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.
W.L. Butler  M. Kitajima 《BBA》1975,396(1):72-85
A model for the photochemical apparatus of photosynthesis is presented which accounts for the fluorescence properties of Photosystem II and Photosystem I as well as energy transfer between the two photosystems. The model was tested by measuring at ?196 °C fluorescence induction curves at 690 and 730 nm in the absence and presence of 5 mM MgCl2 which presumably changes the distribution of excitation energy between the two photosystems. The equations describing the fluorescence properties involve terms for the distribution of absorbed quanta, α, being the fraction distributed to Photosystem I, and β, the fraction to Photosystem II, and a term for the rate constant for energy transfer from Photosystem II to Photosystem I,kT(II→I). The data, analyzed within the context of the model, permit a direct comparison of α andkT(II→I) in the absence (?) and presence (+) of Mg2+:α/?α+= 1.2andk/?T(II→I)k+T(II→I)= 1.9. If the criterion thatα + β = 1 is applied absolute values can be calculated: in the presence of Mg2+,a+ = 0.27 and the yield of energy transfer,φ+T(II→I) varied from 0.065 when the Photosystem II reaction centers were all open to 0.23 when they were closed. In the absence of Mg2+? = 0.32 andφT(II→I) varied from 0.12 to 0.28.The data were also analyzed assuming that two types of energy transfer could be distinguished; a transfer from the light-harvseting chlorophyll of Photosystem II to Photosystem I,kT(II→I), and a transfer from the reaction centers of Photosystem II to Photosystem I,kt(II→I). In that caseα/?α+= 1.3,k/?T(II→I)k+T(II→I)= 1.3 andk/?t(II→I)k+(tII→I)= 3.0. It was concluded, however, that both of these types of energy transfer are different manifestations of a single energy transfer process.  相似文献   

2.
The radioiodinated pindolol analogs 125I-labeled cyanopindolol ([125I]CYP) and 125I-labeled hydroxybenzylpindolol ([125I]HBP) have been used to study binding to human platelet β-adrenergic receptors. [125I]CYP binds to a saturable class of binding sites on platelet membranes with a dissociation constant (Kd) of 14±3 pM and maximal binding capacity (Bmax) of 18±4 fmol/mg protein. Binding of [125I]CYP is reversible and is characterized by forward and reverse rate constants of 1.8·107 s?1·M?1 and 3.8·10?4 s?1, respectively. [125I]HBP binds to a saturable class of platelet membrane sites with a Kd of 50±10 pM and Bmax of 32±6 fmol/mg protein. [125I]HBP also binds to a saturable class of sites on intact platelets with a Kd of 58±14 pM and Bmax of 24±4 molecules per platelet. Binding of [125I]CYP and [125I]HBP is stereospecifically inhibited by propranolol and epinephrine; the (?) stereoisomers are at least 50-times more potent than the (+) stereoisomers. Binding of both radioligands is inhibited by adrenergic ligands with a potency order of propranolol ? isoproterenol > epinephrine > practolol > norepinephrine > phenylephrine. These observations indicate that [125I]CYP and [125I]HBP bind to platelet sites which have the pharmacological characteristics of β-adrenergic receptors but which are not typical of either the β1 or β2 sub-type.  相似文献   

3.
Responses of photosystem I and II activities of Microcystis aeruginosa to various concentrations of Cu2+ were simultaneously examined using a Dual-PAM-100 fluorometer. Cell growth and contents of chlorophyll a were significantly inhibited by Cu2+. Photosystem II activity [Y(II)] and electron transport [rETRmax(II)] were significantly altered by Cu2+. The quantum yield of photosystem II [Y(II)] decreased by 29 % at 100 μg L?1 Cu2+ compared to control. On the contrary, photosystem I was stable under Cu2+ stress and showed an obvious increase of quantum yield [Y(I)] and electron transport [rETRmax(I)] due to activation of cyclic electron flow (CEF). Yield of cyclic electron flow [Y(CEF)] was enhanced by 17 % at 100 μg L?1 Cu2+ compared to control. The contribution of linear electron flow to photosystem I [Y(II)/Y(I)] decreased with increasing Cu2+ concentration. Yield of cyclic electron flow [Y(CEF)] was negatively correlated with the maximal photosystem II photochemical efficiency (F v/F m). In summary, photosystem II was the major target sites of toxicity of Cu2+, while photosystem I activity was enhanced under Cu2+ stress.  相似文献   

