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.
The two-cross technique, a new two-dimensional double-diffusion technique in gelplates, has been applied for simultaneous determination of precipitating titers and diffusion coeffients of antigen and antibody in body fluids. The advantage of this technique is that it works without using any standard solution and ensures conditions of “time-invariant sink”. The theory of the technique has been verified by experimental results on the precipitating system human serum-rabbit anti-human IgG in phosphate-buffered saline solution at pH 7.4. The results obtained using several modes of calculations from experimental parameters have been compared and found satisfactory. The accuracy and reproducibility of the results have been confirmed. It has been found that at 20°C the diffusion coefficient of human IgG in 10-times-diluted serum is (4.4 ± 0.2) × 10?7 cm2 s?1, while the diffusion coefficient of rabbit anti-human IgG in a purified preparation is (2.9 ± 0.2) × 10?7 cm2 s?1. The critical precipitating concentration of human IgG against rabbit anti-human IgG is invariable to concentration and amounts to 0.174 ± 0.03 mg/100 ml at pH 7.4.  相似文献   

2.
We investigated molecular recognition of antibodies to membrane-antigens and extraction of the antigens out of membranes at the single molecule level. Using dynamic force microscopy imaging and enzyme immunoassay, binding of anti-sendai antibodies to sendai-epitopes genetically fused into bacteriorhodopsin molecules from purple membranes were detected under physiological conditions. The antibody/antigen interaction strength of 70-170 pN at loading rates of 2-50 nN/second yielded a barrier width of x = 0.12 nm and a kinetic off-rate (corresponding to the barrier height) of k(off) = 6s(-1), respectively. Bacteriorhodopsin unfolding revealed a characteristic intra-molecular force pattern, in which wild-type and sendai-bacteriorhodopsin molecules were clearly distinguishable in their length distributions, originating from the additional 13 amino acid residues epitope in sendai purple membranes. The inter-molecular antibody/antigen unbinding force was significantly lower than the force required to mechanically extract the binding epitope-containing helix pair out of the membrane and unfold it (126 pN compared to 204 pN at the same loading rate), meeting the expectation that inter-molecular unbinding forces are weaker than intra-molecular unfolding forces responsible for stabilizing native conformations of proteins.  相似文献   

3.
Our study compares the status of human seminal plasma immunoglobulin G (IgG) and IgA secretory component (SC) fucosylation between infertile leukocytospermic and normal, fertile normozoospermic patients. The seminal IgG and SC are decorated with AAL-reactive core fucose, and antennary UEA- and LTA-reactive fucose of Lewisy and Lewisx structures, respectively. However, a correlation between IgG core fucosylation and IgG concentration (r?=??0.52; p?<?0.0003) was observed. The IgG present in leukocytospermic samples is characterized by lower expression of core fucose than in the normal group (0.82?±?0.3 AU and 1.2?±?0.3 AU, respectively; p?<?0.002). In seminal plasma the SC is present in two forms: 78-kDa and 63-kDa. The present study has also shown a higher AAL and LTA specific reactivity of glycans expressed in 63-kDa SC, in comparison to 78-kDa SC, in the normal group. In leukocytospermia, the values of specific lectin reactivity for core fucose, fucose α(1-2)- and α(1-3)- linked, were similar for both SC bands. Moreover, the present study has shown that in leukocytospermic samples the mean concentrations of IgG and S-IgA are twice as high (131.68?±?102.6 mg/l and 36?±?27 mg/l, respectively) as in the normal group (67.68?±?29.2 mg/l; p?<?0.02, and 19?±?18 mg/l, p?<?0.019, respectively). The analysis of IgG and SC fucosylation status and the determination of IgG and S-IgA concentrations in seminal plasma might constitute a valuable diagnosis tools for the evaluation of male infertility associated with leukocytospermia with accompanying inflammation.  相似文献   

4.
Using Langevin modeling, we investigate the role of the experimental setup on the unbinding forces measured in single-molecule pulling experiments. We demonstrate that the stiffness of the pulling device, Keff, may influence the unbinding forces through its effect on the barrier heights for both unbinding and rebinding processes. Under realistic conditions the effect of Keff on the rebinding barrier is shown to play the most important role. This results in a significant increase of the mean unbinding force with the stiffness for a given loading rate. Thus, in contrast to the phenomenological Bell model, we find that the loading rate (the multiplicative value KeffV, V being the pulling velocity) is not the only control parameter that determines the mean unbinding force. If interested in intrinsic properties of a molecular system, we recommend probing the system in the parameter range corresponding to a weak spring and relatively high loading rates where rebinding is negligible.  相似文献   

