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 62 毫秒
1.
《BBA》2023,1864(3):148976
Ultrafast transient absorption (TA) spectroscopy was used to study electron transfer (ET) at 100 K in native (as isolated) reaction centers (RCs) of the green filamentous photosynthetic bacterium Chloroflexus (Cfl.) aurantiacus. The rise and decay of the 1028 nm anion absorption band of the monomeric bacteriochlorophyll a molecule at the BA binding site were monitored as indicators of the formation and decay of the P+BA state, respectively (P is the primary electron donor, a dimer of bacteriochlorophyll a molecules). Global analysis of the TA data indicated the presence of at least two populations of the P excited state, which decay by distinct means, forming the state P+HA (HA is a photochemically active bacteriopheophytin a molecule). In one population (~65 %), P decays in ~2 ps with the formation of P+HA via a short-lived P+BA intermediate in a two-step ET process P → P+BA→ P+HA. In another population (~35 %), P decays in ~20 ps to form P+HA via a superexchange mechanism without producing measurable amounts of P+BA. Similar TA measurements performed on chemically modified RCs of Cfl. aurantiacus containing plant pheophytin a at the HA binding site also showed the presence of two P populations (~2 and ~20 ps), with P decaying through P+BA only in the ~2 ps population. At 100 K, the quantum yield of primary charge separation in native RCs is determined to be close to unity. The results are discussed in terms of involving a one-step P → P+HA superexchange process as an alternative highly efficient ET pathway in Cfl. aurantiacus RCs.  相似文献   

2.
We aimed to find out relations among nonphotochemical quenching (NPQ), gross photosynthetic rate (P G), and photoinhibition during photosynthetic light induction in three woody species (one pioneer tree and two understory shrubs) and four ferns adapted to different light regimes. Pot-grown plants received 100% and/or 10% sunlight according to their light-adaptation capabilities. After at least four months of light acclimation, CO2 exchange and chlorophyll fluorescence were measured simultaneously in the laboratory. We found that during light induction the formation and relaxation of the transient NPQ was closely related to light intensity, light-adaption capability of species, and P G. NPQ with all treatments increased rapidly within the first 1–2 min of the light induction. Thereafter, only species with high P G and electron transport rate (ETR), i.e., one pioneer tree and one mild shade-adapted fern, showed NPQ relaxing rapidly to a low steady-state level within 6–8 min under PPFD of 100 μmol(photon) m?2 s?1 and ambient CO2 concentration. Leaves with low P Gand ETR, regardless of species characteristics or inhibition by low CO2 concentration, showed slow or none NPQ relaxation up to 20 min after the start of low light induction. In contrast, NPQ increased slowly to a steady state (one pioneer tree) or it did not reach the steady state (the others) from 2 to 30 min under PPFD of 2,000 μmol m?2 s?1. Under high excess of light energy, species adapted to or plants acclimated to high light exhibited high NPQ at the initial 1 or 2 min, and showed low photoinhibition after 30 min of light induction. The value of fastest-developing NPQ can be quickly and easily obtained and might be useful for physiological studies.  相似文献   

3.
The somatic and gonad productions of the cirolanid isopod Excirolana armata were analyzed by taking monthly samples from December 2003 to November 2005 on Una beach, S?o Paulo state (24°S), southeastern Brazil. Sampling was performed along three fixed transects established from the base of the foredunes to the waterline. Weight-specific growth rate was used to estimate the E. armata somatic production for 2004 and 2005, separately. The gonad production was estimated based on the monthly reproductive potential (mean number of eggs/embryos per female × monthly abundance of ovigerous females with near-release broods) for 2004. The annual somatic production of E. armata population varied from 15.57 to 17.25?g AFDW m?1?year?1 and the somatic production/biomass ratio (P s/B) from 3.55 to 3.14?year?1 for 2004 and 2005, respectively. The P s/B ratios were higher for males (4.02 and 3.19?year?1 for 2004 and 2005) than for females (3.10?year?1 for both years). The annual gonad production (P g?=?1.07?g AFDW m?1?year?1) contributed about 15 and 6% to the total production (P s?+?P g) of females and the population, respectively. The proportion of gonad to somatic production of females (P g/P s) increased with individual size (ca 90% in the 7.5?mm size class), and the annual weight-specific gonad production (P g/B ratio) was estimated to 0.24?year?1. The high P s/B ratios estimated for E. armata derive from the fast growth of individuals and show the importance of this population to the energy flow on Una beach ecosystem. However, the low percentage of juveniles verified in this population and in other studies of populations of the genus Excirolana is discussed as an important source of underestimation of P s/B ratio.  相似文献   

