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.
An improved protocol for the isolation of DNA from dry material of someHesperis specimens is described. The isolated DNA is suitable for random amplification of polymorphic DNA (RAPD) analysis. Different DNA extraction protocols were examined to determine which might yield DNA from dry leaf tissue ofHesperis specimens. The methods examined include the protocols with hexadecyltrimethylammonium bromide (CTAB) described by Doyle and Doyle (1987); sodium dodecyl sulfate (SDS) by Dellaporta et al. (1983); and CTAB and SDS, the modified minipreparation, by Dellaporta et al (1983). None of these procedures yielded DNA of suitable purity for RAPD assay. We established an improved procedure involving CTAB and enzymatic digestion of proteins and RNA. The recovery of DNA with an average yield of 25 mg/g of leaf material was possible with this procedure. RAPD bands, which could be used to distinguish amongHesperis specimens, were generated.  相似文献   

2.
Summary Pinus radiata is the most important softwood plantation species in Australia and New Zealand. The improtance of this species in forestry has led to an increasing demand to improve the efficiency of selection time of the production population, which currently takes 13 yr by traditional methods. With the application of molecular biology techniques such as random amplified polymorphic DNA (RAPD) the selection period can be reduced to 6 yr. In this study, the conditions for RAPD were optimized and the feasibility of this marker system was investigated with different families ofPinus radiata from Tasmania and South Australia. Best concentrations of Taq-polymerase (1 U), magnesium chloride (2 mM), and template DNA (20 ng) were selected to test different polymerase chain reaction (PCR) thermocycler profiles. Devey's et al. (1996) program was the most effective for production of clear RAPD bands. Best conditions were investigated to screen 10–12 bp arbitrary Breasatec and Operon primers. Both types were found useful at detecting genetic variation between families. Seventy percent and thirty percent of the selected Bresatec and Operon primers, respectively, produced polymorphic bands.  相似文献   

3.
Swensen  Susan M.  Mullin  Beth C. 《Plant and Soil》1997,194(1-2):185-192
Current taxonomic schemes place plants that can participate in root nodule symbioses among disparate groups of angiosperms. According to the classification scheme of Cronquist (1981) which is based primarily on the analysis of morphological characters, host plants of rhizobial symbionts are placed in subclasses Rosidae and Hamamelidae, and those of Frankia are distributed among subclasses Rosidae, Hamamelidae, Magnoliidae and Dilleniidae. This broad phylogenetic distribution of nodulated plants has engendered the notion that nitrogen fixing endosymbionts, particularly those of actinorhizal plants, can interact with a very broad range of unrelated host plant genotypes. New angiosperm phylogenies based on DNA sequence comparisons reveal a markedly different relationship among nodulated plants and indicate that they form a more coherent group than has previously been thought (Chase et al., 1993; Swensen et al., 1994; Soltis et al., 1995). Molecular data support a single origin of the predisposition for root nodule symbiosis (Soltis et al., 1995) and at the same time support the occurrence of multiple origins of symbiosis within this group (Doyle, 1994; Swensen, 1996; Swensen and Mullin, In Press).  相似文献   

4.
Genomic DNA was isolated from frozen needles of maturePinus radiata clones using a modified extraction technique incorporating cetyltrimethylammonium bromide (CTAB) for cell lysis. A high sodium chloride concentration (2 M) was used at 2 stages of the extraction procedure to eradicate polysaccharides, yielding pure genomic DNA suitable for restriction enzyme digestion and PCR amplification. Extractions were scaled down to suit 1.5-mL Eppendorf tubes, allowing easier handling and enhanced sterility.  相似文献   

5.
A methicillin-resistant Staphylococcus aureus (MRSA) strain B-26, isolated clinically in Hiroshima University Hospital, is resistant to bleomycin together with kanamycin. In the present study, we analysed the nucleotide sequence of the 5.1-kb HindIII fragment containing the bleomycin- and kanamycin-resistance genes, which were previously cloned [Bhuiyan et al. (1995) Appl Microbiol Biotechnol 43: 65–69] from the chromosomal DNA of MRSA B-26. The present study found that the DNA sequence contains the duplicated target sequence (GATTAGAT) consisting of 8 bp for transposase and the entire nucleotide sequence of the plasmid pUB110, together with the sequence of inverted repeats (16 bp), designated IR-r and IR-l in IS431mec. The 8-bp duplication sequence, produced by the transposable element, was first found by us. We proposed that bleomycin resistance in MRSA B-26 is attributed to the IS431mec-mediated integration of pUB110 into the chromosome. Received: 13 September 1995/Received last revision: 20 February 1996/Accepted: 30 March 1996  相似文献   

