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 140 毫秒
1.
The complex Pt(bph) (CO)2 crystallizes in the space group Cmcm with a = 18.647(6), B = 9.566(2) and C = 6.4060(5) Å. The geometry of the molecule is slightly distorted from square planar with a Pt---C(CO) bond distance of 1.98(2) Å and a Pt---C(bph) bond distance of 2.04(2) Å. The Pt(bph)(CO)2 complex serves as a precursor for the preparation of a wide variety of Pt(bph)X2 complexes, where X = monodentate ligands such as acetonitrile, pyridine, etc., and X2 = bidentate ligands such as bypyridine, 1,10-phenanthroline, etc. In the solid state, the complex exhibits a green color, but when ground with an alkali metal salt turns deep blue to purple. In CH2Cl2, the color disappears but optical transitions are observed at 271 nm (2.7 × 104 M−1 cm−1), 303 nm (1.1 × 104 M−1 cm−1) and 330 nm (5.5 × 103 M−1 cm−1). The complex is a weak emitter exhibiting a structured spectrum in CH2Cl2 at r.t. with maxima located at 562 and 594 nm and an emission lifetime of 3.1 μs when excited at 337 nm.  相似文献   

2.
Electron self-exchange in solutions of the ‘blue’ copper protein plastocyanin is catalysed by the redox-inert multivalent cations Mg2+ or Co(NH3)3+6. Measurements of specific 1H-NMR line broadening with 50% reduced solutions in the presence of these cations show that electron exchange proceeds through encounters of cation-protein complexes which dissociate at high ionic strength. In the presence of 8mM (5 equivalents/total protein) Co(NH3)3+6, with 10 mM cacodylate (pH*6.0) as background electrolyte, the bimolecular rate constant at 25°C is 7 × 104 M−1·s−1. For comparison, the ‘electrostatically screened’ rate constant measured in 0.1 M KCl in the absence of added multivalent cations is ˜ 4 × 103 M1·s−1.

Plastocyanin Electron self-exchange NMR Protein-protein interaction Multivalent cation Blue copper protein  相似文献   


3.
The syntheses and structures of [Ni(H2O)6]2+[MF6]2− (M = Ti,Zr,Hf) and Ni3(py)12F6·7H2O are reported. The former three compounds are isostructural, crystallizing in the trigonal space group (No. 148) with Z = 3. The lattice parameters are a = 9.489(4), C = 9.764(7) Å, with V = 761(1) Å3 for Ti; a = 9.727(2), C = 10.051(3) Å, with V = 823.6(6) Å3 for Zr; and a = 9.724(3), C = 10.028(4)Å, with V = 821.2(8)Å3 for Hf. The structures consist of discrete [Ni(H2O)6]2+ and [MF6]2− octahedra joined by O---HF hydrogen bond Large single crystals were grown in an aqueous hydrofluoric acid solution. Ni3(py)12F6·7H2O crystallizes in the monoclinic space group I2/a (No. 15) with Z = 4. The lattice parameters are a = 16.117(4), B = 8.529(3), C = 46.220(7) Å, β = 92.46(2)°, and V = 6348(5) Å3. The structure consists of discrete Ni(py)4F2 octahedra linked through H---O---HF and H---O---HO hydrogen bonding interactions. Single c were grown from a (HF)x·pyridine/pyridine/water solution.  相似文献   

4.
Compounds of formula [Al(CH3CN)6][MCl6]3(CH3CN)3 (M=Ta (1); Nb (2); Sb (3)) have been synthesized from the reactions of MCl5 and AlCl3 in acetonitrile and characterized by X-ray crystallography. Complex 1 crystallizes in the tetragonal space group P4/mbm with a = B = 10.408(2), C = 7.670(3) Å, V = 830.9(4) Å3 and Z = 2/3. Complex 2 crystallizes in the tetragonal space group P4/mnc with a = B = 330(a), C = 15.320(3) Å3 V = 1634.8(4) Å3 and Z = 4/3. Complex 3 also crystallizes in the tetragonal space group P4/mnc with a = B = 10.313(1), C = 15.238(2) Å, V = 1621.0(1) Å3 and Z = 4/3. The non-integer Z values for complexes 1–3 result unusual problems of disorder and/or twinning in these crystal structures due to their high symmetry. The M---Cl distances range from 2.329(3) Å in the Ta complex to 2.355(1) Å in the Sb complex, while the Al---N distances are similar in all three complexes, ranging from 1.92(1) to 1.97(1) Å, respectively. Complexes 1–3 are the first structurally characterized complexes that contain a (hexaacetonitrile)aluminum(III) cation.  相似文献   