4.
The effect of para vs meta substitution on the biological behavior of an intact antibody and an F(ab′)2 fragment was investigated. Paired-label studies were performed using 81C6 IgG and OC 125 F(ab′)2 labeled using the N-succinimidyl esters of both p-[125I]- and m-[131I]iodobenzoate as well as with the potential catabolites, p-[125I]- and m-[131I]iodobenzoic acid. In all 3 studies, up to 55% lower uptake of 131I in thyroid and stomach was observed, suggesting that the m-substituted species were more inert to dehalogenation in vivo.  相似文献   

5.
Z Chen  G Wang  X Zhai  Y Hu  D Gao  L Ma  J Yao  X Tian 《Cell death & disease》2014,5(4):e1164
Apoptosis is a major mode of cell death occurring during ischemia–reperfusion (I/R) induced injury. The p66Shc adaptor protein, which is mediated by PKCβ, has an essential role in apoptosis under oxidative stress. This study aimed to investigate the role of PKCβ2/p66Shc pathway in intestinal I/R injury. In vivo, ischemia was induced by superior mesenteric artery occlusion in mice. Ruboxistaurin (PKCβ inhibitor) or normal saline was administered before ischemia. Then blood and gut tissues were collected after reperfusion for various measurements. In vitro, Caco-2 cells were challenged with hypoxia–reoxygenation (H/R) to simulate intestinal I/R. Translocation and activation of PKCβ2 were markedly induced in the I/R intestine. Ruboxistaurin significantly attenuated gut damage and decreased the serum levels of tumor necrosis factor-α (TNF-α) and interleukin-6 (IL-6). Pharmacological blockade of PKCβ2 suppressed p66Shc overexpression and phosphorylation in the I/R intestine. Gene knockdown of PKCβ2 via small interfering RNA (siRNA) inhibited H/R-induced p66Shc overexpression and phosphorylation in Caco-2 cells. Phorbol 12-myristate 13-acetate (PMA), which stimulates PKCs, induced p66Shc phosphorylation and this was inhibited by ruboxistaurin and PKCβ2 siRNA. Ruboxistaurin attenuated gut oxidative stress after I/R by suppressing the decreased expression of manganese superoxide dismutase (MnSOD), the exhaustion of the glutathione (GSH) system, and the overproduction of malondialdehyde (MDA). As a consequence, ruboxistaurin inhibited intestinal mucosa apoptosis after I/R. Therefore, PKCβ2 inhibition protects mice from gut I/R injury by suppressing the adaptor p66Shc-mediated oxidative stress and subsequent apoptosis. This may represent a novel therapeutic approach for the prevention of intestinal I/R injury.  相似文献   

6.
Two ferredoxins from nitrogen-fixing cells of the phototrophic bacterium Rhodopseudomonas capsulata, strain B10, are purified to a homogeneous state and characterized. The molecular mass of ferredoxin I is about 12 kDa and that of ferredoxin II, 18 kDa. Ferredoxin I contains 8 Fe2+ and 8 S2?; ferredoxin II has 4 Fe2+ and 4 S2? per molecule. The redox potential of ferredoxin I is about ?270 mV and that of ferredoxin II ?419 mV. Ferredoxin I is more labile to the action of O2, O?2, H2O2 and heating. The ferredoxins are also different in their absorption and EPR spectra, amino acid composition and electron-transfer activity to Rps. capsulata nitrogenase: both C2H2 reduction and H2 evolution by Rps. capsulata nitrogenase proceed faster in the presence of ferredoxin I than in case of ferredoxin II. Synthesis of ferredoxin I takes place only in Rps. capsulata nitrogen-fixing cells grown in light under anaerobic conditions whereas ferredoxin II formation does not depend on the source of nitrogen or the growth medium, though the amount of ferredoxin II varies with the growth conditions. Its highest level has been found in the cells grown in lactate-limited medium in the presence of CO2 and light or in the presence of glutamate in darkness under anaerobic conditions.  相似文献   