5.
In the present study we have evaluated the effect of a single hemodialysis session on the brain-derived neurotrophic factor levels in plasma [BDNF]pl and in serum [BDNF]s as well as on the plasma isoprostanes concentration [F2 isoprostanes]pl, plasma total antioxidant capacity (TAC) and plasma cortisol levels in chronic kidney disease patients. Twenty male patients (age 69.8?±?2.9?years (mean?±?SE)) with end-stage renal disease undergoing maintenance hemodialysis on regular dialysis treatment for 15?C71?months participated in this study. A single hemodialysis session, lasting 4.2?±?0.1?h, resulted in a decrease (P?=?0.014) in [BDNF]s by ~42?% (2,574?±?322 vs. 1,492?±?327?pg?ml?1). This was accompanied by an increase (P?<?10?4) of [F2-Isoprostanes]pl (38?±?3 vs. 116?±?16?pg?ml?1), decrease (P?<?10?4) in TAC (1,483?±?41 vs. 983?±?35 trolox equivalents, ??mol?l?1) and a decrease (P?=?0.004) in plasma cortisol level (449.5?±?101.2 vs. 315.3?±?196.3?nmol?l?1). No changes (P?>?0.05) in [BDNF]pl and the platelets count were observed after a single dialysis session. Furthermore, basal [BDNF]s in the chronic kidney disease patients was significantly lower (P?=?0.03) when compared to the age-matched control group (n?=?23). We have concluded that the observed decrease in serum BDNF level after hemodialysis accompanied by elevated [F2-Isoprostanes]pl and decreased plasma TAC might be caused by enhanced oxidative stress induced by hemodialysis.  相似文献   

6.
In this article, a technique for accurate direct measurement of protein‐to‐protein interactions before and after the introduction of a drug candidate is developed using atomic force microscopy (AFM). The method is applied to known immunosuppressant drug candidate Echinacea purpurea derived cynarin. T‐cell/CD28 is on‐chip immobilized and B‐cell/CD80 is immobilized on an AFM tip. The difference in unbinding force between these two proteins before and after the introduction of cynarin is measured. The method is described in detail including determination of the loading rates, maximum probability of bindings, and average unbinding forces. At an AFM loading rate of 1.44 × 104 pN/s, binding events were largely reduced from 61 ± 5% to 47 ± 6% after cynarin introduction. Similarly, maximum probability of bindings reduced from 70% to 35% with a blocking effect of about 35% for a fixed contact time of 0.5 s or greater. Furthermore, average unbinding forces were reduced from 61.4 to 38.9 pN with a blocking effect of ~37% as compared with ~9% by SPR. AFM, which can provide accurate quantitative measures, is shown to be a good method for drug screening. The method could be applied to a wider variety of drug candidates with advances in bio‐chip technology and a more comprehensive AFM database of protein‐to‐protein interactions. Biotechnol. Bioeng. 2012; 109: 2460–2467. © 2012 Wiley Periodicals, Inc.  相似文献   

7.
Summary Atomic force microscopy (AFM) was used to measure the morphology and physicochemical properties of rhizobia and to probe cell-surface polymers with tips modified with soybean agglutinin (SBA). AFM measurements of the length, width, and height of Sinorhizobium fredii CCRC15769 were 1580±450, 870±70, and 270±50 nm, respectively (means±SD). Different AFM image modes revealed the morphology, adhesion, viscoelasticity, and surface roughness of rhizobia in air using the tapping operation. Force–distance relationships between SBA-terminated AFM probes and Bradyrhizobium japonicum USDA110 were recorded and the retraction curves showed an unbinding force of 106±48 pN with a loading rate of 1 nN/s in PBS containing 0.1 mM Mn2+ and 0.1 mM Ca2+ (pH 7.2). The technique of AFM provides information about the morphology and molecular interaction forces of rhizobia under physiological conditions.  相似文献   