5.
The effects of ambient temperature and humidity, month, age and genotype on sperm production and semen quality in AI bulls in Brazil were evaluated. Data from two consecutive years were analyzed separately. Seven Bos indicus and 11 Bos taurus bulls from one artificial insemination (AI) center were evaluated in Year 1 and 24 B. indicus and 16 B. taurus bulls from three AI centers were evaluated in Year 2. Ambient temperature and humidity did not significantly affect sperm production and semen quality, probably because there was little variation in these variables. Month accounted for less than 2% of the variation in sperm production and semen quality. Increased bull age was associated with decreased sperm motility (P<0.10) and increased minor sperm defects (P<0.001) in Year 1. B. indicus bulls had greater (P<0.005) sperm concentration than B. taurus bulls in both years (1.7×109/ml versus 1.2×109/ml in Year 1 and 1.6×109/ml versus 1.2×109/ml in Year 2, respectively). Ejaculate volume was not significantly affected by genotype in Year 1 (6.6 ml versus 6.9 ml in B. indicus and B. taurus bulls, respectively), but B. indicus bulls had greater (P<0.05) total (11.4×109 versus 8.2×109) and viable (6.7×109 versus 4.9×109) numbers of spermatozoa in the ejaculate than B. taurus bulls. In Year 2, B. taurus bulls had greater (P<0.05) ejaculate volume than B. indicus bulls (8.2 ml versus 6.7 ml, respectively) and total and viable number of spermatozoa in the ejaculate were not significantly different between genotypes (10.3×109 versus 9.1×109 and 6.1×109 versus 5.4×109 in B. indicus and B. taurus bulls, respectively). Sperm motility was not significantly affected by genotype (mean, 59%). In Year 1, B. indicus bulls tended (P<0.10) to have more major sperm defects and had more (P<0.05) total sperm defects than B. taurus bulls (11.8% versus 8.7% and 13.6% versus 10.0%, respectively). In Year 2, B. indicus bulls tended (P<0.10) to have more total sperm defects than B. taurus bulls (16.2% versus 13.3%, respectively). In conclusion, neither ambient temperature and humidity nor month (season) significantly affected sperm production and semen quality. B. indicus bulls had significantly greater sperm concentration and B. taurus bulls had significantly fewer morphologically defective spermatozoa.  相似文献   

6.
Cobalt(III) complexes with a thiolate or thioether ligand, t-[Co(mp)(tren)]+ (2), t-[Co(mtp)(tren)]2+ (1Me) and t-[Co(mta)(tren)]2+ (2Me), (mp = 3-mercaptopropionate, MA = 3-(methylthio)propionate and MTA = 2-(methylthio)acetate) have been prepared in aqueous solutions. The crystal structures of 1, 2, 1Me and 2Me were determined by X-ray diffraction methods. The crystal data are as follows, t-[Co(mp)(tren)]ClO4 (1CIO4): monoclinic, P21/n, A = 10.877(8), B = 11.570(4), c = 12.173(7) Å, β = 92.20(5)°, V = 1531(1) Å3, Z = 4 and R = 0.060; t-[Co(ma)(tren)]Cl·3H2O (2Cl·3H2O): monoclinic, P21/n, a = 7.7688(8), B = 27.128(2), C = 7.858(1) Å, β = 100.63(1)°, V = 1627.7(3) Å3, Z = 4 and R = 0.066; (+)465CD-t-[Co(mtp)(tren)](ClO4)2 ((+)465CD-1Me(ClO4)2): orthorhombic, P212121, A = 10.6610(7), B = 11.746(1), C = 15.555(1) Å, V = 1947.9(3) Å3, Z = 4 and R = 0.068; (+)465CD-t-[Co(mta)(tren)](ClO4)2 ((+)465CD-2Me(ClO4)2): orthorhombic, P212121, a = 10.564(1), B = 11.375(1), C = 15.434(2) Å, V = 1854.7(4) Å3, Z = 4 and R = 0.047. All central Co(III) atoms have approximately octahedral geometry, coordinated by four N, one O, and one S atoms. All of the complexes are only isomer, of which the sulfur atom in the didentate-O,S ligands are located at the trans position to the tertiary amine nitrogen atom of tren. 1 and 1Me contain six-membered chelate ring, and 2 and 2Me do five-membered chelate ring in the didentate ligand. The chirality of the asymmetric sulfur donor atom in (+)465CD-1Me is the S configuration and that in (+)465CD-2Me is the R one. The 1H NMR, 13C NMR and electronic absorption spectral behaviors and electrochemical properties of the present complexes are discussed in relation to their stereochemistries.  相似文献   