7.
Four bacteria isolated from peat biofilters, Thiobacillus thioparus DW44, Thiobacillus sp. HA43, Xanthomonas sp. DY44 and Hyphomicrobium sp. I55, were selected to enhance the removal ratios of hydrogen sulfide (H2S), methanethiol (MT) and dimethyl sulfide (DMS) in a mixed gas system. Two bacteria, DW44 and I55, which degrade H2S, MT, DMS and dimethyl disulfide (DMDS), were mixed with DY44 or HA43 which degrade only H2S and MT. Although DMS removal was significantly inhibited by the presence of H2S and MT in a peat biofilter inoculated with the single bacterium, enhanced removability of H2S, MT and DMS was observed by mixing Hyphomicrobium sp. I55 either with Thiobacillus sp. HA43 or Xanthomonas sp. DY44. The removal rate (g-S-kg-dry peat−1·d−1) by I55 after 8 d was 0.664 in total sulfur load, 0.827 g-S·kg-dry g-S·-kg-dry peat−1·d−1, but the rates by the mixed cultures of I55 plus HA43, and I55 plus DY44 were 0.760 and 0.801, respectively. In particular, DMS removability in mixed gases by a mixed culture of I55 and DY44 was almost equivalent to that by I55 when only DMS was supplied, suggesting that removal of H2S and MT, which inhibited DMS removal, was preferentially conducted by DY44 and led to improved DMS removability by I55.  相似文献   

8.
A time-resolved spectroscopic study of the isolated photosynthetic reaction center (RC) from Heliobacterium modesticaldum reveals that thermal equilibration of light excitation among the antenna pigments followed by trapping of excitation and the formation of the charge-separated state P800 +A0 occurs within ~25 ps. This time scale is similar to that reported for plant and cyanobacterial photosystem I (PS I) complexes. Subsequent electron transfer from the primary electron acceptor A0 occurs with a lifetime of ~600 ps, suggesting that the RC of H. modesticaldum is functionally similar to that of Heliobacillus mobilis and Heliobacterium chlorum. The (A0  ? A0) and (P800 + ? P800) absorption difference spectra imply that an 81-OH-Chl a F molecule serves as the primary electron acceptor and occupies the position analogous to ec3 (A0) in PS I, while a monomeric BChl g pigment occupies the position analogous to ec2 (accessory Chl). The presence of an intense photobleaching band at 790 nm in the (A0  ? A0) spectrum suggests that the excitonic coupling between the monomeric accessory BChl g and the 81-OH-Chl a F in the heliobacterial RC is significantly stronger than the excitonic coupling between the equivalent pigments in PS I.  相似文献   