8.
Leaf respiration (R L) of evergreen species co-occurring in the Mediterranean maquis developing along the Latium coast was analyzed. The results on the whole showed that the considered evergreen species had the same R L trend during the year, with the lowest rates [0.83 ± 0.43 μmol(CO2) m?2 s?1, mean value of the considered species] in winter, in response to low air temperatures. Higher R L were reached in spring [2.44 ± 1.00 μmol(CO2) m?2 s?1, mean value] during the favorable period, and in summer [3.17 ± 0.89 μmol(CO2) m?2 s?1] during drought. The results of the regression analysis showed that 42% of R L variations depended on mean air temperature and 13% on total monthly rainfall. Among the considered species, C. incanus, was characterized by the highest R L in drought [4.93 ± 0.27 μmol(CO2) m?2 s?1], low leaf water potential at predawn (Ψpd= ?1.08 ± 0.18 MPa) and midday (Ψmd = ?2.75 ± 0.11 MPa) and low relative water content at predawn (RWCpd = 80.5 ± 3.4%) and midday (RWCmd = 67.1 ± 4.6%). Compared to C. incanus, the sclerophyllous species (Q. ilex, P. latifolia, P. lentiscus, A. unedo) and the liana (S. aspera), had lower R L [2.72 ± 0.66 μmol(CO2) m?2 s?1, mean value of the considered species], higher RWCpd (91.8 ± 1.8%), RWCmd (82.4 ± 3.2%), Ψpd (?0.65 ± 0.28 MPa) and Ψmd (?2.85 ± 1.20 MPa) in drought. The narrow-leaved species (E. multiflora, R. officinalis, and E. arborea) were in the middle. The coefficients, proportional to the respiration increase for each 10°C rise (Q10), ranging from 1.49 (E. arborea) to 1.98 (A. unedo) were indicative of the different sensitivities of the considered species to air temperature variation.  相似文献   

9.
There is concern that shear could cause protein unfolding or aggregation during commercial biopharmaceutical production. In this work we exposed two concentrated immunoglobulin‐G1 (IgG1) monoclonal antibody (mAb, at >100 mg/mL) formulations to shear rates between 20,000 and 250,000 s?1 for between 5 min and 30 ms using a parallel‐plate and capillary rheometer, respectively. The maximum shear and force exposures were far in excess of those expected during normal processing operations (20,000 s?1 and 0.06 pN, respectively). We used multiple characterization techniques to determine if there was any detectable aggregation. We found that shear alone did not cause aggregation, but that prolonged exposure to shear in the stainless steel parallel‐plate rheometer caused a very minor reversible aggregation (<0.3%). Additionally, shear did not alter aggregate populations in formulations containing 17% preformed heat‐induced aggregates of a mAb. We calculate that the forces applied to a protein by production shear exposures (<0.06 pN) are small when compared with the 140 pN force expected at the air–water interface or the 20–150 pN forces required to mechanically unfold proteins described in the atomic force microscope (AFM) literature. Therefore, we suggest that in many cases, air‐bubble entrainment, adsorption to solid surfaces (with possible shear synergy), contamination by particulates, or pump cavitation stresses could be much more important causes of aggregation than shear exposure during production. Biotechnol. Bioeng. 2009;103: 936–943. © 2009 Wiley Periodicals, Inc.  相似文献   

10.
Pulse radiolysis-kinetic spectrometry has been used to investigate the reaction of hydrated electrons with ferricytochrome c in dilute aqueous solution at pH 6.5–7.0. Time resolutions from 2·10?7 to 1 s were employed. Transient spectra from 320 to 580 nm were characterized with a wavelength resolution of ±0.5 nm. 1 In neutral salt-free solution, k(ferricytochrome c+e?aq)=(6.0±0.9)·1010 M?1·s?1 and k(ferricytochrome c+H)=(1.2±0.2)·1010 M?1·s?1. The reaction of ferricytochrome c with hydrated electrons is sensitive to ionic strength; in 0.1 M NaClO4, k(ferricytochrome c+e?aq)=(2.4±0.4)·1010 M?1·s?1. In contrast, k(ferricytochrome c+H) is insensitive to ionic strength. Time resolution of three spectral stages has been accomplished. The primary spectrum is the first observable spectrum detectable after irradiation and is formed in a second-order process. Its rate of formation is indisting-uishable from the rate of disappearance of the electron spectrum. The secondary spectrum is generated in a true first order intramolecular process, k(p→s)=(1.2±0.1)·105 s?1. The tertiary spectrum is also generated in a true first-order process, k(s→t)=(1.3±0.2)·102 s?1. The specific rates of both transformations are independent of the wavelength of measurement. The tertiary spectrum, observable 50 ms after initial reaction and remaining unchanged thereafter for at least 1 s, shows that relaxed ferrocytochrome c is the only detectable product. This product is not autoxidizable, as expected for native reduced enzyme. It is more probable that the intramolecular changes responsible for the p→s and s→t spectral transformations involve the influence of conformational relaxation of ferrocytochrome c upon electronic energy states then that they are intramolecular transmission of reducing equivalents from primary sites of electron attachment.  相似文献   

