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.
Xyloglucan oligosaccharides were isolated with various degreesof polymerization (DP) and reduced with tritiated sodium borohydride.The 3H-oligosaccharides were tested for their ability to bindto amorphous and microcrystalline celluloses and to cellulosefilter paper. The time course of binding indicated that theradiolabeled oligosaccharides continued to be bound for at least1 h after heating at 120°C. The binding probably requiredthe organization of the oligosaccharides and celluloses by gradualannealing after heating. Although neither pentasaccharide (glucose:xylose, 3 : 2), heptasaccharide (glucose: xylose, 4 : 3) andnonasaccharide (glucose : xylose : galactose : fucose, 4 : 3: 1 : 1) failed to bind to the celluloses, binding occurredwith oligosaccharides with DP equivalent to more than four consecutive1,4-ß-glucosyl residues. The extent of binding tothe celluloses increased gradually from octasaccharide (glucose:xylose, 5 : 3) to hendecosanosaccharide (glucose/xylose, 12: 9), with the increase in the DP of 1,4-ß-glucosylresidues. The binding of reduced cello-dextrins to celluloserequired at least 4 consecutive 1,4-ß-glucosyl residues.The extent of binding of cellopentitol or cellohexitol to cellulosewas similar to that of hendecosanosaccharide, showing lowerbinding for xyloglucan oligosaccharides in spite of longer chainsof 1,4-ß-glucosyl residues. These findings suggestthat the mode of binding to cellulose of xyloglucan oligosaccharidesis different from that of cello-oligosaccharides. (Received February 18, 1994; Accepted June 1, 1994)  相似文献   

2.
The survivorships and growth rates ofPlanorbarius corneus andPlanorbis planorbis were measured during the whole life spanof the snails at constant temperatures of 5, 10, 15, 20 and25°C. Life expectancy tables were constructed. The maximumlongevities for P. corneus (231 weeks) and P. planorbis (175weeks) occurred at 15°C and 10°C respectively. The extremetemperatures were particularly unfavourable. Snail growth which was expressed logarithmically was most sensitiveto temperature during the exponential stage. However, if weconsider a given species duringjts entire life span, there wereno significant differences between temperature groups whichwere in order of decreasing suitability: —P. corneus: 25, 20 and 15°C; 10°C; 5°C; —P.planorbis: 20 and 15°C; 25 and 10°C; 5°C. Using the growth constants from the Von Berta-lanffy's modeladjusted to give a gaussian curve, it was calculated that theoptimum temperatures for the growth of P. planorbis and P. corneuswere 19 and 20.5°C respectively. A mathematical model basedon spline functions makes it possible to predict the growthof snails over a wide range of temperatures. (Received 2 June 1993; accepted 18 November 1993)  相似文献   

3.
The temperatures of the roots, the apical meristem, and theshoots of Zea mays plants were varied independently of eachother and the rates of leaf extension were measured. When thetemperature of the apical meristem and region of cell expansionat the base of the leaf was kept at 25 °C, changes of leafextension in response to changes of root and shoot temperatureswere less pronounced. When the temperature of the meristematicregion was changed by increments of 5 or 10 °C from 0 to40 °C, and the root and shoot temperatures were kept at25 °C, rapid changes in leaf extension occurred. It was concluded that the rates of leaf extension were controlledat root-zone temperatures of 5 to 35 °C by heating or coolingof the meristematic region. Changes in rates of leaf extensionin response to changes in air temperature were attributed todirect effects on the temperature of the meristematic regionand on the physiology of the leaf.  相似文献   

4.
Sand and nest temperatures were monitored during the 2002–2003 nesting season of the green turtle, Chelonia mydas, at Heron Island, Great Barrier Reef, Australia. Sand temperatures increased from ∼ 24°C early in the season to 27–29°C in the middle, before decreasing again. Beach orientation affected sand temperature at nest depth throughout the season; the north facing beach remained 0.7°C warmer than the east, which was 0.9°C warmer than the south, but monitored nest temperatures were similar across all beaches. Sand temperature at 100 cm depth was cooler than at 40 cm early in the season, but this reversed at the end. Nest temperatures increased 2–4°C above sand temperatures during the later half of incubation due to metabolic heating. Hatchling sex ratio inferred from nest temperature profiles indicated a strong female bias.  相似文献   

