23 °C/min (ejaculated sperm in the absence of CPAs), ∼29 °C/min (ejaculated sperm in the presence of glycerol), ∼24 °C/min (epididymal sperm in the absence of CPAs) and ∼24 °C/min (epididymal sperm in the presence of glycerol). In conclusion, the subzero water transport response and consequently the subzero water transport parameters are not significantly different between the ejaculated and epididymal macaque spermatozoa under corresponding cooling conditions.  相似文献   

5.
Granulocytes are currently transfused as soon as possible after collection because they rapidly deteriorate after being removed from the body. This short shelf life complicates the logistics of granulocyte collection, banking, and safety testing. Cryopreservation has the potential to significantly increase shelf life; however, cryopreservation of granulocytes has proven to be difficult. In this study, we investigate the membrane permeability properties of human granulocytes, with the ultimate goal of using membrane transport modeling to facilitate development of improved cryopreservation methods. We first measured the equilibrium volume of human granulocytes in a range of hypo- and hypertonic solutions and fit the resulting data using a Boyle-van’t Hoff model. This yielded an isotonic cell volume of 378 μm3 and an osmotically inactive volume of 165 μm3. To determine the permeability of the granulocyte membrane to water and cryoprotectant (CPA), cells were injected into well-mixed CPA solution while collecting volume measurements using a Coulter Counter. These experiments were performed at temperatures ranging from 4 to 37 °C for exposure to dimethyl sulfoxide, glycerol, ethylene glycol, and propylene glycol. The best-fit water permeability was similar in the presence of all of the CPAs, with an average value at 21 °C of 0.18 μm atm−1 min−1. The activation energy for water transport ranged from 41 to 61 kJ/mol. The CPA permeability at 21 °C was 6.4, 1.0, 8.4, and 4.0 μm/min for dimethyl sulfoxide, glycerol, ethylene glycol, and propylene glycol, respectively, and the activation energy for CPA transport ranged between 59 and 68 kJ/mol.  相似文献   

6.
Fourier transform infrared spectroscopy (FTIR) and cryomicroscopy were used to define the process of cellular injury during freezing in LNCaP prostate tumor cells, at the molecular level. Cell pellets were monitored during cooling at 2 °C/min while the ice nucleation temperature was varied between − 3 and − 10 °C. We show that the cells tend to dehydrate precipitously after nucleation unless intracellular ice formation occurs. The predicted incidence of intracellular ice formation rapidly increases at ice nucleation temperatures below − 4 °C and cell survival exhibits an optimum at a nucleation temperature of − 6 °C. The ice nucleation temperature was found to have a great effect on the membrane phase behavior of the cells. The onset of the liquid crystalline to gel phase transition coincided with the ice nucleation temperature. In addition, nucleation at − 3 °C resulted in a much more co-operative phase transition and a concomitantly lower residual conformational disorder of the membranes in the frozen state compared to samples that nucleated at − 10 °C. These observations were explained by the effect of the nucleation temperature on the extent of cellular dehydration and intracellular ice formation. Amide-III band analysis revealed that proteins are relatively stable during freezing and that heat-induced protein denaturation coincides with an abrupt decrease in α-helical structures and a concomitant increase in β-sheet structures starting at an onset temperature of approximately 48 °C.  相似文献   

7.
Overwintering success is one of the key aspects affecting the development and outbreaks of the spruce bark beetle, Ips typographus (L.) populations. This paper brings detailed analysis of cold tolerance, and its influence on overwintering success, in two Central European populations of I. typographus during two cold seasons. Evidence for a supercooling strategy in overwintering adults is provided. The lower lethal temperature corresponds well to the supercooling point that ranges between −20 and −22 °C during winter months. The supercooled state is stabilized by the absence of internal ice nucleators and by seasonal accumulation of a mixture of sugars and polyols up to the sum concentration of 900 mM. The cryoprotective function of accumulated metabolites is probably based on increasing the osmolality and viscosity of supercooled body fluids and decreasing the relative proportion of water molecules available for lethal formation of ice nuclei. No activity of thermal hysteresis factors (stabilizers of supercooled state) was detected in hemolymph. Lethal times for 50% mortality (Lts50) in the supercooled state at −5, −10 or −15 °C are weeks (autumn, spring) or even months (winter), suggesting relatively little mortality caused by chill injury. Lts50 at −15 °C are significantly shorter in moist (6.9 days) than in dry (>42 days) microenvironment because there is higher probability of external ice nucleation and occurrence of lethal freezing in the moist situation.  相似文献   