9.
With the aim to develop beneficial tracers for cerebral tumors, we tested two novel 5-iodo-2′-deoxyuridine (IUdR) derivatives, diesterified at the deoxyribose residue. The substances were designed to enhance the uptake into brain tumor tissue and to prolong the availability in the organism. We synthesized carrier added 5-[125I]iodo-3′,5′-di-O-acetyl-2′-deoxyuridine (Ac2[125I]IUdR), 5-[125I]iodo-3′,5′-di-O-pivaloyl-2′-deoxyuridine (Piv2[125I]IUdR) and their respective precursor molecules for the first time. HPLC was used for purification and to determine the specific activities. The iodonucleoside tracer were tested for their stability against human thymidine phosphorylase. DNA integration of each tracer was determined in 2 glioma cell lines (Gl261, CRL2397) and in PC12 cells in vitro. In mice, we measured the relative biodistribution and the tracer uptake in grafted brain tumors. Ac2[125I]IUdR, Piv2[125I]IUdR and [125I]IUdR (control) were prepared with labeling yields of 31–47% and radiochemical purities of >99% (HPLC). Both diesterified iodonucleoside tracers showed a nearly 100% resistance against degradation by thymidine phosphorylase. Ac2[125I]IUdR and Piv2[125I]IUdR were specifically integrated into the DNA of all tested tumor cell lines but to a less extend than the control [125I]IUdR. In mice, 24 h after i.p. injection, brain radioactivity uptakes were in the following order Piv2[125I]IUdR>Ac2[125I]IUdR>[125I]IUdR. For Ac2[125I]IUdR we detected lower amounts of radioactivities in the thyroid and stomach, suggesting a higher stability toward deiodination. In mice bearing unilateral graft-induced brain tumors, the uptake ratios of tumor-bearing to healthy hemisphere were 51, 68 and 6 for [125I]IUdR, Ac2[125I]IUdR and Piv2[125I]IUdR, respectively. Esterifications of both deoxyribosyl hydroxyl groups of the tumor tracer IUdR lead to advantageous properties regarding uptake into brain tumor tissue and metabolic stability.  相似文献   

10.
The mechanisms of the ultrafast charge separation in reaction centers of photosystem I (PS I) complexes are discussed. A kinetic model of the primary reactions in PS I complexes is presented. The model takes into account previously calculated values of redox potentials of cofactors, reorganization energies of the primary P700+A 0 - and secondary P700+A 1 - ion-radical pairs formation, and the possibility of electron transfer via both symmetric branches A and B of redox-cofactors. The model assumes that the primary electron acceptor A0 in PS I is represented by a dimer of chlorophyll molecules Chl2A/Chl3A and Chl2B/Chl3B in branches A and B of the cofactors. The characteristic times of formation of P700+A 0 - and P700+A 1 - calculated on the basis of the model are close to the experimental values obtained by pump-probe femtosecond absorption spectroscopy. It is demonstrated that a small difference in the values of redox potentials between the primary electron acceptors A0A and A0B in branches A and B leads to asymmetry of the electron transfer in a ratio of 70: 30 in favor of branch A. The secondary charge separation is thermodynamically irreversible in the submicrosecond range and is accompanied by additional increase in asymmetry between the branches of cofactors of PS I.  相似文献   

11.
J.H. Golbeck  B.R. Velthuys  B. Kok 《BBA》1978,504(1):226-230
Absorption changes accompanying the formation of light-induced P-700+ were investigated in a highly enriched Photosystem I preparation where an intermediate electron acceptor preceding P-430 could be detected. In an enriched Photosystem I particle, light-induced reversible absorption changes observed at 700 nm in the presence of dithionite resembled those previously seen at 703 nm and 820 nm [9], thus indicating the presence of a backreaction between P-700+ and A?2. After this same Photosystem I particle was treated to denature the bound iron-sulfur centers, the photochemical changes that could be attributed to P-700 A2 were completely lost. These results provide evidence that the intermediate electron acceptor, A2, is a bound iron-sulfur protein. Additional studies in the 400–500 nm region with Photosystem I particles prepared by sonication indicate that the spectrum of A2 is different from that of P-430.  相似文献   

12.
We have separated and purified two forms of Met-tRNAf deacylase (or two separate enzymes), an activity that mediates in part the suppression of polypeptide chain initiation that occurs in heme deficiency or with double-stranded RNA, 1000-fold from the 0.5 M KCl wash of rabbit reticulocyte ribosomes. Deacylase I is a minor activity with an S20,w of 5.9, D20,w of 4.9 and Mr of 110 000, while deacylase II is the major activity with an S20,w of 3.3, D20,w of 7.1 and Mr of 43 000. Both convert crude reticulocyte or pure yeast, wheat germ, and E. coli [35S]Met-tRNAf to [35S]methionine and tRNAMetf and have no effect on reticulocyte [35S]fMet-tRNAf, [3H]Ala-tRNA or [3H]Lys-tRNA. However, while deacylase I has similar activity throughout the pH range of 6.1–8.1, deacylase II has a sharp pH optimum at 7.9 and is almost completely inactive at 6.1. In addition, deacylase II shows a much greater affinity for pure Met-tRNAf than deacylase I (Km of 1.5–3 nM vs. 100 nM), and, while deacylase II is selectively inhibited by tRNAMetf, deacylase I is inhibited similarly by any added tRNA.  相似文献   

