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.
Microencapsulation of bixin using high-amylose corn starch was carried out by the acidification method. Bixin powders were characterized by differential scanning calorimetry (DSC), X-ray diffractometry (XRD), FT-IR spectrometry, color parameters, encapsulation efficiency, bixin release profile. In addition, the effect of whey protein (WP) on the microencapsulation process was investigated. The results obtained from DSC, X-ray diffraction and FT-IR spectrometry indicated that only in the samples prepared at 90 °C (B0WP90°C, B10WP90°C, and B20WP90°C) there was formation of crystalline structures, with melting temperatures at 117.2°, 105° and 104 °C, respectively. The possible interactions between bixin, WP and amylose starch are also discussed.  相似文献   

2.
Sorghum grain (two varieties) was modified to different water contents (12% to 16% wet weight basis) and heated to 60°C, 70°C and 80°C for periods of 4, 8 and 12 min. Germination, seedling vigour, seedling dry matter, free fatty acid (FFA) content, fungal contamination and infestation with the insect pest Rhyzopertha dominica (F.) were all markedly affected by heat treatment. The effectiveness of the heat treatment was also influenced by the size of the sample used. Heat treatment at 60°C and 70°C resulted in germination being unaffected or stimulated while at 80°C and the higher water contents significantly reduced, when compared to untreated controls. The dry matter of seedlings, and seedling vigour was positively correlated with germination and heat treatment. Heat treatment had no effect on FFA. All stages of the insect pest, R. dominica, were destroyed at 70°C and an 8 min exposure time. However, the water content of the sorghum was critical in determining the efficacy against this pest. The percentage fungal contamination of grain was reduced from 90% to about 25% by heat treatment. However, some grain fungi, particularly Eurotium spp., Aspergillus niger and Penicillium spp. could still be isolated from sorghum grain treated at 80°C for up to 12 minutes.  相似文献   

3.
Fractionation of the cell wall material of parenchyma of mature runner beans with and without chlorite-HOAc treatment, clearly showed that at least two main types of wall proteins were present. One relatively rich in hydroxyproline (HP) associated with α′-cellulose, from which most (90%) of it could be readily liberated by chlorite-HOAc treatment and the other relatively poor in HP associated with hemicellulose A. The chlorite HOAC solubilized “glycoprotein” contained a high proportion of arabinose and galactose. It was purified by PhOH-H2O fractionation and the molar ratios of HP, arabinose, galactose, xylose, rhamnose, glucose and uronic acid in the purified glycoprotein (“glycoprotein X”) were 1:2·6:2·4:0·2:0·2:0·1:0·3. The principal amino acids of glycoprotein X were HP (43·5 mol%), serine and proline which together comprised 66 mol% of the total. These results suggest that the HP-rich wall glycoprotein is associated with cellulose microfibrils and approximates in conformation to polyhydroxyproline carrying arabinose and galactose oligosaccharide side chains.  相似文献   

4.
In order to improve the gas evolution rate during anaerobic digestion of coffee waste by two-phase methane fermentation with slurry-state liquefaction, the liquefaction and gasification processes were separately investigated. In the liquefaction process (including the acidification process), treatment at a pH above 6 had no major effects on the liquefaction and acidification rates. However, the VFA production rates were 880 and 320 mg/l·d during mesophilic (37°C) and thermophilic (53°C) liquefaction, respectively. Mesophilic conditions were superior to thermophilic conditions in the liquefaction. With respect to the gasification process, a high TOC volumetric loading rate of 21 g/l·d was achieved during thermophilic gasification. However, the mesophilic gasification did not yield stable data, even at a low TOC volumetric loading rate of 2 g/l·d. The gas yield was 1.7 l/g TOC consumed during thermophilic gasification. The thermophilic liquefaction and thermophilic gasification reactors were connected in series and a two-phase experiment was conducted with the reactors at various volumetric ratios. The maximum gas evolution rate of 1.43 l/l·d was achieved with a combination of a gasification reactor with a 0.45l working volume and liquefaction reactor with a 2l working volume. This rate was 1.7 times higher than the rate obtained in a previous study.  相似文献   