7.
Nauplii batch cultures of Balanus amphitrite were reared with four different diatoms (Skeletonema costatum, Thalassiosira pseudonana, Chaetoceros gracilis, silicate-limited C. gracilis) at three different cells concentrations: 1×105, 5×105, and 1×106 cells ml−1. The cyprid energy reserves were quantified as the ratio of triacylglycerols (TAG) to DNA. Energy reserves of larvae fed on different diatoms at a concentration of 1×106 cells ml−1 were ranked in the order: silicate-limited C. gracilis>C. gracilis>T. pseudonana>S. costatum. There was a significant linear relationship between the TAG content of the diet and cyprid energy reserves. The effect of cyprid energy reserves on metamorphosis to polystyrene surface in the presence and the absence of conspecific settlement factor (SF) was studied after 12, 24, and 48 h of incubation. A strong positive correlation between energy reserves and percent metamorphosis was observed in the absence of SF (r12 h=0.88, r24 h=0.82, r48 h=0.68, P<0.05). A weak positive correlation was observed in the presence of SF (r12 h=0.43, r24 h=0.48, r48 h=0.50, P<0.05). In both treatments, more than 80% of the cyprids with high energy reserves metamorphosed within 24 h. In contrast, a high proportion of cyprids with low energy reserves metamorphosed in response to SF in 24 h. Our results indicate that discriminatory metamorphic behavior of cyprids is closely linked to their TAG/DNA ratio, a proxy for energy reserve.  相似文献   

8.
The synthesis of the tetradentate pendant arm macrocycles 1,4,7-triazacyclononane-N-acetate (L1) and N-(2-hydroxybenzyl)-1,4,7-triazacyclononane (HL2) and their coordination chemistry with vanadium(IV) and (V) are reported. The following mononuclear species have been prepared and characterized by UV-Vis, IR spectroscopy: [L1VIVO(NCS)] (1), [L1VO2]·H2O (2), [L2VO(NCS)] (3), [L2VO(NCS)]Cl (4), and [L2VO2] (5). In addition, the dinuclear, mixed valent complexes [L21V2O3]Br (6), [L22V2O3](ClO4)·0.5acetone (7), and the homovalent complex [L22V2O3](ClO4)2 (8) have been synthesized. Complexes 2, 3, 6 and 7 have been characterized by single crystal X-ray crystallography. Crystal data: 2, space group P21c,a=9.944(4),b=6.701(3),c=18.207(8)Å, β=102.88(3)°, V=1182.7 Å3, Z=4, Dcalc=1.51 g cm−3, R=0.049 based on 4760 reflections; 3, space group Pbca, A=11.003(6), b=14.295(7), C=20.21(1) Å, V=3178.8 Å3, Z=8, Dcalc=1,50 g cm−3, R=0.057 based on 1049 reflections; 6, space Pbcn, a=12.922(3), B=13.852(3), C=12.739(3) Å, V=2280.3 Å3, Z=4, Dcalc=1,75 g cm−3, R=0.047 based on 1172 reflections; 7, space group C2/c, A=23.553(9), B=13.497(5), C=20.951(8) Å, β=90.03(3)°, V=6660.2 Å3, Z=8, Dcalc=1.49 g cm−3, R=0.053 based on 3698 reflections. Complexes 6 and 7 are mixed valent V(IV)/(V) complexes containing the [OV---O---VO]3+ core. In the solid state 6 belongs to class III (delocalized) and 7 to class I (localized) according to the Robin and Day classification of mixed valent compounds. A rationale for these differing electronic structures is given.  相似文献   

