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 quasielastic light scattering method was used to study the ionic strength dependence of the mutual diffusion coefficient of sodium polystyrene sulfonate (NaPSS) as a function of NaCl and CaCl2 concentrations. The results indicate a splitting in the relaxation times that depends on the ratio Cp/Cs, where Cp and Cs are the polyion and added salt concentrations. A universal relationship taking into account Manning's theory of condensation and the Debye screening due to the added salt is proposed to characterize the fast–slow relaxation time transition.  相似文献   

2.
Quasielastic light scattering and electrophoretic light scattering experiments were performed on chicken erythrocyte polynucleosome solutions at various temperatures and ionic strengths. The apparent diffusion coefficient, Dapp, was found to depend on the scattering vector K. In general, Dapp can be described as a damped oscillatory function of K in the ionic strength range of 10 to 60 mM and over the temperature range of 10 to 40°C. Electrophoretic light scattering studies on total digest chromatin samples indicate the apparent charge on the polynucleosomes increases as the ionic strength is lowered from 10 to 1 mM. These data are interpreted in terms of fluctuations in the surface charge distribution of the polyion and subsequent inducement of an asymmetric distribution of small ions about the polyion. These fluctuation components lead to the formation of “clusters” of polyions.  相似文献   

3.
Quasi-elastic light scattering studies on some polyelectrolyte systems exhibit a somewhat “bizarre” behavior in the profile of the apparent diffusion coefficient Dapp as a function of the salt concentration Cs. As Cs is decreased, Dapp first increases in accordance with polyelectrolyte theories, and then undergoes a precipitous drop in value by over an order of magnitude at a well-defined critical value Cs = C. This “transition” from Cs > C (ordinary) to Cs < C (extraordinary) is referred to as the “ordinary-extraordinary” (o-e) transition. Ghosh, Peitzsch, and Reed [(1992) Biopolymers, Vol. 32, pp. 1105–1122] proposed a “filterable aggregate” (FA) and “other particle” interpretation for the o-e transition and its reversibility in regard to ionic strength changes. The present communication examines in detail the FA model as applied to the o-e transition. It is shown that the FA model fails to account of the established characteristics of the o-e transition. © 1993 John Wiley & Sons, Inc.  相似文献   

4.
Activities of Mg and Ca in solutions of heparin with added sodium or potassium chloride have been determined by means of a previously described dye spectrophotometric method. The polyion concentration, Cp (expressed as mol univalent anionic charges/I) is 0.001 in all cases, solution total ionic strengths are 0.005, 0.0075, 0.01, 0.02 and 0.04 mol/l. Divalent metal ion concentrations are varied between 0 and 1.0 Cp. The results for the metal ion activities are expressed in the form of binding isotherms, θ2 vs. C2/Cp (θ2=C2b/Cp; C2b=bound divalent metal ion concentration) and Scatchard plots, K2 vs. θ2/(C2-C2b, at different ionic strengths. Results are correlated with a theoretical treatment of the mixed counterion system, called the ‘two variable theory’, recently developed by Manning. The ionic strength dependence of θ2 and K2 is found to agree with the prediction of the two variable theory. Mg interaction with heparin appears to be independent of the nature of the charged groups on the polyion and is in very good agreement with the two variable theory. Ca binding was found to be considerably higher than Mg binding and is in excess of theoretical predictions, suggesting a localized or specific interaction of Ca with heparin.  相似文献   