5.
Aims: To determine thermal resistance, the effect of pasteurization temperature variations (c. 2°C) in a continuous system in the number of decimal reductions (n) of a Byssochlamys strain in clarified apple juice (CAJ). Methods and Results: Thermal destruction kinetics of Byssochlamys fulva IOC 4518 in thermal death tubes were determined at 85°, 90°, 92° and 95°C by using Weibull distribution frequency model. Three processes with different heating and holding temperatures (A: 94°, 92°C; B: 95°, 93°C; C: 96°, 94°C, respectively) were performed in a continuous system. Process time was 30 s. δ (time of first decimal reduction) values were: 42·98, 8·10, 3·62 and 1·81 min. Variable n ranged from 0·16 to >4·78 for process B (equivalent to industrial). Variable n (0·95–2·66 log CFU ml?1) were obtained in CAJ bottles processed under condition B, while process A resulted in total heat‐resistant mould (HRM) survival and process C in total HRM destruction. Conclusions: This study demonstrates that small variations in temperature during the CAJ pasteurization could result in elimination or survival of HRM due to its nonlogarithmic behaviour. Significance and Impact of the Study: This was the first study to use Weibull frequency method to model inactivation of HRM in fruit juices. Temperature variations could culminate in the presence of HRM in pasteurized juices even when low counts (<10 spores per 100 ml) were present in the raw materials.  相似文献   

6.
Bacillus licheniformis MB-2, isolated from a hot spring water in Manado, Indonesia, secreted a unique chitosanase. Media consisted of 0.24% chitosan, 0.25% casiton, 1% MgSO4, 1.4% K2HPO4, 0.02% CaCl2·2H2O, 0.002% FeSO4·7H2O (w/v) was used for enzyme production. Purification of the enzyme through the hydrophobic interaction chromatography system (butyl Sepharose 4 FF) resulted in two major active fractions; the F2 fraction was shown as a single band at both sodium dodecyl sulfate-polyacrylamide gel electrophoresis and zymogram analysis with apparent molecular mass of 75 kDa. The enzyme worked best at 70°C and pH between 6.0 and 7.0. When incubated at 70, 80, and 90°C, the t1/2 values were 26.56, 18.44, and 16.74 min, respectively with the k constant being at 0.026, 0.037, and 0.04/min. When heated at 90°C, the enzyme retained its activity up to 8 h in the presence of 1mM MnCl2. The enzyme's activity was unaffected by the presence of 1 M NaCl and 6 M urea but was decreased by 2 M of guanidine hydrochloride. Albeit the enzyme did not degrade colloidal and glycol chitin, it hydrolyzed glycol chitosan up to 0.8% and colloidal chitosan up to 11%. The 85% deacetylated (DDA) soluble chitosan was the most susceptible to this enzyme, followed by 90% and 100% DDA chitosan. The K m app values of the 85, 90, and 100% DDA soluble chitosans were found as 0.23, 0.24, and 0.58 mg/mL, whereas the Vmax values were 843, 668, and 261 U/mg, respectively. The hydrolysis products of F2 chitosanase at 24 h incubation (70°C) were pentasaccharide (GlcN)5 and hexasaccharide (GlcN)6. The prelimiaary test showed inhibitory effect of chitooligosaccharides resulted from enzymatic degradation toward Pseudomonas aeruginosa, Salmonella typhimurium. Listeria monocytogenes, Bacillus cereus, Escherichia coli, and Staphylococcus aureus.  相似文献   

7.
8.
The critical thermal maxima (TMAX) of threadfin shad Dorosoma petenense exposed to standardized stress (30 s handling in a dip‐net), simulating stressors endured during fish loading before transport, were measured over a range of holding temperatures (15, 20 and 25° C). Dorosoma petenense TMAX showed a significant thermal effect, displaying mean ±s.d . critical thermal maxima of 26·5 ± 1·6, 30·9 ± 1·2 and 33·3 ± 1·4° C, when tested at temperatures of 15, 20 and 25° C, respectively. Dorosoma petenense TMAX levels were also affected by stress, with handled fish showing significantly lower values than control fish exposed to 15 (mean ±s.d . TMAX = 25·6 ± 2·0° C), 20 (27·6 ± 2·8° C) and 25° C (32·0 ± 2·6° C). In addition to providing basic information on D. petenense thermal tolerance, experimental results suggest that fishery managers should consider the whole suite of potential stressors, such as air exposure during handling and fish loading, when developing management criteria.  相似文献   