6.
It is shown that step-scan Fourier transform infrared spectroscopy can be applied to resolve the QA QB QAQB transition in Rhodobacter sphaeroides reaction centres with a 5 µs time resolution. In the mid-infrared region (1900 – 1200 cm–1), transient signals previously assigned to QA/B and QA/B vibrations, respectively (Brudler et al. 1994; Brudler et al. 1995; Breton and Nabedryk 1996), can be resolved with this new technique. In addition, the three small positive bands in the spectral region of the carboxylic C=O stretching modes of acidic amino acid side chains are also resolved at 1730, 1719 and 1704 cm–1. A global fit analysis yields two exponentials with half-times of 150 µs and 1.2 ms in agreement with IR spectroscopic studies at single wavenumbers (Hienerwadel et al. 1995), in the UV/VIS and near IR (Tiede et al. 1996, Li et al. 1996). The establishement of the step-scan technique enables a new approach to elucidate the molecular mechanism of this transition.  相似文献   

7.
We have used a [3H] thymidine incorporation assay and microscopic observation in order to reassess recently published data dealing with the response of tobacco protoplasts to phytohormones, lipochitooligosaccharides and peptides ( Harling et al . 1997 ; Hayashi et al . 1992 ; Miklashevichs et al . 1996 ; Miklashevichs et al . 1997 ; Röhrig et al . 1995 ; Röhrig et al . 1996 ; van de Sande et al . 1996 ; Walden et al . 1994 ). These proliferation assays reveal that, in contrast to published data, isolated cells of the investigated mutant plant lines axi159 ( Hayashi et al . 1992 ; Walden et al . 1994 ), axi4/1 ( Harling et al . 1997 ) and cyi1 ( Miklashevichs et al . 1997 ), which were generated by activation T-DNA tagging, were unable to grow in the absence of auxin or cytokinin. Furthermore, lipochitooligosaccharides which play a key role in the induction of nodules on roots of legumes were unable to promote auxin- or cytokinin-independent cell division in tobacco protoplasts as claimed by Röhrig et al . (1995 , 1996 ). The finding of van de Sande et al . (1996 ) that ENOD40 confers tolerance of high auxin concentration to wild-type tobacco protoplasts was also reinvestigated. The results of our investigations show that we were unable to reproduce the proliferation data presented in this study, which were obtained by counting tobacco protoplast-derived cells undergoing division. In total, none of the published data on phytohormone-independent division of tobacco cells could be reproduced.  相似文献   

8.
Summary Ethanol was produced by a strain ofPichia stipitis adapted to an inhibitory acid wood hydrolysate ofPinus radiata. The best ethanol productivity for batch cultures was 0.21 g/l h at 0.7% ethanol. Varying culture conditions increased ethanol concentration to 0.76%, however the productivity decreased to 0.18 g/l h. A decrease in ethanol concentration in the culture fluid was noted late in the batch which suggested ethanol catabolism. Values of kinetic parameters (K m,K s, max, andV max) were evaluated for this system. The use of calcium alginate immobilized cells in a continuous-flow stirred tank reactor lead to enhanced fermentative performance, namely a maximum productivity of 0.27 g/l h and 1.13% ethanol yield. The immobilized cells in continuous flow reactors represent an attractive option for fermenting sugars released by sulphuric acid hydrolysis ofP. radiata wood.  相似文献   

9.
Summary The kinetics of thermal deactivation for thermostable DNA polymerase enzymes were investigated by using the experimental data published elsewhere (Nielson et al. 1996. Strategies. 9, 7–8). The order of deactivation (a) and the deactivation rate constants (k d) were determined for different Taq DNA polymerase enzymes and were found to be of first order.  相似文献   

