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 78 毫秒
1.
Abstract The freshwater Charophyte Chora corallina dies when subjected to 70 molm?3 NaCl if the Ca2+ concentration is 0.1 mol m ?3. This stress is accompanied by a depolarization of the cell to a membrane potential more positive than EK, a net influx of Na+ into the vacuole, and a net loss of K+ from the vacuole. Raising the Ca2+ concentration to 7 mol m ?3 in the presence of elevated Na+ restores the Na+ to Ca2+ ratio to 10: 1 as in the control solution, and results in enhanced survival even though turgor is not regulated. Mg2+ is not a good substitute for Ca2+. It is suggested that the main reason that C. corallina fails to occupy saline habitats is its failure to regulate turgor, not sensitivity to Na +, since the latter is similar to that seen in C. buckellii, which is found in saline habitats.  相似文献   

2.
1. The effect of light intensity on photosynthesis and the fate of newly fixed organic carbon was compared for three characean algae collected at the same depth (10 m) but differing in their depth distributions. For each species we determined photosynthesis–irradiance (P–E) responses, the partitioning of newly fixed carbon into four intracellular pools (low molecular‐weight compounds, polysaccharides, lipids and proteins) and the extracellular organic carbon (EOC) release at a range of photon flux densities (PFD) 0–60 μmol m–2 s–1. 2. The P–E responses differed between the three species, with the light compensation point (Ec) and dark respiration rate highest in the shallowest species (Chara fibrosa), intermediate in the mid‐range species (C. globularis) and lowest in the deepest species (C. corallina). Photosynthetic efficiency (α) and photosynthesis: respiration ratios were lowest in C. fibrosa and highest in C. corallina. 3. In all three species, the low molecular weight pool was the principal photosynthetic product (>60% of fixed C) at 3 μmol m–2 s–1 PFD, but its proportional contribution decreased rapidly with increasing irradiance. Polysaccharide rose to become the major product (>35% of fixed C) at saturating PFD (35 μmol m–2 s–1). 4. Protein synthesis was saturated at 5 μmol m–2 s–1 in all species and was consistently a lower proportion of the fixed carbon in C. corallina than the other species. The fraction incorporated in the lipid pool increased slightly with irradiance but was always less than 10% of fixed C, while the proportion lost as EOC was unaffected by light, being significantly higher in C. fibrosa than the other species. 5. A kinetic experiment with C. fibrosa at 35 μmol m–2 s–1 PFD revealed a continued increase in net polysaccharide, protein and lipid synthesis during a 22.5‐h light period, whereas the net size of the low molecular weight pool remained constant. In a subsequent dark period, protein and lipid synthesis continued at the expense of the polysaccharide and low‐molecular‐weight pools. The EOC release rose to a constant low release in the light, then peaked slightly immediately after the dark–light transition before returning to the same rate as in the light. Extrapolating these data over 24 h suggests that the proportion of fixed carbon lost as EOC may be as high as 10% in this species. 6. The interspecific differences in carbon acquisition between the three species reflected their depth distributions, with the deeper species having more efficient photosynthetic metabolism, lower P:R ratios and less EOC release, although no apparent differences in internal partitioning of photosynthate.  相似文献   