8.
Typically, subzero permeability measurements are experimentally difficult and infrequently reported. Here we report an approach we have applied to mouse oocytes. Interrupted cooling involves rapidly cooling oocytes (50 °C/min) to an intermediate temperature above the intracellular nucleation zone, holding them for up to 40 min while they dehydrate, and then rapidly cooling them to −70 °C or below. If the intermediate holding temperature and holding time are well chosen, high post thaw survival of the oocytes is possible because the freezable water is removed during the hold. The length of time required for the exit of the freezable water allows the water permeability at that temperature to be determined. These experiments used 1.5 M ethylene glycol in PBS and included a transient hold of 2 min for equilibration at −10 °C, just below the extracellar ice formation temperature. We obtain an Lp = 1.8 × 10−3 μm min−1 atm−1 at −25 °C based on a hold time of 30 min yielding 80% survival and the premise that most of the freezable water is removed during the 30 min hold. If we assume that the water permeability is a continuous function of temperature and that its Ea changes at 0 °C, we obtain a subzero Ea of 21 kcal/mol; higher than the suprazero value of 14 kcal/mol. A number of assumptions are required for these water loss calculations and the resulting value of Lp can vary by up to a factor of 2, depending on the choices make.  相似文献   

9.
Expression of urea transporter UT-B confers high urea permeability to mammalian erythrocytes. Erythrocyte membranes also permeate various urea analogues, suggesting common transport pathways for urea and structurally similar solutes. In this study, we examined UT-B-facilitated passage of urea analogues and other neutral small solutes by comparing transport properties of wildtype to UT-B-deficient mouse erythrocytes. Stopped-flow light-scattering measurements indicated high UT-B permeability to urea and chemical analogues formamide, acetamide, methylurea, methylformamide, ammonium carbamate, and acrylamide, each with Ps > 5.0 × 10− 6 cm/s at 10 °C. UT-B genetic knockout and phloretin treatment of wildtype erythrocytes similarly reduced urea analogue permeabilities. Strong temperature dependencies of formamide, acetamide, acrylamide and butyramide transport across UT-B-null membranes (Ea > 10 kcal/mol) suggested efficient diffusion of these amides across lipid bilayers. Urea analogues dimethylurea, acryalmide, methylurea, thiourea and methylformamide inhibited UT-B-mediated urea transport by > 60% in the absence of transmembrane analogue gradients, supporting a pore-blocking mechanism of UT-B inhibition. Differential transport efficiencies of urea and its analogues through UT-B provide insight into chemical interactions between neutral solutes and the UT-B pore.  相似文献   

10.
The structure and water permeability of bilayers composed of the ether-linked lipid, dihexadecylphosphatidylcholine (DHPC), were studied and compared with the ester-linked lipid, dipalmitoylphosphaditdylcholine (DPPC). Wide angle X-ray scattering on oriented bilayers in the fluid phase indicate that the area per lipid A is slightly larger for DHPC than for DPPC. Low angle X-ray scattering yields A = 65.1 Å2 for DHPC at 48 °C. LAXS data provide the bending modulus, KC = 4.2 × 10−13 erg, and the Hamaker parameter H = 7.2 × 10−14 erg for the van der Waals attractive interaction between neighboring bilayers. For the low temperature phases with ordered hydrocarbon chains, we confirm the transition from a tilted Lβ′ gel phase to an untilted, interdigitated LβI phase as the sample hydrates at 20 °C. Our measurement of water permeability, Pf = 0.022 cm/s at 48 °C for fluid phase DHPC is slightly smaller than that of DPPC (Pf = 0.027 cm/s) at 50 °C, consistent with our triple slab theory of permeability.  相似文献   