10.
Now is a very exciting time for researchers in the area of the primary reactions of purple bacterial photosynthesis. Detailed structural information is now available for not only the reaction center (Lancaster et al. 1995, in: Blankenship RE et al. (eds) Anoxygenic Photosynthetic Bacteria, pp 503–526), but also LH2 from Rhodopseudomonas acidophila (McDermott et al. 1995, Nature 374: 517–521) and LH1 from Rhodospirillum rubrum (Karrasch et al. 1995. EMBO J 14: 631–638). These structures can now be integrated to produce models of the complete photosynthetic unit (PSU) (Papiz et al., 1996, Trends Plant Sci, in press), which opens the door to a much more detailed understanding of the energy transfer events occurring within the PSU.Abbreviations Bchl bacteriochlorophyll - LH light-harvesting - PSU photosynthetic unit Division of Biochemistry and Molecular Biology, Institute of Biomedical and Life Sciences  相似文献   

11.
Fluorochrome C-banding ofPinus radiata andP. taeda metaphase chromosomes showed many pericentromeric DAPI bands and interstitial CMA bands inP. radiata, and centromeric and interstitial CMA bands inP. taeda. Giemsa C-band patterns differed between the species with centromeric bands inP. radiata but no consistent bands inP. taeda. A karyotype ofP. radiata was developed based on banding patterns that distinguished all but two of the 12 pairs of chromosomes. In situ hybridization (ISH) using probes for high-copy ribosomal DNA (rDNA) showed 10 pairs of 18S–25S sites and two pairs of 5S sites in both species. Most of the sites were interstitial or centromeric.  相似文献   

12.
Moderate doses of ethanol (1–2 g/kg) markedly increase locomotor activity in some inbred mouse strains, for example, the DBA/2J (D2), but have relatively little effect in other strains, for example, the C57BL/6J (B6). In the present study, we conducted a genome-wide search in a B6D2 F2 intercross (N = 925) for quantitative trait loci (QTLs) associated with the locomotor response. A QTL with a LOD score of 8.4 was detected on Chromosome (Chr) 2; this QTL accounted for 11.4% of the phenotypic variance and approximately 30% of the genetic variance. The QTL on Chr 2 is in the same general region as QTLs previously described for ethanol preference/consumption (Rodriguez et al. Alcohol Clin Exp Res 19, 367, 1995; Melo et al. Nat Genet 13, 147, 1996; Phillips et al. Mamm Genome, in press), acute ethanol withdrawal (Buck et al. J. Neurosci 17, 3946, 1997) and nitrous oxide withdrawal severity (Belknap et al. Behav Genet 23, 213, 1993). A logical candidate gene in the region of interest is the enzyme which synthesizes GABA, glutamic acid decarboxylase 1 (GadI). Received: 15 September 1998 / Accepted: 8 October 1998  相似文献   

13.
 A method to isolate mutants with derepressed expression of cytochrome oxidases and better symbiotic performance is presented. A mutant of Rhizobium etli, CFN030, isolated by its azide-resistant phenotype, was obtained by transposon Tn5-mob mutagenesis. This mutant has a derepressed expression of cytochrome aa3, higher respiratory activities when cultured microaerobically and an improved symbiotic nitrogen fixation capacity. This phenotype was similar to the previously described mutant CFN037, which was isolated by its increased capacity to oxidize N,N,N′,N′-tetramethyl-p-phenylenediamine (TMPD) [Soberón M et al. (1990) J Bacteriol 172:1676–1680]. We show here that although both mutants have a similar symbiotic phenotype, they are affected in different genes. Strain CFN030 has the Tn5 inserted in the chromosome while in strain CFN037 the transposon was located in plasmid b. Cytochrome spectral analysis of both mutant strains in the post-exponential phase of growth, showed the expression of an additional terminal oxidase (cbb3) that is not expressed in the wild-type strain. Received: 10 April 1995/Received revision: 21 August 1995/Accepted: 7 September1995  相似文献   

14.
The Azospirillum brasilense trpE gene has been isolated by DNA hybridization and by genetic complementation of an Escherichia coli trpE deletion mutant. DNA sequence analysis of a 3.1-kb PstI restriction fragment of A. brasilense revealed the presence of an open reading frame encoding a putative TrpE(G) fusion protein. Previously an A. brasilense clone containing trpGDC was identified (Zimmer et al. Mol Gen Genet 229:41–51, 1991). It can, therefore, be concluded that A. brasilense contains two trpG genes. A putative leader peptide is found upstream of trpE(G), containing three consecutive tryptophan residues. Putative terminator and anti-terminator loops have also been identified. The LLESX10S motif, which is responsible for feedback inhibition by tryptophan in other TrpE proteins, is absent in the A. brasilense TrpE(G) fused protein. Received: 29 May 1996 / Accepted: 5 July 1996  相似文献   