11.
Pentaammineruthenium(III) complexes of deoxyinosine (dIno) and xanthosine (Xao) ([RuIII(NH3)5(L)], L?is?dIno, Xao) in basic solution were studied by UV?Cvis spectroscopy, liquid chromatography/electrospray ionization mass spectrometry, and high-performance liquid chromatography. Both RuIII complexes disproportionate to RuII and RuIV. Disproportionation followed the rate law d[RuII]/dt?=?(k o?+?k 1[OH?])[RuIII]. k o and k 1 of disproportionation at 25?°C were 2.1 (±0.1)?×?10?3?s?1 and 21.4?±?3.2?M?1 s?1, respectively, for [RuIII(NH3)5(dIno)], and 3.5 (±0.7)?×?10?4?s?1 and 59.7?±?3.6?M?1?s?1, respectively, for [RuIII(NH3)5(Xao)]. The [RuIII(NH3)5(Xao)] complex disproportionates at a faster rate than [RuIII(NH3)5(dIno)] owing to the stronger electron-withdrawing effect of exocyclic oxygen in Xao. The activation parameters ??H ? and ??S ? for k 1 of [RuIII(NH3)5(dIno)] were 80.2?±?15.2?kJ?mol?1 and 47.6?±?9.8?J?K?1 mol?1, respectively, indicating that the disproportionation of RuIII to RuII and RuIV is favored owing to the positive entropy of activation. The final products of both complexes in basic solution under Ar were compared with those under O2. Under both conditions [Ru(NH3)5(8-oxo-L)] was produced, but via different mechanisms. In both aerobic and anaerobic conditions, the deprotonation of highly positively polarized C8-H of Ru-L by OH? initiates a two-electron redox reaction. For the next step, we propose a one-step two-electron redox reaction between L and RuIV under anaerobic conditions, which differentiates from Clarke??s mechanism of two consecutive one-electron redox reactions between L, RuIII, and O2.  相似文献   

12.
This study addresses factors governing nitrification and denitrification rates, along with the abundance of the bacterial groups likely involved in these activities, in Kongsfjorden, an Arctic fjord at Ny-Ålesund, Svalbard. The fjord was sampled three times during the month of March 2008 as day length and direct solar radiation increased. Although initially well mixed, cooler and more saline, the fjord became stratified, warmer and less saline during late March. The concentrations of NH4 + (4.4?±?1.6 to 6?±?1.6 μM) and NO2 ? (1?±?0.3 to 1.2?±?0.4 μM) increased progressively with the decrease in NO3 ? (6.1?±?1.3 to 3.8?±?1.5 μM), reflecting the onset of primary productivity. Nitrification rates and the culturable population of nitrifiers decreased significantly from 1.6?±?0.9 to 0.4?±?0.1 ng at NH4 +-N l?1 h?1 and 5.1?±?0.3?×?102 to 29?±?14 cells l?1, respectively. In contrast, denitrification rates increased (2.4?±?0.5 to 4.6?±?1.3 ng-at NO3 ?-N l?1 h?1), although the abundance of culturable denitrifiers did not vary significantly. A significant correlation of nitrifiers with NO3 ? during early March (p?<?0.01, r?=?0.51) indicated that nitrifiers may play an important role in regulating the NO3 ? pool and thereby in controlling the abundance of denitrifiers. However, the contribution of nitrification to the total NO3 ? pool decreased with time. Experimental simulations were also set up to understand the impact of change in duration of light and progressive increase in temperature on these processes. The application of 24 h light inhibited nitrification, suggesting that during peak Arctic summer the contribution of nitrification to the nitrate pool is minimal. It was also observed that a brief exposure to light (≤6 h) was enough to hamper nitrification rates. Experimental simulations suggested that a gradual increase in temperature in the fjord may enhance the magnitude of nitrification and denitrification in the fjord.  相似文献   