11.
Bilayers composed of phosphatidylcholine (PC), sphingomyelin (SM), and cholesterol (CHOL) are commonly used as systems to model the raft-lipid domain structure believed to compartmentalize particular cell membrane proteins. In this work, micropipette aspiration of giant unilamellar vesicles was used to test the elasticities, water permeabilities, and rupture tensions of single-component PC, binary 1:1 PC/CHOL, and 1:1 SM/CHOL, and ternary 1:1:1 PC/SM/CHOL bilayers, one set of measurements with dioleoyl PC (DOPC; C18:1/C18:1 PC) and the other with stearoyloleoyl PC (SOPC; C18:0/C18:1 PC). Defining the elastic moduli (KA), the initial slopes of the increase in tension (σ) versus stretch in lipid surface area (αe) were determined for all systems at low (15°C) and high (32-33°C) temperatures. The moduli for the single-component PC and binary phospholipid/CHOL bilayers followed a descending hierarchy of stretch resistance with SM/CHOL > SOPC/CHOL > DOPC/CHOL > PC. Although much more resistant to stretch than the single-component PC bilayers, the elastic response of vesicle bilayers made from the ternary phospholipid/CHOL mixtures showed an abrupt softening (discontinuity in slope), when immediately subjected to a steady ramp of tension at the low temperature (15°C). However, the discontinuities in elastic stretch resistance at low temperature vanished when the bilayers were held at ∼1 mN/m prestress for long times before a tension ramp and when tested at the higher temperature 32-33°C. The elastic moduli of single-component PC and DOPC/CHOL bilayers changed very little with temperature, whereas the moduli of the binary SOPC/CHOL and SM/CHOL bilayers diminished markedly with increase in temperature, as did the ternary SOPC/SM/CHOL system. For all systems, increasing temperature increased the water permeability but decreased rupture tension. Concomitantly, the measurements of permeability exhibited a prominent correlation with the rupture tension across all the systems. Together, these micromechanical tests of binary and ternary phospholipid/CHOL bilayers demonstrate that PC hydrocarbon chain unsaturation and temperature are major determinants of the mechanical and permeation properties of membranes composed of raft microdomain-forming lipids.  相似文献   

12.
When cooled at rapid rates to temperatures between −10 and −30°C, the incidence of intracellular ice formation was less in protoplasts enzymically isolated from cold acclimated leaves of rye (Secale cereale L. cv Puma) than that observed in protoplasts isolated from nonacclimated leaves. The extent of supercooling of the intracellular solution at any given temperature increased in both nonacclimated and acclimated protoplasts as the rate of cooling increased. There was no unique relationship between the extent of supercooling and the incidence of intracellular ice formation in either nonacclimated or acclimated protoplasts. In both nonacclimated and acclimated protoplasts, the extent of intracellular supercooling was similar under conditions that resulted in the greatest difference in the incidence of intracellular ice formation—cooling to −15 or −20°C at rates of 10 or 16°C/minute. Further, the hydraulic conductivity determined during freeze-induced dehydration at −5°C was similar for both nonacclimated and acclimated protoplasts. A major distinction between nonacclimated and acclimated protoplasts was the temperature at which nucleation occurred. In nonacclimated protoplasts, nucleation occurred over a relatively narrow temperature range with a median nucleation temperature of −15°C, whereas in acclimated protoplasts, nucleation occurred over a broader temperature range with a median nucleation temperature of −42°C. We conclude that the decreased incidence of intracellular ice formation in acclimated protoplasts is attributable to an increase in the stability of the plasma membrane which precludes nucleation of the supercooled intracellular solution and is not attributable to an increase in hydraulic conductivity of the plasma membrane which purportedly precludes supercooling of the intracellular solution.  相似文献   

