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 15 毫秒
1.
Genetic analysis of unusual complex-heterozygous genotypes in populations ofO. grandiflora from Alabama (USA) has shown that these strains are composed of a typicalgrandiflora (B) complex and an altered B complex (designated as BA) which probably contains genetic elements derived from an A genotype such as the beta complex ofO. biennis group 1. Analysis of the meiotic configurations of artificial hybrids between the new strains and a series of complexes of known segmental arrangement allowed determination of the arrangements of the unknown complexes. These data are evidence for origin of the altered B complexes.  相似文献   

2.
The osmotic pressure equation for nonideal, associating systems of the type nA +mB ? AnBm, has been derived, by using the assumption yA nB m/yA nyB m = 1. This treatment can also be applied to related associations such as nA + mB ? AB + AB2 + A2B + …. From osmotic pressure experiments on the pure reactants it is possible to obtain the molecular weights (MA and MB) of the reactants and also the virial coefficients (BAA and BBB) of the reactants. The osmotic pressure of a nonreacting mixture of A and B can be calculated from these measurements. It can be used along with osmotic pressure measurements on equilibrium mixtures of A and B to obtain expressions containing the equilibrium constant (or constants) and the cross-virial coefficients (BAB and BBA). Several procedures are described for the evaluation of the equilibrium constant (or constants) and the BAB or BBA terms. It appears that this procedure is a general one which is applicable to associations of the type nA + mB ? AB + A2B + AB2 + …. By correcting for nonideal behavior, one should then be able to apply it to any method available for analyzing ideal associations of the types considered here. In addition it is possible, subject to certain restrictions, to analyze associations of the type 3A + B ? A2 + AB.  相似文献   

3.
Summary The finding of two duplicated C4A haplotypes in a normal French family led to a detailed study of their C4 polymorphism. The father had an extremely rare A*6A*11, B* QO haplotype inherited by all of his children and the mother had the more common A*3A*2, B*QO haplotype. Two HLA identical daughters only have four C4A alleles. The father's A11 allotype expresses Ch: 1 (Chido) rather than Rg:1 (Rodgers) and represents a new Ch phenotype Ch: 1,-2,-3,-4,-5,-6. In order to clarify the genetic background in this unusual family, DNA studies of restriction fragment length polymorphisms (RFLPs) were undertake. The father's rare haplotype, which expresses two C4A allotypes, results from a long and a short C4 gene normally associated with the A*6, B*1 that also exhibits the BglII RFLP. As it travels in an extended MHC haplotype HLA A2, B57 (17), C2*C, BF*S, DR7 that is most frequently associated with A*6, B*1, we postulate that the short C4B has been converted in the chain region to a C4A gene which produces a C4A protein. This report of a short C4A gene is the first example in the complex polymorphism of C4.  相似文献   

4.
Syoyu Kobayasi 《Biopolymers》1971,10(5):915-922
Analytical and numerical calculations of the decay processes of the electric birefringence in an isomerizing system have been performed. Two modes of isomerization are considered: in the first mode, the direction of the axis (of the optical anisotropy) of a molecule is conserved during isomerization; in the second mode, it is not. In the first mode, if GAis not equal to GB, and that, if kAopA2 is not equal to kBopB2 (when the orienting electric field is weak), in which GA, GB and pA, pb are the optical anisotropy and permanent electric dipole moments, of the molecule in two states A and B of the isomerization, and kAo and kBo are the rate constants for the transitions A → B and B → A, respectively, and if diffusion rates are very much slower than the chemical rates, the relaxation due to the chemical reaction can be detected in the decay of the electric birefringence. In the second mode, if diffusion rates are somewhat slower than the chemical rates, the relaxation due to the chemical reaction appears in the decay, even though there are no differences in optical anisotropy and electric moments in the two states. When the rotary diffusion coefficients are different in the two states, the decay process becomes almost one-component, bringing the chemical rate about ten times higher to diffusion coefficients in both modes.  相似文献   