3.
Lipid profiles of three strains (Mexico, Australia, Japan) of Chattonella marina (Subrahmanyan) Hara et Chihara were studied under defined growth (phosphate, light, and growth phase) and harvest (intact and ruptured cells) conditions. Triacylglycerol levels were always <2%, sterols <7%, free fatty acids varied between 2 and 33%, and polar lipids were the most abundant lipid class (>51% of total lipids). The major fatty acids in C. marina were palmitic (16:0), eicosapentaenoic (EPA, 20:5ω3), octadecatetraenoic (18:4ω3), myristic (14:0), and palmitoleic (16:1ω7c) acids. Higher levels of EPA were found in ruptured cells (21.4–29.4%) compared to intact cells (8.5–25.3%). In general, Japanese N‐118 C. marina was the highest producer of EPA (14.3–29.4%), and Mexican CMCV‐1 the lowest producer (7.9–27.1%). Algal cultures, free fatty acids from C. marina, and the two aldehydes 2E,4E‐decadienal and 2E,4E‐heptadienal (suspected fatty acid‐derived products) were tested against the rainbow trout fish gill cell line RTgill‐W1. The configuration of fatty acids plays an important role in ichthyotoxicity. Free fatty acid fractions, obtained by base saponification of total lipids from C. marina showed a potent toxicity toward gill cells (median lethal concentration, LC50 (at 1 h) of 0.44 μg · mL?1 in light conditions, with a complete loss of viability at >3.2 μg · mL?1). Live cultures of Mexican C. marina were less toxic than Japanese and Australian strains. This difference could be related to differing EPA content, superoxide anion production, and cell fragility. The aldehydes 2E,4E‐decadienal and 2E,4E‐heptadienal also showed high impact on gill cell viability, with LC50 (at 1 h) of 0.34 and 0.36 μg · mL?1, respectively. Superoxide anion production was highest in Australian strain CMPL01, followed by Japanese N‐118 and Mexican CMCV‐1 strains. Ruptured cells showed higher production of superoxide anion compared to intact cells (e.g., 19 vs. 9.5 pmol · cell?1 · hr?1 for CMPL01, respectively). Our results indicate that C. marina is more ichthyotoxic after cell disruption and when switching from dark to light conditions, possibly associated with a higher production of superoxide anion and EPA, which may be quickly oxidized to produce more toxic derivates, such as aldehydes.  相似文献   

4.
Carbon transport across the plasma membrane, and carbon fixation were measured in perfused Chara internodal cells. These parameters were measured in external media of pH 5·5 and pH 8·5, where CO2 and HCO3- are, respectively, the predominant carbon species in both light and dark conditions. Cells perfused with medium containing ATP could utilize both CO2 and HCO3- from the external medium in the light. Photosynthetic carbon fixation activity was always higher at pH 5·5 than at pH 8·5. When cells were perfused either with medium containing hexokinase and 2-deoxyglucose to deplete ATP from the cytosol (HK medium) or with medium containing vanadate, a specific inhibitor of the plasma membrane H+-ATPase (V medium), photosynthetic carbon fixation was strongly inhibited at both pH 5·5 and 8·5. Perfusion of cells with medium containing pyruvate kinase and phosphoenolpyruvate (PEP) to maximally activate the H+-ATPase (PK medium), stimulated the photosynthetic carbon fixation activities. Oxygen evolution of isolated chloroplasts and the carbon fixation of cells supplied 14C intracellularly were not inhibited by perfusion media containing either hexokinase and 2-deoxyglucose or vanadate. The results indicate that Chara cells possess CO2 and HCO3- transport systems energized by ATP and sensitive to vanadate in the light. In the dark, intact cells also fix carbon. By contrast, in cells perfused with medium containing ATP, no carbon fixation was detected in 1 mol m -3 total dissolved inorganic carbon (TDIC) at pH 8·5. By increasing TDIC to 10 mol m-3, dark fixation became detectable, although it was still lower than that of intact cells at 1mol m-3 TDIC. Addition of PEP or PEP and PEP carboxylase to the perfusion media significantly increased the dark-carbon fixation. Perfusion with vanadate had no effect on the dark-carbon fixation.  相似文献   

5.
Phosphate uptake in the freshwater charophyte plant Chara corallina was found to be strongly dependent on the presence of Na in the external medium. Based on the reciprocal stimulations of 32Pi uptake by Na and 22Na uptake by Pi, the logical mechanism for Pi uptake appears to be a nNa/Pi symport with a half‐maximal stimulation (Km) for Na of approximately 300 μM and a Km for Pi of approximately 10 μM . Comparison of the stimulations of 32Pi and 22Na influxes at pH 6 gives a stoichiometry of Na : Pi of 5·68. The reduction in Pi influx with increasing pH is consistent with the transported species being the monovalent H2PO4?. In voltage‐clamp experiments, currents elicited by Pi in the presence of Na were equivalent to an influx of positive charge which exceeded the measured influxes of 32P by a factor of 6·26. Intracellular perfusion was used to examine the dependence of Pi influx on ATP and Na. In perfused cells, Pi influx was low when ATP was absent from the internal medium or Na was absent from the external medium. Addition of ATP alone had little effect whereas addition of Na alone increased the 32Pi influx slightly. Addition of both ATP and Na together restored Pi influx to rates comparable to those of intact cells. It is suggested that the ATP is required for membrane hyperpolarization which in turn drives the highly electrogenic flux of Pi with up to 6 Na. However, consideration of the electrochemical potential differences for Na and Pi at pH less than 6 shows that nNa/Pi would not be feasible. It is suggested that at low pH, H+ may substitute for Na.  相似文献   