4.
A capacitor microphone was used to measure the enthalpy and volume changes that accompany the electron transfer reactions, PQAhv P+Q?A and PQAQBhv P+QAQ?B, following flash excitation of photosynthetic reaction centers isolated from Rhodopseudomonas sphaeroides. P is a bacteriochlorophyll dimer (P-870), and QA and QB are ubiquinones. In reaction centers containing only QA, the enthalpy of P+Q?A is very close to that of the PQA ground state (ΔHr = 0.05 ± 0.03 eV). The free energy of about 0.65 eV that is captured in the photochemical reaction evidently takes the form of a substantial entropy decrease. In contrast, the formation of P+QAQ?B in reaction centers containing both quinones has a ΔHr of 0.32 ± 0.02 eV. The entropy change must be near zero in this case. In the presence of o-phenanthroline, which blocks electron transfer between Q?A and QB, ΔHr for forming P+Q?AQB is 0.13 ± 0.03 eV. The influence of flash-induced proton uptake on the results was investigated, and the ΔHr values given above were measured under conditions that minimized this influence. Although the reductions of QA and QB involve very different changes in enthalpy and entropy, both reactions are accompanied by a similar volume decrease of about 20 ml/mol. The contraction probably reflects electrostriction caused by the charges on P+ and Q?A or Q?B.  相似文献   

5.
Primary charge separation dynamics in four mutant reaction centers (RCs) of the purple bacterium Rhodobacter sphaeroides with increased midpoint potential of the primary electron donor P (M160LH, L131LH, M197FH, and M160LH + L131LH + M197FH) have been studied by femtosecond transient absorption spectroscopy at room temperature. The decay of the excited singlet state in the wild-type and mutant RCs is complex and has two main exponential components, which indicates heterogeneity of electron transfer rates or the presence of reverse electron transfer reactions. The radical anion band of monomeric bacteriochlorophyll BA at 1020 nm was first observed in transient absorbance difference spectra of single mutants. This band remains visible, although with somewhat reduced amplitude, even at delays up to tens of picoseconds when stimulated emission is absent and the reaction centers are in the P+H A ? state. The presence of this band in this time period indicates the existence of thermodynamic equilibrium between the P+B A ? HA and P+BAH A ? states. The data give grounds for assuming that the value of the energy difference between the states P*, P+B A ? HA, and P+BAH A ? at early times is of the same order of magnitude as the energy kT at room temperature. Besides, monomeric bacteriochlorophyll BA is found to be an immediate electron acceptor in the single mutant RCs, where electron transfer is hampered due to increased energy of the P+B A ? state with respect to P*.  相似文献   

6.
The intermediates, U and V, of a simple model reaction system (the so-called Brusselator) are shown to have both uniform and patterned steady states in a membrane. The patterned distribution of U and V has the symmetry of hexagons in the plane and is stable for B #62; Bm, B being the controlled concentration of one of the parametric species. The uniform solution is stable for B < Bc, and, since Bc #62; Bm, the increase and subsequent decrease of B over an interval that includes (Bm,Bc) can produce a hysteresis in which the transitions from mixed (i.e. uniform) to unmixed (i.e. hexagonlly patterned) states happen quite suddenly at the two critical values. Since the transfer or reactive properties of the membrane might be different in the two states, this model provides a bistable membrane element that might have prototypical value in a theory of biochemical control.  相似文献   