13.
The light brown apple moth, Epiphyas postvittana, a leafroller native to southeastern Australia was discovered in California in 2006. The highly polyphagous nature of this pest adds to the importance of being able to predict the potential distribution of this invader across the North American continent. The spread of ectothermic species that lack winter diapause, such as E. postvittana, can be limited by their ability to tolerate cold temperature extremes. In this study we examined the cold hardiness of 4th to 6th instar E. postvittana, the only life stages known to overwinter in California, through a combination of supercooling point (SCP) and mortality at low temperatures. Our results showed that the mean SCP for E. postvittana ranged from −14.1 °C for 6th instars to −16.0 °C for 4th instars. Lethal time leading to 50% mortality (LT50) for the three instars combined were 2.5 h at −10.5 °C, 41 h at −6.5 °C and 198 h at −0.9 °C. At 3 °C, the LT50 of 4th instars was significantly lower at 775 h than that for 5th and 6th instars combined at 1029 h. The cold hardiness characteristics of later-instar E. postvittana larvae were comparable to those of pink bollworm, Pectinophora gossypiella, a diapausing invasive with a geographic distribution restricted to southern California. Slightly greater cold hardiness is shown by the indigenous non-diapausing leafroller Argyrotaenia franciscana, which is restricted to the Pacific Coast of North America. We therefore conclude that the moderate cold hardiness of E. postvittana will substantially limit its spread into northern temperate regions of North America.  相似文献   

14.
15.
Cucujus clavipes puniceus is a freeze avoiding beetle capable of surviving the long, extremely cold winters of the Interior of Alaska. Previous studies showed that some individuals typically supercool to mean values of approximately − 40 °C, with some individuals supercooling to as low as − 58 °C, but these non-deep supercooling (NDSC) individuals eventually freeze if temperatures drop below this. However, other larvae, especially if exposed to very cold temperatures, supercool even further. These deep supercooling (DSC) individuals do not freeze even if cooled to − 100 °C. In addition, the body water of the DSC larvae vitrifies (turns to a glass) at glass transition temperatures of − 58 to − 70 °C. This study examines the proteomes of DSC and NDSC larvae to assess proteins that may contribute to or inhibit the DSC trait. Using high throughput proteomics, we identified 138 proteins and 513 Gene Ontology categories in the DSC group and 104 proteins and 573 GO categories in the NDSC group. GO categories enriched in DSC include alcohol metabolic process, cellular component morphogenesis, monosaccharide metabolic process, regulation of biological quality, extracellular region, structural molecule activity, and antioxidant activity. Proteins unique to DSC include alpha casein precursor, alpha-actinin, vimentin, tropomyosin, beta-lactoglobulin, immunoglobulins, tubulin, cuticle proteins and endothelins.  相似文献   

16.
Experiments were carried out to clarify the effects of an electrostatic field on ice formation. Distilled water was used as the test sample and was kept in a special container and cooled down to a constant temperature of −30 °C. The strength of the electrostatic field used in the experiments was in the range of 1.0 × 103-1.0 × 105 V/m. The results indicated that the electrostatic field was capable of inducing nucleation in water supercooled at a relatively high temperature and raising the temperature of supercooling by up to 1.6 °C. The analysis indicated that dipole polarization of the water molecules by the electrostatic field is the primary factor in this phenomenon. Under an electrostatic field, water molecules have a tendency to align with the electrostatic field. Water molecules with dipole moments along the direction of the electrostatic field are the most stable, and have the maximum value for the Boltzmann distribution function. These properties are conducive to nucleation. A special method was used to determine the dependence of the phase transformation time on the application of an electrostatic field. It was found that the phase transformation time was unaffected by the application of an electrostatic field and only the supercooling temperature was affected.  相似文献   

17.
The cold tolerance of first-instar nymphs of the Australian plague locust, Chortoicetes terminifera, was examined using measures of total body water content, supercooling point and mortality for a range of sub-zero temperature exposure regimes. The supercooling points for starved and fed nymphs were −13.1 ± 0.9 and −12.6 ± 1.6 °C, and freezing caused complete mortality. Above these temperatures, nymphs were cold tolerant to different degrees based on whether they were starved or given access to food and water for 24 h prior to exposure. The rate of cooling also had a significant effect on mortality. Very rapid cooling to −7 °C caused 84 and 87% mortality for starved and fed nymphs respectively, but this significantly decreased for starved nymphs if temperature declined by more ecologically realistic rates of 0.5 and 0.1 °C min−1. These results are indicative of a rapid cold hardening response and are discussed in terms of the likely effects of cold nights and frost on first-instar nymphal survival in the field.  相似文献   