15.
DNA sequence analysis of chloroplast genomes has revealed many short nucleotide repeats analogous to nuclear microsatellites, or simple sequence repeats (SSRs). We designed PCR primers flanking five of these regions identified in the chloroplast sequence from Pinus thunbergii and tested them for amplification in Pinus radiata, P. elliotii, P. taeda, P. strobus, Pseudotsuga menziesii, Cupressus macrocarpa, four New Zealand native conifer species (Podocarpus totara, Podocarpus hallii, Podocarpus nivalis, Agathis australis), and four angiosperms (Vitex lucens, Nestegis cunninghamii, Actinidia chinensis, and Arabidopsis thaliana). A PCR product in the expected size range was amplified from all species and interspecific polymorphism was detected at all five loci. Intraspecific polymorphism was detected in P. radiata with four of the five primer pairs. One of these polymorphic chloroplast SSR (cpSSR) was then used to determine the inheritance of chloroplasts in 206 progeny from four control-pollinated, full-sibling P. radiata families. Approximately 99% of the progeny had the cpSSR variant of the pollen parent indicating that in Pinus radiata, like most other conifers, chloroplasts are typically inherited from the paternal parent. These results suggest that polymorphic chloroplast SSRs will be a valuable tool for studying chloroplast diversity, cyto-nuclear disequilibrium, and plastid inheritance in a range of species, and for the analysis of gene flow via pollen and paternity in species with paternal transmission of chloroplasts.  相似文献   

16.
A comparison of cellulose synthesized in vitro from primary walls of etiolated mung bean (Vigna radiata) seedlings and secondary walls of cotton fibers (Gossypium hirsutum) was made by applying conditions found to be essential for in vitro cellulose I assembly from cotton (Kudlicka et al., 1995, Plant Physiology, vol. 107, pp. 111–123). Mung bean fractions including the plasma membrane (PM), the first solubilized fraction (SE1), and the second solubilized fraction (SE2), incorporated more radioactive UDP-Glc into the total product than the same fractions from secondary walls. A significant difference was found with the mild digitonin solubilized fraction (SE1), which produced eight times more total product than the SE1 fraction of cotton. However, the SE1 fraction from cotton produced a larger quantity of cellulose (32.1%) than from mung bean (6.9%). Treatment of the in vitro product by acetic/nitric acid reagent (AN) for varying periods of time demonstrated that cellulose synthesized in vitro from mung bean was more easily degraded than cellulose from cotton fibers. This would suggest that cellulose I produced in vitro from the cotton SE1 fraction may have a higher crystallinity and DP than cellulose I produced in vitro from mung bean. The fibrils of cellulose produced by the SE, fraction of mung bean were loosely associated and not arranged into a compact bundle as in case of cellulose I synthesized by the cotton SE1 fraction. The electron diffraction patterns (ED) of both products show reflections characteristic for cellulose I. Products from the SE2 fraction of mung bean and cotton reveal similarities with the cellulose II allomorph synthesized, as well as abundant β-1,3-glucan.  相似文献   