5.
Deuterated Chlorella (D-Chlorella) cells were obtained by cultivationin a medium that contained D2O. The modification of cell componentsby deuterium (D) increased the heat sensitivity of cells, whichdepended on the extent of deuteration. To elucidate the mechanismof the D-induced increase in the heat sensitivity, the effectof incorporation of D on the denaturation of proteins was investigated.Proteins obtained from D-Chlorella cells, when preheated at37°C, were aggregated to a greater extent than those fromH-Chlorella against the heating at 45°C. The rate of synthesisof heat-shock proteins (hsps) in D-Chlorella was consistentlylower than that in Hcells. Furthermore, an experiment in vitroindicated that deuterated proteins denatured more rapidly thannormal proteins upon heating at 60°C. (Received August 12, 1993; Accepted November 30, 1993)  相似文献   

6.
Transformation of the water cluster distribution in wet potato starch (with a water content of 27 to 45%) at temperatures that ranged from–50 to +80°C was studied by differential scanning calorimetry. A significant difference was observed between the transformations in the temperature ranges below and above 0°C. Both cooling and heating at T < 0°C enabled a reorganization of the initial size distribution of water clusters characteristic for room temperature. These changes could lead to an increase of the average cluster size during both crystallization and melting. The transformation intensity depended on the water content and scanning rate and differed between the native and amorphous states of starch. In this case, the cluster-size distribution remained unimodal. However, heating of wet native starch to temperatures close to the point of transition into the amorphous state (75–80°C) induced a bimodal distribution due to the emergence of large water clusters; thus, the heterogeneity of the water distribution within the native granules increased.  相似文献   

7.
The chlorophyll fluorescence (F) temperature curves in a linear time-temperature heating/cooling regime were used to study heat-induced irreversible F changes in primary green leaves of spring barley (Hordeum vulgare L. cv. Akcent). The leaf segments were heated in a stirred water bath at heating rates of 0.0083, 0.0166, 0.0333, and 0.0500 °C s−1 from room temperature up to maximal temperature T m and then linearly cooled to 35 °C at the same rate. The F intensity was measured by a pulse-modulated technique. The results support the existence of the two critical temperatures of irreversible F changes postulated earlier, at 45–48 and 53–55 °C. The critical temperatures are slightly dependent on the heating rate. Two types of parameters were used to characterize the irreversibility of the F changes: the coefficient of irreversibility μ defined as the ratio of F intensity at 35 °C at the starting/ending parts of the cycle and the slopes of tangents of linear parts of the F temperature curve. The dependence of μ on T m revealed a maximum, which moved from 54 to 61 °C with the increasing heating/cooling rate v from 0.0083 to 0.0500 °C s−1, showing two basic phases of the irreversible changes. The Arrhenius and Eyring approaches were applied to calculate the activation energies of the initial increase in μ. The values varied between 30 and 50 kJ mol−1 and decreased slightly with the increasing heating rate.  相似文献   

8.
M. Zeroni  J. Gale 《Plant and Soil》1987,104(1):93-98
Rose plants (Rosa hybrida ‘Sonia’=‘Sweet Promise’) were grown in heated (minimum night temperature 17°C), and unheated greenhouses with or without root heating to 21°C. These trials covered 6 growth cycles extending over two winter seasons. In the heated greenhouse, root heating did not increase yield, flower quality or plant development. In the unheated greenhouse, root-heated plants grew as well as those in the air-heated greenhouse as long as the air temperature did not fall below 6°C. When minimum night temperatures fell below 6°C, growth, yield and quality were reduced, irrespective of root temperature. Daytime plant water relations were studied in plants growing at 6 different root temperatures in the unheated greenhouse. Leaf resistance to water diffusion was lowest at optimal root temperature. Total leaf water potential was not significantly affected by root temperature.  相似文献   