5.
clpC ofBacillus subtilis is part of an operon containing six genes. Northern blot analysis suggested that all genes are co-transcribed and encode stress-inducible proteins. Two promoters (PA and PB) were mapped upstream of the first gene. PA resembles promoters recognized by the vegetative RNA polymerase EσA. The other promoter (PB) was shown to be dependent on σB, the general stress σ factor in B. subtilis, suggesting that clpC, a potential chaperone, is expressed in a σB-dependent manner. This is the first evidence that σB in B, subtilis is involved in controlling the expression of a gene whose counterpart, clpB, is subject to regulation by σ32 in Escherichia coli, indicating a new function of σB-dependent general stress proteins. PB deviated from the consensus sequence of σB promoters and was only slightly induced by starvation conditions. Nevertheless, strong induction by heat, ethanol, and salt stress occurred at the σB-dependent promoter, whereas the vegetative promoter was only weakly induced under these conditions. However, in a sigB mutant, the σA-like promoter became inducible by heat and ethanol stress, completely compensating for sigB deficiency. Only the downstream σA-like promoter was induced by certain stress conditions such as hydrogen peroxide or puromycin. These results suggest that novel stress-induction mechanisms are acting at a vegetative promoter. Involvement of additional elements in this mode of induction are discussed.  相似文献   

6.
Zhu XG  Govindjee  Baker NR  deSturler E  Ort DO  Long SP 《Planta》2005,223(1):114-133
Chlorophyll a fluorescence induction (FI) is widely used as a probe for studying photosynthesis. On illumination, fluorescence emission rises from an initial level O to a maximum P through transient steps, termed J and I. FI kinetics reflect the overall performance of photosystem II (PSII). Although FI kinetics are commonly and easily measured, there is a lack of consensus as to what controls the characteristic series of transients, partially because most of the current models of FI focus on subsets of reactions of PSII, but not the whole. Here we present a model of fluorescence induction, which includes all discrete energy and electron transfer steps in and around PSII, avoiding any assumptions about what is critical to obtaining O J I P kinetics. This model successfully simulates the observed kinetics of fluorescence induction including O J I P transients. The fluorescence emission in this model was calculated directly from the amount of excited singlet-state chlorophyll in the core and peripheral antennae of PSII. Electron and energy transfer were simulated by a series of linked differential equations. A variable step numerical integration procedure (ode15s) from MATLAB provided a computationally efficient method of solving these linked equations. This in silico representation of the complete molecular system provides an experimental workbench for testing hypotheses as to the underlying mechanism controlling the O J I P kinetics and fluorescence emission at these points. Simulations based on this model showed that J corresponds to the peak concentrations of Q A QB (QA and QB are the first and second quinone electron acceptor of PSII respectively) and Q A Q B and I to the first shoulder in the increase in concentration of Q A Q B 2− . The P peak coincides with maximum concentrations of both Q A Q B 2− and PQH2. In addition, simulations using this model suggest that different ratios of the peripheral antenna and core antenna lead to differences in fluorescence emission at O without affecting fluorescence emission at J, I and P. An increase in the concentration of QB-nonreducing PSII centers leads to higher fluorescence emission at O and correspondingly decreases the variable to maximum fluorescence ratio (F v/F m).  相似文献   

7.
Benzidine staining of starch gels after electrophoresis of sera to which haematin was added revealed polymorphism of haemopexin in sheep, mouflon and goat. In sheep three phenotypes were observed, Hpx A, Hpx AB and Hpx B. Pedigree data support the hypothesis of codominant inheritance from a single locus by two alleles, HpxA and HpxB. Neuraminidase treatment of haemopexin preparations showed that Hpx B covered two variants, B1 and B2, thus indicating genetic control by three alleles (HpxA, HpxB1 and HpxB2). In sheep populations the frequency of HpxB is low. In mouflon, in addition to the two variants that are like those of sheep, absence of haemopexin was observed in some animals, by using starch gel electrophoresis as well as immunoelectrophoresis. In goat, three phenotypes were detected, Hpx A, Hpx AB and Hpx B, differing in migration from those of sheep. Haemopexins of the studied species are heterogeneous. Sialic acid is responsible for electrophoretic heterogeneity of sheep haemopexin. Chemical. composition (amino acid and carbohydrate), molecular weight (56 060) and N-terminal sequence (Leu-Pro-Pro-) of sheep haemopexin were also determined.  相似文献   

8.
The loci in the major histocompatibility complex (MHC) of the rat which code for class I and class II antigens—RT1.A and RT1.B, respectively — have previously been separated by laboratory-derived recombinants and by observations in inbred and wild rats. Closely linked to the MHC is the growth and reproduction complex (Grc) which contains genes influencing body size (dw-3) and fertility (ft). These phenotypic markers were used in this study to orient the A and B loci of the MHC. Two recombinants were used for mapping. The BIL(R1) animal is a recombinant between the MHC and Grc, and it carries the haplotype RT1.A lBlGrc+. The r10 animal is an intra-MHC recombinant, and it has the haplotype RT1.A nB1 Grc. These recombinants were characterized serologically, by mixed lymphocyte reactivity, by immune responsiveness to poly (Glu52Lys33Tyr15) and by the presence of the dw-3 gene. The data demonstrate that the gene order of the loci is: dw-3-RT1.B-RT1.A.  相似文献   

