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.
Hydrobiologia 361: 169–189, 1998. Due to technical problems, Figures 73–84 and Figures 85–94 were unfortunately omitted from pages 182 and 183, respectively, of the above volume, duplicates of Figures 1–6 appearing in their place. We now present the correct Figures with their captions, on pages 214 and 215 of this volume. Additionally there are four corrections to Table 1, which appeared on pages 186-188  相似文献   

2.
Srebrow A, Friedmann Y, Ravanpay A, Daniel CW, Bissell MJ (1998): Expression of Hoxa-1 and Hoxb-7 is regulated by extracellular matrix-dependent signals in mammary epithelial cells. J Cell Biochem 69:377–391. In Figure 3 on pages 384 and Figure 4 on page 385, two labels were misprinted. The top label on the right side of Figure 3B should have been Hoxb-7 instead of Hoxb-1, and the center label of Figure 4B should have been Hoxb-7 instead of Hoxa-7. The corrected figures are reprinted on the following pages. The Publisher apologizes for the error.  相似文献   

3.
Using a combination of laser photolysis and temperature-jump techniques, the kinetics of hemoglobin S polymerization have been studied over a wide range of delay times (10(-3) to 10(5)s), concentrations (0.2 to 0.4 g/cm3) and temperatures (5 to 50 degrees C). A slow temperature-jump technique was used to induce polymerization in samples with delay times between 10(2) seconds and 10(5) seconds by heating a solution of completely deoxygenated hemoglobin S. For samples with shorter delay times, polymerization was induced by photodissociating the carbon monoxide complex in small volumes (10(-9) cm3) using a microspectrophotometer equipped with a cw argon ion laser. The photolysis technique is described in some detail because of its importance in studying hemoglobin S polymerization at physiological concentrations and temperatures. In order, to establish conditions for complete photodissociation with minimal laser heating, a series of control experiments on normal human hemoglobin was performed and theoretically modeled. The concentration dependence of the tenth time is found to decrease with increasing hemoglobin S concentration. In the range 0.2 to 0.3 g/cm3, the tenth time varies as the 36th power of the hemoglobin S concentration, while in the range 0.3 to 0.4 g/cm3 it decreases to 16th power. As the tenth times become shorter, the progress curves broaden, with the onset of polymerization becoming less abrupt. For tenth times greater than about 30 seconds, measurements with the laser photolysis technique on small volumes yield highly irreproducible tenth times, but superimposable progress curves, indicating stochastic behavior. The initial part of the progress curves from both temperature-jump and laser photolysis experiments is well fit with an equation for the concentration of polymerized monomer, delta (t) = A[cosh (Bt) -1], which results from integration of the linearized rate equations for the double nucleation mechanism described in the accompanying paper (Ferrone et al., 1985). The dependence of the parameters A and B on temperature and concentration is obtained from fitting over 300 progress curves. The rate B has a large concentration dependence, varying at 25 degrees C from about 10(-4) S-1 at 0.2 g/cm3 to about 100 s-1 at 0.4 g/cm3.  相似文献   

4.
Component 7c is one of the four homologous type II intermediate-filament proteins that, by association with the complementary type I proteins, form the microfibrils or intermediate filaments in wool. Component 7c was isolated as the S-carboxymethyl derivative from Merino wool and its amino acid sequence was determined by manual and automatic sequencing of peptides produced by chemical and enzymic cleavage reactions. It is an N-terminally blocked molecule of 491 residues and Mr (not including the blocking group) of 55,600; the nature of the blocking group has not been determined. The predicted secondary structure shows that component 7c conforms to the now accepted pattern for intermediate-filament proteins in having a central rod-like region of approximately 310 residues of coiled-coil alpha-helix flanked by non-helical N-and C-terminal regions. The central region is divided by three non-coiled-coil linking segments into four helical segments 1A, 1B, 2A and 2B. The N-and C-terminal non-helical segments are 109 and 71 residues respectively and are rich in cysteine. Details of procedures use in determining the sequence of component 7c have been deposited as a Supplementary Publication SUP 50152 (65 pages) at the British Library Document Supply Centre, Boston Spa, Wetherby, West Yorkshire LS23 7BQ, U.K., from whom copies can be obtained on the terms indicated in Biochem. J. (1989) 257,5. The information comprises: (1) details of chemical and enzymic methods used for cleavage of component 7c, peptides CN1, CN2 and CN3, and various other peptides, (2) details of the procedures used for the fractionation and purification of peptides from (1), including Figures showing the elution profiles from the chromatographic steps used, (3) details of methods used to determine the C-terminal sequence of peptide CN3, and (4) detailed evidence to justify a number of corrections to the previously published sequence.  相似文献   

