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 photosynthesis and growth responses of Sargassum thunbergii germlings to different light intensities (10, 60, and 300 μmol photons m?2 s?1) were investigated. Maximum photochemical efficiency (F v/F m), rapid light curves (RLCs), and photochemical and non-photochemical quenching (qP and NPQ) were estimated by a pulse amplitude-modulated fluorometer. The photosynthesis of S. thunbergii germlings exhibited different properties to optimize light capture and utilization. The excitation pressure (1???qP) was rapidly increased to approximately 0.27 showing that germlings responded to high light by chronic photoinhibition with an accumulation of closed reaction centers, which ultimately resulted in a slow growth. This was accompanied by a reduced F v/F m with time and a development of high capacity for NPQ. Although F v/F m in moderate-light germlings did not fully recover overnight, germlings demonstrated a less severe chronic photoinhibition considering the reduced degree of excitation pressure accumulation of approximately 0.15. The relative stability of photosynthetic capacity (rETRmax, E k, and α) could endow germlings with the highest relative growth rate (RGR) of approximately 9.3 % day?1 in moderate light. By contrast, low-light germlings demonstrated high F v/F m and F o, corresponding high α collectively suggested greater efficiency of light absorption and energy transformation. Sustained increases in electron transport capacity (rETRmax and E k) occurred in low-light germlings, which resulted in a stable RGR of over 8.2 % day?1. Consequently, S. thunbergii germlings are considered to prefer low light regimes and have a relative capacity of moderate and high light tolerance. However, the light acclimation to oversaturating conditions is at the cost of slow growth to maintain survival.  相似文献   

2.
Cement plants account for significant emissions of CO2 and other pollutants into the atmosphere. As a means for its mitigation, we tested the effect of a cement industry-based flue gas simulation (FGS — 18% CO2, 9% O2, 300 ppm NO2, 140 ppm SO2) on the green alga, Chlorella sorokiniana. Culture pH, cell density, cell viability and productivity, specific growth rates, photosynthetic performance, and biochemical composition were monitored. The treatments consisted of different FGS volumes (0.1, 0.3, 0.8, 1.5, 6, and 48 L day?1) that were applied in a series of laboratory-scale semi-continuous batch cultures under controlled conditions. Controls were exposed to 18% CO2 enriched air. Cell density showed that C. sorokiniana was able to grow in all treatments, but compared to the controls, low pH (~ 5.0) caused by 48 L FGS day?1 led to 27% decrease in specific growth rate. Increasing FGS exposure decreased maximum and operational quantum yields obtained by pulse amplitude modulated fluorometry, while photochemical quenching remained constant (~ 0.93). The α and rETR max parameters calculated from rapid light curves decreased with increasing FGS exposure. Total proteins and carbohydrates (per cell basis) increased after 6 and 48 L FGS day?1, which can be advantageous for biotechnological applications, but cell productivity (cells L?1 day?1) decreased. Despite the effects in physiology, C. sorokiniana could withstand a pH range of 6.0–5.0 imposed by 48 L FGS day?1. Overall, C. sorokiniana can be considered a robust species in flue gas bioremediation.  相似文献   

3.
This research examined the application of the maximum quantum efficiency (F v/F m) and relative electron transport rate versus irradiance curves (rETR) as a rapid, sensitive assessment of Lake Erie phytoplankton nutrient status. I evaluated the potential benefits of the variable fluorescence parameters by comparing these parameters with chemical and physiological nutrient status assays. I tested the hypothesis that F v/F m and rETR curves could diagnose nutrient status in natural lake phytoplankton and be capable of discriminating which inorganic nutrient is limited temporally and spatially. F v/F m was on average highest in the more eutrophic west basin (WB) and lowest in the more oligotrophic central basin (CB). According to the chemical and physiological indicators, P deficiency was most severe in the CB during summer stratification and N deficiency was strongest in the WB during isothermal conditions. Like F v/F m, rETR at light saturation (rETRmax) and the initial slope of the rETR versus irradiance curve (α) decreased as the severity of N and P deficiency increased. Amendment with N or P stimulated increased F v/F m, rETRmax, and α in N- and P-limited samples, respectively, and abolished the photoinhibition apparent in rETR curves of nutrient-limited samples. These results supported the view that the N and P deficiency assays, and corresponding variations of variable fluorescence parameters, were valid indicators of widely variable N and P deficiency in the phytoplankton, and could be used to provide a promising tool in determining phytoplankton nutrient status. Contrary to my hopes, it did not appear that rETR–irradiance curves could discriminate between N and P deficiency. Identification of the most limiting nutrient still demanded additional information beyond the variable fluorescence measurements.  相似文献   