5.
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   

6.
A novel porphyrin‐C60 dyad (PCD1) is designed and synthesized to investigate and manipulate the supramolecular structure where geometrically isotropic [such as [60]fullerene (C60)] and anisotropic [such as porphyrin (Por)] units coexist. It is observed that PCD1 possesses an enantiomeric phase behavior. The melting temperature of the stable PCD1 thermotropic phase is 160 °C with a latent heat (ΔH) of 18.5 kJ mol?1. The phase formation is majorly driven by the cooperative intermolecular Por–Por and C60–C60 interactions. Structural analysis reveals that this stable phase possesses a supramolecular “double‐cable” structure with one p‐type Por core columnar channel and three helical n‐type C60 peripheral channels. These “double‐cable” columns further pack into a hexagonal lattice with a = b = 4.65 nm, c = 41.3 nm, α = β = 90°, and γ = 120°. The column repeat unit is determined to possess a 12944 helix. With both donor (D; Pro) and acceptor (A; C60) units having their own connecting channels as well as the large D/A interface within the supramolecular “double‐cable” structure, PCD1 has photogenerated carriers with longer lifetimes compared to the conventional electron acceptor [6,6]‐phenyl‐C61‐butyric acid methyl ester. A phase‐separated columnar morphology is observed in a bulk‐heterojunction (BHJ) material made by the physical blend of a low band‐gap conjugated polymer, [poly[2,6‐(4,4‐bis‐(2‐ethylhexyl)‐4H‐cyclopenta [2,1‐b;3,4‐b′]‐dithiophene)‐alt‐4,7‐(2,1,3‐benzothia‐diazole)] (PCPDTBT), and PCD1. With a specific phase structure in the solid state and in the blend, PCD1 is shown to be a promising candidate as a new electron acceptor in high performance BHJ polymer solar cells.  相似文献   

7.
Many macromolecular interactions, including protein‐nucleic acid interactions, are accompanied by a substantial negative heat capacity change, the molecular origins of which have generated substantial interest. We have shown previously that temperature‐dependent unstacking of the bases within oligo(dA) upon binding to the Escherichia coli SSB tetramer dominates the binding enthalpy, ΔHobs, and accounts for as much as a half of the observed heat capacity change, ΔCp. However, there is still a substantial ΔCp associated with SSB binding to ssDNA, such as oligo(dT), that does not undergo substantial base stacking. In an attempt to determine the origins of this heat capacity change, we have examined by isothermal titration calorimetry (ITC) the equilibrium binding of dT(pT)34 to SSB over a broad pH range (pH 5.0–10.0) at 0.02 M, 0.2 M NaCl and 1 M NaCl (25°C), and as a function of temperature at pH 8.1. A net protonation of the SSB protein occurs upon dT(pT)34 binding over this entire pH range, with contributions from at least three sets of protonation sites (pKa1 = 5.9–6.6, pKa2 = 8.2–8.4, and pKa3 = 10.2–10.3) and these protonation equilibria contribute substantially to the observed ΔH and ΔCp for the SSB‐dT(pT)34 interaction. The contribution of this coupled protonation (∼ −260 to −320 cal mol−1 K−1) accounts for as much as half of the total ΔCp. The values of the “intrinsic” ΔCp,0 range from −210 ± 33 cal mol−1 °K−1 to −237 ± 36 cal mol−1K−1, independent of [NaCl]. These results indicate that the coupling of a temperature‐dependent protonation equilibria to a macromolecular interaction can result in a large negative ΔCp, and this finding needs to be considered in interpretations of the molecular origins of heat capacity changes associated with ligand‐macromolecular interactions, as well as protein folding. Proteins 2000;Suppl 4:8–22. © 2000 Wiley‐Liss, Inc.  相似文献   

8.
Hard carbons (HCs) are promising anodes of sodium‐ion batteries (SIBs) due to their high capacity, abundance, and low cost. However, the sodium storage mechanism of HCs remains unclear with no consensus in the literature. Here, based on the correlation between the microstructure and Na storage behavior of HCs synthesized over a wide pyrolysis temperature range of 600–2500 °C, an extended “adsorption–insertion” sodium storage mechanism is proposed. The microstructure of HCs can be divided into three types with different sodium storage mechanisms. The highly disordered carbon, with d002 (above 0.40 nm) large enough for sodium ions to freely transfer in, has a “pseudo‐adsorption” sodium storage mechanism, contributing to sloping capacity above 0.1 V, together with other conventional “defects” (pores, edges, heteroatoms, etc.). The pseudo‐graphitic carbon (d‐spacing in 0.36–0.40 nm) contributes to the low‐potential (<0.1 V) plateau capacity through “interlayer insertion” mechanism, with a theoretical capacity of 279 mAh g?1 for NaC8 formation. The graphite‐like carbon with d002 below 0.36 nm is inaccessible for sodium ion insertion. The extended “adsorption–insertion” model can accurately explain the dependence of the sodium storage behavior of HCs with different microstructures on the pyrolysis temperature and provides new insight into the design of HC anodes for SIBs.  相似文献   