9.
Although the chloroplast movement can be strongly affected by ambient temperature, the information about chloroplast movement especially related to high temperatures is scarce. For detailed investigation of the effects of heat stress (HS) on tobacco leaves (Nicotiana tabacum L. cv. Samsun), we used two different HS treatments in dark with wide range of elevated temperatures (25–45°C). The leaf segments were either linearly heated in water bath at heating rate of 2°C min−1 from room temperature up to maximal temperature (T m) and then linearly cooled down to 25°C or incubated for 5 min in water bath at the same T m followed by 5 min incubation at 25°C (T-jump). The changes in light-induced chloroplast movement caused by the HS pretreatment were detected after the particular heating regime at 25°C using a method of time-dependent collimated transmittance (CT) and compared with the chlorophyll O–J–I–P fluorescence rise (FLR) measurements. The inhibition of chloroplast movement started at about 40°C while the fluorescence parameters responded generally at higher T m. This difference in sensitivity of CT and FLR was higher for the T-jump than for the linear HS indicating importance of applied heating regime. A possible influence of chloroplast movement on the FLR measurement and a physiological role of the HS-impaired chloroplast movement are discussed.  相似文献   

10.
Xyloglucanase from an extracellular culture filtrate of alkalothermophilic Thermomonospora sp. was purified to homogeneity with a molecular weight of 144 kDa as determined by SDS-PAGE and exhibited specificity towards xyloglucan with apparent K m of 1.67 mg/ml. The enzyme was active at a broad range of pH (5–8) and temperatures (40–80°C). The optimum pH and temperature were 7 and 70°C, respectively. The enzyme retained 100% activity at 50°C for 60 h with half-lives of 14 h, 6 h and 7 min at 60, 70 and 80°C, respectively. The kinetics of thermal denaturation revealed that the inactivation at 80°C is due to unfolding of the enzyme as evidenced by the distinct red shift in the wavelength maximum of the fluorescence profile. Xyloglucanase activity was positively modulated in the presence of Zn2+, K+, cysteine, β-mercaptoethanol and polyols. Thermostability was enhanced in the presence of additives (polyols and glycine) at 80°C. A hydrolysis of 55% for galactoxyloglucan (GXG) from tamarind kernel powder (TKP) was obtained in 12 h at 60°C and 6 h at 70°C using thermostable xyloglucanases, favouring a reduction in process time and enzyme dosage. The enzyme was stable in the presence of commercial detergents (Ariel), indicating its potential as an additive to laundry detergents.  相似文献   

11.
Since xyloglucan is believed to bind to cellulose microfibrils in the primary cell walls of higher plants and, when isolated from the walls, can also bind to cellulose in vitro, the binding mechanism of xyloglucan to cellulose was further investigated using radioiodinated pea xyloglucan. A time course for the binding showed that the radioiodinated xyloglucan continued to be bound for at least 4 hours at 40°C. Binding was inhibited above pH 6. Binding capacity was shown to vary for celluloses of different origin and was directly related to the relative surface area of the microfibrils. The binding of xyloglucan to cellulose was very specific and was not affected by the presence of a 10-fold excess of (1→2)-β-glucan, (1→3)-β-glucan, (1→6)-β-glucan, (1→3, 1→4)-β-glucan, arabinogalactan, or pectin. When xyloglucan (0.1%) was added to a cellulose-forming culture of Acetobacter xylinum, cellulose ribbon structure was partially disrupted indicating an association of xyloglucan with cellulose at the time of synthesis. Such a result suggests that the small size of primary wall microfibrils in higher plants may well be due to the binding of xyloglucan to cellulose during synthesis which prevents fasciation of small fibrils into larger bundles. Fluorescent xyloglucan was used to stain pea cell wall ghosts prepared to contain only the native xyloglucan:cellulose network or only cellulose. Ghosts containing only cellulose showed strong fluorescence when prepared before or after elongation; as predicted, the presence of native xyloglucan in the ghosts repressed binding of added fluorescent xyloglucan. Such ghosts, prepared after elongation when the ratio of native xyloglucan:cellulose is substantially reduced, still showed only faint fluorescence, indicating that microfibrils continue to be coated with xyloglucan throughout the growth period.  相似文献   

12.
Two experiments were conducted to assess the response of cauliflower (Brassica oleracea L. var. botrytis) cv. “Nautilus” F1 hybrid to different constant temperatures after curd initiation by keeping the plants in six different temperature-controlled glasshouse compartments with heating set point temperatures of 6, 10, 14, 18, 22, and 26 °C (±4 °C) at the School of Plant Sciences, The University of Reading, UK during winter 1998–1999 and summer 1999. Many of the growth parameters increased with increasing mean growing temperature up to an optimum temperature and then declined with further increases in temperature. Therefore, cauliflower’s growth and development after curd initiation could be resolved into linear or curvilinear function of effective temperatures calculated with optimum temperatures between 19 and 23 °C. It is suggested that future warmer climates will be beneficial for winter cauliflower production rather than summer cauliflower production.  相似文献   