4.
Ligand exchange reactions of oxorhenium(V) precursors with bidentate SN and tridentate Schiff bases derived from the condensation of ketones or aldehydes with dithiocarbazic acid methyl ester (H2NNHC(S)SCH3) produce novel ‘3+2’ mixed-ligand complexes carrying the SNO/SN donor atom set. Thus, reactions of either [NBu4][ReOCl4] or Na[ReO(Gluconate)2] with SNO ligands (H2Ln) or a mixture of bidentate SN (HLm) and tridentate SNO (H2Ln) in methanol solutions lead, respectively, to the six-coordinated mixed ligand oxorhenium(V) compounds of types [ReO(Ln)(HLn)] and [ReO(Ln)(Lm)], combining one tridentate dianionic SNO donor Schiff base (L) and one bidentate anionic SN donor ligand (HL). Coordination geometry around rhenium is distorted octahedral with the two SN donor atom sets of each ligand defining the equatorial plane, while apical positions are occupied by the oxo group and the oxygen atom of the tridentate SNO ligand (L), as shown by single-crystal X-ray diffraction structure of [ReO(L1)(HL1)] 1.  相似文献   

5.
A protocol for in vitro propagation of cineraria (Senecio cruentus) was developed. The highest frequency of shoot proliferation was obtained from nodal explants cultured on Murashige and Skoog (MS) medium supplemented with 2.0?mg L?1 6-benzyladenine (BA) and 0.5?mg L?1 ??-naphthalene acetic acid (NAA), with a mean number of 14 shoots per explant. A high concentration of BA (4.0?mg L?1) and repeated subcultures resulted in hyperhydric shoots. Decreasing the BA concentration to 1.0?mg L?1 in the culture medium eliminated hyperhydricity. The concentration of ammonium nitrate (NH4NO3) and temperature had marked effects on somaclonal variation. Variation was observed when the cultures were maintained at 15?°C but not at 25?°C. Variants with blue-colored leaves and stems were identified; whereas, normal plants maintained their green-colored leaves and stems. The highest frequency of variation (67.5?%), with a mean number of 3.0 variant shoots per explants, was obtained on shoot proliferation medium (MS?+?2.0?mg L?1 BA and 0.5?mg L?1 NAA) devoid of NH4NO3. The best rooting (100?%), with the highest number of roots per shoot (10.8) and the greatest root length (6.8?cm) was obtained on medium supplemented with 0.1?mg L?1 NAA. In vitro-grown plantlets were successfully acclimatized in a greenhouse, and transferred to the field.  相似文献   