9.
The studies presented here confirm earlier reports that an actin-like protein is abundant in brain. However, when the traditional procedures for isolating muscle actin are applied to brain, many different proteins are extracted. Tubulin, a major protein in brain with properties similar to actin, is the major constituent. A method is described for isolating the “brain actin” to a purity of 90–95%. The isolation method begins with an extraction of bovine brain in low ionic strength buffer with ATP and sucrose. The extract is treated with NH4SO4, MgCl, and KCl and incubated at 37°C. A precipitate is formed which contains primarily tubulin and brain actin. Resolubilization of the brain actin is achieved with a low ionic strength buffer solution with sucrose and ATP. Further purification is accomplished by a cycle of polymerization—depolymerization. This “brain actin” shares with muscle actin the following properties: (1) Similar molecular weight and molecular charge as determined by SDS polyacrylamide gel and ordinary disc electrophoresis; (2) Polymerization to a filamentous form under the same conditions; (3) Contains 3-methylhistidine; (4) Vinblastine sulfate will induce filament formation.  相似文献   

10.
Artificial ribonucleases of AnBCL series were synthesized by solid‐phase method. They consist of a hydrophobic alkyl radical A (n = 3–12 carbon atoms), an “RNA‐binding domain” B (bisquaternary salt of 1,4‐diazabicyclo[2.2.2]octane), a “catalytic domain” C (histidine residue) and a “linker” L that joins the domains B and C. The effect of the alkyl radical on the catalytic properties of the chemical catalyst was studied using three activated phosphate ester substrates: p‐nitrophenyl phosphate, bis‐p‐nitrophenyl phosphate, and thymidine‐3′‐p‐nitrophenyl phosphate.  相似文献   

11.
Fumio Oosawa 《Biopolymers》1968,6(1):135-144
A theoretical analysis is given on the effect of low molecular salts on the titration behavior of partially neutralized linear polyacids or polybases. On the basis of the additivity law experimentally and theoretically established on the activity of counterions in polyelectrolyte solutions containing salts, the chemical potential of alkali added for neutralization of polyacids is derived as a function of the salt concentration. The relation between the pH and the salt, concentration at constant degree of neutralization is expressed by pH(cs) ? pH(0) = ({[? (γpCa)/?Ca]/γos} ? 1) ln [1 + (γosCspCA)] where Ca is the concentration of alkali, C8, the concentration of salts, γp is the activity coefficient of counterions in the absence of salts and γos is that in the absence of polyelectrolytes. This relation is found to be in good agreement with experimental data obtained in synthetic linear polyelectrolytes as well as in rodlike charged proteins. Therefore, it is concluded that the titration behavior of linear polyions can be understood from the special feature of the integrated coulomb interaction between linear polyions and small ions which was essential for the additivity law. In addition, the screening effect of salts on the electric potential around the polyion is discussed.  相似文献   