9.
Experimental evidence is provided that selenomethionine oxide (MetSeO) is more readily reducible than its sulfur analogue, methionine sulfoxide (MetSO). Pulse radiolysis experiments reveal an efficient reaction of MetSeO with one-electron reductants, such as e-aq (k = 1.2 × 1010M-1s-1), CO·-2 (k = 5.9 × 108 M-1s-1) and (CH3)2) C·OH (k = 3.5 × 107M-1s-1), forming an intermediate selenium-nitrogen coupled zwitterionic radical with the positive charge at an intramolecularly formed Se N 2σ/1σ* three-electron bond, which is characterized by an optical absorption with λmax at 375 nm, and a half-life of about 70 μs. The same transient is generated upon HO· radical-induced one-electron oxidation of selenomethionine (MetSe). This radical thus constitutes the redox intermediate between the two oxidation states, MetSeO and MetSe. Time-resolved optical data further indicate sulfur-selenium interactions between the Se N transient and GSH. The Se N transient appears to play a key role in the reduction of selenomethionine oxide by glutathione.  相似文献   

10.
The molecular structure of the title complexes [Fe(H2O)4][Fe(Hedta)(H2O)]2 · 4H2O (I) and [Fe(H[2edta)(H2O)] · 2H2O (II) have been determined by single-crystal X-ray analyses. The crystal data are as follows: I: monoclinic, P21/n, A = 11.794(2), B = 15.990(2), C = 9.206(2) Å, β = 90.33(1)°, V = 1736.1(5) Å3, Z = 2 and R = 0.030; II: monoclinic, C2/c, A = 11.074(2), B = 9.856(2), C = 14.399(2) Å, β = 95.86(1)°, V = 1563.3(4) Å3, Z = 4 and R = 0.025. I is found to be isomorphous with the MnII analog reported earlier and to contain a seven-coordinate and approximately pentagonal-bipyramidal (PB) [FeII(Hedta)(H2O] unit in which Hedta acts as a hexadentate ligand. The [FeII(H2edta)(H2O)] unit in II has also a seven-coordinate PB structure with the two protonated equatorial glycine arms both remaining coordinated, and thus bears a structural resemblance to the seven-coordinate [CoII(H2edta)(H2O)] reported previously.  相似文献   

11.
To clarify the radical-scavenging activity of butylated hydroxytoluene (BHT), a food additive, stoichiometric factors (n) and inhibition rate constants (kinh) were determined for 2,6-di-tert-butyl-4-methylphenol (BHT) and its metabolites 2,6-di-tert-butyl-p-benzoquinone (BHT-Q), 3,5-di-tert-butyl-4-hydroxybenzaldehyde (BHA-CHO) and 3,5-di-tert-butyl-4-hydroperoxy-4-methyl-2,5-cyclohexadiene-1-one (BHT-OOH). Values of n and kinh were determined from differential scanning calorimetry (DSC) monitoring of the polymerization of methyl methacrylate (MMA) initiated by 2,2′-azobis(isobutyronitrile) (AIBN) or benzoyl peroxide (BPO) at 70 °C in the presence or absence of antioxidants (BHT-related compounds). The n values declined in the order BHT (1–2) > BHT-CHO, BHT-OOH (0.1–0.3) > BHT-Q (0). The n value for BHT with AIBN was approximately 1.0, suggesting dimerization of BHT. The kinh values declined in the order BHT-Q ((3.5–4.6)×104 M−1 s−1) > BHT-OOH (0.7–1.9×104 M−1 s−1) > BHT-CHO ((0.4–1.7)×104 M−1 s−1) > BHT ((0.1–0.2)×104 M−1 s−1). The kinh for metabolites was greater than that for the parent BHT. Growing MMA radicals initiated by BPO were suppressed much more efficiently by BHT or BHT-Q compared with those initiated by AIBN. BHT was effective as a chain-breaking antioxidant.  相似文献   

12.
对植物光合和后光合分馏进行分析,有助于提升对植物生理和水分管理等的认识。本研究通过测定大气、侧柏叶片和枝条韧皮部可溶性化合物的δ13C,探讨了光合作用时大气和叶片间碳同位素的分馏(ΔCa-leaf)和光合作用后叶片到枝条间的碳同位素分馏(ΔCleaf-phlo)对土壤含水量(SWC)和大气CO2浓度(Ca)的响应。结果表明: ΔCa-leaf在SWC为田间持水量(FC)的95%~100%(95%~100%FC)且Ca为400 μmol·mol-1时达到最大值(13.06‰),在SWC为35%~45%FC且Ca为800 μmol·mol-1时达到最小值(8.63‰);气孔导度和叶肉细胞导度均与ΔCa-leaf呈显著线性正相关,相关系数分别为0.43和0.44;而ΔCleaf-phlo并未受到SWC和Ca的显著影响。本研究不仅可以提高对碳同位素的分馏机制的认识,而且可以为植物对未来气候变化的生存适应性提供理论依据。  相似文献   