9.
Abstract

ID NOE 1H NMR spectroscopy at 500 MHz was employed to examine the structure of poly(dA)·poly(dT) in solution. NOE experiments were conducted as a function of presaturation pulse length (50, 30, 20 and 10 msec) and.power (19 and 20 db) to distinguish the primary NOEs from spin diffusion. The 10 msec NOE experiments took 49 hrs and over 55,000 scans for each case and the difference spectra were almost free from diffusion.

The spin diffused NOE difference spectra as well as difference NOE spectra in 90% H2O + 10% D2O in which TNH3 was presaturated enabled to make a complete assignment of the base and sugar protons. It is shown that poly(dA) ·poly(dT) melts in a fashion in which single stranded bubbles are formed with increasing temperature.

Extremely strong primary NOEs were observed at H2′/H2″ when AH8 and TH6 were presaturated. The observed NOEs at AH2′ and that AH2″ were very similar as were the NOEs at TH2′ and TH2″. The observed NOEs at AH2′ and AH2″when AH8 was presaturated were very similar to those observed at TH2′ and TH2″ when TH6 was presaturated. In addition, presaturation of H1′ of A and T residues resulted in similar NOEs at AH2′/H2″ and TH2′/H2″ region and these NOEs at H2′ and H2″ were distinctly asymmetric as expected in a C2′-endo sugar pucker. There was not a trace of NOE at AH8 and TH6 when AH3′ and TH3′ were presaturated indicating that C3′-endo, × = 30–40° conformation is not valid for this DNA. From these NOE data, chemical shift shielding calculations and stereochemistry based computer modellings, we conclude that poly(dA)·poly(dT) in solution adopts a right- handed B-DNA duplex in which both dA and dT strands are conformationally equivalent with C2′-endo sugar pucker and a glycosyl torsion, ×, of ?73°, the remaining backbone torsion angles being φ′ = 221°, ω′ = 212°, ω = 310°, φ = 149°, ψ = 42°, ψ′ = 139°. The experimental data are in total disagreement with the heteronomous DNA model of Arnott et. al. proposed for the fibrous state. (Arnott, S., Chandrasekaran, R., Hall, I.H., and Puigjaner, L.C., Nucl. Acid Res. 11, 4141, 1983).  相似文献   

10.
The effect of incubation temperature (2, 4, 6, 8 and 10° C) on haddock Melanogrammus aeglefinus development and growth during the embryonic period and in subsequent ontogeny in a common post‐hatch thermal environment (6° C) was investigated. Hatching times were inversely proportional to incubation temperature and ranged from 20·3 days at 2° C to 9·1 days at 10° C. Growth rates were directly proportional to incubation temperature during both the embryonic and larval periods. There was a significant decline in growth rates following hatch in all temperature groups. Compared to the endogenously feeding embryos, growth rates in the exogenous period declined by 4·4‐fold at 4° C to 3·9‐fold at 8° C, indicative of the demarcation between the endogenous and exogenous feeding periods. Yolk utilization varied from 17 days at 2° C to 6 days at 10° C and followed a three‐stage sigmoidal pattern with the initial lag period inversely proportional to incubation temperature. Time to 50% yolk depletion varied inversely with temperature but occurred 1–1·5 days post‐hatch at all temperatures. Additionally, the period between 10 and 90% yolk depletion also decreased with increased temperature. Overall developmental rate was sequential with and directly proportional (2·3‐fold increase) to incubation temperature while the time spent in each developmental stage was inversely proportional to temperature. Larger embryos tended to be produced at lower temperatures but this pattern reversed following hatch, as larvae from higher temperature groups grew more rapidly than those from other temperature groups. Larvae from all temperatures achieved a similar length (c.total length 4·5 mm) upon complete yolk absorption. The study demonstrated the significant impact that temperature has upon developmental and growth rates in both endogenous and exogenous feeding periods. It also illustrated that temperature changes during embryogenesis had significant and persistent effects on growth in subsequent ontogeny.  相似文献   