7.
In order to develop a multi-microbe probiotic preparation of Lactobacillus reuteri G8-5 and Bacillus subtilis MA139 in solid-state fermentation, a series of parameters were optimized sequentially in shake flask culture. The effect of supplementation of B. subtilis MA139 as starters on the viability of L. reuteri G8-5 was also explored. The results showed that the optimized process was as follows: water content, 50 %; initial pH of diluted molasses, 6.5; inocula volume, 2 %; flask dry contents, 30~35 g/250 g without sterilization; and fermentation time, 2 days. The multi-microbial preparations finally provided the maximum concentration of Lactobacillus of about 9.01?±?0.15 log CFU/g and spores of Bacillus of about 10.30?±?0.08 log CFU/g. Compared with pure fermentation of L. reuteri G8-5, significantly high viable cells, low value of pH, and reducing sugar in solid substrates were achieved in mixed fermentation in the presence of B. subtilis MA139 (P?<?0.05). Meanwhile, the mixed fermentation showed the significantly higher antimicrobial activity against E. coli K88 (P?<?0.05). Based on the overall results, the optimized process enhanced the production of multi-microbe probiotics in solid-state fermentation with low cost. Moreover, the viability of L. reuteri G8-5 could be significantly enhanced in the presence of B. subtilis MA139 in solid-state fermentation, which favored the production of probiotics for animal use.  相似文献   

8.
In the classical tadpole assay employed for Meyer-Overton type correlations, tricaine and benzocaine are 7–8 times more potent than n-alkanols of equivalent lipid solubility. Median anesthetic concentrations for loss of the righting reflex, AC50(RR)s, were 0.165 and 0.103 mM for tricaine and benzocaine, and log Ps (octanol: water) were 1.81 and 1.98, respectively. Tadpoles receiving a half AC50(RR) each of tricaine and ethanol showed less than additive effects, suggesting a substantial difference in their mechanism(s) of action. AC50(RR)s for both tricaine and benzocaine increased two fold coincident with metamorphosis, reflecting a change from an equilibrium to a nonequilibrium steady state.  相似文献   

9.
It is shown that large-amplitude stationary solitary electrostatic waves can exist in a symmetric plasma (e ? e + or C 60 ? C 60 + ), and the relevant parameter ranges (i.e., the range of the Mach numbers and the degree to which the plasma should be nonequilibrium) are determined. The basic requirement for the existence of such waves, specifically, that the symmetric plasma be in a nonequilibrium state, can easily be satisfied in low-density collisionless ideal plasmas under laboratory conditions.  相似文献   

10.
In framework of the continuum electrostatics theory, the reorganization energies of the electron transfers QA?–QB (fast phase), Bph?–QA, P+–QA?, and P+–QB? in the photosynthetic bacterial reaction center have been calculated. The calculations were based on the static dielectric permittivity spatial distribution derived from the data on the electrogenesis, with the corresponding characteristic times relatively close to the reaction times of QA?–QB (fast phase) and Bph?–QA but much shorter than those times of the latter two recombination reactions. The calculated reorganization energies were reasonably close to the experimental estimates for QA?–QB (fast phase) and Bph?–QA but substantially lower than those of P+–QA? and P+–QB?. A higher effective dielectric permittivity contributes to this effect, but the dominant contribution is most probably made by a non-dielectric relaxation, especially for the P+–QB? recombination influenced by the proton transfer. This situation calls for reconsidering of the current electron transfer rate estimates.  相似文献   

11.
The small-signal admittance of membranes associated with channel conduction is derived for a general channel model. A general channel model is represented by a set of chemical reactions with each species of the reactions representing a channel state. The membrane admittance is shown to be related to the phenomenological relaxation matrix of the reactions. If the kinetic reactions are at a non-equilibrium steady state, the relaxation matrix may have complex eigenvalues and the equivalent circuit of the membrane admittance may contain RLC or RLC-like branches. For equilibrium kinetic systems, on the other hand, the equivalent circuit contains only RL or RC branches. Thus, the membrane admittance of equilibrium channels is quite different from that of non-equilibrium channels. In particular, we show that the low frequency feature in the admittance of squid axons as observed by Fishman, Poussart, Moore &; Siebenga (1977) can be obtained easily from a non-equilibrium cycling steady-state model.  相似文献   

12.
It is experimentally shown that slow chilling of photosynthetic reaction centers from purple bacteria Rhodobacter sphaeroides to <230 K under intense illumination leads to appearance of long-lived chargeseparated states (P+Q A ? ). This longevity implies that the recombination is blocked or the charge-separated state is stabilized. The longevity effect is caused by structural relaxation of reaction centers to a new equilibrium state that diminishes the free energy difference of recombination. The possible mechanism of such relaxation involves orientation of the polar water molecules in the semiquinone local electrostatic field. Detailed analysis of the longevity effect has been carried out, and its outcome supports the idea that many electron transfer reactions in biological systems are non-equilibrium.  相似文献   