13.
In platelets, the glycoprotein (GP) Ib-IX receptor complex senses blood shear flow and transmits the mechanical signals into platelets. Recently, we have discovered a juxtamembrane mechanosensory domain (MSD) within the GPIbα subunit of GPIb-IX. Mechanical unfolding of the MSD activates GPIb-IX signaling into platelets, leading to their activation and clearance. Using optical tweezer-based single-molecule force measurement, we herein report a systematic biomechanical characterization of the MSD in its native, full-length receptor complex and a recombinant, unglycosylated MSD in isolation. The native MSD unfolds at a resting rate of 9 × 10?3 s?1. Upon exposure to pulling forces, MSD unfolding accelerates exponentially over a force scale of 2.0 pN. Importantly, the unfolded MSD can refold with or without applied forces. The unstressed refolding rate of MSD is ~17 s?1 and slows exponentially over a force scale of 3.7 pN. Our measurements confirm that the MSD is relatively unstable, with a folding free energy of 7.5 kBT. Because MSD refolding may turn off GPIb-IX’s mechanosensory signals, our results provide a mechanism for the requirement of a continuous pulling force of >15 pN to fully activate GPIb-IX.  相似文献   

14.
To reduce CO2 emissions from alcoholic fermentation, Arthrospira platensis was cultivated in tubular photobioreactor using either urea or nitrate as nitrogen sources at different light intensities (60 μmol m?2 s?1?≤?I?≤?240 μmol m?2 s?1). The type of carbon source (pure CO2 or CO2 from fermentation) did not show any appreciable influence on the main cultivation parameters, whereas substitution of nitrate for urea increased the nitrogen-to-cell conversion factor (Y X/N ), and the maximum cell concentration (X m ) and productivity (P X ) increased with I. As a result, the best performance using gaseous emissions from alcoholic fermentation (X m ?=?2,960?±?35 g m?3, P X ?=?425?±?5.9 g m?3 day?1 and Y X/N ?=?15?±?0.2 g g?1) was obtained at I?=?120 μmol m?2 s?1 using urea as nitrogen source. The results obtained in this work demonstrate that the combined use of effluents rich in urea and carbon dioxide could be exploited in large-scale cyanobacteria cultivations to reduce not only the production costs of these photosynthetic microorganisms but also the environmental impact associated to the release of greenhouse emissions.  相似文献   

15.
In CD3CN solutions the kinetic parameters characterising rotation about the CNMe2 and CNH2 bonds in [UO2(1,1-DMU)5]2+ (1,1-DMU = 1,1- dimethylurea) were determined as: k(265 K) = 39.1 ± 0.4 and 2960 ± 60 s?1, ΔH3 = 49.1 ± 0.76 and 61.1 ±0.5 kJ mol?1, ΔS2 = ?28.3 ± 2.7 and 53.1 ± 2.2 J K?1 mol?1 respectively from 1H NMR studies. Resonances arising from the three isomeric 1,3-DMU (= 1,3-dimethylurea) ligands were observed for [UO2(1,3-DMU)5]2+ in CD3CN solution and the kinetic parameters characterising their isomerisations were also determined. The three isomers of 1,3-DMU have not previously been detected in solution and it appears that coordination of 1,3-DMU to UO22+ increases the barrier to rotation about the carbon nitrogen bond, as is also shown to be the case for 1,1-DMU.  相似文献   

16.
The semiarid and arid zones cover a quarter of the global land area and support one‐fifth of the world's human population. A significant fraction of the global soil–atmosphere exchange for climatically active gases occurs in semiarid and arid zones yet little is known about these exchanges. A study was made of the soil–atmosphere exchange of CH4, CO, N2O and NOx in the semiarid Mallee system, in north‐western Victoria, Australia, at two sites: one pristine mallee and the other cleared for approximately 65 years for farming (currently wheat). The mean (± standard error) rates of CH4 exchange were uptakes of ?3.0 ± 0.5 ng(C) m?2 s?1 for the Mallee and ?6.0 ± 0.3 ng(C) m?2 s?1 for the Wheat. Converting mallee forest to wheat crop increases CH4 uptake significantly. CH4 emissions were observed in the Mallee in summer and were hypothesized to arise from termite activity. We find no evidence that in situ growing wheat plants emit CH4, contrary to a recent report. The average CO emissions of 10.1 ± 1.8 ng(C) m?2 s?1 in the Mallee and 12.6 ± 2.0 ng(C) m?2 s?1 in the Wheat. The average N2O emissions were 0.5 ± 0.1 ng(N) m?2 s?1 from the pristine Mallee and 1.4 ± 0.3 ng(N) m?2 s?1 from the Wheat. The experimental results show that the processes controlling these exchanges are different to those in temperate systems and are poorly understood.  相似文献   