5.
1. The hemoglobin of the sting-ray, Dasyatis sabina, is both polymorphic and heterogeneous; three components predominate. 2. One major component has two kinds of polypeptide chain, of which one, presumably an alpha-chain, has a blocked NH2-terminus and an arginyl COOH-terminus, whereas carboxypeptidases A and B release tyrosine and histidine from the COOH-terminus of the beta-chain. 3. The amino acid sequence of the beginning NH2-terminal segment of the beta-chain of the major component has been determined. 4. The hemoglobin of the sting-ray, Dasyatis sabina, is highly resistant to urea and does not dissociate readily into subunits. 5. Oxygen binding by the hemoglobin is not affected by organic phosphates or high concentrations of either NaCl or urea. 6. The hemoglobin does not polymerize beyond tetramers. 7. Cooperativity, as monitored by n in the Hill equation, is pH-dependent and maximal between pH 8.5 and 9.0. 8. The hemoglobin has a large Bohr effect; the oxygen affinity is 16 times higher at pH 10 than at pH 6.5.  相似文献   

6.
The symmetrical trifunctional cross-linking reagent trimesoyl tris(methyl phosphate) (3), reacts selectively with amino groups (beta 1Val and beta 82Lys) in the diphosphoglycerate binding site of human hemoglobin A, producing cross-linked tetrameric species in good yield. A major species is triply linked, alpha alpha beta 1(82) greater than B beta 82, where B symbolizes benzene-1,3,5-tricarbonyl. Both this triply linked species and the doubly linked species, alpha alpha beta 1B beta 82, produced from deoxyhemoglobin have a considerably lower oxygen affinity than does native hemoglobin while maintaining a high degree of cooperativity (n50 = 2.4), making them potentially useful as red cell substitutes, in principle delivering twice as much oxygen as whole blood between pO2 = 100 and = 40 Torr. The yield of products indicates that triply and doubly linked species form in parallel so that there are independent routes to each. It is proposed that differences in routes are due to stereoisomerism about the amide bonds which form from reaction of the reagent with the protein.  相似文献   

7.
The 3-fold increase in the carbamylation rate of Val-1 (alpha) of hemoglobin upon deoxygenation described earlier is now shown to be a sensitive probe of conformational change. Thus, whereas this residue in methemoglobin A is carbamylated at the same rate as in liganded hemoglobin, upon addition of inositol hexaphosphate its carbamylation rate is enhanced 30% as much as the total change in the rate between the CO and deoxy states. For CO-hemoglobin Kansas in the presence of the organic phosphate, the relative increase in the carbamylation rate of this residue is about 50%. These results indicate that methemoglobin A and hemoglobin Kansas in the presence of inositol hexaphosphate do not assume a conformation identical with deoxyhemoglobin but rather form either a mixture of R and T states or an intermediate conformation in the region around Val-1 (alpha). Studies on the mechanism for the rate enhancement in deoxyhemoglobin suggest that the cyanate anion binds to groups in the vicinity of Val-1 (alpha) prior to proton transfer and carbamylation of this NH2-terminal residue. Thus, specific removal with carboxypeptidase B of Arg-141 (alpha), which is close to Val-1 (alpha) in deoxyhemoglobin, abolishes the enhancement in carbamylation. Chloride, which has the same valency as cyanate, is a better competitive inhibitor of the carbamylation of deoxyhemoglobin (Ki = 50 mM) compared with liganded hemoglobin. Nitrate and iodide are also effective inhibitors of the carbamylation of Val-1 (alpha) of deoxyhemoglobin (Ki = 35 mM); inorganic phosphate, sulfate, and fluoride are poor competitive inhibitors. The change in pKa of Val-1 (alpha) upon deoxygenation may be due to its differential interaction with chloride.  相似文献   