13.
The formation of chlorophyll triplet states during illumination of Photosystem I reaction center samples depends upon the redox state of P-700, X and ferredoxin Centers A and B. When the reaction centers are in the states P-700+A1XFdBFd?A and P-700 A1XFd?BFd?A prior to illumination, we observe electron paramagnetic resonance (EPR) spectra from a triplet species which has zero-field splitting parameters (|D| and |E|) larger than those of either the chlorophyll a or chlorophyll b monomer triplet, and a polarization which results from population of the triplet spin sublevels by an intersystem crossing mechanism. We interpret this triplet as arising from photoexcited chlorophyll antenna species associated with reaction centers in the states P-700+Fd?A and P-700+X?, respectively, which undergo de-excitation via intersystem crossing. When the reaction centers are in the states P-700A1XFd?BFd?A and P-700A1X?Fd?BFd?A prior to illumination, we observe a triplet EPR signal with a polarization which results from population of the triplet spin sublevels by radical pair recombination, and which has a |D| value similar to that of chlorophyll a monomer. We interpret this triplet (the radical pair-polarized triplet) as arising from 3P-700 which has been populated by the process P-700+A?13P-700A1. We observe both the radical pair-polarized triplet and the chlorophyll antenna triplet when the reaction centers are in the state P-700 A1XFd?BFd?A, presumably because the processes P-700+A?1X → P-700+A1X? and P-700+A?1X3P-700 A1X have similar rate constants when Centers A and B are reduced, i.e., the forward electron transfer time from A?1 to X is apparently much slower in the redox state P-700 A1XFd?BFd?A than it is in state P-700 A1XFdBFdA. The amplitude of the radical pair-polarized triplet EPR signal does not decrease in the presence of a 13.5-G-wide EPR signal centered at g 2.0 which was recorded in the dark prior to triplet measurements in samples previously frozen under intense illumination. This g 2.0 signal, which has been attributed to phototrapped A?1 (Heathcote, P., Timofeev, K.N. and Evans, M.C.W. (1979) FEBS Lett. 101, 105–109), corresponds to as many as 12 spins per P-700 and can be photogenerated during freezing without causing any apparent attenuation of the radical pair-polarized triplet amplitude. We conclude that species other than A?1 contribute to the g 2.0 signal.  相似文献   

14.
D. Kleinfeld  M.Y. Okamura  G. Feher 《BBA》1984,766(1):126-140
The electron-transfer reactions and thermodynamic equilibria involving the quinone acceptor complex in bacterial reaction centers from R. sphaeroides were investigated. The reactions are described by the scheme: We found that the charge recombination pathway of D+QAQ?B proceeds via the intermediate state D+Q?AQB, the direct pathway contributing less than approx. 5% to the observed recombination rate. The method used to obtain this result was based on a comparison of the kinetics predicted for the indirect pathway (given by the product kAD-times the fraction of reaction centers in the Q?AQB state) with the observed recombination rate, kobsD+ →D. The kinetic measurements were used to obtain the pH dependence (6.1 ? pH ? 11.7) of the free energy difference between the states Q?AQB and QAQ?B. At low pH (less than 9) QAQ?B is stabilized relative to Q?AQB by 67 meV, whereas at high pH Q?AQB is energetically favored. Both Q?A and Q?B associate with a proton, with pK values of 9.8 and 11.3, respectively. The stronger interaction of the proton with Q?B provides the driving force for the forward electron transfer.  相似文献   

15.
A theoretical analysis of non-stationary states in membrane pores is given in this paper, which is based on the rate-theory treatment of transport processes. The principal aim of this study is to give a basis for the interpretation of relaxation experiments in which an external parameter, such as the voltage across the membrane, is suddenly displaced. From the time course of the membrane current information about the microscopic properties of the pore may be obtained. The pore is considered as a sequence of binding sites, separated by energy barriers over which the ion has to jump. It is found that under certain conditions damped oscillations occur after the initial perturbations of the membrane. In all other cases the approach towards the steady state may be described by a discrete spectrum of n relaxation times, where n is the number of binding sites within the pore. In the case of a pore with regular energy profile (internal barriers of identical height) the relaxation times may be obtained as the roots of Tchebycheff polynomials for arbitrary n. It is shown that the present treatment becomes identical with the continuum analysis of transport processes in the limit of large n.  相似文献   