6.
Summary The presence of a Ca2+ channel in the plasmalemma of tonoplast-freeNitellopsis obtusa cells was demonstrated and its characteristics were studied using current- and voltage-clamp techniques. A long-lasting inward membrane current (I m ), recorded using a step voltage clamp, consisted of a single component without time-dependent inactivation. Increasing either [Ca2+] o or [Cl] o 1) enhanced the maximum amplitude of inwardI m ((I m ) p ) and 2) shifted the peak voltage ((V m ) p ) at(I m ) p to more positive values under ramp-shaped voltage clamping and 3) depolarized the peak value of action potentials. This behavior is consistent with predictions based on the Nernst equation for Ca2+ but not for Cl. DIDS (4,4-diisothiocyano-2,2-disulfonic acid stilbene) did not suppress(I m ) p in tonoplast-free cells, in contrast with its effect on normal cells. La3+ and nifedipine blocked(I m ) p irreversibly. On the other hand, Ca2+ channel agonist, BAY K 8644 irreversibly enhanced(I m ) p . Both Sr2+ influx and K+ efflux increased upon excitation. The charge carried by Sr2+ influx was compensated for by K+ efflux. It is concluded that only the Ca2+ channel is activated during plasmalemma excitation in tonoplast-free cells. In terms of the magnitude of(I m ) p , Sr2+ could replace Ca2+, but Mn2+, Mg2+ and Ba2+ could not. External pH affected(I m ) p and the membrane conductance (g m ) at(I m ) p ((g m ) p ). Increasing the external ionic strength caused increases in both(I m ) p and(g m ) p , and shifted(V m ) p to positive values. At the same time, Sr2+ influx increased. Thus Ca2+ channel activation seems to be enhanced by increasing external ionic strength. The possible involvement of surface potential is discussed.  相似文献   

7.
Summary The effects of various agents on active sodium transport were studied in the toad bladder in terms of the equivalent circuit comprising an active conductanceK a, an electromotive forceE Na, and a parallel passive conductanceK p. For agents which affectK a, but notE Na orK p, the inverse slope of the plot of total conductance against short-circuit currentI 0 evaluatesE Na, and the intercept representsK p. Studies employing 5×10–7 m amiloride to depressK a indicate a changingE Na, invalidating the use of the slope technique with this agent. An alternative suitable technique employs 10–5 m amiloride, which reducesI 0 reversibly to near zero without effect onK p. Despite curvilinearity of the -I0 plot under these conditions,K p may therefore be estimated fairly precisely from the residual conductance. It then becomes possible to follow the dynamic behavior ofK a andE Na (in the absence of 10–5 m amiloride) by frequent measurements of andI 0, utilizing the relationshipsK a=K-K p, andK Na=I O/(K-K p). 2-deoxy-d-glucose (7.5×10–3 m) depressedK a without affectingE Na. Amiloride (5×10–7 m) depressedK a and enhancedE Na. Vasopressin (100 mU/ml) enhancedK a markedly and depressedE Na slightly. Ouabain (10–4 m) depressed bothK a andE Na. All of the above effects were noted promptly;K p was unaffected. The electromotive force of Na transportE Na appears not to be a pure energetic parameter, but to reflect kinetic factors as well, in accordance with thermodynamic considerations.  相似文献   