13.
Mapability of Very Close Markers of Bacteriophage λ   总被引:3,自引:0,他引:3       下载免费PDF全文
Recombinant frequency was compared with nucleotide distance in crosses involving markers in either the PRM or the cy region of phage λ. For each pair of markers, we performed reciprocal four-factor crosses of the following types: (I) A+m1 +m2-B- x A-m1 -m2+B+; and (II) A+m1 -m2+B- x A-m1 +m2-B+. In crosses of type I, the frequency of A+m1 +m2+B+ recombinants among total (selected) A+B+ progeny was directly proportional to nucleotide distance between m1 and m2 in the range from 3 to 160 nucleotides. When less than three nucleotides separated m1 and m2, the measured yields of m1+m2+ recombinants were significantly depressed.

We also found that the frequency of A+m1 +m2+B+ recombinants among total A+B+ progeny was significantly lower (about 10-fold on the average) in crosses of type II than in the corresponding crosses of type I. Since mismatch correction should yield A+m1 +m2+B+ recombinants with approximately equal frequencies in type I and II crosses, we suggest: (1) that most m1+m2+ recombinants produced in type I crosses must arise from the formation of heteroduplex structures with a discontinuity (in the source of genetic information) between sites m1 and m2, and (2) that mismatch correction is not a major pathway for production of recombinants for close markers in normal λ infection.

  相似文献   

14.
Sweet flag (Acorus calamus L.) and yellow flag (Iris pseudacorus L.) have been used increasingly in constructed wetlands (CWs) for treatment of eutrophic wastewater. In order to properly match plant species with the type of wastewater being treated, it is important to know the performance of plant species under different NO3/NH4+ ratios. We investigated dry matter (DW) production and N content of A. calamus and I. pseudacorus under five NO3/NH4+ ratios (100/0, 75/25, 50/50, 25/75, and 0/100) in a hydroponic system. Results showed that the two species exhibited different preferences for NO3 and NH4+. Total DW, shoot DW, and N content were greater with NO3/NH4+ ratios of 50/50 and 75/25 than otherwise for A. calamus, but these parameters were only higher under the sole NO3 treatment in I. pseudacorus. We conclude that A. calamus could be best used for treating wastewater in constructed wetlands with NO3/NH4+ ratios between 50/50 and 75/25, while I. pseudacorus for treating wastewater with NO3 only to achieve the highest biomass production and efficiency in the removal of N.  相似文献   

15.
E. J. Eisen 《Genetics》1975,79(2):305-323
Long-term response to within full-sib family selection for increased postweaning gain was evaluated in lines having different effective population sizes (Ne) and selection intensities (i). Line designations were I4(4), I8(2), I16(2), M4(4), M8(2) and M16(2), where I and M indicate selection of the top 50% and 25%, respectively; 4, 8 and 16 represent the number of parental pairs per replicate and number of replicates is given in parentheses. Realized within full-sib family heritabilities (hR2) in the first phase of selection (0-14 generations) were larger in 16-pair lines than in 4- and 8-pair lines. In the second phase of selection (>14 generations), hR2 declined significantly (P<.01) in all lines, and only the I16 and M16 lines had hR2 values significantly (P<.01) greater than zero. Realized genetic correlations involving number born, 12-day litter weight, weaning weight and six-week weight tended to decline in the second phase of selection. The I16, M16 and control (C16) replicates were crossed in all combinations at generation 14. Crosses were then selected within litters for high postweaning gain. The hR2 values in the crossbred lines were all larger than those in the second selection phase for M16-1, M16-2 and I16-1, but not for I16-2. Within each Ne level, total response was significantly (P<.01) less for I lines compared with M lines. Total response increased as Ne increased, within each level of i. Relatively small differences in realized i values among Ne lines could not account for this result. The difference in total response among the Ne lines at a given selection intensity may be due to inbreeding depression and a combination of interactions involving "drift" and selection. By crossing replicates of the M lines with the C16 control, the effects of inbreeding depression were removed. Inbreeding depression and genetic drift, as defined herein, were equally important in accounting for differences among Ne lines in total response.  相似文献   