16.
The above data relating to the reaction between 16 hour cultures of S. aureus and antistaphylococcus bacteriophage in nutrient broth of pH 7.6 at 36°C. and with mechanical shaking to maintain a uniform B suspension, bring out the following points: (a) B growth in P-B mixtures does not differ from growth in controls without P except in the case of a very high initial P/B ratio as noted below. There is no evidence that lytic destruction of B begins shortly after mixing P and B nor that B growth is stimulated by P, for the B growth curves in the presence of ordinary [P]''s and in controls are identical. Only at the sudden onset of the rapid lytic process does the B curve of a P-B mixture deviate from the control curve. (b) B growth is an essential conditioning factor for P formation. (c) Both B growth and P production exhibit short lags. During this time P diffuses into or becomes adsorbed to B so rapidly that by the end of the lag period only 10 to 30 per cent of the total P present is extracellular, the remainder being associated with the B. (d) During the logarithmic B growth phase, P formation is also logarithmic but proceeds at a much faster rate. That is, d P/d t is proportional to a power of d B/d t. Consequently the statement that each time a B divides a certain amount of P is formed is not correct. (e) As B growth enters the phase of positive acceleration equilibrium between the extracellular and intracellular P fractions becomes established and is maintained up to the onset of lysis, extracellular [P] representing a small constant percentage of total [P]. The distribution of P on a constant percentage basis suggests the manner in which a relatively simple chemical compound would be distributed and is not at all typical of the distribution one would expect if P were a complex organized parasite. (f) When the value of log P/B = 2.1 lysis begins. Obviously, this limiting value for any initial [B] is reached sooner the higher the initial [P]. When log P/B at the time of mixing P and B is already 2.1 or greater, there is no growth of B and lysis soon occurs. (g) While there is good evidence that lysis is brought about by the attainment of a particular [P] per B and not by a certain [P] per ml., it is not clear at this time which of the ratios intracellular P/B, extracellular P/B or total P/B is the major conditioning factor for B lysis. (h) Experimentally the maximal [P]''s of lysates made by mixing a constant initial [B] with widely varying Po''s fall within a relatively narrow range. This fact is explained by the large value of d log P/d t as compared to d log B/d t. That is, the loci of points at which log P = 2.1 + log B (maxima-lysis begins) on the curves of log P against t originating in various [Po]''s will lie at a nearly constant level above the abscissa. Because of this same relationship the maximal [P]''s of such a series will be in the reverse order of magnitude of the Po''s, i.e., the larger the Po the smaller will be the maximal [P] attained during the reaction (cf. Fig, 16). (i) The lytic destruction of B is logarithmic with time, in this respect being similar to most death rate processes. The value -d log B/d t for a particular initial [B] is constant for various initial values of [P]. There is good evidence that cells need not be growing in order to undergo lysis. (j) During B lysis a considerable percentage of the total maximal P formed is destroyed, the chief loss probably occurring in the intracellular fraction. The major portion (70 to 90 per cent) of the final P present after the completion of bacteriophagy is set free during the brief phase of bacterial dissolution. (k) When the entire process of bacteriophagy is completed the lysates are left with certain [P]''s determined by the foregone P-B reaction. The destruction of P during lysis is sufficiently regular to maintain the relationship established at the maximal [P]''s. Therefore the final [P]''s have the same points in common that were noted in "h" as applying to the maximal [P]''s. That is, they all are grouped within a narrow range of [P] values, those having been made with high Po''s being of lower titre than those made with low initial [P]''s. (1) There is a significant difference in the temperature coefficients of P and B formation. Further, the temperature coefficients of P and B destruction during lysis differ in almost the same ratio. Consequently, while all experimental evidence postulates B growth as an essential conditioning factor for P formation, the temperature coefficient data suggest that the two processes are basically separate reactions. A similar interpretation holds in the case of B dissolution and P inactivation. (m) The major events in the complete process of "bacteriophagy" are mathematically predictable. The [B] at which lysis occurs under certain standard conditions for given values of Bo and Po may be calculated from the equation: See PDF for Equation Substitution of this value for log B in the equation: See PDF for Equation gives satisfactory agreement with observed values for t (lysis). (n) The kinetic analysis of the P-B reaction predicts that the values of log Po plotted against t (lysis) for a constant Bo will give a straight line. This plot is employed in a method for the quantitative estimation of P described in an earlier paper on the basis of experimental observation alone. Its use is made more rational by the facts given above.  相似文献   