6.
There has been considerable interest in cultivation of green microalgae (Chlorophyta) as a source of lipid that can alternatively be converted to biodiesel. However, almost all mass cultures of algae are carbon-limited. Therefore, to reach a high biomass and oil productivities, the ideal selected microalgae will most likely need a source of inorganic carbon. Here, growth and lipid productivities of Tetraselmis suecica CS-187 and Chlorella sp were tested under various ranges of pH and different sources of inorganic carbon (untreated flue gas from coal-fired power plant, pure industrial CO2, pH-adjusted using HCl and sodium bicarbonate). Biomass and lipid productivities were highest at pH 7.5 (320?±?29.9 mg biomass L?1 day?1and 92?±?13.1 mg lipid L?1 day?1) and pH 7 (407?±?5.5 mg biomass L?1 day?1 and 99?±?17.2 mg lipid L?1 day?1) for T. suecica CS-187 and Chlorella sp, respectively. In general, biomass and lipid productivities were pH 7.5?>?pH 7?>?pH 8?>?pH 6.5 and pH 7?>?pH 7.5?=?pH 8?>?pH 6.5?>?pH 6?>?pH 5.5 for T. suecica CS-187 and Chlorella sp, respectively. The effect of various inorganic carbon on growth and productivities of T. suecica (regulated at pH?=?7.5) and Chlorella sp (regulated at pH?=?7) grown in bag photobioreactors was also examined outdoor at the International Power Hazelwood, Gippsland, Victoria, Australia. The highest biomass and lipid productivities of T. suecica (51.45?±?2.67 mg biomass L?1 day?1 and 14.8?±?2.46 mg lipid L?1 day?1) and Chlorella sp (60.00?±?2.4 mg biomass L?1 day?1 and 13.70?±?1.35 mg lipid L?1 day?1) were achieved when grown using CO2 as inorganic carbon source. No significant differences were found between CO2 and flue gas biomass and lipid productivities. While grown using CO2 and flue gas, biomass productivities were 10, 13 and 18 %, and 7, 14 and 19 % higher than NaHCO3, HCl and unregulated pH for T. suecica and Chlorella sp, respectively. Addition of inorganic carbon increased specific growth rate and lipid content but reduced biomass yield and cell weight of T. suecica. Addition of inorganic carbon increased yield but did not change specific growth rate, cell weight or content of the cell weight of Chlorella sp. Both strains showed significantly higher maximum quantum yield (Fv/Fm) when grown under optimum pH.  相似文献   

7.
A 14-membered tetraaza macrocycle, 2,13-bis(2-carbomethoxyethyl)-5,16-dimethyl-2,6,13,17-tetraazatricyclo[16.4.0.1.1807.12]docosane (L2) bearing two N-CH2CH2COOMe groups, and its nickel(II) and copper(II) complexes have been prepared and characterized. The nickel(II) and copper(II) complexes of 2-(2-carbomethoxyethyl)-5,16-dimethyl-2,6,13,17-tetraazatricyclo[16.4.0.1.1807.12]docosane (L3) containing one N-CH2CH2COOMe group have also been prepared. The crystal structure of [NiL2](ClO4)2 shows that the complex has a slightly distorted trans-octahedral coordination geometry with two relatively short axial Ni-O (N-CH2CH2COOMe group) bonds (2.136(3) Å). In various solvents, however, a considerable proportion of [NiL2]2+ exists as a square-planar form, in which the functional pendant arms are not involved in coordination. The proportion of the square-planar isomer varies with solvents in the order of nitromethane ? acetonitrile < H2O < DMF ? DMSO. In the case of [CuL2](ClO4)2, only one N-CH2CH2COOMe group is involved in coordination. The N-CH2CH2COOMe group of [NiL3](ClO4)2 is not directly involved in coordination even in the solid state, though the functional group of [CuL3](ClO4)2 is coordinated to the metal ion.  相似文献   

8.
High salinity wastewaters have limited treatment options due to the occurrence of salt inhibition in conventional biological treatments. Using recirculating marine aquaculture effluents as a case study, this work explored the use of Constructed Wetlands as a treatment option for nutrient and salt loads reduction. Three different substrates were tested for nutrient adsorption, of which expanded clay performed better. This substrate adsorbed 0.31 mg kg?1 of NH4 +?N and 5.60 mg kg?1 of PO4 3??P and 6.9 mg kg?1 dissolved salts after 7 days of contact. Microcosms with Typha latifolia planted in expanded clay and irrigated with aquaculture wastewater (salinity 2.4%, 7 days hydraulic retention time, for 4 weeks), were able to remove 94% NH4 +?N (inlet 0.25 ± 0.13 mg L?1), 78% NO2 ??N (inlet 0.78 ± 0.62 mg L?1), 46% NO3 ??N (inlet 18.83 ± 8.93 mg L?1) whereas PO4 3??P was not detected (inlet 1.41 ± 0.21 mg L?1). Maximum salinity reductions of 52% were observed. Despite some growth inhibition, plants remained viable, with 94% survival rate. Daily treatment dynamics studies revealed rapid PO4 3??P adsorption, unbalancing the N:P ratio and possibly affecting plant development. An integrated treatment approach, coupled with biomass valorization, is suggested to provide optimal resource management possibilities.  相似文献   