8.
Growth patterns of larval sardine Sardinops melanostictus were studied in a coastal nursery area, in southern Japan for four monthly hatch cohorts of larvae (November, December, January and February) for the 2003–2004 and 2004–2005 seasons. Laird–Gompertz models were fitted to each cohort using both total length (LT)‐at‐age at capture and mean LT‐at‐age data derived from backcalculations. In both approaches, the absolute daily growth rates (GR) and absolute daily growth rates at the inflection point (GXO) were estimated. In parallel, individual growth rates (GI) were derived from backcalculated LT (LB). Growth showed the following general common patterns irrespective of hatch month, season and methods: (1) significant Laird–Gompertz fits, (2) asymptotic growth, (3) a decrease in GR after the inflexion point, except for February for the 2003–2004 season that showed an apparent constant growth pattern, (4) six in eight cohorts showed GXO ranging from 0·8 to 1·2 mm day?1 and (5) a decreasing tendency of GI from 1·75 to 0·24 mm day?1, from first feeding through the first month of larval life. The contrasting pattern between the 2003–2004 and the 2004–2005 seasons were: (1) allometric v. logarithmic (ln) LT and otolith radius relationships, (2) low GXOv. high GXO, (3) high GRv. low GR when growth turned asymptotic, (4) low GXOv. high GXO when monthly hatch cohorts were combined and (5) LB and GI not differing among monthly hatch cohorts. The differences in growth patterns and growth rates between seasons seemed to be linked to the influx of warmer and oligotrophic waters of the Kuroshio Current that triggered an increase of 3° C in the coastal area for the 2003–2004 seasons. In the overall context, however, the high GXO, within cohorts and seasons reported in the current study, suggests that either sea surface temperature (SST) or food availability, or both are in the optimal range of preferences for S. melanostictus larvae. Consequently, nearshore coastal areas seem to be playing an important role as a nursery area for the larval stage of this species.  相似文献   

9.
Abstract: To examine the possibility that NaF enhances phosphoinositide-specific phospholipase C (PIC) activity in neural tissues by a mechanism independent of a guanine nucleotide binding protein (Gp), we have evaluated the contribution of Gp activation to NaF-stimulated phosphoinositide hydrolysis in human SK-N-SH neuroblastoma cells. Addition of NaF to intact cells resulted in an increase in the release of inositol phosphates (450% of control values; EC50 of ~ 8 mM). Inclusion of U-73122, an aminosteroid inhibitor of guanine nucleotide-regulated PIC activity in these cells, resulted in a dose-dependent inhibition of NaF-stimulated inositol lipid hydrolysis (IC50 of ~ 3.5 μM). When added to digitonin-permeabilized cells, NaF or guanosine-5′-O-thiotriphosphate (GTPγS) resulted in a three- and sevenfold enhancement, respectively, of inositol phosphate release. In the combined presence of optimal concentrations of NaF and GTPγS, inositol phosphate release was less than additive, indicative of a common site of action. Inclusion of 2–5 mM concentrations of guanosine-5′-O-(2-thiodiphosphate) (GDPβS) fully blocked phosphoinositide hydrolysis elicited by GTPγS, whereas that induced by NaF was partially inhibited (65%). However, preincubation of the cells with GDPβS resulted in a greater reduction in the ability of NaF to stimulate inositol phosphate release (87% inhibition). Both GTPγS and NaF-stimulated inositol phosphate release were inhibited by inclusion of 10 μM U-73122 (54–71%). The presence of either NaF or GTPγS also resulted in a marked lowering of the Ca2+ requirement for activation of PIC in permeabilized cells. These results indicate that in SK-N-SH cells, little evidence exists for direct stimulation of PIC by NaF and that the majority of inositol phosphate release that occurs in the presence of NaF can be attributed to activation of Gp.  相似文献   