17.
S.K. Chamorovsky  R. Cammack 《BBA》1982,679(1):146-155
When spinach Photosystem I particles, frozen in the dark with ascorbate, are illuminated at low temperatures, one electron is transferred from P-700 to either iron-sulphur centre A or B. It was found that the proportion of centre A or B reduced depended on the temperature of illumination. At 25 K, reduction of centre A, as detected by ESR spectroscopy, was strongly preferred. At higher temperatures, at about 150K, there was an increased proportion of reduced centre B. Reduction of B was more strongly preferred in particles frozen in 50% glycerol. The kinetics of dark reoxidation of A? and B? at various temperatures were followed by observing the radical signal of P-700+, and also by periodically cooling to 25 K to measure the ESR spectra of the iron-sulphur centres. The recombination of A? and P-700+ occurred at lower temperatures than that at of B?; at 150–200 K, centre B was the more stable electron trap. Dark reoxidation of both centres was more rapid in samples that were illuminated at 25 K than in samples illuminated at 150–215 K. In no case was net electron transfer between centres A and B observed. Differences in g values of the ESR spectra in particles illuminated at 25 and 200 K indicate that the iron-sulphur centres are in altered conformational states. It is concluded firstly that, in the frozen state, the rates of dark electron transfer decrease in the sequence A?P-700+ > B?P-700+ > B? → A; secondly, that when centres A or B are photoreduced, a temperature-dependent conformational change takes place which slows down the rate of recombination with P-700+.  相似文献   

18.
The maintenance requirement(s) exhibited by Bacillus sphaericus1593M was clearly established from the data on the steady state continuous cultures wherein the growth yield was found to be increasing with increase in dilution rate (4). In the present study we have attempted to estimate the maintenance coefficient of B. sphaericus 1593M by achieving zero growth rate in a total cell retention culture (TCRC). Zero growth rate was achieved at a cell density of 5.57?g?l?1 which remained constant reflecting a steady state in the TCRC. Based on the substrate flux calculations, maintenance coefficient was found to be 0.087?h?1. Further, using this data of maintenance coefficient, steady state cell density and substrate profiles in continuous cultures of B. sphaericus 1593M were predicted by simulation, which showed a close relationship to that of the experimental data.  相似文献   

19.
The second messenger nucleotide cyclic diadenylate monophosphate (c-di-AMP) has been identified in several species of Gram positive bacteria and Chlamydia trachomatis. This molecule has been associated with bacterial cell division, cell wall biosynthesis and phosphate metabolism, and with induction of type I interferon responses by host cells. We demonstrate that B. burgdorferi produces a c-di-AMP synthase, which we designated CdaA. Both CdaA and c-di-AMP levels are very low in cultured B. burgdorferi, and no conditions were identified under which cdaA mRNA was differentially expressed. A mutant B. burgdorferi was produced that expresses high levels of CdaA, yet steady state borrelial c-di-AMP levels did not change, apparently due to degradation by the native DhhP phosphodiesterase. The function(s) of c-di-AMP in the Lyme disease spirochete remains enigmatic.  相似文献   

20.
The genetic relationships and diversity within the European and Asiatic Buxus species were analysed using AFLP, genome size analysis and chromosome counts. Based on these results two major clusters could be defined. One genetic cluster contained B. sempervirens and B. balearica, European species, and B. colchica, an Asiatic species but with leaf morphology similar to B. sempervirens. Species in this cluster were characterised by a genome size between 1.38 and 1.69?pg?2C?1 and a chromosome number of 2n?=?2x?=?28 (diploid). Only four B. sempervirens cultivars within this cluster were triploid. A second cluster contained the Asiatic Buxus species B. microphylla, B. harlandii, B. hyrcana, B. myrica, B. henryi, B. bodinieri and B. wallichiana. Within this second genetic cluster three different ploidy levels could be observed. B. harlandii, B. hyrcana and nine B. microphylla cultivars were tetraploid (2n?=?4x?=?56) with a genome size of >2.5?pg?2C?1. Fifteen other B. microphylla cultivars were triploid (2n?=?3x?=?42). The other Asiatic Buxus species, B. henryi, B. bodinieri and eight B. microphylla cultivars, were diploid with a genome size of ca. 1.5?pg?2C?1.  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