9.
The primary act of charge separation was studied in P+BA and P+HA states (P, primary electron donor; BA and HA, primary and secondary electron acceptor) of native reaction centers (RCs) of Rhodobacter sphaeroides R-26 using femtosecond absorption spectroscopy at low (90 K) and room temperature. Coherent oscillations were studied in the kinetics of the stimulated emission band of P* (935 nm), of absorption band of BA (1020 nm) and of absorption band of HA (760 nm). It was found that in native RCs kept in heavy water (D2O) buffer the isotopic decreasing of basic oscillation frequency 32 cm –1 and its overtones takes place by the same factor 1.3 in the 935, 1020, and 760 nm bands in comparison with the samples in ordinary water H2O. This suggests that the femtosecond oscillations in RC kinetics with 32 cm –1 frequency may be caused by rotation of hydrogen-containing groups, in particular the water molecule which may be placed between primary electron donor PB and primary electron acceptor BA. This rotation may appear also as high harmonics up to sixth in the stimulated emission of P*. The rotation of the water molecule may modulate electron transfer from P* to BA. The results allow for tracing of the possible pathway of electron transfer from P* to BA along a chain consisting of polar atoms according to the Brookhaven Protein Data Bank (1PRC): Mg(PB)-N-C-N(His M200)-HOH-O = BA. We assume that the role of 32-cm –1 modulation in electron transfer along this chain consists of a fixation of electron density at BA during a reversible electron transfer, when populations of P* and P+BA states are approximately equal.  相似文献   

10.
Glucose-6-phosphate dehydrogenase (E. C.: 1.1.1.49) phenotypes and 6-phosphogluconate dehydrogenase (E. C.: 1.1.1.44) phenotypes were determined by starch-gel electrophoresis of red cell hemolysates of Galago crassicaudatus subspp., Propithecus verreauxi, Lemur spp., Hapalemur griseus, and Macaca mulatta. A single glucose-6-phosphate dehydrogenase (G6PD) phenotype was found in each species. A single 6-phosphogluconate dehydrogenase (6PGD) phenotype was found in Lemur spp., Hapalemur griseus, and Galago crassicaudatus argentatus. In a group of six Propithecus verreauxi, three 6PGD phenotypes, PGD A, PGD AB, and PGD B, were found. Three phenotypes, PGD A, PGD AB, and PGD B, were found in 38 G. c. crassicaudatus. The three phenotypes in each species are apparently the products of two codominant autosomal alleles, PGDA and PGDB. The frequency of PGDA in G. c. crassicaudatus is 0.263. A population of 260 free-ranging macaques displays a polymorphism at the 6PGD locus. Three phenotypes, PGD A, PGD AB, and PGD B, were found. These also appear to be controlled by two codominant autosomal alleles, PGDA and PGDB the frequency of PGDA = 0.913. Additional analysis of three well-defined troops within the macaque population indicated that there are no significant differences between the troops or within the population at the 6PGD locus.  相似文献   

11.

Exosomes, small-sized extracellular vesicles, carry components of the purinergic pathway. The production by cells of exosomes carrying this pathway remains poorly understood. Here, we asked whether type 1, 2A, or 2B adenosine receptors (A1Rs, A2ARs, and A2BRs, respectively) expressed by producer cells are involved in regulating exosome production. Preglomerular vascular smooth muscle cells (PGVSMCs) were isolated from wildtype, A1R?/?, A2AR?/?, and A2BR?/? rats, and exosome production was quantified under normal or metabolic stress conditions. Exosome production was also measured in various cancer cells treated with pharmacologic agonists/antagonists of A1Rs, A2ARs, and A2BRs in the presence or absence of metabolic stress or cisplatin. Functional activity of exosomes was determined in Jurkat cell apoptosis assays. In PGVSMCs, A1R and A2AR constrained exosome production under normal conditions (p?=?0.0297; p?=?0.0409, respectively), and A1R, A2AR, and A2BR constrained exosome production under metabolic stress conditions. Exosome production from cancer cells was reduced (p?=?0.0028) by the selective A2AR agonist CGS 21680. These exosomes induced higher levels of Jurkat apoptosis than exosomes from untreated cells or cells treated with A1R and A2BR agonists (p?=?0.0474). The selective A2AR antagonist SCH 442416 stimulated exosome production under metabolic stress or cisplatin treatment, whereas the selective A2BR antagonist MRS 1754 reduced exosome production. Our findings indicate that A2ARs suppress exosome release in all cell types examined, whereas effects of A1Rs and A2BRs are dependent on cell type and conditions. Pharmacologic targeting of cancer with A2AR antagonists may inadvertently increase exosome production from tumor cells.