12.
 The effects of the ionic atmosphere on the enthalpic and entropic contributions to the reduction potential of native (state III) beef heart cytochrome c have been determined through variable-temperature direct electrochemistry experiments. At neutral or slightly alkaline pH values, from 5 to 50  °C, the reduction enthalpy and entropy become less negative with decreasing ionic strength. The reduction entropy extrapolated at null ionic strength is approximately zero, indicating that, in the absence of the screening effects of the salt ions on the network of the electrostatic interactions at the protein-solvent interface, the solvation properties and the conformational flexibility of the two redox states are comparable. The moderate decrease in E°′ observed with increasing ionic strength [ΔE°′IS =(E°′) I =0.1 M–(E°′) I =0 M=–0.035 V at 25  °C], once the compensating enthalpic and entropic effects of the salt-induced changes in the hydrogen bonding within the hydration sphere of the molecule in the two redox states are factorized out, results in being ultimately determined by the stabilizing enthalpic effect of the negatively charged ionic atmosphere on the ferri form. At pH 9, the ionic strength dependence of the reduction termodynamics of cytochrome c follows distinctive patterns, possibly as a result of specific binding of the hydroxide ion to the protein. A decrease in ionic strength at constant pH, as well as a pH increase at constant ionic strength, induces a depression of the temperature of the transition from the low-T to high-T conformer of cytochrome c, which suggests that a temperature-induced decrease in the pK a for a residue deprotonation is the key event of this conformational change. Received: 7 April 1999 / Accepted: 19 July 1999  相似文献   

13.
A new method of isolating nuclei and chromosomes of salivary gland cells is described. — The influence of ionic strength and pH of the medium on the state of decondensation of chromosomal bands is studied. In the isolation medium (a modified Ringer solution), all the bands are in a condensed state; as the ionic strength is increased the bands decondense. This reaction of the bands to increasing ionic strength is dependent on the pH which determines: 1) the range of ionic strengths which causes decondensation of the bands; i.e., the lower the pH, the higher the ionic strength is required for decondensation (at pH 7.3, 150–350 mM NaCl, at pH 4.3, 500–800 mM NaCl), and 2) the extent of structural changes caused by increasing ionic strength; that is, at neutral pH the bands become diffuse (“fading”) and at moderate acidic pH (optimum 4.3) the bands unravel to yield pufflike structures (“swelling”). — All ion species tested induce decondensation of bands, but each one is effective differently; specifically, Mg+ is more effective than Na+ and K+, and ClO4 ? is more effective than Cl?. — “Swelling” as induced at pH 4.3 by high ionic strength cannot be reversed by a mere lowering of ionic strength (to 150 mM NaCl) and a subsequent raise of pH (to 7.5); it can be reversed only by an addition of histones. The various histone fractions act differently on the recondensation process. — “Swelling” is correlated with an increase in template activity as evidenced by an increased incorporation of 3H-UTP, measured in the presence of ATP, CTP, GTP and exogeneous RNA polymerase. — The individual bands differ in their sensitivity to an increasing ionic strength. This differential sensitivity expresses itself only if one of the following conditions is met: 1) a moderately acidic pH (optimum 4.3) or 2) the presence of divalent cations at neutral pH. — In a few bands the sensitivity to an increasing ionic strength is dependent on the ionic species (Na+, K+, Mg++ and Ca++). — It is attempted to explain the above reactions on the basis of the physico-chemical properties of chromosomes.  相似文献   

14.
Pleckstrin homology (PH) domains mediate protein–membrane interactions by binding to phosphatidylinositol phosphate (PIP) molecules. The structural and energetic basis of selective PH–PIP interactions is central to understanding many cellular processes, yet the molecular complexities of the PH–PIP interactions are largely unknown. Molecular dynamics simulations using a coarse-grained model enables estimation of free-energy landscapes for the interactions of 12 different PH domains with membranes containing PIP2 or PIP3, allowing us to obtain a detailed molecular energetic understanding of the complexities of the interactions of the PH domains with PIP molecules in membranes. Distinct binding modes, corresponding to different distributions of cationic residues on the PH domain, were observed, involving PIP interactions at either the “canonical” (C) and/or “alternate” (A) sites. PH domains can be grouped by the relative strength of their C- and A-site interactions, revealing that a higher affinity correlates with increased C-site interactions. These simulations demonstrate that simultaneous binding of multiple PIP molecules by PH domains contributes to high-affinity membrane interactions, informing our understanding of membrane recognition by PH domains in vivo.  相似文献   