10.
Early life history traits of young‐of‐the‐year (YOY) round herring Etrumeus teres, caught in Tosa Bay (south‐western Japan), were studied using otolith microstructure analysis for the 2000–2003 year classes. Hatch dates ranged from October to March, and were restricted to either autumn or winter within each year class. YOY ranged from 50 to 123 mm total length (LT) and from 57 to 192 days in age. The relationship of LT to otolith radius was linear. Individual growth rates (GI) were backcalculated between the 70th and 150th days (the size range of most YOY caught) using the biological intercept method. GI ranged from 0·3 to 1·4 mm day?1 and decreased in most cases as season progressed irrespective of year class, although GI in winter cohorts were significantly higher than in autumn cohorts. Otolith growth rates (GO) ranged from 2·13 to 12·25 μm day?1 for autumn spawned YOY and from 3·12 to 12·41 μm day?1 for YOY spawned in winter. The GO trajectories followed three consistent patterns: (1) an increase in increment widths after first feeding through the second week of larval life, then (2) a plateau in increment spacing before increment widths increased again until reaching the maximum growth rate, followed by (3) a gradual decrease in increment widths until the end of the fifth month. The three stages occurred irrespective of spawning season, although YOY spawned in October and December had higher GO during stages (1) and (2) than YOY spawned in February and March, whereas higher GO was observed for late‐winter cohorts in stage (3). Otolith growth from YOY spawned in December and January showed an intermediate pattern between YOY hatched in the early autumn (October to December) and late winter (February to March). The GO trajectories were cross‐matched to the calendar date to estimate time series of otolith growth rates (GOTS) for each year. A parabolic trend was found with maximum GOTS in autumn and spring and minimum values in winter. This trend was significantly correlated to daily sea surface temperature variations. The differences in otolith growth trajectories suggest that the otolith microstructure of E. teres may be used as a natural tag for identifying autumn and winter spawned cohorts.  相似文献   

11.
Thermal denaturation of Na- and Li-DNA from chicken erythrocytes was studied by means of scanning microcalorimetry in salt-free solutions at DNA concentrations (Cp) from 4.5 · 10?2 to 1 · 10?3 moles of nucleotides/liter (M). Linear dependencies of DNA melting temperature (Tm) vs lgCp were obtained: ((1)) ((2)) for Na- and Li-DNA, respectively. Microcalorimetry data were compared with the results of spectrophotometric studies at 260 nm of DNA thermal denaturation in Me-DNA + MeCl solutions at Cp ? (6–8) · 10?5 M and Cs = 0–40 mM (Me is Na or Li, Cs is salt concentration). It was found that Eqs. (1) and (2) are valid in DNA salt-free solutions over the Cp range 6 · 10?5?4.5 · 10?2M. Protonation of DNA bases due to the absorption of CO2 from air in Na-DNA + NaCl solutions affects DNA melting parameters at Cs < 4 mM. Linear dependence of Tm on lga+ is found in Na-DNA + NaCl at Cs > 0.4 mMin the absence of contact of solutions with CO2 from air (a+ is cation activity). A dependence of [dTm/dlga+] on Li+ activity was observed in Li-DNA + LiCl solutions at Cs < 10 mM: [dTm/dlga+] increases from 17°–18° at Cs > 10 mM to 28°–30° at Cs ? 0.2–0.4 mM. Spectrophotometric measurements at 282 nm show that this effect was caused by protonation of bases in fragments of denatured DNA in neutral solutions. The Poisson–Boltzmann (PB) equation was solved for salt-free DNA at the melting point. The linear dependence of Tm vs lgCp was interpreted in terms of Manning's condensation theory. PB and Manning's theories fit the experimental data if charge density parameter (ξ) of denatured DNA is in the range 1.8–2.1 (assuming for native DNA ξ = 4.2). Specificity of Li ions in interactions with DNA is discussed. © 1994 John Wiley & Sons, Inc.  相似文献   