  相似文献   

12.
The fourth component of complement (C4) has two classes of protein, C4A and C4B, both of which have many allelic forms. The serological determinants Rodgers (Rg1, Rg2) and Chido (Ch1, Ch2, Ch3) are generally associated with C4A and C4B, respectively. The C4B3 allotype has been detected in a single Canadian family that expresses a novel Ch phenotype, Ch:–1, 2, –3. There was no information for the Rg determinants, as the C4A * 2B * 3 haplotype would normally express Rg on the C4A protein. Other C4B3 allotypes in informative families have different Ch phenotypes, and the relationships of these within extended major histocompatibility complex haplotypes are discussed in this paper.  相似文献   

13.
The major histocompatibility complex (MHC) in natural populations of rats is composed of genetic phenotypes that are similar, if not identical, to those seen in inbred laboratory strains. Examination of individual wild rats from a single location in the city of Pittsburgh, Pennsylvania, resulted in the identification of seven different RT1.A histocompatibility serotypes and three RT1.B mixed lymphocytes responses. In this population of animals there is a significant association (p < 0.005) between four RT1.A and RT1.B phenotypic pairs: RT1.A8B1, RTl.AkBn, RTl.AdBa and RT1.A1Ba. The observed values for linkage disequilibrium (0.211,0.076,0.070 and 0.085, respectively) are very high and are close to the maximum expected, given the individual allelic frequencies. Although the animals included in this study were obtained from one location, agreement with Hardy-Weinberg equilibrium is demonstrated for other loci in the same population. The demonstration of equilibrium suggests that significant inbreeding is not affecting this population of rats. Not enough is known about the allelic frequencies in surrounding rat populations to determine how important the effect of migration is on these disequilibrium values. The large linkage disequilibria may indicate that, in the rat, environmental selective forces are operating to ensure the nonrandom association of separate components of the MHC.  相似文献   

14.
The Pm blood group system in Peromyscus maniculatus was originally described by Rasmussen as consisting of two co-dominant alleles, PmA and PmB, responsible for the antigenic characteristics A and B. This paper describes two additional antisera, anti-C and anti-X, which expand this system and show it to be serologically complex. Anti-X is comprised of antibodies directed against at least two antigenic characteristics X‘ and X“. Four and possibly six alleles are described by the four antisera. Different subspecies exhibit different spectra of alleles at the Pm locus. A segregation distortion is evidenced.  相似文献   

15.
Reaction center-B875 pigment-protein complexes were purified from Rhodocyclus gelatinosus. The proteic components consist of 7–8 polypeptides among which some were identified by their apparent molecular weights: the light harvesting B875 polypeptides and of 8 and 6 kDa, reaction center L (23 kDa), M (28 kDa) and H (34 kDa), cytochrome c (43 kDa). Four c-type hemes were found per reaction center. Flash-induced absorbance changes showed the presence of both QA and QB in the complex. Charge recombination times were determined to be: 1.16±0.2 (n=30) for P+QAQB - and 7–10 ms for P+QA - in presence of herbicides. From quinone analysis on one hand and kinetics of charge recombination on the other hand, we proposed that in the reaction center of Rhodocyclus gelatinosus QA is menaquinone 8 and QB is ubiquinone 8.  相似文献   