13.
Among amphibians, the ability to compensate for the effects of temperature on the locomotor system by thermal acclimation has only been reported in larvae of a single species of anuran. All other analyses have examined predominantly terrestrial adult life stages of amphibians and found no evidence of thermal acclimatory capacity. We examined the ability of both tadpoles and adults of the fully aquatic amphibian Xenopus laevis to acclimate their locomotor system to different temperatures. Tadpoles were acclimated to either 12 °C or 30 °C for 4 weeks and their burst swimming performance was assessed at four temperatures between 5 °C and 30 °C. Adult X. laevis were acclimated to either 10 °C or 25 °C for 6 weeks and their burst swimming performance and isolated muscle performance was determined at six temperatures between 5 °C and 30 °C. Maximum swimming performance of cold-acclimated X. laevis tadpoles was greater at cool temperatures and lower at the highest temperature in comparison with the warm-acclimated animals. At the test temperature of 12 °C, maximum swimming velocity of tadpoles acclimated to 12 °C was 38% higher than the 30 °C-acclimation group, while at 30 °C, maximum swimming velocity of the 30 °C-acclimation group was 41% faster than the 12 °C-acclimation group. Maximum swimming performance of adult X. laevis acclimated to 10 °C was also higher at the lower temperatures than the 25 °C acclimated animals, but there was no difference between the treatment groups at higher temperatures. When tested at 10 °C, maximum swimming velocity of the 10 °C-acclimation group was 67% faster than the 25 °C group. Isolated gastrocnemius muscle fibres from adult X. laevis acclimated to 10 °C produced higher relative tetanic tensions and decreased relaxation times at 10 °C in comparison with animals acclimated to 25 °C. This is only the second species of amphibian, and the first adult life stage, reported to have the capacity to thermally acclimate locomotor performance. Accepted: 28 October 1999  相似文献   

14.
Specimens of the Arctic Collembolon Onychiurus arcticus were exposed to desiccation at several subzero temperatures over ice and at 0.5 °C over NaCl solutions. The effects of desiccation on water content (WC), body fluid melting point (MP), supercooling point (SCP) and survival were studied at several acclimation temperatures and relative humidities. Exposure to temperatures down to −19.5 °C caused a substantial and increasing dehydration. At the lowest exposure temperature unfrozen individuals lost 91.6% of the WC at full hydration but more than 80% of the individuals survived when rehydrated. Exposure at 0.5 °C to decreasing relative humidities (RH) from 100% to 91.3% caused increasing dehydration and increasing mortality. Survival of equally dehydrated individuals was higher at subzero temperatures than at 0.5 °C. Concurrent with the decline in WC a lowering of the MP was observed. Animals exposed to −3 °C and −6 °C over ice for 31 days had a MP of −3.8 and < −7.5 °C, respectively. Specimens from a laboratory culture had a mean SCP of −6.1 °C, and acclimation at 0 or −3 °C had little effect on SCPs. Exposure at −8.2 °C over ice for 8 days, however, caused the mean SCP to decline to −21.8 °C due to the severe dehydration of these individuals. Dehydration at 0.5 °C in 95.1 and 93.3% RH also caused a decline in SCPs to about −18 °C. Individuals that had been acclimated over ice at −12.4 °C or at lower temperatures apparently did not freeze at all when cooled to −30 °C, probably because all freezeable water had been lost. These results show that O. arcticus will inevitably undergo dehydration when exposed to subzero temperatures in its natural frozen habitat. Consequently, the MP and SCP of the Collembola are substantially lowered and in this way freezing is avoided. The increased cold hardiness by dehydration is similar to the protective dehydration mechanism described in earthworm cocoons and Arctic enchytraeids. Accepted: 5 January 1998  相似文献   