9.
New molybdenum complexes were prepared by the reaction of [MoVIO2(acac)2] or (NH4)2[MoVOCl5] with different N-substituted pyridoxal thiosemicarbazone ligands (H2L1 = pyridoxal 4-phenylthiosemicarbazone; H2L2 = pyridoxal 4-methylthiosemicarbazone, H2L3 = pyridoxal thiosemicarbazone). The investigation of monomeric [MoO2L1(CH3OH)] or polymeric [MoO2L1-3] molybdenum(VI) complexes revealed that molybdenum is coordinated with a tridentate doubly-deprotonated ligand. In the oxomolybdenum(V) complexes [MoOCl2(HL1-3)] the pyridoxal thiosemicarbazonato ligands are tridentate mono-deprotonated. Crystal and molecular structures of molybdenum(VI) [MoO2L1(CH3OH)]·CH3OH, and molybdenum(V) complexes [MoOCl2(HL1)]·C2H5OH, as well as of the pyridoxal thiosemicarbazone ligand methanol solvate H2L3·MeOH, were determined by the single crystal X-ray diffraction method.  相似文献   

10.
Reactions of chloro(3-thiapentane-1,5-dithiolato)oxorhenium(V) [ReO(SSS)Cl] with N-methyl-1H-imidazole-2-thiol (HL1) and 2-pyrimidinethiol (HL2) have been studied to form ‘3+1’ oxorhenium(V) complexes. In the absence of triethylamine, [Re(SSS)(HL1)]Cl (1a) was formed, while in the presence of triethylamine [Re(SSS)L1] (1b) and [Re(SSS)L2] (2) were produced. Molecular structures of complexes 1a and 2 were determined to be distorted square pyramidal by single crystal X-ray analytical method. From cyclic voltammetric studies, furthermore, it was proposed that complexes 1b and 2 are irreversibly oxidized to Re(VI) at around 0.84 and 1.01 V versus Ag/AgNO3, respectively, and are reduced to Re(IV) at −1.55 and −1.51 V with the dissociation of L1 or L2, followed by the quasi-reversible reductions to Re(III) at around −1.69 V, respectively.  相似文献   

11.
The Pd(II) and Pt(II) complexes with triazolopyrimidine C-nucleosides L1 (5,7-dimethyl-3-(2′,3′,5′-tri-O-benzoyl-β-d-ribofuranosyl-s-triazolo)[4,3-a]pyrimidine), L2 (5,7-dimethyl-3-β-d-ribofuranosyl-s-triazolo[4,3-a]pyrimidine) and L3 (5,7-dimethyl[1,5-a]-s-triazolopyrimidine), [Pd(en)(L1)](NO3)2, [Pd(bpy)(L1)](NO3)2, cis-Pd(L3)2Cl2, [Pd2(L3)2Cl4] · H2O, cis-Pd(L2)2Cl2 and [Pt3(L1)2Cl6] were synthesized and characterized by elemental analysis and NMR spectroscopy. The structure of the [Pd2(L3)2Cl4] · H2O complex was established by X-ray crystallography. The two L3 ligands are found in a head to tail orientation, with a Pd?Pd distance of 3.1254(17) Å. L1 coordinates to Pd(II) through N8 and N1 forming polymeric structures. L2 coordinates to Pd(II) through N8 in acidic solutions (0.1 M HCl) forming complexes of cis-geometry. The Pd(II) coordination to L2 does not affect the sugar conformation probably due to the high stability of the C-C glycoside bond.  相似文献   

12.
The rate of phosphate hydrolysis of ATP in the substitution-inert complex Co(NH3)4ATP-has been examined in the presence and absence of [Co(cyclen)(H2O)2]3+. The rate of hydrolysis of Co(NH3)4ATP- in the absence of [Co(cyclen)(H2O)2]3+ is essentially independent of pH in the range 6.0 to 9.0, and the rate constant is 2.6 × 10?5 sec ?1 at pH 9.0, 40°C, and 1.0 M ionic strength Rate constants for the hydrolysis of Co(NH3)4ATP- in the presence of [Co(cyclen)(H2O)2]3+ are sharply dependent upon pH in the same range. The rate constants at pH 8.0, 8.6, and 9.0 are 8, 63, and 95 times larger than the rate constant at pH 7.0. At pH 9 the rate constant is 1.2 × 10?3 sec?1 for 16 mM Co(NH3)4ATP- in the presence of 10 mM [Co(cyclen)(H2O)2]3+. The proposed mechanism for hydrolysis involves the coordination of a phosphate group of Co(NH3)4ATP- by [Co(cyclen)(H2O)2]3+ to form a dinuclear species, followed by internal attack of coordinated hydroxide on the phosphate chain.  相似文献   