8.
The alpha-helical antibiotic peptide (P18: KWKLFKKIPKFLHLAKKF-NH2) designed from the cecropin A(1-8)-magainin 2 (1-12) hybrid displayed strong bactericidal and tumoricidal activity without inducing hemolysis. The effect of the Pro9 residue at central position of P18 on cell selectivity was investigated by Pro9 --> Leu or Pro9 --> Ser substitution. Either substitution markedly reduced the antibacterial activity of P18 and increased hemolysis, although it did not significantly affect cytotoxicity against human transformed tumor and normal fibroblast cells. These results suggest that a proline kink in alpha-helical antibiotic peptide P18 serves as a hinge region to facilitate ion channel formation on bacterial cell membranes and thus plays an important role in providing high selectivity against bacterial cells. Furthermore, to investigate the structure-antibiotic activity relationships of P18, a series of N- or C-terminal deletion and substitution analogs of P18 were synthesized. The C-terminal region of P18 was related to its antibiotic activity and alpha-helical conformation on lipid membranes rather than N-terminal one. Higher alpha-helicity of the peptides was involved in the hemolytic and antitumor activity rather than antibacterial activity. Except for [L9]-P18 and [S9]-P18, all the designed peptides containing a Pro residue showed potent antibacterial activity, although they did not induce a cytolytic effect against human erythrocyte and normal fibroblast cells at the concentration required to kill bacteria. In particular, P18 and some analogs (N-1, N-2, N-3, N-3L and N-4L) with potent bactericidal and tumoricidal activity and little or no normal cell toxicity may serve as an attractive candidate for the development of novel anti-infective or antitumor agents.  相似文献   

9.
The subunit structure, dissociation, and unfolding of the hemoglobin of the earthworm, Lumbricus terrestris, were investigated by light scattering molecular weight methods and changes in optical rotatory dispersion (at 233 nm) and absorption in the Soret region. Urea and the alkylureas, methyl-, ethyl-, propyl-, and butylurea, were employed as the reagents to cause both dissociation and unfolding of the protein. Analysis of the light scattering data suggests that the dissociation patterns as a function of hemoglobin concentration in the various dissociating solvents can be described in quantitative terms, either as an equilibrium mixture consisting of parent duodecamers and hexamers of 3 x 10(6) and 1.5 x 10(6) molecular weight (in 1-3 M urea, 1-2 M methyl- and ethylurea, and 1 M propylurea), as a mixture of hexamers and monomers, the latter with a molecular weight of 250000 (i.e., in 4 M urea), or as a mixture of all three species of duodecamers, hexamers, and monomers, seen in 2 M propylurea. Parallel studies by optical rotation and absorption measurements indicate that there is little or no unfolding of the subunits at urea and alkylurea concentrations where complete dissociation to hexamers and extensive dissociation to monomers can be achieved. Further splitting of the monomers (A subunits) to smaller fragments of one-third to one-quarter of the molecular weight of the monomers (B subunits) is seen in the presence of 7 and 8 M urea (pH 7) and in alkaline urea to propylurea solutions. Analysis of the dissociation data of duodecamers to monomers, based on equations used in studies of the urea and amide dissociation of human hemoglobin A from our laboratory, suggests few urea and alkylurea binding sites at the areas of hexamer contacts in the associated duodecameric form of L. terrestris hemoglobin. This suggests that hydrophobic interactions are not the dominant forces that govern the state of association of L. terrestris hemoglobin relative to polar and ionic interactions. The unfolding effects of the ureas, at concentrations above the dissociation transitions, are closely similar to their effects on other globular proteins, suggesting that hydrophobic interactions play an important role in the maintenance of the folded conformation of the subunits. Use of the Peller-Flory equation, with binding constants based on free energy transfer data of hydrophobic amino acid side chains and denaturation data used in previous denaturation studies, gave a relatively good acount of the observed denaturation midpoints obtained with the various ureas supporting these conclusions.  相似文献   