13.
The reactions of [(H5C6)3P]2ReH6 with (CH3CN)3Cr(CO)3, (diglyme)Mo(CO)3 or (C3H7CN)3W(CO)3 led to the formation of [(H5C6)3P]2ReH6M(CO)3 (M = Cr, Mo, W) complexes. These have been characterized by IR and NMR spectroscopies, as well as elemental analyses. A single crystal X-ray diffraction study has also been carried out for the M = Cr complex as a K(18-crown-6)+ salt. The complex crystallizes as a THF monosolvate in the monoclinic space group P21/n with a = 22.323(6), B = 9.523(2), C = 27.502(5) Å, β = 104.98(2)0 and V = 5648 Å3 for Z = 4. The Re---Cr separation is 2.5745(12) Å, and the two phosphine ligands are oriented unsymmetrically. Although the hydride ligands were not found, the presence of three bridging hydrides and a dodecahedral coordination geometry about rhenium could be inferred. Low temperature 1H and 31P NMR spectroscopic studies did not reveal the low symmetry of the solid state structure.  相似文献   

14.
New mixed metal complexes SrCu2(O2CR)3(bdmap)3 (R = CF3 (1a), CH3 (1b)) and a new dinuclear bismuth complex Bi2(O2CCH3)4(bdmap)2(H2O) (2) have been synthesized. Their crystal structures have been determined by single-crystal X-ray diffraction analyses. Thermal decomposition behaviors of these complexes have been examined by TGA and X-ray powder diffraction analyses. While compound 1a decomposes to SrF2 and CuO at about 380°C, compound 1b decomposes to the corresponding oxides above 800°C. Compound 2 decomposes cleanly to Bi2O3 at 330°C. The magnetism of 1a was examined by the measurement of susceptibility from 5–300 K. Theoretical fitting for the susceptibility data revealed that 1a is an antiferromagnetically coupled system with g = 2.012(7), −2J = 34.0(8) cm−1. Crystal data for 1a: C27H51N6O9F9Cu2Sr/THF, monoclinic space group P21/m, A = 10.708(6), B = 15.20(1), C = 15.404(7) Å, β = 107.94(4)°, V = 2386(2) Å3, Z = 2; for 1b: C27H60N6O9Cu2Sr/THF, orthorhombic space group Pbcn, A = 19.164(9), B = 26.829(8), C = 17.240(9) Å, V = 8864(5) Å3, Z = 8; for 2: C22H48O11N4Bi2, monoclinic space group P21/c, A = 17.614(9), B = 10.741(3), C = 18.910(7) Å, β = 109.99(3)°, V = 3362(2) Å3, Z = 4.  相似文献   

15.
In vivo glycerolipid metabolism was studied in sciatic nerves of normal and Trembler mice. The results showed that two kinetically independent pathways were implicated in the labeling of diacylglycerophospholipids from [3H]palmitate: the Kennedy pathway and a ‘direct acylation’ pathway. In normal nerves, 45% of the glycerophospholipids were labeled, with a rate constant k3 = 3.9 × 10−3 min−1, from phosphatidic acid and diacylglycerol intermediates, themselves formed with a rate constant of k1 = 0.24 min−1 from a free 3H-fatty acid pool, FFA1, that represents 45% of the total injected label. The remaining 55% of the glycerophospholipids were labeled from a kinetically distinct free 3H-fatty acid pool, FFA2, with a rate constant of k4 = 9.8 × 10−2, via a process that does not implicate a detectably labeled metabolic intermediate (‘direct acylation’). Glycerophospholipid labeling via the Kennedy pathway in the Trembler mouse sciatic nerves was reduced to 75% of the normal level, while labeling via the ‘direct acylation’ pathway was increased 1.4-fold. The values of the rate constants for free 3H-fatty acid utilisation (k1 and k4) were both increased about 2.5-fold, while that of glycerophospholipid formation from diacylglycerol (k3) was close to normal. Copyright © 1996 Elsevier Science Ltd  相似文献   