15.
Tomato seedlings were grown in a 12-hour day at constant andalternating day and night temperatures ranging from 10°to 30° C. The pattern of results was similar at light intensitiesof 400 and 800 f.c. The maximum rate of dryweight accumulationoccurred at a constant temperature close to 25° C. The effectsof day and night temperatures on total dry weight showed a considerabledegree of independence. The optimum day temperature was 25°C irrespective of the night temperature; the optimum night temperatureincreased from 18° to 25° C over the whole range ofday temperature. On average, day temperature affected totaldry weight twice as much as night temperature. High night temperaturesto some extent compensated for low day temperatures. The optimumday and night temperatures for leaf growth were both 25°C. On average day temperature affected leaf growth one and ahalf times as much as night temperature. By 12-hourly sampling it was shown that the cotyledons and leavesgrow throughout both day and night and that high night temperatureaccelerates nocturnal growth (cotyledons by cell expansion,young leaves by cell multiplication). Plants having receivedonly one night at 25° C, as compared with 15° C, showa slightly greater assimilation during the following light period,apparently as a consequence of increased photosynthetic surface.The respiratory loss in dry weight during darkness was not significantlyaffected by temperature over the range 15–25° C.  相似文献   

16.
The genes encoding the catalytic domains (CD) of the three endoglucanases (EG I; Cel7B, EG II; Cel5A, and EG III; Cel12A) from Trichoderma reesei QM9414 were expressed in Escherichia coli strains Rosetta-gami B (DE3) pLacI or Origami B (DE3) pLacI and were found to produce functional intracellular proteins. Protein production by the three endoglucanase transformants was evaluated as a function of growth temperature. Maximal productivity of EG I-CD at 15°C, EG II-CD at 20°C and EG III at 37°C resulted in yields of 6.9, 72, and 50 mg/l, respectively. The endoglucanases were purified using a simple purification method based on removing E. coli proteins by isoelectric point precipitation. Specific activity toward carboxymethyl cellulose was found to be 65, 49, and 15 U/mg for EG I-CD, EG II-CD, and EG III, respectively. EG II-CD was able to cleave 1,3–1,4-β-d-glucan and soluble cellulose derivatives. EG III was found to be active against cellulose, 1,3–1,4-β-d-glucan and xyloglucan, while EG I-CD was active against cellulose, 1,3–1,4-β-d-glucan, xyloglucan, xylan, and mannan.  相似文献   

17.
Continuous and simultaneous measurements of CO2 exchange andtranspiration rates of whole soybean plants were made undercontrasting, controlled environmental conditions for periodsof up to 3 d. Daytime temperatures and vapour pressure deficits(VPD) were 27.5 °C/12 mb; 27.5 °C/5 mb; 22.5 °C/12mb, and 22.5 °C/5 mb. Night temperatures were 5°C lowerthan day temperatures and night VPD was 2.7 mb and 3.5 mb atthe higher and lower temperature respectively. The experimentalconditions were virtually the same as those under which theplants had been grown. Transpiration rates were higher at the higher VPD but were alsoinfluenced by temperature. At 12 mb VPD the rates were 16 percent lower at 22.5 °C than at 27.5 °C. Temperature hadno effect on the transpiration rate at 5 mb VPD. Photosynthesis rates were lower at 5 mb VPD than at 12 mb VPDat both temperatures: the difference was substantially greater(c. 70 per cent) at 22.5 °C. Under all treatments meso-phyllresistance (r'm) appeared to have a major effect on the photosyntheticrate, and varied more than twofold between treatments. r'm washighest in plants grown at 22.5 °C/5 mb VPD and lowest at27.5 °C/12 mb VPD.  相似文献   