13.
Although sea‐ice represents a harsh physicochemical environment with steep gradients in temperature, light, and salinity, diverse microbial communities are present within the ice matrix. We describe here the photosynthetic responses of sea‐ice microalgae to varying irradiances. Rapid light curves (RLCs) were generated using pulse amplitude fluorometry and used to derive photosynthetic yield (ΦPSII), photosynthetic efficiency (α), and the irradiance (Ek) at which relative electron transport rate (rETR) saturates. Surface brine algae from near the surface and bottom‐ice algae were exposed to a range of irradiances from 7 to 262 μmol photons · m?2 · s?1. In surface brine algae, ΦPSII and α remained constant at all irradiances, and rETRmax peaked at 151 μmol photons · m?2 · s?1, indicating these algae are well acclimated to the irradiances to which they are normally exposed. In contrast, ΦPSII, α, and rETRmax in bottom‐ice algae reduced when exposed to irradiances >26 μmol photons · m?2 · s?1, indicating a high degree of shade acclimation. In addition, the previous light history had no significant effect on the photosynthetic capacity of bottom‐ice algae whether cells were gradually exposed to target irradiances over a 12 h period or were exposed immediately (light shocked). These findings indicate that bottom‐ice algae are photoinhibited in a dose‐dependent manner, while surface brine algae tolerate higher irradiances. Our study shows that sea‐ice algae are able to adjust to changes in irradiance rapidly, and this ability to acclimate may facilitate survival and subsequent long‐term acclimation to the postmelt light regime of the Southern Ocean.  相似文献   

14.
Complexes of FeCl2 with the known bis(3-methyl-2-thione-imidazolyl)methane (L1) and the new bis(3-tert-butyl-2-thione-imidazolyl)methane (L2) are reported. For both [L1FeCl2]n (3) and [L2FeCl2]n (4) X-ray crystallography reveals that 1D-polymeric chain structures are present in the solid state, with the two mercaptoimidazolyl units of L1 and L2 coordinating to different metal ions. Complexes 3 and 4 are further characterized by Mössbauer spectroscopy and SQUID magnetometry. NMR spectroscopy suggests that the complexes largely dissociate in polar solvents. X-ray structures of L2 and its precursor bis(imidazolium) salt are also reported.  相似文献   

15.
Malus baccata is widely used as a rootstock in cold regions of the world because of its cold hardiness. In this study, a highly efficient Agrobacterium rhizogenes strain 8196 transformation system was developed using in vitro-derived stem segments of M. baccata. Approximately 37?% agro-infected explants produced hairy roots when they were incubated on Murashige and Skoog (MS) medium without plant growth regulators. A total of 95?% of hairy roots exhibited glucuronidase activity. Calli were induced from putatively-transformed hairy roots, and subsequently shoots were observed within 4?weeks of culture. The influence of 6-benzyladenine (BA), indole-3-butyric acid (IBA), thidiazuron (TDZ), and gibberellic acid 3 (GA3) on regeneration were investigated using an L9 (34) orthogonal experiment. About 73?% of shoots were regenerated when callus was incubated on MS medium along with 2.0?mg?L?1 BA, 0.5?mg?L?1 IBA, 0.3?mg?L?1 GA3, and 0.5?mg?L?1 TDZ. Moreover, hairy root regenerants showed higher rooting ability and exhibited morphological aberrations such as shortened stem, etiolated, wrinkled and clustered leaves than those of control.  相似文献   