17.
Synechococcus R-2 (PCC 1942) actively accumulates sulphate in the light and dark. Intracellular sulphate was 1.35 ± 0.23 mol m?3 (light) and 0.894 ± 0.152 mol m?3 (dark) under control conditions (BG-11 media: pHo, 7.5; [SO42?]o, 0.304 mol m?3). The sulphate transporter is different from that found in higher plants: it appears to be an ATP-driven pump transporting one SO42?/ATP [ΔμSO42?i,o=+ 27.7 ± 0.24 kJ mol?1 (light) and + 24 ± 0.34 kj mol?1 (dark)]. The rate of metabolism of SO42?at pHo, 7.5 was 150 ± 28 pmol m?2 s?1 (n = 185) in the light but only 12.8 ± 3.6 pmol m?2 s?1 (n = 61) in the dark. Light-driven sulphate uptake is partially inhibited by DCMU and chloramphenicol. Sulphate uptake is not linked to potassium, proton, sodium or chloride transport. The alga has a constitutive over-capacity for sulphate uptake [light (n= 105): Km= 0.3 ± 0.1 mmol m?3, Vmax, = 1.8 ± 0.6 nmol m?2 s?1; dark (n= 56): Km= 1.4 ± 0.4 mmol m?3, Vmax= 41 ± 22 pmol m?2 s?1]. Sulphite (SO32?) was a competitive inhibitor of sulphate uptake. Selenate (SeO42?) was an uncompetitive inhibitor.  相似文献   

18.
Synechococcus R-2 (PCC 7942) actively accumulated Cl? in the light and dark, under control conditions (BG-11 media: pHo, 7·5; [Na+]o, 18 mol m?3; [Cl?]o, 0·508 molm?3). In BG-11 medium [Cl?], was 17·2±0·848 mol m?3 (light), electrochemical potential of Cl? (ΔμCl?i,o) =+211±2mV; [Cl?]i= 1·24±0·11 mol m?3(dark), ΔμCl?i,o=+133±4mV. Cl? fluxes, but not permeabilities, were much higher in the light: ?Cl?i,o= 4·01±5·4 nmol m?2 s?1, PCl?i,o= 47±5pm s?1 (light); ?Cl?i,o= 0·395±0·071 nmol m?2 s?1, PCl?i,o= 69±14 pm s?1 (dark). Chloride fluxes are inhibited by acid pHo (pHo 5; ?Cl?i,o= 0·14±0·04 nmol m?2 s?1); optimal at pHo 7·5 and not strongly inhibited by alkaline pHo (pHo 10; ?Cl?1i,o= 1·7±0·14 nmol m?2 s?1). A Cl?in/2H+in coporter could not account for the accumulation of Cl? alkaline pHo. Permeability of Cl? is very low, below 100pm s?1 under all conditions used, and appears to be maximal at pHo 7·5 (50–70 pm s?1) and minimal in acid pHo (20pm s?1). DCCD (dicyclohexyl-carbodiimide) inhibited ?Cl?i,o in the light about 75% and [Cl?]i fell to 2·2±0·26 (4) mol m?3. Valinomycin had no effect but monensin severely inhibited Cl? uptake ([Cl?]i= 1·02±0·32 mol m?3; ?Cl?i,o= 0·20±0·1 nmol m?2 s?1). Vanadate (200 mmol m?3) accelerated the Cl? flux (?Cl?i,o= 5·28±0·64 nmol m?2 s?1) but slightly decreased accumulation of Cl? ([Cl?], = 13·9±1·3 mol m?3) in BG-11 medium but had no significant effect in Na+-free media. DCMU (dichlorophenyldimethylurea) did not reduce [Cl?], or ?Cl?i,o to that found in the dark ([Cl?]i= 8·41±0·76 mol m?3; ?Cl?i,o= 2·06±0·36 nmol m?2 s?1). Synechococcus also actively accumulated Cl? in Na+-free media, [Cl?]i was lower but ΔΨi,o hyperpolarized in Na+-free media and so the ΔμCl?i,o was little changed ([Cl?]i= 7·98±0·698 mol m?3; ΔμCl?i,o=+203±3 mV). Net Cl? uptake was stimulated by Na+; Li+ acted as a partial analogue for Na+. Synechococcus has a Na+ activated Cl? transporter which is probably a primary 2Cl?/ATP pump. The Cl? pump is voltage sensitive. ΔμCl?i,o is directly proportional to ΔΨi,o(P»0·01%): ΔμCl?i,o= -1·487 (±0·102) ×ΔΨi,o, r= -0·983, n= 31. The ΔμCl?i,o increased (more positive) as the Δμi,o became more negative. The ΔμCl?i,o has no known function, but might provide a driving force for the uptake of micronutrients.  相似文献   