16.
[125I]LSD (labeled at the 2 position) has been introduced as the first 125I-labeled ligand for serotonin 5-HT2 (S2) receptors. In the present study we examined the binding of [125I]LSD and its non-radioactive homologue, 2I-LSD, to bovine caudate homogenates. The binding of [125I]LSD is saturable, reversible, stereospecific and is destroyed by boiling the membranes. The specific to total binding ratio in this tissue is 75–80% and Scatchard plots of the binding data reveal Kd = 1.1 nM, Bmax = 9.6 fmol/mg wet weight tissue. The association and dissociation rate constants are highly temperature dependent. At 0°C the net dissociation is less than 5% after 1 h and the association rate is proportionately slow. IC50 values for a variety of compounds show a clear 5-HT2 (S2) serotonergic pattern at this [125I]LSD site. Blockage of this primary 5-HT2 (S2) caudate binding site by 0.3 μM mianserin reveals the presence of a weaker [125I]LSD binding site with a Kd = 9.1 nM, Bmax = 7.6 fmol/mg tissue. This secondary site is a D3 dopaminergic receptor site, as shown by the relative abilities of various displacers to inhibit this binding. Binding studies with nonradioactive 2I-LSD reveal a clear preference for D2 over D3 dopamine receptor sites. [125I]LSD is a sensitive and selective label for 5-HT2 (S2) serotonin receptor sites in both rat frontal cortex and bovine caudate membranes. Blockage of the primary bovine caudate [125I]LSD binding site with mianserin allows the high sensitivity of [125I]LSD to be applied to D2 dopamine receptor studies as well.  相似文献   

17.
The dimer [Ir(μ-Cl)(C8H14)2]2 reacts with the ligands (S)-(C5H4CH2CH(Ph)PPh2)Li and (R)-(C5H4CH(Cy)CH2PPh2)Li to give (S)-[Ir(η5-C5H4CH2CH(Ph)PPh2P)(C8H14)] and (R)-[Ir(η5-C5H4CH(Cy)CH2PPh2P)(C8H14)], which upon treatment with CH3I at room temperature afford the cationic iridium(III) compounds (S,SIr)-[Ir(η5-C5H4CH2CH(Ph)PPh2P)(CH3)(C8H14)][I] as a single diastereomer, and (R)-[Ir(η5-C5H4CH(Cy)CH2PPh2P)(CH3)(C8H14)][I] as a 9:1 mixture of two diastereomers. If the oxidative addition reaction is performed at reflux in methylene chloride, the starting complexes convert to the neutral compounds (S)-[Ir(η5-C5H4CH2CH(Ph)PPh2P)(CH3)(I)] and (R)-[Ir(η5-C5H4CH(Cy)CH2PPh2P)(CH3)(I)] as 1.6:1 and 3.3:1 mixtures of diastereoisomers, respectively. Carbonyl iridium complexes are synthesized by reacting [IrCl(CO)(PPh3)2] with the ligands to afford (S)-[Ir(η5-C5H4CH2CH(Ph)PPh2P)(CO)] and (R)-[Ir(η5-C5H4CH(Cy)CH2PPh2P)(CO)]. They give upon treatment with CH3I the cationic species (S)-[Ir(η5-C5H4CH2CH(Ph)PPh2P)(CH3)(CO)][I] and (R)-[Ir(η5-C5H4CH(Cy)CH2PPh2P)(CH3)(CO)][I] as 1.6:1 and 3:1 mixture of diastereomers, respectively. No migratory-insertion of the methyl group into the carbonyl-metal bond has been observed even after prolonged heating.  相似文献   