10.
Babesia divergens multiplication cycle involves erythrocyte invasion, intracellular division, and erythrocyte lysis with the simultaneous liberation of hemoglobin. We have decided to set up a spectrophotometric protocol based on hemoglobin concentration in the culture supernatants to monitor B. divergens in vitro growth. After the selection of 405 nm as the most appropriate endpoint hemoglobin wavelength in our conditions (hemoglobin concentration in the supernatant), cultures were standardized [1 x 10(9) red blood cell (RBC)/ml, 1-2.5 x 10(5) infected red blood cell (iRBC)/ml] to allow their monitoring over 3 days. The protocol was then compared to the most commonly used growth measurement methods: parasitemia counting and [(3)H]hypoxanthine incorporation. An excellent correlation was demonstrated between A(405) of the culture supernatant and parasitemia of the iRBC, whatever the RBC concentration used in the medium. This correlation was also evidenced between A(405) and [(3)H]hypoxanthine incorporation for [(3)H]hypoxanthine concentrations lower than 4 microCi/ml. Our assays also highlighted the inhibitory effect of [(3)H]hypoxanthine on B. divergens growth even when used at low concentrations (0.8 microCi/ml) and for a short incorporation duration (24 h). This effect was confirmed by both A(405) and parasitemia counting. In conclusion, A(405) measurement of B. divergens culture supernatant represents a simple, rapid, safe, and reliable way to measure the in vitro growth of this parasite. Generation times of three different B. divergens strains were then determined by the protocol described here and varied between 8 h 36 min and 13 h 8 min.  相似文献   

11.
1. On the basis of their molecular masses, four types of polypeptides (A, B, C, D) were obtained by SDS-PAGE of the extracellular hemoglobin of the polychaete annelid Arenicola marina. 2. On 2-dimensional polyacrylamide gel electrophoresis, the erythrocruorin dissociated into six different types of polypeptide chains; A1, A2, B1, B2, C and D. 3. A1 and B1 migrate in 2-dimensional electrophoresis at the same position as alpha and beta chains of human hemoglobin.  相似文献   

12.
The hemoglobins of three snake species: Liophis miliaris, Bothrops alternatus and Boa constrictor present a single ATP binding site per tetramer. The ATP association constant values for the deoxyhemoglobins at pH 7.5 were about KD ≅ 106 M−1 (107 M−1 for B. contrictor), three to four orders of magnitude higher than the respective values for oxyhemoglobin of about KO ≅ 102 M−1. The deoxyhemoglobin constant values markedly decrease as a function of pH, becoming, at pH 8.5, about KD ≅ 103 M−1 whereas for the oxyhemoglobin the constants remain of about the same, KO ≅ 102 M−1, at the pH range studied. The high ATP binding affinity constants, compared to those of human hemoglobin A, were explained from a molecular structural standpoint, considering L. miliaris hemoglobin, whose complete primary sequence is known. Two distinct amino acid residue differences were found in the β-chain, one being Trp (NA3) (more hydrophobic) in the snake hemoglobin which substitutes the Leu (NA3) in human hemoglobin, and the second being Val 101 β (G3) instead of Glu 101 β (G3). The substitutions could provide an un-neutralized, positively charged, residue Lys-104β and, taking into account its high pK value, the pH dependence of ATP binding affinity for the snake hemoglobin would originate from pH-dependent ionization of phosphate groups of the allosteric effector. The physiological implications of the high ATP binding constant, as well as the possible protective role of the nucleotide binding against the effect of high environmental temperatures on the oxygen dissociation curves, are discussed.  相似文献   

13.
Responses (chemotaxis and changes in membrane potential) of Tetrahymena, Physarum, and Nitella against aqueous solution of homologous series of n-alcohols, n-aldehydes and n-fatty acids were studied for clarifying the hydrophobic character of chemoreceptive membranes. Results were: (1) All organisms studied responded to homologous compounds examined when the concentration of these chemicals exceeded their respective threshold, Cth, and the response, R, were expressed approximately as R=alpha log (C/Cth) for C greater than Cth. (2) Increase of the length of hydrocarbon chain in homologues decreased Cth. Plots of log Cth against the number of carbon atoms, n, in n-alcohols, n-aldehydes and n-fatty acids showed linear relationships as represented by long Cth=-An+B. A and B are positive constants for respective functional end groups of the chemicals and biological membranes used. The above empirical equation was interpreted in terms of the partition equilibrium of methylene groups between bulk solution and membrane phase. Parameter A was shown to be a measure of hydrophobicity of the membrane, and B represented the sensitivity of chemoreception of the membrane. (3) Thresholds, Cth, for various hydrophobic reagents were compared with those of human olfactory reception, T. Plots of log T against log Cth fell on straight lines for respective organisms with different slopes which were proportional to parameter A.  相似文献   