17.
Most receptor-like protein tyrosine phosphatases (PTPases) display a high degree of homology with cell adhesion molecules in their extracellular domains. We studied the functional significance of processing for the receptor-like PTPases LAR and PTPσ. PTPσ biosynthesis and intracellular processing resembled that of the related PTPase LAR and was expressed on the cell surface as a two-subunit complex. Both LAR and PTPσ underwent further proteolytical processing upon treatment of cells with either calcium ionophore A23187 or phorbol ester TPA. Induction of LAR processing by TPA in 293 cells did require overexpression of PKCα. Induced proteolysis resulted in shedding of the extracellular domains of both PTPases. This was in agreement with the identification of a specific PTPσ cleavage site between amino acids Pro821 and Ile822. Confocal microscopy studies identified adherens junctions and desmosomes as the preferential subcellular localization for both PTPases matching that of plakoglobin. Consistent with this observation, we found direct association of plakoglobin and β-catenin with the intracellular domain of LAR in vitro. Taken together, these data suggested an involvement of LAR and PTPσ in the regulation of cell contacts in concert with cell adhesion molecules of the cadherin/catenin family. After processing and shedding of the extracellular domain, the catalytically active intracellular portions of both PTPases were internalized and redistributed away from the sites of cell–cell contact, suggesting a mechanism that regulates the activity and target specificity of these PTPases. Calcium withdrawal, which led to cell contact disruption, also resulted in internalization but was not associated with prior proteolytic cleavage and shedding of the extracellular domain. We conclude that the subcellular localization of LAR and PTPσ is regulated by at least two independent mechanisms, one of which requires the presence of their extracellular domains and one of which involves the presence of intact cell–cell contacts. A key element in the regulation of cell–cell and cell– matrix contacts is the tyrosine phosphorylation of proteins that are localized in focal adhesions and at intercellular junctions (for reviews see Kemler, 1993; Clark and Brugge, 1995). While much is known about the protein tyrosine kinases involved in the phosphorylation of cell adhesion components, very little information exists about the identity of protein tyrosine phosphatases (PTPases),1 which are responsible for the dephosphorylation and thereby regulation of these structural complexes. Probable candidates are those receptor-like PTPases that contain cell adhesion molecule-like extracellular domains and could therefore regulate their intrinsic phosphatase activity in response to cell contact. Recent reports suggest that some PTPases do, in fact, possess properties that resemble those of classical cell adhesion molecules (for review see Brady-Kalnay and Tonks, 1995). A direct involvement in cell–cell contact has so far been demonstrated for PTPμ (Brady-Kalnay et al., 1993; Gebbink et al., 1993) and PTPκ (Sap et al., 1994), for which a homophilic interaction between their extracellular domains was found. The localization of PTPμ (Brady-Kalnay et al., 1995; Gebbink et al., 1995), PTPκ (Fuchs et al., 1996), and PCP-2 (Wang et al., 1996) was restricted to sites of cell–cell contact and surface expression of PTPμ (Gebbink et al., 1995), and PTPκ (Fuchs et al., 1996) was increased in a cell density-dependent manner. Moreover, a direct association of PTPκ (Fuchs et al., 1996) and PTPμ (Brady-Kalnay et al., 1995) with members of the cadherin/catenin family suggests that proteins of the cell adhesion complex represent physiological substrates for these PTPases. A possible regulatory function in cell–matrix adhesion has been proposed for LAR, another receptor-like PTPase, which associated with focal cell–substratum adhesions via the newly identified LAR interacting protein 1, LIP-1 (Serra-Pages et al., 1995).PTPμ (Gebbink et al., 1991), PTPκ (Jiang et al., 1993; Fuchs et al., 1996), PTPδ (Krueger et al., 1990; Mizuno et al., 1993, Pulido et al., 1995a), PCP-2 (Wang et al., 1996), and LAR (Streuli et al., 1988, Pot et al., 1991) are members of the so-called type II receptor-like PTPases. The extracellular domains of these PTPases contain a variable number of Ig-like and fibronectin type III-like (FNIII) domains (for review see Charbonneau and Tonks, 1992). With the exception of PCP-2 (Wang et al., 1996), these PTPases also share characteristics in their biosynthesis. They all underwent proteolytic processing by a furin-like endoprotease and were expressed at the cell surface in two subunits which were not covalently linked (Streuli et al., 1992; Yu et al., 1992; Jiang et al., 1993; Brady-Kalnay and Tonks, 1994; Gebbink et al., 1995; Pulido et al., 1995a; Fuchs et al., 1996). It was shown for LAR that the E subunit, which contains the cell adhesion molecule-like extracellular domain, was shed from the cell surface when cells were grown to a high density (Streuli et al., 1992). This shedding of the E subunit of LAR was the result of an additional proteolytic processing step that could also be induced by treatment of the cells with the phorbol ester TPA (Serra-Pages et al., 1995). An accumulation of E subunits in the supernatant of cells was also observed for PTPμ (Gebbink et al., 1995) and PTPδ (Pulido et al., 1995a), and this suggests a common mechanism in the regulation of type II PTPases. However, the effect of proteolytic processing on either the catalytic activity, the substrate specificity, or the cellular localization of these PTPases has not yet been determined. We report here that PTPσ, a recently identified new member of the family of receptor-like type II PTPases (Pan et al., 1993; Walton et al., 1993; Yan et al., 1993; Ogata et al., 1994; Zhang et al., 1994), underwent biosynthesis and proteolytic processing in a manner that resembled that of the most closely related PTPase LAR. Moreover, further proteolytic processing of PTPσ as well as of LAR could be induced by treatment of the cells with TPA or the calcium ionophore A23187. Transient expression studies indicated that TPA-induced processing of LAR, but not PTPσ, was dependent on the coexpression of PKCα. Inducible processing of both PTPases took place in the extracellular segment of the P subunit in a juxtamembrane position and led to the shedding of the E subunit. Both LAR and PTPσ were predominantly localized in regions of cell–cell contact and accumulated in dot-like structures that could be identified as adherens junctions and desmosomes by colocalization with plakoglobin (Cowin et al., 1986). Moreover, plakoglobin and β-catenin, another component of E-cadherin–containing cell adhesion complexes in adherens junctions, associated directly with the intracellular domain of LAR in vitro. The inducible shedding of the E subunit of LAR and PTPσ was followed by a redistribution of the PTPases within the cell membrane and by an internalization of the cleaved P subunits. It therefore represents a mechanism through which the phosphatase activity of these PTPases could be regulated in response to cell–cell contact. The cell adhesion molecule-like character of LAR and PTPσ was further supported by the fact that the internalization of LAR and PTPσ occurred independently of the proteolytic processing if cells were grown in calcium-depleted growth medium. The analogies in specific localization as well as internalization behavior of PTPσ and LAR, with molecules of the cadherin/catenin family, strongly suggest a direct involvement of PTPσ and LAR in the formation or maintenance of intercellular contacts.  相似文献   