19.
Biofuels derived from non-crop sources, such as microalgae, offer their own advantages and limitations. Despite high growth rates and lipid accumulation, microalgae cultivation still requires more energy than it produces. Furthermore, invading organisms can lower efficiency of algae production. Simple environmental changes might be able to increase algae productivity while minimizing undesired organisms like competitive algae or predatory algae grazers. Microalgae are susceptible to pH changes. In many production systems, pH is kept below 8 by CO2 addition. Here, we uncouple the effects of pH and CO2 input, by using chemical pH buffers and investigate how pH influences Nannochloropsis salina growth and lipid accumulation as well as invading organisms. We used a wide range of pH levels (5, 6, 7, 8, 9, and 10). N. salina showed highest growth rates at pH 8 and 9 (0.19?±?0.008 and 0.19?±?0.011, respectively; mean ± SD). Maximum cell densities in these treatments were reached around 21 days into the experiment (95.6?×?106?±?9?×?106 cells mL?1 for pH 8 and 92.8?×?106?±?24?×?106 cells mL?1 for pH 9). Lipid accumulation of unbuffered controls were 21.8?±?5.8 % fatty acid methyl esters content by mass, and we were unable to trigger additional significant lipid accumulation by manipulating pH levels at the beginning of stationary phase. Ciliates (grazing predators) occurred in significant higher densities at pH 6 (56.9?±?39.6?×?104 organisms mL?1) than higher pH treatments (0.1–6.8?×?104 organisms mL?1). Furthermore, the addition of buffers themselves seemed to negatively impact diatoms (algal competitors). They were more abundant in an unbuffered control (12.7?±?5.1?×?104 organisms mL?1) than any of the pH treatments (3.6–4.7?×?104 organisms mL?1). In general, pH values of 8 to 9 might be most conducive to increasing algae production and minimizing invading organisms. CO2 addition seems more valuable to algae as an inorganic carbon source and not as an essential mechanism to reduce pH.  相似文献   

20.
Genome methylation plays a key role in regulating gene expression, but limited knowledge exists concerning the link between DNA methylation and economic traits in forest trees. We measured photosynthetic characteristics and growth traits in 130 intraspecific hybrids of Chinese white poplar (Populus tomentosa Carr.) and detected their genome methylation. The phenotypic data were normally distributed, and each trait had a significant difference among the hybrids. The net photosynthetic rate (Pn, 14.83?±?3.76???mol?m?2?s?1), stomatal conductance (Gs, 0.29?±?0.09?mol?m?2?s?1), and intercellular CO2 concentration (Ci, 264.50?±?30.94???mol?mol?1) showed similar trends. Positive correlations were found between Pn and height (H, 133.59?±?50.44?cm) and basal diameter (D, 16.29?±?5.20?mm), respectively. Using methylation-sensitive amplification polymorphism (MSAP) analysis, 32 primer-pair combinations generated 715 polymorphic markers. Positive correlations between photosynthetic characteristics, such as Pn and Gs, and total relative methylation level and relative hemimethylation (CNG methylation) level were investigated. Eighty-one candidate markers were associated with Pn, Gs, or Ci, 13 of which were also associated with growth traits using single MSAP molecular marker association. Sequencing and BLAST analysis showed that candidate markers were linked to genes encoding protochlorophyllide reductase and proteins of cytochrome P450 CYP4/CYP19/CYP26 subfamilies, and linked to genes taking part in, e.g., photosystem II. Therefore, the regions defined by the MSAP candidate markers are linked to genes that are essential for photosynthetic characteristics that respond to DNA methylation and subsequently affect growth traits.  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