12.
C. Fu    D. Li    W. Hu    Y. Wang  † Z. Zhu   《Journal of fish biology》2007,70(2):347-361
The growth and energy budget for F2‘all‐fish’ growth hormone gene transgenic common carp Cyprinus carpio of two body sizes were investigated at 29·2° C for 21 days. Specific growth rate, feed intake, feed efficiency, digestibility coefficients of dry matter and protein, gross energy intake (IE), and the proportion of IE utilized for heat production (HE) were significantly higher in the transgenics than in the controls. The proportion of IE directed to waste products [faecal energy (FE) and excretory energy loss (ZE+UE) where ZE is through the gills and UE through the kidney], and the proportion of metabolizable energy (ME) for recovered energy (RE) were significantly lower in the transgenics than in the controls. The average energy budget equation of transgenic fish was as follows: 100 IE= 19·3 FE+ 6·0 (ZE+UE) + 45·2 HE+ 29·5 RE or 100 ME= 60·5 HE+ 39·5 RE. The average energy budget equation of the controls was: 100 IE= 25·2 FE+ 7·4 (ZE+UE) + 35·5 HE+ 31·9 RE or 100 ME= 52·7 HE+ 47·3 RE. These findings indicate that the high growth rate of ‘all‐fish’ transgenic common carp relative to their non‐transgenic counterparts was due to their increased feed intake, reduced lose of waste productions and improved feed efficiency. The benefit of the increased energy intake by transgenic fish, however, was diminished by their increased metabolism.  相似文献   

13.
The ability to improve fitness via adaptive evolution may be affected by environmental change. We tested this hypothesis in an in vitro experiment with the plant pathogen Rhizoctonia solani Anastomosis Group 3 (AG-3), assessing genetic and environmental variances under two temperatures (optimal and higher than optimal) and three fungicide concentrations (no fungicide, low and high concentration of a copper-based fungicide). We measured the mean daily growth rate, the coefficient of variation for genotypic (I G) and environmental variance (I E) in growth, and broad-sense heritability in growth. Both higher temperature and increased fungicide concentration caused a decline in growth, confirming their potential as stressors for the pathogen. All types of standardized variances in growth—I G, phenotypic variance, and I E as a trend—increased with elevated stress. However, heritability was not significantly higher under enhanced stress because the increase in I G was counterbalanced by somewhat increased I E. The results illustrate that predictions for adaptation under environmental stress may depend on the type of short-term evolvability measure. Because mycelial growth is linked to fitness, I G reflects short-term evolvability better than heritability, and it indicates that the evolutionary potential of R. solani is positively affected by stress.  相似文献   

14.
Store-operated Ca2+ influx, suggested to be mediated via store-operated cation channel (SOC), is present in all cells. The molecular basis of SOC, and possible heterogeneity of these channels, are still a matter of controversy. Here we have compared the properties of SOC currents (I SOC) in human submandibular glands cells (HSG) and human parotid gland cells (HSY) with I CRAC (Ca2+ release-activated Ca2+ current) in RBL cells. Internal Ca2+ store-depletion with IP3 or thapsigargin activated cation channels in all three cell types. 1 μM Gd3+ blocked channel activity in all cells. Washout of Gd3+ induced partial recovery in HSY and HSG but not RBL cells. 2-APB reversibly inhibited the channels in all cells. I CRAC in RBL cells displayed strong inward rectification with E rev(Ca) = >+90 mV and E rev (Na) = +60 mV. I SOC in HSG cells showed weaker rectification with E rev(Ca) = +25 mV and E rev(Na) = +10 mV. HSY cells displayed a linear current with E rev = +5 mV, which was similar in Ca2+- or Na+-containing medium. pCa/pNa was >500, 40, and 4.6 while pCs /pNa was 0.1,1, and 1.3, for RBL, HSG, and HSY cells, respectively. Evidence for anomalous mole fraction behavior of Ca2+/Na+ permeation was obtained with RBL and HSG cells but not HSY cells. Additionally, channel inactivation with Ca2+ + Na+ or Na+ in the bath was different in the three cell types. In aggregate, these data demonstrate that distinct store-dependent cation currents are stimulated in RBL, HSG, and HSY cells. Importantly, these data suggest a molecular heterogeneity, and possibly cell-specific differences in the function, of these channels.This revised version was published online in June 2005 with a corrected cover date.  相似文献   