14.
An attempt was made to characterize the hemolymph of Biomphalaria glabrata with reference to "normal" intra-specific variation, i.e., both inter- and intra-strain differences. Total protein concentration, per cent hemoglobin, pH, and osmolarity were studied. Seven geographic strains of B, glabrata were examined. In addition, observations were made on the hemolymph of Biomphalaria straminea, several strains of Helisoma caribaeum, and on B. glabrata subjected to infection with Schistosoma mansoni or to periods of starvation. Intra-strain differences in total protein concentration and total hemoglobin concentration in B. glabrata appeared to be more closely related with snail size than with absolute age. Inter-strain variation in B. glabrata was also noted, but the differences were of the same magnitude as those from intra-strain samples. Significant differences in total protein concentration were observed, however, between the means of similar size B. glabrata, B. straminea and H. caribaeum. The osmolatity of the hemolymph from different size B. glabrata was similar as were the osmolalities of the hemolymph from similar size snails of different strains. However, all B. glabrata strains exhibited hemolymph osmolalities lower than observed in strains of H. caribaeum. Infection with S. mansoni reduced the protein concentration of B. glabrata hemolymph. Differences were noted as early as 1.5-24 hr post-infection, with significant alterations occurring at about 11 days post-infection. To a lesser extent, starvation also depleted the protein content of the hemolymph.  相似文献   

15.
1. Mechanoreceptors monitoring tension in working muscles are described in the Decapoda Crustacea.

2. The receptors are associated with apodemes of muscles in the walking leg and are well‐developed in the extensor and flexor of the meropodite (Figures 1, 2).

3. The unbranched dendrites of the receptor neurones innervate the tissues surrounding the insertions of the muscle fibres (Figures 3, 4, 5(A)).

4. The receptors show spontaneous activity with the M‐C joint at resting position and this activity increases when the muscle is stretched by holding the joint at a different position (Figure 7).

5. Isometric tension increase in the muscle recruits sensory units (Figures 8, 10(A)) and increases the activity of units firing (Figure 9).

6. Apodeme receptors may be an entirely distinct input channel from chordotonal organs (Figure 10(B,C)). Joint movements produced by a standard muscle stimulus against increasing loads reveal very different responses (Figure 11).

7. Attempts to determine whether chordotonal organs (CP1, Figures 5(B), 6) monitor isometric muscle tension (Figure 12) suggest possible complexities in their dynamic responses.

8. Abbreviations used in this paper are FASN flexor apodeme sensory nerve, EASN extensor apodeme sensory nerve, BASN bender apodeme sensory nerve, and OASN opener apodeme sensory nerve.  相似文献   

16.
Experiments on cross-immunity reactions between three viruses attacking Theobroma cacao L. on the Gold Coast are described. A field trial involving 3 acres of graft-inoculated trees revealed some degree of protection afforded by Theobroma virus 1B against infection with virus 1A. The protection appeared to be more effective against insect inoculation than against graft transmission, being only temporary in the latter. Virus 1C (probably unrelated but not to be called Theobroma virus 2 until more evidence is available) conferred no protection against virus 1A.
The latent periods for these viruses were calculated from this experiment, which also provided data on their effects on yield. Virus 1A reduced yield by 50% in the first year after inoculation and killed the trees in the second. Virus 1B had no appreciable effect on yield; virus 1C reduced yield by 50 % in the third year after inoculation but there was no further decline in the fourth.
The rates of spread of these viruses were compared and significant differences demonstrated.  相似文献   

17.
During the re-isolation of the lead compound nodulisporic acid A (1a) and targeted chemical screening for related compounds, we discovered a series of 1'-deoxy congeners named herein nodulisporic acids B (1b), B1 (2b), and B2 (3b). In comparison with nodulisporic acid A, these compounds were less active and were chemically unstable resulting into formation of delta23 dehydro derivatives. Therefore, these compounds were stabilized and isolated as sodium salts and methyl ester. Nodulisporic acid B is 100-fold less active than nodulisporic acid A against fleas. The isolation, structure elucidation, and biological activities of these compounds are described.  相似文献   