15.
The “extraordinary” diffusional phase (EP) at low ionic strength, and the conditions for 1 its removability by filtration were investigated for dilute solutions of the following linear polyelectrolytes: poly(L -lysine), heparin, chondroitin-6-sulfate, hyaluronate, polystyrene sulfonate, and variably ionized polyacrylamide. The EP was not present for all the different types studied, and for heparin, for example, the phase was present only for samples from certain sources. In all cases the phase was removable by filtration through sufficiently small pore-size membranes. Once filtered, the EP remained absent for over one week. It is concluded that the extraordinary diffusional phase consists of fairly stable polyelectrolyte aggregates, and sometimes also includes other very small particulate impurities. These aggregates and other small particles are thought to be present, or at least nascent, in the dry polyelectrolyte material, so that their properties may depend critically on the manner in which such dry material is produced. Tests for “reversibility” of the EP by cycling between high and low Cs by dialysis further confirm these conclusions. The evidence is thus against the EP representing any type of temporal aggregates or local ordering, at least for the linear polyelectrolytes studied in this work Rather, due to the extremely feeble scattering of ordinary polyelectrolytes at low ionic strength, the weak scattering from residual aggregates and other particles, not removed by ordinary filtration and centrifugation procedures, give autocorrelable scattering signals with long decay times. The “loss” of the extraordinary phase as ionic strength increases appears to be due simply to the weak EP scattering signal getting buried in the sharply increasing scattering from the ordinary polyelectrolyte phase. Model calculations based on experimental data support this latter conclusion. © 1992 John Wiley & Sons, Inc.  相似文献   

16.
The origin and early development of procambium and associated ground meristem of major and minor veins have been examined in the leaf blades of seven C4 grass species, representing different taxonomic groups and the three recognized biochemical C4 types (NAD-ME, PCK, and NADP-ME). Comparisons were made with the C3 species, Festuca arundinacea. In “double sheath” (XyMS+) species (Panicum effusum, Eleusine coracana, and Sporoboìus elongatus), the procambium of major veins gives rise to xylem, phloem, and a mestome sheath; associated ground meristem differentiates into PCA (“C4 mesophyll”) tissue and the PCR (“Kranz”) sheath. Development in the C3 species parallels this pattern, except that associated ground meristem differentiates into mesophyll and a parenchymatous bundle sheath. In contrast, major vein procambium of “single sheath” (XyMS–) species (Panicum bulbosum, Digitaria brownii, and Cymbopogon procerus) differentiates into xylem, phloem and a PCR sheath; associated ground meristem gives rise to PCA tissue. These observations of major vein development support W. V. Brown's hypothesis that the PCR sheaths of “double sheath” (XyMS+) C4 grasses are homologous with the parenchymatous bundle sheaths of C3 grasses, while in “single sheath” (XyMS–) C4 species they are homologous with the mestome sheath. Although there are some similarities in the development of the major and minor vascular bundle procambium in the C4 species examined, the ontogeny of the smaller minor veins is characterized by a precocious delineation of the PCR sheath layer that may even precede the appearance of the distinctive cytological features of ground meristem and procambium. This contracted development in minor veins appears to be related to their close spacing in mature leaves and to their comparatively late appearance during leaf ontogeny.  相似文献   