16.
Inhibition of electron transport and damage to the protein subunits by ultraviolet-B (UV-B, 280–320 nm) radiation have been studied in isolated reaction centers of the non-sulfur purple bacterium Rhodobacter sphaeroides R26. UV-B irradiation results in the inhibition of charge separation as detected by the loss of the initial amplitude of absorbance change at 430 nm reflecting the formation of the P+(QAQB) state. In addition to this effect, the charge recombination accelerates and the damping of the semiquinone oscillation increases in the UV-B irradiated reaction centers. A further effect of UV-B is a 2 fold increase in the half- inhibitory concentration of o-phenanthroline. Some damage to the protein subunits of the RC is also observed as a consequence of UV-B irradiation. This effect is manifested as loss of the L, M and H subunits on Coomassie stained gels, but not accompanied with specific degradation products. The damaging effects of UV-B radiation enhanced in reaction centers where the quinone was semireduced (QB ) during UV-B irradiation, but decreased in reaction centers which lacked quinone at the QB binding site. In comparison with Photosystem II of green plant photosynthesis, the bacterial reaction center shows about 40 times lower sensitivity to UV-B radiation concerning the activity loss and 10 times lower sensitivity concerning the extent of reaction center protein damage. It is concluded that the main effect of UV-B radiation in the purple bacterial reaction center occurs at the QAQB quinone acceptor complex by decreasing the binding affinity of QB and shifting the electron equilibration from QAQB to QA QB. The inhibitory effect is likely to be caused by modification of the protein environment around the QB binding pocket and mediated by the semiquinone form of QB. The UV-resistance of the bacterial reaction center compared to Photosystem II indicates that either the QAQB acceptor complex, which is present in both types of reaction centers with similar structure and function, is much less susceptible to UV damage in purple bacteria, or, more likely, that Photosystem II contains UV-B targets which are more sensitive than its quinone complex.Abbreviations Bchl bacteriochlorophyll - P Bchl dimer - QA primary quinone electron acceptor - QB secondary quinone electron acceptor - RC reaction center - UV-B ultraviolet-B  相似文献   

17.
Abstract: Competition [3H]RX821002 ([3H]2-methoxyidazoxan) binding experiments with α2-adrenoceptor subtype-specific antagonists—BRL 44408 (α2A selective), ARC 239 (α2B selective), and others—were performed to delineate through rigorous computer modeling receptor subtypes in the postmortem human brain. In the hippocampus, hypothalamus, cerebellum, and brainstem the whole population of α2-adrenoceptors appears to belong to the α2A subtype (100%; Bmax = 34–90 fmol/mg of protein). In the frontal cortex, the predominant receptor was the α2A subtype (87%; Bmax = 53 fmol/mg of protein), although a small population of the α2B/C subtype (13%; Bmax = 8 fmol/mg of protein) was also detected. In the caudate nucleus, a mixed population of α2A (64%; Bmax = 9 fmol/mg of protein) and α2B/C (36%; Bmax = 5 fmol/mg of protein) subtypes was detected. In the cortex and caudate and in the presence of ARC 239 (to mask the α2B/C-adrenoceptors), competition experiments with the agonist guanoxabenz clearly modeled the high- and low-affinity states of the α2A subtype. In the presence of ARC 239 and the GTP analogue guanylyl-5′-imidodiphosphate together with NaCl and EDTA (to eliminate the high-affinity α2A-adrenoceptor) guanoxabenz only recognized the low-affinity α2A-adrenoceptor. The results indicate that in the human brain the predominant α2-adrenoceptor is of the α2A subtype and that this functionally relevant receptor subtype is not heterogeneous in nature.  相似文献   

18.
Anther morphology of various species of the genus Triticum is consistent with previous evidence that different biotypes of T. boeoticum (AA) and T. urartu (BB) contributed the respective A and B genomes to the emmer (AeAeBeBe) and timopheevi (AtAtBtBt) tetraploid complexes. Anther length of T. dicoccoides (2.8 mm) and T. araraticum (3.0 mm) is mid-way between that of T. boeoticum (3.6) and T. urartu (2.2) and the same as that of the sterile hybrid T. boeoticum X T. urartu. With respect to mode of dehiscence, post anthesis reflexion of anther lobes and twisting of the anthers, T. dicoccoides resembles T. boeoticum, whereas T. araraticum resembles T. urartu. With respect to anther apex, T. dicoccoides resembles neither T. boeoticum nor T. urartu but is identical with their F1 hybrid. Among amphiploids involving four different Aegilops species and T. boeoticum lines, none could similarly account for the length or other diagnostic attributes of the tetraploid anthers. Anther morphology is consistent with the presumed genomic composition of the cultivated tetraploid and hexaploid wheats and seems to distinguish effectively the different genomic groups of the genus Triticum.  相似文献   