18.
25 strains of Clostridium perfringens were screened for hydroxysteroid dehydrogenase activity; 19 contained NADP-dependent 3alpha-hydroxysteroid dehydrogenase and eight contained NAD-dependent 12alpha-hydroxysteroid dehydrogenase active against conjugated and unconjugated bile salts. All strains containing 12alpha-hydroxysteroid dehydrogenase also contained 3alpha-hydroxysteroid dehydrogenase although 12alpha-hydroxysteroid dehydrogenase was invariably in lesser quantity than the 3alpha-hydroxysteroid dehydrogenase. In addition, 7alpha-hydroxysteroid dehydrogenase activity was evident only when 3alpha, 7alpha, 12alpha-trihydroxy-5beta-cholanoate was substrate but notably absent when 3alpha, 7alpha-dihydroxy-5beta-cholanoate was substrate. The oxidation product 12alpha-hydroxy-3, 7-diketo-5beta-cholanoate is rapidly further degraded to an unknown compound devoid of either 3alpha- or 7alpha-OH groups. Group specificity of these enzymes was confirmed by thin-layer chromatography studies of the oxidation products. These enzyme systems appear to be constitutive rather than inducible. In contrast to C. perfringens. Clostridium paraputrificum (five strains tested) contained no measurable hydroxysteroid dehydrogenase activity. pH studies of the C. perfringens enzymes revealed a sharp pH optimum at pH 11.3 and 10.5 for the 3alpha-OH- and 12alpha-OH-oriented activities, respectively. Kinetic studies gave Km estimates of approx. 5 X 10(-5) and 8 X 10(-4) M with 3alpha, 7a-dihydroxy-5beta-cholanoate and 3alpha, 12alpha-dihydroxy-5beta-cholanoate as substrates for two respective enzymes. 3alpha-hydroxysteroid dehydrogenase was active against 3alpha-OH-containing steroids such as androsterone regardless of the sterochemistry of the 5H (Both A/B cis and A/B trans steroides were substrates). There was no activity against 3beta-OH-containing steroids. The 3alpha- and 12alpha-hydroxysteroid dehydrogenase activities, although differing in cofactor requirements cannot be distinguished by their appearance in the growth curve, their mobility on disc gel electrophoresis, elution volume on passage through Sephadex G-200 or heat inactivation studies.  相似文献   

19.
We have investigated in vitro the metabolic capability of 3 extrahepatic cytochromes P-450, CYP1A1, 1B1 and 2J2, known to be over-expressed in various tumors, to biotransform 5 tyrosine kinase inhibitors (TKI): dasatinib, imatinib, nilotinib, sorafenib and sunitinib. Moreover, mRNA expression of CYP1A1, 1B1, 2J2 and 3A4 in 6 hepatocellular and 14 renal cell carcinoma tumor tissues and their surrounding healthy tissues, was determined.Our results show that CYP1A1, 1B1 and especially 2J2 can rapidly biotransform the studied TKIs with a metabolic efficiency similar to that of CYP3A4. The mRNA expression of CYP1A1, 1B1, 2J2 and 3A4 in tumor biopsies has shown i) the strong variability of CYP expression and ii) distinct outliers showing high expression levels (esp. CYP2J2) that are compatible with high intratumoral CYP activity and tumor-specific TKI degradation.CYP2J2 inhibition could be a novel clinical strategy to specifically increase the intratumoral rather than plasma TKI levels, improving TKI efficacy and extending the duration before relapse. Such an approach would be akin to beta-lactamase inhibition, a classical strategy to avoid antibiotic degradation and resistance.  相似文献   

20.
Echinocardium cordatum had long been considered as cosmopolitan, but molecular data revealed it is a complex of cryptic species, with two non-hybridizing species (B1 & B2) in the Mediterranean Sea living in syntopy with Echinocardium mediterraneum. Histological analyses of the gonads from a 17-month sampling period revealed a statistically significant time lag between the Maturity Indices of E.?cordatum and E.?mediterraneum. The main environmental stimulus may be different for the two nominal species, possibly seawater temperature for E.?cordatum and chlorophyll a concentration for E.?mediterraneum. Within the E.?cordatum complex, spawning timing and synchrony are different according to major geographic areas (Atlantic/Pacific/Mediterranean) and/or the corresponding genetic subdivision [A/P/(B1 & B2)]. In contrast, the effects of temperature on the reproductive cycle seem rather to mirror the genetic lineages than environmental similarities of the different localities. Between the sister species (B1 & B2) no differences could be detected, maybe due to small sample sizes.  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