17.
Poly(Lys(HBr)-Gly-Pro-Pro-Gly-Pro) has been synthesized and studied by circular dichroism (CD) spectroscopy. It is apparently the first polyhexapeptide collagen model reported with an ionizable side chain. The monomer (ε-(p-nitrobenzyloxycarbonyl)-Lys-Gly-Pro-Pro-Gly-Pro-p-nitrophenyl-ester) was prepared by a stepwise strategy employing active esters. Polymerization in N,N-dimethyl formamide, followed by removal of the Lys side chain protection with HBr/acetic acid, gave a polydisperse product. Fractionation was accomplished by gel filtration chromatography. The polydisperse material had a molecular weight (Mr = 5–17,000). High molecular weight fractions from triple helices under concentrated conditions at 2°C. The triple helical structure gives a CD pattern very similar to that of collagen and its triple helical analogs. However, unlike collagen, the polyhexapeptide undergoes spontaneous dissociation at temperatures substantially below the melting temperature from a triple helical form to single chains. This process is promoted at low concentrations, high temperature, neutral pH, and low molecular weight, and is apparently due, in large part, to unfavorable ionic side-chain interactions. In addition to this relatively slow “ionic” dissociation the triple helical polypeptide may be thermally dissociated in a manner similar to collagen. The thermal denaturation is a relatively fast process compared with ionic dissociation. A high molecular weight fraction (3 × Mr = 48,000) was found to melt at 42°C at neutral pH but increased to 54°C at pH 12 where the lysyl side chains are predominantly deprotonated. Furthermore, reconstitution of triple helices appeared to be more readily achieved at high pH. Thus it is concluded that ionic repulsion between side chains causes destabilization of the triple helix and hinders reconstitution.  相似文献   

18.
Twelve healthy male subjects each undertook two bouts of moderate exercise (70% VO2max for 30 minutes) in the morning (08:00) and late afternoon (18:00) at least 4 days apart. Measurements were made of heart rate, core (rectal) temperature, sternum skin temperature, and forearm skin blood flow during baseline conditions, during the bout of exercise, and throughout a 30-minute recovery period. Comparisons were made of the changes of heart rate, temperature, and skin blood flow produced by the exercise at the two times of day. Student t tests indicated that baseline values for core temperature (37.15°C ±. 06°C vs. 36.77°C ± 0.06°C) and sternum temperature (33.60°C ± 0.29°C vs. 32.70°C ± 0.38°C) were significantly (p <. 05) higher in the late afternoon than the early morning. Two-way analysis of variance (ANOVA) indicated that the increases in core and sternum temperatures during exercise were significantly less (p =. 0039 and. 0421, respectively) during the afternoon bout of exercise compared with the morning, even though the work loads, as determined by changes in heart rate, were not significantly different (p =. 798) at the two times of testing. There were also tendencies for resting forearm skin blood flow to be higher in the afternoon than in the morning and for exercise to produce a more rapid rise in this variable in the afternoon. The possible mechanisms producing these responses to exercise are discussed in terms of those that are responsible for the normal circadian rhythm of core temperature. It is concluded that the body's ability to remove a heat load is less in the early morning, when the circadian system is in a “heat gain” mode, than in the late afternoon, when heat gain and “heat loss” modes are balanced more evenly. (Chronobiology International, 17(2), 197–207, 2000)  相似文献   

19.
Asthma is a “seasonal disease” with symptoms either aggravated by environmental changes during specific seasons or prevalent at certain times of the year for other reasons. We examined whether the heart rate variability (HRV) of asthmatic children changes by season. The HRV during a portion of one night (00:00–04:00) and day (12:00–16:00) and the entire 24h period (00:00–24:00) during each of the four seasons was analyzed. The data of 95 children with asthma and 106 healthy children, as controls, were assessed. In children with asthma during the 24h period, seasonal variation in the low-frequency (LF) band (0.04–0.15 Hz) and the high-frequency (HF) band (0.15–0.4 Hz) were detected (HF: F = 6.81, p =.0003; LF: F = 4.18, p =.008. The HF value in the summer was significantly higher than in autumn and spring (Scheffe test: autumn vs. summer, s = 4.46, p<.001; spring vs. summer, s = 2.86, p<.05), while the LF value in autumn was significantly lower than in summer (s = 3.42, p <.01). In the control group, no seasonal variation in HF, LF, or LF/HF was detected. The findings infer the HRV, a surrogate measure of autonomic nervous system function, of asthmatic children is more susceptible to seasonal changes brought about by either endogenous annual rhythms or environmental weather phenomena. (Chronobiology International, 17(4), 503–511, 2000)  相似文献   

20.
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   

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