19.
Abstract

A single-point substitution of the O4′ oxygen by a CH2 group at the sugar residue of A 6 (i.e. 2′-deoxyaristeromycin moiety) in a self-complementary DNA duplex, 5′- d(C1G2C3G4A5A6T7T8C9G10C11G12)2 ?3, has been shown to steer the fully Watson-Crick basepaired DNA duplex (1A), akin to the native counterpart, to a doubly A 6:T7 Hoogsteen basepaired (1B) B-type DNA duplex, resulting in a dynamic equilibrium of (1A)→←(1B): Keq = k1/k-1 = 0.56±0.08. The dynamic conversion of the fully Watson-Crick basepaired (1A) to the partly Hoogsteen basepaired (1B) structure is marginally kinetically and thermodynamically disfavoured [k1 (298K) = 3.9± 0.8 sec?1; δH°? = 164±14 kJ/mol;-TδS°? (298K) = ?92 kJ/mol giving a δG298°? of 72 kJ/mol. Ea (k1) = 167±14 kJ/mol] compared to the reverse conversion of the Hoogsteen (1B) to the Watson-Crick (1A) structure [k-1 (298K) = 7.0±0.6 sec-1, δH°? = 153±13 kJ/mol;-TδS°? (298K) = ?82 kJ/mol giving a δG298°? of 71 kJ/mol. Ea (k-1) = 155±13 kJ/mol]. A comparison of δG298°? of the forward (k1) and backward (k-1) conversions, (1A)→←(1B), shows that there is ca 1 kJ/mol preference for the Watson-Crick (1A) over the double Hoogsteen basepaired (1B) DNA duplex, thus giving an equilibrium ratio of almost 2:1 in favour of the fully Watson-Crick basepaired duplex. The chemical environments of the two interconverting DNA duplexes are very different as evident from their widely separated sets of chemical shifts connected by temperature-dependent exchange peaks in the NOESY and ROESY spectra. The fully Watson-Crick basepaired structure (1A) is based on a total of 127 intra, 97 inter and 17 cross-strand distance constraints per strand, whereas the double A 6:T7 Hoogsteen basepaired (1B) structure is based on 114 intra, 92 inter and 15 cross-strand distance constraints, giving an average of 22 and 20 NOE distance constraints per residue and strand, respectively. In addition, 55 NMR-derived backbone dihedral constraints per strand were used for both structures. The main effect of the Hoogsteen basepairs in (1B) on the overall structure is a narrowing of the minor groove and a corresponding widening of the major groove. The Hoogsteen basepairing at the central A 6:T7 basepairs in (1B) has enforced a syn conformation on the glycosyl torsion of the 2′- deoxyaristeromycin moiety, A 6, as a result of substitution of the endocyclic 4′-oxygen in the natural sugar with a methylene group in A 6. A comparison of the Watson-Crick basepaired duplex (1A) to the Hoogsteen basepaired duplex (1B) shows that only a few changes, mainly in α, σ and γ torsions, in the sugar-phosphate backbone seem to be necessary to accommodate the Hoogsteen basepair.  相似文献   

20.
Thermoluminescence and delayed luminescence investigations of the autotrophically and photoheterotrophically cultivated green alga, Chlamydobotrys stellata, demonstrated that both the thermoluminescence and delayed luminescence yields are much lower in the photoheterotophic algae than in the autotrophic ones due to an efficient luminescence quenching of unknown mechanism. The relative contributions of the so called Q (S2Q?A charge recombination) and B (S2Q?B and S3Q?B charge recombinations) thermoluminescence bands to the glow curve as well as the QA(S2Q?B charge recombination) and QB (S2Q?B and S3Q?B charge recombinations) delayed luminescence components to the delayed luminescence decay of autotrophically and photoheterotrophically cultivated Chl. stellata were compared using a computer assisted curve resolution method. It was found that, while in the autotrophic cells the area of the B band was considerably larger than of the Q band, in photoheterotrophic cells the Q band was more effectively charged than the B band. In the delayed luminescence decay curves measured in the seconds to minutes time region the amplitude of the QA component relative to that of the QB component was larger in the photoheterotrophic cells than in the autotrophic ones. These observations demonstrate that, after light-induced charge separation in the photosystem II reaction centers of autotrophic cells, electrons are “quasipermanently” stored mainly in the secondary quinone acceptor pool, QB but in the nonquenched photosystem II reaction centers of photoheterotrophic cells the main reservoir of electrons is the primary quinone acceptor, QA. This behaviour indicates an inhibition of electron transport in the photoheterotrophic alga at the level of the secondary quinone acceptor, QB.  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