15.
The proliferation kinetics of cells of the line NHIK 1922 grown in vitro and as solid tumours in the athymic mutant nude mouse has been studied. In vitro, growth curves were determined for exponentially growing populations and for populations synchronized by mitotic selection. The phase durations for these populations were determined by flow cytofluorometric measurements of DNA-histograms and pulsed incorporation of [3H]TdR respectively. The generation time and the phase durations for synchronized populations were found to be about equal to those for exponentially growing populations. The duration of the phases G1, S and G2+ M was found to be 8·5–9·5, 11·0–12·0 and 6·0–6·5 hr respectively, i.e. the generation time was 26·5–27·0 hr. The proliferation kinetics in vivo were studied by flow cytofluorometry and by the technique of percentage labelled mitoses. The median duration of S-phase and (G2+ M)-phase in vivo was found to be approximately the same as that observed in vitro, while the median duration of G1-phase was found to be approximately 5 hr longer in vivo than under the present in vitro growth conditions. The growth fraction in vivo was estimated to be approximately 50%. The non-proliferative compartment of the tumour cells was found to consist mainly of cells with the DNA-content of cells in G1-phase. It is concluded that the reduced rate of proliferation of NHIK 1922 cells in vivo is correlated with alterations in the duration of G1-phase and, hence, the proportion of cells in G1-phase.  相似文献   

16.
Chaoborus is of great interest to many freshwater ecologists. The adults can become pests in certain areas in North America and the larvae are an important food source for fish. In this preliminary study, we identified variable microsatellite loci in three species: Chaoborus astictopus (HE = 0.52–0.76), Chaoborus americanus (HE = 0.46–0.80) and Chaoborus punctipennis (HE = 0.66–0.81). Using a biotin/streptavidin capture technique of repetitive sequences in a 96‐well format, we obtained microsatellite‐enriched genomic libraries for all three species and identified six polymorphic microsatellite markers for each species. None of the primers did yield a polymerase chain reaction fragment in a cross‐species test.  相似文献   

17.
Photosynthesis and respiration of three Alaskan Porphyra species, P. abbottiae V. Krishnam., P. pseudolinearis Ueda species complex (identified as P. pseudolinearis” below), and P. torta V. Krishnam., were investigated under a range of environmental parameters. Photosynthesis versus irradiance (PI) curves revealed that maximal photosynthesis (Pmax), irradiance at maximal photosynthesis (Imax), and compensation irradiance (Ic) varied with salinity, temperature, and species. The Pmax of Porphyra abbottiae conchocelis varied between 83 and 240 μmol O2 · g dwt?1 · h?1 (where dwt indicates dry weight) at 30–140 μmol photons · m?2 · s?1 (Imax) depending on temperature. Higher irradiances resulted in photoinhibition. Maximal photosynthesis of the conchocelis of P. abbottiae occurred at 11°C, 60 μmol photons · m?2·s?1, and 30 psu (practical salinity units). The conchocelis of P. “pseudolinearis” and P. torta had similar Pmax values but higher Imax values than those of P. abbottiae. The Pmax of P. “pseudolinearis” conchocelis was 200–240 μmol O2 · g dwt?1 · h?1 and for P. torta was 90–240 μmol O2 · g dwt?1 · h?1. Maximal photosynthesis for P. “pseudolinearis” occurred at 7°C and 250 μmol photons · m?2 · s?1 at 30 psu, but Pmax did not change much with temperature. Maximal photosynthesis for P. torta occurred at 15°C, 200 μmol photons · m?2 · s?1, and 30 psu. Photosynthesis rates for all species declined at salinities <25 or >35 psu. Estimated compensation irradiances (Ic) were relatively low (3–5 μmol · photons · m?2 · s?1) for intertidal macrophytes. Porphyra conchocelis had lower respiration rates at 7°C than at 11°C or 15°C. All three species exhibited minimal respiration rates at salinities between 25 and 35 psu.  相似文献   