18.
The effects of temperature on the polar movement of IAA through6-mm and 12-mm segments of Zea mays roots have been investigatedover the range from 1 to 50°C. At all temperatures an acropetal polar movement of IAA predominated,although at low temperatures and at 50°C the 6-mm segmentsshowed a transient basipetal polarity, before the persistentacropetal polarity developed. At 1°C the differences betweenacropetal and basipetal movement of IAA were less distinct thanat the other temperatures. There is, however, a marked metabolically-dependentacropetal movement of IAA through the tissues at 1°C, becausewhen the segments were deprived of oxygen the acropetal movementwas severely reduced while the basipetal movement was reducedto a smaller extent. At 1°C and at 5°C there was alwaysa persistent basipetal polarity of IAA movement through 6-mmand 12-mm segments under anaerobic conditions. The velocity of acropetal movement (mm h–1) was the samethrough the 6-mm and the 12-mm segments and was markedly affectedby temperature. It increased from 1°C to a maximum valueof 8 mm h–1 at 31°C and then decreased again at 40and 50°C. The velocity of basipetal movement could be assessedonly at 1 and 5°C at which temperatures it was greater thanthe velocity of acropetal movement, and virtually independentof segment length. The acropetal flux of IAA (cpm h–1) was much less through12-mm segments than through 6-mm segments. For both lengthsof segment, however, the flux showed a complex relationshipwith ambient temperature, increasing from 1°C to a maximumat 10–15°C, declining to a minimum value at 31°Cand then rising again at 40 and 50°C. The basipetal fluxof IAA could be astimated only at 1 and 5°C at which itwas very much smaller than the acropetal flux. The amount of IAA in the receiver blocks increased linearlywith time at the lower temperatures. At temperatures withinthe range 15°C to about 31°C, however, the amount ofIAA in the receiver blocks began to decline if the transportperiods exceeded a certain length. The time at which this declinein the IAA in the receiver block began was related to the ambienttemperature. Chromatographic analysis indicated one radioactive substancein receiver blocks at the apical end of segments supplied withIAA-1-14C at the basal end after transport periods of 6 h at25°C, and 72 h at 5°C. The Rf of this substance wasclosely similar to that of the radioactive IAA supplied in thedonor blocks.  相似文献   

19.
Stimulation of the rate of photosynthesis at 2·0 kPaO2 in comparison with 21 kPa O2 and carbohydrate accumulationover 4h were measured during exposure of sunflower (Helianthusannuus L.) and rape (Brassica napus L.), grown at 30 °Cand 13 °C, to temperatures between 7 °C and 35 °C.The effect of reducing source: sink ratio by shading on theresponse of photosynthetic rate to temperature was also determined.Stimulation of photosynthesis by 2·0 kPa O2 in comparisonwith 21 kPa O2 decreased over 4 h at cool temperatures in sunflowerplants grown at 30 °C but not in rape grown at 30 °C.Stimulation did not decrease over 4 h in plants grown at 13CC. Sucrose was the main carbohydrate accumulated over 4 h;its accumulation increased with decreasing temperature. Starchaccumulation either decreased or remained the same with decreasingtemperature. In plants grown at 30 °C more carbohydrateaccumulated between 8 °C and 21 °C in sunflower thanin rape, but more carbohydrate accumulated at 30 °C in rapethan in sunflower. In plants grown at 13 °C much less carbohydrateaccumulated between 13 °C and 23 °C than in plants grownat 30 °C. Photosynthetic rate in plants grown at 30 °Cexposed to between 20 °C and 35 °C over 32 h (14 h light-10h dark-8 h light), declined over 32 h at 20 °C and 25 °Cin sunflower and at 20 °C in rape. This fall over 32 h,especially at 20 °C in sunflower, was significantly reducedby shading the rest of the plant. Shading had little effecton photosynthetic rate above 25 °C. The work confirms thatlow temperature imposes a sink-limitation on photosynthesiswhich occurs at higher temperatures in sunflower than in rape.This limitation may be relieved by decreasing the source:sinkratio. Key words: Sunflower, rape, photosynthesis, carbohydrates, sink demand, temperature  相似文献   

20.
The hypothesis that enhanced phytoplankton growth explainedthe increase in phytc-plankton biomass observed at a frontalstructure off the Antarctic Peninsula during the summer of 1993was examined by analysing the phytoplankton cells undergoingmitosis. The frequency of dividing cells (FDC) of an unidentifiedphototrophic species of Cryptophyceae, which dominated phytoplanktonbiomass at the front, varied between 1.2 and 31.6%, with higherpercentages associated with the frontal structure. FDC valuesincreased as phytoplankton biomass increased, suggesting thatactive growth, rather than passive accumulation, was responsiblefor the enhanced phytoplankton biomass observed. The low temperatures(mean ± SE = 1.11 ± 0.52°C) that characterizedthe Antarctic waters sampled imposed an upper limit to the maximumFDC reached by the Cryptophyceae, but the relationship betweenFDC and temperature suggests a clear response of the maximalgrowth rate of this species to small changes in temperature.The stabilization of the water column, resulting in higher lightavailability and heating of the surface water in this frontalarea, appeared to promote the growth of the phototrophic speciesof Cryptophyceae and emphasizes the importance of mesoscaleprocesses as determinants of phytoplankton growth dynamics.  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