16.
The hydrolysis of 2,4-dinitrophenylphosphate (DNPP) to orthophosphate and 2,4-dinitrophenolate (DNP) is accelerated in the presence of excess tn2Co(H2O)23+ or trpnCo(H2O)23+ at rates which maximize at pHs close to those at which the hydroxoaquatetraaminecobalt(III) complex concentrations peak (tn2, pH 6.4; trpn, pH 6.0; tn = trimethylenediamine; trpn = 3,3′,3″-triaminotripropylamine). For dilute DNPP solutions (10−4 M) the hydrolysis rates (25°C, 0.50 M NaClO4) increase with increasing Co/DNPP ratio in ways that are qualitatively as well as quantitatively different for the two systems (trpn: steady increase moving toward rate saturation, higher rates; tn2: ‘S’-shaped curve with very low rates at low ratios, lower rates compared to trpn for comparable ratios). For the trpn system the results are interpreted on the basis of pre-equilibrium formation of the 1:1 monodentate-DNPP cobalt complex by substitution of the labile water on cobalt, and rate-determining attack by the cis-coordinated hydroxide on the phosphorus center to affect hydrolysis. For the tn2 system the main path to hydrolysis is through a 2:1 cobalt to DNPP complex in which attack by a cis-coordinated hydroxide is again involved. The more complex rate behavior and the slower hydrolysis rates observed for tn2 system result from the formation of cis and trans isomers in which trans arrangements of coordinated DNPP and hydroxide leave the latter unavailable to participate in intramolecular hydrolysis. Computer fitting of the observed rate data provides values of equilibrium and rate constants for the two systems. Detailed mechanistic schemes are proposed. For the trpn system at pH 6.0 and a 25:1 cobalt to DNPP ratio (5 × 10−5 M DNPP) the observed acceleration over hydrolysis in the absence of the cobalt complex is 3 × 103; the calculated specific rate constant for hydrolysis in the reactive 1:1 complex (k 0.2 s−1) represents an acceleration over the unpromoted rate of 3 × 104.  相似文献   

17.
Three new crystalline tin selenide salts have been prepared from the reactions of [PPh4]2[Sn(Se43] in supercritical solvents. The starting material pyrolyzes in supercritical acetonitrile to form [PPh4]4[Sn6Se21] (I), and it also reacts with SnSe in supercritical ammonia leading to a mixture of [PPh4]4[Sn3Se11]2 (II). and [PPh4]2[Sn(Se4)(Se6)2] (III). All three compounds have been characterized by single crystal X-ray diffraction. Crystallographic data: for I, C96H90P4Se21Sn6, space group triclinic, P-1, A = 18.763(3), B = 24.600(4), C = 13.137(1) Å, = 102.63(1), β = 93.66(1), γ = 108.72(1)°, V = 5544(1) Å3, Z = 2, R = 0.0350, RW = 0.0317: for II, C96H80P4Se22Sn6, space group monoclinic P21/c, A = 31.500(4), B = 16.572(3), C = 22.352(3) Å, β = 103.53(1)°, V = 11344(3) Å3, Z = 4, R = 0.0771, RW = 0.0664: for III, C48H40P2Se16Sn, space group monoclinic, C2/c, A = 25.381(2), B = 13.934(4), C = 19.465(3) Å, β = 121.587(8)°, V = 5867(2) Å3, Z = 4, R = 0.0807, RW = 0.0650. One of the compounds, [PPh4]2[Sn(Se4(Se62], is a molecular cluster while the other two complexes [PPh4]4[Sn3Se11]2 and [PPh4]4[Sn6Se21], are one dimensional tin selenide chains. The structures of the two chains are related and consits of tetrahedral and distorted trigonal bipyramidal tin(IV) centers bridged by Se2−, Se22− and Se32− chains.  相似文献   

18.
镉胁迫对吊兰及银边吊兰生长及镉富集特性的影响   总被引:2,自引:0,他引:2  
选择吊兰和银边吊兰为试验材料,采用水培法研究其在不同Cd2+处理浓度(0、20、80、200 μmol·L-1)下生长及生理特性的变化。结果表明: 20 μmol·L-1镉对两种吊兰的影响较小,单叶面积、总叶面积、叶绿素(Chl)a含量、总叶绿素[Chl (a+b)]含量、类胡萝卜素含量、Chl a/Chl b值、胞间二氧化碳浓度(Ci)和蒸腾速率(Tr)与对照(CK)基本无显著差异;80 μmol·L-1镉胁迫下两种吊兰叶片初始荧光(Fo)和非光化学淬灭系数(NPQ)升至最高水平;200 μmol·L-1镉胁迫下,两种吊兰生物量、叶绿素含量、最大净光合速率(Pn)、气孔导度(gs)、最大光化学量子产量(Fv/Fm)、实际光化学量子产量Y(II)、转移系数(TF)以及各部分生物量均降至最低水平,而两种吊兰的过氧化物酶(POD)、抗坏血酸过氧化物酶(APX)及过氧化氢酶(CAT)活性和银边吊兰的丙二醛(MDA)含量均有不同程度的增加。随着镉处理浓度的增加,两种吊兰各器官Cd含量持续升高,且主要富集在根部;吊兰各器官Cd含量及富集系数(BCF)在胁迫处理下均高于银边吊兰。研究表明,两种吊兰对镉具有一定的耐性,其中吊兰对Cd的耐受能力强于银边吊兰,可考虑作为绿化植物用于修复镉污染水体或土壤。  相似文献   