18.
A biolistic particle delivery system was used to genetically transform embryogenic tissue of Pinus radiata. The introduced DNA contained a uidA reporter gene under the control of either the tandem CaMV 35S or the artificial Emu promoter, and the npt II selectable marker controlled by the CaMV 35S promoter. The average number of stable, geneticin-resistant lines recovered was 0.5 per 200 mg fresh weight bombarded tissue. Expression of the uidA reporter gene was detected histochemically and fluorimetrically in transformed embryogenic tissue and in derived mature somatic embryos and regenerated plants. The integration of uidA and npt II genes into the Pinus radiata genome was demonstrated using PCR amplification of the inserts and Southern hybridisation analysis. The expression of both genes in transformed tissue was confirmed by Northern hybridisation analysis. More than 150 transgenic Pinus radiata plants were produced from 20 independent transformation experiments with four different embryogenic clones. Received: 9 May 1997 / Revision received: 18 September 1997 / Accepted: 18 October 1997  相似文献   

19.
New Books     
Factor , J. R. (ed) 1995: Biology of the lobster Homarus americanus. Hölldobler, B. & Wilson, E. O. 1995: Ameisen. Terman , M. R. 1996: Messages from an owl. Williams , T. D. 1995: The penguins. With contributions by R. P. Wilson , P. D. Boersma & D. L. Stokes . Leyhausen , P. 1996: Katzenseele. Wesen und Sozialverhalten. Alterman, L., Doyle, G. A. & Izard , M. K. (eds) Heymer, A. 1995: Die Pygmäen. Menschenforschung im afrikanischen Regenwald.  相似文献   

20.
A new record of a fossil Monterey pine (Pinus radiata D. Don) is recorded from rocks of the Plio-Pleistocene age, southern California, and the affinities of other presumed fossil closed-cone pines are revised. This has led to the recognition of two new pine species, Pinus storeyana Axelrod of subsect. Oocarpae and Pinus verdiana Axelrod of subsect. Ponderosae. Geologic evidence indicates that forerunners of the present California and Mexican species of Oocarpae were in proximity in Mexico during the Miocene. They were isolated following opening of the Gulf of California, the northward movement of Baja and Alta California on the San Andreas fault system, and the spread of desert climate. New populations of Pinus radiata developed during the Pliocene (5–2 million years) and later as environmental diversity increased. This also enabled Pinus remorala Mason and P. muricata D. Don, which the fossil record suggests were distinct into the Late Pleistocene, to hybridize in the new topographic-climatic Holocene environments.  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