18.
In a companion paper, we demonstrated that dynamic range limitations can confound measurement of the osmotically inactive volume using electrical sensing zone instruments (e.g., Coulter counters), and presented an improved parameter estimation method in which a lognormal function was fit to the cell volume distribution to allow extrapolation beyond the bounds of the data. Presently, we have investigated the effect of dynamic range limitations on measurement of the cell membrane water permeability (Lp), and adapted the lognormal extrapolation method for estimation of Lp from transient volume data. An alternative strategy (the volume limit adjustment method, in which the measured isotonic volume distribution is used to generate model predictions for curve fitting, and the bounds of the dynamic range are adjusted such that extrapolation is not required) was also developed. The performance of these new algorithms was compared to that of a conventional parameter estimation method. The best-fit Lp values from in vitro experiments with mouse insulinoma (MIN6) cells differed significantly for the different parameter estimation techniques (< 0.001). Using in silico experiments, the volume limit adjustment method was shown to be the most accurate (relative error 0.4 ± 3.2%), whereas the conventional method underestimated Lp by 19 ± 2% for MIN6 cells. Parametric analysis revealed that the error associated with the conventional method was sensitive to the dynamic range and the width of the volume distribution. Our initial implementation of the lognormal extrapolation method also yielded significant errors, whereas accuracy of this algorithm improved after including a normalization scheme.  相似文献   

19.
Shinsuke Seki 《Cryobiology》2010,61(1):155-157
When cells that have been subjected to supposedly innocuous freezing or vitrification procedures are used as the source material for subsequent experiments, it is important that they possess or exhibit the same relevant properties as fresh cells. In this study, we compared the temperatures of intracellular ice formation (IIF) in previously vitrified mouse oocytes/embryos with those in fresh intact ones. In the case of MII oocytes, 2-cell embryos, 4-6-cell embryos, and morulae, there are no significant differences (p > 0.05); namely, -33.3 °C (fresh) vs. -35.4 °C (vitrified) with MII oocytes, -40.6 °C (fresh) vs. -38.7 °C (vitrified) with 2-cell embryos, -38.0 °C (fresh) vs. -39.4 °C (vitrified) with 4-6-cell embryos, -24.5 °C (fresh) vs. -24.2 °C (vitrified) with morulae. But, in 8-cell embryos, there is a significant difference (p < 0.05) between fresh (−37.9 °C) and vitrified (−32.9 °C). If we include this significant difference, the overall IIF temperature of fresh cells is 0.74 °C lower than that of previously vitrified cells. If we exclude it, the IIF temperature for fresh cells is 0.32 °C higher than that for previously vitrified cells. Our conclusion then is that there is no difference between the IIF temperatures of fresh and previously vitrified cells.  相似文献   

20.
Control of choline oxidation in rat kidney mitochondria   总被引:1,自引:0,他引:1  
Niaobh O'Donoghue 《BBA》2009,1787(9):1135-1087
Choline is a quaternary amino cationic organic alcohol that is oxidized to betaine in liver and kidney mitochondria. Betaine acts as an intracellular organic osmolyte in the medulla of the kidney. Evidence is provided that kidney mitochondria have a choline transporter in their inner membrane. The transporter has a Km of 173 ± 64 μM and a Vmax of 0.4 ± 0.1 nmol/min/mg mitochondrial protein (at 10 °C). Uptake of choline is not coupled to betaine efflux. Transporter activity demonstrates a dependence on membrane potential and choline transport is inhibited by hemicholinium-3. Steady-state oxygen consumption due to choline oxidation in kidney mitochondria was measurable at 37 °C (125 ± 6 pmolO2/min/mg mitochondrial protein), in the absence of other mitochondrial electron transport chain substrates and the choline transporter was shown to be the major site of control (96 ± 4%) over choline oxidation flux in isolated kidney mitochondria. We conclude that the choline transporter in rat kidney mitochondria is the major site of control over the production of the organic osmolyte, betaine.  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