18.
Radioactive gangliosides, N-[14C]-acetylneuraminylgalactosylglucosylceramide ([14C]GM3) and N- [14C]-acetylneuraminylgalactosyl-N-acetylgalactosaminyl- [N-acetylneuraminyl]-galactosylglucosylceramide ([14C]GD1a), were synthesized from CMP-[14C]sialic acid and the appropriate precursor glycolipid using specific sialyltransferase activities. These compounds were isolated and used as substrates to assay sialidase activity in HeLa cells. Although sodium butyrate added to the culture medium increased GM3 biosynthesis in HeLa cells, sialidase activity, as well as that of other glycohydrolases, was the same in control and butyrate-treated HeLa cells. The same sialidase activity appeared to hydrolyze both [14C]GM3 and [14C]GD1a, but not fetuin; the enzyme had a pH optimum of 5.0 and a Km of 75 μm for the ganglioside substrates. Although the cells contained a high sialidase activity (4–7 nmol/mg of protein/h) and could bind exogenously added [14C]GM3, no “ecto”-sialidase activity would be detected in intact cells under conditions where a close to physiological pH is maintained. The results indicate that ganglioside sialidase is not involved directly in the morphological and biochemical differentiation induced in HeLa cells by exposure to sodium butyrate.  相似文献   

19.
The red seaweed Gracilariopsis is an important crop extensively cultivated in China for high‐quality raw agar. In the cultivation site at Nanao Island, Shantou, China, G. lemaneiformis experiences high variability in environmental conditions like seawater temperature. In this study, G. lemaneiformis was cultured at 12, 19, or 26°C for 3 weeks, to examine its photosynthetic acclimation to changing temperature. Growth rates were highest in G. lemaneiformis thalli grown at 19°C, and were reduced with either decreased or increased temperature. The irradiance‐saturated rate of photosynthesis (Pmax) decreased with decreasing temperature, but increased significantly with prolonged cultivation at lower temperatures, indicating the potential for photosynthesis acclimation to lower temperature. Moreover, Pmax increased with increasing temperature (~30 μmol O2 · g?1FW · h?1 at 12°C to 70 μmol O2 · g?1FW · h?1 at 26°C). The irradiance compensation point for photosynthesis (Ic) decreased significantly with increasing temperature (28 μmol photons · m?2 · s?1 at high temperature vs. 38 μmol photons · m?2 · s?1 at low temperature). Both the photosynthetic light‐ and carbon‐use efficiencies increased with increasing growth or temperatures (from 12°C to 26°C). The results suggested that the thermal acclimation of photosynthetic performance of G. lemaneiformis would have important ecophysiological implications in sea cultivation for improving photosynthesis at low temperature and maintaining high standing biomass during summer. Ongoing climate change (increasing atmospheric CO2 and global warming) may enhance biomass production in G. lemaneiformis mariculture through the improved photosynthetic performances in response to increasing temperature.  相似文献   

20.
We compared autotrophic growth of the dinoflagellate Karlodinium micrum (Leadbeater et Dodge) and the cryptophyte Storeatula major (Butcher ex Hill) at a range of growth irradiances (Eg). Our goal was to determine the physiological bases for differences in growth–irradiance relationships between these species. Maximum autotrophic growth rates of K. micrum and S. major were 0.5 and 1.5 div.·d?1, respectively. Growth rates were positively correlated with C‐specific photosynthetic performance (PPC, g C·g C?1·h?1) (r2=0.72). Cultures were grouped as light‐limited (LL) and high‐light (HL) treatments to allow interspecific comparisons of physiological properties that underlie the growth–irradiance relationships. Interspecific differences in the C‐specific light absorption rate (EaC, mol photons·g C?1·h?1) were observed only among HL acclimated cultures, and the realized quantum yield of C fixation (φC(real.), mol C·mol photons?1) did not differ significantly between species in either LL or HL treatments. The proportion of fixed C that was incorporated into new biomass was lower in K. micrum than S. major at each Eg, reflecting lower growth efficiency in K. micrum. Photoacclimation to HL in K. micrum involved a significant loss of cellular photosynthetic capacity (Pmaxcell), whereas in S. major, Pmaxcell was significantly higher in HL acclimated cells. We conclude that growth rate differences between K. micrum and S. major under LL conditions relate primarily to cell metabolism processes (i.e. growth efficiency) and that reduced chloroplast function, reflected in PPC and photosynthesis–irradiance curve acclimation in K. micrum, is also important under HL conditions.  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