19.
Common cocklebur has several biotypes including multiple seeded cocklebur (MSC), NCC-TX, and NCC-MS. Alternaria helianthi applied at 2.5×104 conidia mL-1 in a 50% micro-emulsion of unrefined corn oil (MESUCO) or 0.2% Silwet L 77 caused 60-75% mortality on NCC-TX and MSC. Increasing the conidial concentration to 5×104 mL-1 increased mortality to 100% on MSC and NCC-TX, and 75% on NCC-MS. At 10×104 conidia mL-1, A. helianthi caused 100% mortality in all three biotypes. No mortality occurred in any biotype at inoculation rates of 2.5 and 5×104 conidia mL-1 when applied in water. Increasing the dew period from 0 to 12 h increased mortality from 0 to 100% on all three biotypes at a rate of 2.5×104 conidia mL-1 in Silwet and MESUCO. MSC appears to be the most sensitive biotype.  相似文献   

20.
《植物生态学报》2017,41(6):670
Aims Anthropogenic pollutants cause an increase in ground-level ozone concentration, which is a known threat to plant growth and yield and has been extensively observed worldwide. Since ozone is only slightly soluble in water, it is deposited mainly through dry deposition in terrestrial ecosystem. The object of this study was to analyze the characteristics of ozone dry deposition and to estimate the contribution of stomatal and non-stomatal ozone deposition pathways to total ozone deposition in a winter wheat field.Methods The research site was a winter wheat (Triticum aestivum) field located in Yongfeng experimental station of Nanjing University of Information Science & Technology. The data used in this study were collected from March 16, 2016 to May 30, 2016. We observed ozone dry deposition with an eddy-covariance system. This system mainly included a 3D sonic anemometer, an open-path infrared absorption spectrometer, a fast-response ozone chemiluminescent analyzer and a slow-response ozone monitor. We simultaneously measured meteorological data including solar radiation (SR), air temperature (T), air relativity humidity (RH), wind speed, net radiation, and rainfall. All raw data were recorded with data-logger and averaged every 30 min.Important findings Half hourly means of ozone concentrations (CO3), ozone flux (FO3) and ozone dry deposition velocity (Vd) in the winter wheat field were 32.9 nL·L-1, -5.09 nmol·m-2·s-1, 0.39 cm·s-1, and the ranges of them were 16-58 nL·L-1, -2.9- -11.7 nmol·m-2·s-1, 0.17-0.63 cm·s-1, respectively. FO3 and CO3/Vd were found to be mismatched with phase peaks occurring at different time intervals. The ecosystem was more effective on ozone dry deposition, under conditions of moderate to high SR (SR ≥ 400 W·m-2), moderate T and humility (T = 18 °C and RH > 40%). The relationship between Vdmax and SR was this function (y = 1.06 -exp (-0.0094 - x)). Vdmax increased with SR When SR < 400 W·m-2, and Vdmax reached its maximum when SR =400 W·m-2. Vdmax maintained its maximum when SR ≥ 400 W·m-2. The relationship between Vdmax and T was “bell” curve (y = 1.06 - (x - 18)2/169). Vdmax reached its maximum when T = 18 °C. Vdmax decreased with RH when RH < 40 % (y = 0.030x - 0.106). The variation of Vd might uncertainty when RH was high. There was a liner positive relationship between friction velocity (u*) and Vd, but this relationship was not significant. The mean day-to-day and daytime contributions of stomatal and non-stomatal ozone deposition pathway to total ozone deposition were 32%, 68% and 42%, 58%, respectively, during the whole experimental period.  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