18.
C10H26N10ONiZn, tris(1,2-diaminoethane) zinc(II) tetrakis(cyano)niccolate(II) monohydrate (I), orthorhombic, Pbca, a = 1.1680(4), b = 1.5844(3), c = 1.9981(6) nm, Z = 8 d(meas) = 1.54, d(calc) = 1.53 g cm?3. C10H24N10NiZn, tris(1,2-diaminoethane) zinc(II) terakis(cyano)niccolate(II), (II), monoclinic, P21/n, a = 0.7957(2), b = 1.5170(5), c = 1.4932(4) nm, β = 96.41(2)°, Z = 4, d(meas) = 1.49, d(calc) = 1.51 g cm?3. Both the structures (I) and (II) have been solved by the heavy atom method and refined by full-matrix least-squares to R(I) = 0.086 for 1890 independent reflections and R(II) = 0.058 for 1689 independent reflections, respectively. In the case of (II) the superlattice structure problem was solved. The crystal structure of (I) consists of [Zn(en)3]2+ cations, [Ni(CN)4]2? anions and water molecules. Two of the cyano groups in trans positions are bonded to water molecules by hydrogen bonds, the distances CN?O being 0.289 and 0.291 nm, respectively. The crystal structure of (II) is constituted by [Zn(en)3]2+ cations and [Ni(CN)4]2? anions.  相似文献   

19.
Rabbit articular chondrocytes in suspension culture synthesize Type II colagen [3α1(II)] in the absence of extracellular Ca2+ and Type Icollagen [2α1?(I)·α2] in the complete medium. As a result of pre-treatment in monolayer culture with calcitonin or parathyroid hormone in the complete medium, an influx of Ca2+ into the cells occurs. These cells produce mainly Type I collagen when transferred to suspension cultures in the medium devoid of CaCl2. If added directly to the suspension culture medium containing no CaCl2, calcitonin stimulates an active efflux of Ca2+ from the cells into the medium and leads the cells to synthesize Type I collagen. Under similar conditions, parathyroid hormone does not change the collagen-phenotype.  相似文献   

20.
Organic and water extracts of Isochrysis galbana T-ISO (=Tisochrysis lutea), Tetraselmis sp. and Scenedesmus sp. were evaluated for their antioxidant activity, acetylcholinesterase (AChE) inhibition, cytotoxicity against tumour cell lines, and fatty acids and total phenolic content (TPC). I. galbana T-ISO had the highest TPC (3.18 mg GAE g?1) and radical scavenging activity, with an IC50 value of 1.9 mg mL?1 on the acetone extract. The extracts exhibited a higher ability to chelate Fe2+ than Cu2+, and the maximum Fe2+ chelating capacity was observed in the hexane extract of Scenedesmus sp. (IC50=0.73 mg mL?1) and Scenedesmus sp. (IC50?=?0.73 mg mL?1). The highest ability to inhibit AChE was observed in the water and ether extracts of Scenedesmus sp., with IC50 values of 0.11 and 0.15 mg mL?1, respectively, and in the water extract of I. galbana (IC50?=?0.16 mg mL?1). The acetone extract of I. galbana T-ISO significantly reduced the viability of human hepatic carcinoma HepG2 cells (IC50?=?81.3 μg mL?1) as compared to the non-tumour murine stromal S17 cell line, and displayed a selectivity index of 3.1 at the highest concentration tested (125 μg mL?1). All species presented a highly unsaturated fatty acids profile. Results suggest that these microalgae, particularly I. galbana T-ISO, could be a source of biomolecules for the pharmaceutical industry and the production of functional food ingredients and can be considered as an advantageous alternative to several currently produced microalgae.  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