16.
The effects of five (5 000, 10 000, 15 000, 20 000, 24 000 kg ha?1 year?1) different doses of organic fertilizer (cow dung) were studied on pond productivity in terms of plankton production and fish biomass in freshwater fish ponds. The grow out period was 60 days. Physico-chemical factors of pond waters were also monitored. With an increase in the fertilizer dose, biochemical oxygen demand (BOD) (1.7 ± 0.1 – 10.35 ± 0.05 mg L?1), O-PO4 (0.04 ± 0.0 – 0.77 ± 0.02 mg L?1) and NH4-N (0.03 ± 0.02 – 0.32 ± 0.02 mg L?1) increased significantly (P < 0.05). Alkalinity (79.0 ± 1.6 – 164.0 ± 3.8 mg L?1) also increased with the increase in fertilizer dose, declining after 60 and 75 days (48.8 ± 1.13 – 67.9 ± 2.1 mg L?1). NO3-N was maximum (1.66 ± 0.2 mg L?1) in the ponds which received cow dung at 15 000 kg ha?1 year?1, and declined (0.94 ± 0.5 mg L?1) at higher doses. Dissolved oxygen (DO) remained significantly high (4.7 mg L?1) up to the third (15 000 kg ha?1 year?1) treatment. Highest plankton population (phytoplankton 17 350.0 ± 1 250.0 no L?1), zooplankton (373.0 ± 22.0 no L?1), species diversity (phytoplankton 3.0, zooplankton 2.3), fish biomass (4.45 kg) and specific growth rate (SGR) (2.36 % body weight (BW) d?1) were also observed in ponds which were treated with fertilizer at 15 000 kg ha?1 year?1. However, at higher doses, a decline in these parameters (phytoplankton, 0.0 – 8 810.0 ± 690.0 no L?1; zooplankton, 0.0 – 205.0 ± 25.0 no L?1; fish biomass, 2.3 kg; SGR, 1.25 % body weight (BW) d?1) was observed. Furthermore, with a decrease in the water temperature from 24 °C (on day 60) to 21 °C (on day 75), a decline in nutrient release, plankton population L?1 and species diversity was observed. Sediment analysis indicated that with an increase in the fertilizer dosage, a significant and progressive increase in the accumulation of organic carbon (0.787 ± 0.006 – 0.935 ± 0.01), total nitrogen (0.877 ± 0.071 – 1.231 ± 0.03), NH4-N (54.4 ± 0.57 – 68.95 ± 0.81), NO3-N (78.5 ± 1.21 – 98.5 ± 0.35), total P (140.0 ± 0.50 – 151.0 ± 1.27) and soluble P (7.15 ± 0.18 – 10.1 ± 0.56) took place; similarly, electrical conductivity (EC) values of sediment also increased progressively (from 200.0 ± 7.1–300.0 ± 10.63 μ mhos cm?1).  相似文献   

17.
Imidazole-2-thiol derivatives H2L1-3 (H2L1 = 1H-benzoimidazole-2-thiol, H2L2 = 5-methyl-1H-benzoimidazole-2-thiol, and H2L3 = 1H-imidazole-2-thiol) act as neutral monodentate ligands in a number of technetium and rhenium complexes. Disubstituted M(V) (M = Tc, Re) complexes of the type [AsPh4]{[MOCl2(H2Ln)2(H2O)]Cl2} are formed when [MOCl4] react with H2L1-3 in 1:2 stoichiometric ratio. Single crystal X-ray structure determinations were carried out on [AsPh4]{[TcOCl2(H2L1)2(H2O)]Cl2}. The coordination sphere is pseudo-octahedral in which the sulfur atoms of two ligands sit in the equatorial plane and a water molecule is in trans to the TcO multiple bond. All the complexes react with an excess of the corresponding ligand to form tetrasubstituted cationic species {[MO(H2Ln)4]Cl3}. These complexes can be also isolated by reaction of [MOCl4] with an excess of ligand. No complex is obtained with benzothiazole-2-thiol (HL4) and benzoxazole-2-thiol (HL5). Ligand exchange reactions of [ReOCl3(PPh3)2] with HL4,5 have also been investigated. Treating the oxo-precursor with HL4 no product is isolated, while with HL5 the chelate oxo-compound [ReOCl2(L5)(PPh3)] is formed as two isomers. An interesting organometallic complex of Re(IV) [ReCl3(L5∗)(PPh3)2] is obtained when a slight excess of HL5 reacts with [ReOCl3(PPh3)2] in refluxing benzene solution and in air. Geometry about the Re atom is approximately octahedral in which the equatorial plane contains three Cl atoms and the carbon atom of the benzoxazole ligand anion, the apical positions are occupied by two PPh3. The reaction with O-ethyl S-hydrogen p-tolyl carbonothioimidate HL6 which contains the same heteroatoms of HL5 does not form an organometallic species, but forms the chelate oxo-Re(V) complex [ReOCl2(L6)(PPh3)]. The solid-state structure has been authenticated by X-ray crystallography.  相似文献   

18.
Reactions of ligands 1-ethyl-5-methyl-3-phenyl-1H-pyrazole (L1) and 5-methyl-1-octyl-3-phenyl-1H-pyrazole (L2) with [PdCl2(CH3CN)2 and K2PtCl4 gave complexes trans-[MCl2(L)2] (L = L1, L2). The new complexes were characterised by elemental analyses, conductivity measurements, infrared, 1H and 13C{1H} NMR spectroscopies and X-ray diffraction. The NMR study of the complex [PdCl2(L1)2], in CDCl3 solution, is consistent with a very slow rotation of ligands around the Pd-N bond, so that two conformational isomers can be observed in solution (syn and anti). Different behaviour is observed for complexes [PdCl2(L2)2] and [PtCl2(L)2] (L = L1, L2), which present an isomer in solution at room temperature (anti). The crystal structure of [PdCl2(L1)2] complex is described, where the Pd(II) presents a square planar geometry with the ligands coordinated in a trans disposition.  相似文献   

19.
Rates of net photosynthesis and respiration were determined for Pithophora oedogonia (Mont.) Wittr. acclimatized to 56 combinations of light (7–1200 μE m?2 s?1) and temperature (5–35°C). Conditions for maximum net photosynthesis were estimated to be 26°C and 970 μE m?2 s?1. The rate of net photosyntheses varied considerably with temperature, with the maximum measured value (9.67 mg O2 h?1 g dry wt.?1) occurring at 25°C. Respiration rate increased with temperature and the light received just prior to measurement. The maximum respiration rate (7.05 mg O2 g?1 h?1) occurred at 30°C and 1200 μE m?2 s?1. Exposure of Pithophora to light levels of 600 or 1200 μE m?2 s?1 prior to determination of the respiration rate resulted in significantly elevated levels of oxygen consumption at temperatures ≥ 15°C. The relationship between light, temperature and photosynthesis and respiration were summarized as three-dimensional response surfaces.  相似文献   

20.
This plant tissue study of micropropagation identifies the selective medium saving for rapid propagation in cultivated Thermopsis turcica, an endangered germplasm of the family Fabaceae. The aim is to obtain the optimum growth medium of T. turcica by enabling the in vitro propagation of this endemic. In this study, the leaves and stems of T. turcica were cultured on a Murashige and Skoog’s medium supplemented with various concentrations (0.5, 1.0 and 2.0 mg L?1 1-Naphthaleneacetic acid) of auxin and (0.2 and 0.5 mg L?1 Zeatin) (1.0 and 2.0 mg L?1 Benzylaminopurine) of cytokinins. Previous research focused on the regeneration from the seed of T. turcica Eber population; we concentrated upon the regeneration of different plant parts (leaf and stem) of T. turcica Aksehir population. In addition, according to the literature on T. turcica that to date the effects of Zeatin on the regeneration has not been performed. The most promising regeneration and growth were obtained from leaf explants cultured on the media with 2.0 mg L?1 1-Naphthaleneacetic acid and 0.5 mg L?1Zeatin (93.3%). The regenerated plantles were rooted on the media containing 2.0 mg L?1 Indole-3-butyric acid. Rooted plantlets were transplanted into potting of sterilized soil. The present study reports on the sufficient in vitro regeneration protocol through organogenesis in T. turcica. The findings presented here have implications for in vitro protection and use of this endemic endangered species in further biotechnological research.  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