11.
The metastable state silk I structures of Bombyx mori silk fibroin in the solid state were studied on the basis of 15N- and 13C-nmr chemical shifts of Ala, Ser, and Gly residues. The 15N cross-polarization magic angle spinning (CP/MAS) nmr spectra of the precipitated fraction after chymotrypsin hydrolysis of B. mori silk fibroin with the silk I and silk II forms were measured to determine the 15N chemical shifts of Gly, Ala, and Ser residues. For comparison, 15N CP/MAS nmr chemical shifts of Ala were measured for [15N] Ala Philosamia cynthia ricini silk fibroin with antiparallel β-sheet and α-helix forms. The 13C CP/MAS nmr chemical shifts of Ala, Ser, and Gly residues of B. mori silk fibroin with the silk I and silk II forms, as well as 13C CP/MAS nmr chemical shifts of Ala residue of P. c. ricini silk fibroin with β-sheet and α-helix forms, are used for the examination of the silk I structure. Both silk I and α-helix peaks are shifted to a lower field than silk II (β-sheet) for the Cα carbons of the Ala residues, while both Cβ carbon peaks are shifted to higher field. However, the silk I peak of the 15N nucleus of the Ala residue is shifted to lower field than the silk II peak, but the α-helix peak is shifted to high field. Thus, the difference in the structure between the silk I and α-helix is reflected in a different manner between the 13C and 15N chemical shifts. The Cα and Cβ chemical shift contour plots for Ala and Ser residues, and the Cα plot for the Gly residue, were prepared from the Protein Data Bank data obtained for 12 proteins and used for discussing the silk I structure quantitatively from the conformation-dependent chemical shifts. The plots reported by Le and Oldfield for 15N chemical shifts were also used for the purpose. All these chemical shift data support Fossey's model (Ala: ϕ = −80°, φ = 150°, Gly: ϕ = −150°, φ = 80°) and do not support Lotz and Keith's model (Ala: ϕ = −104.6°, φ = 112.2°, Gly: ϕ = 79.8°, φ = 49.7° or Ala: ϕ = −124.5°, φ = 88.2°, Gly: ϕ = −49.8°, φ = −76.1°) as the silk I structure. © 1997 John Wiley & Sons, Inc.  相似文献   

12.
The thermoregulatory responses of ten paraplegic (PA; T3/4-L4) and nine able-bodied (AB) upper body trained athletes were examined at rest and during prolonged arm-cranking exercise and passive recovery. Exercise was performed for 90 min at 80% peak heart rate, and at 21.5 (1.7)°C and 47.0 (7.8)% relative humidity on a Monark cycle ergometer (Ergomedic 814E) adapted for arm exercise. Mean peak oxygen uptake values for the PA and AB athlete groups were 2.12 (0.41) min−1 and 3.19 (0.38) l · min−1, respectively (P<0.05). At rest, there was no difference in aural temperature between groups [36.2 (0.4)°C for both groups]. However, upper body skin temperatures for the PA athletes were approximately 1.0 °C warmer than for the AB athletes, whereas lower body skin temperatures were cooler than those for the AB athletes (1.3 °C and 2.7 °C for the thigh and calf, respectively). Upper and lower body skin temperatures for the AB athletes were similar. During exercise, blood lactate peaked after 15 min of exercise for both groups [3.33 (1.26) mmol · l−1 and 4.30 (1.03) mmol · l−1 for the PA and AB athletes, respectively, P<0.05] and decreased throughout the remainder of the exercise period. Aural temperature increased by 0.7 (0.5)°C and 0.6 (0.4)°C for the AB and PA athletes, respectively. Calf skin temperature for the PA athletes increased during exercise by 1.4 (2.8)°C (P<0.05), whereas a decrease of 0.8 (2.0)°C (P<0.05) was observed for the AB athletes. During the first 20 min of recovery from exercise, the calf skin temperature of the AB athletes decreased further [−2.6 (1.3)°C; P<0.05]. Weight losses and changes in plasma volume were similar for both groups [0.7 (0.5) kg and 0.7 (0.4) kg; 5.4 (4.9)% and 9.7 (6.2)% for the PA and AB athletes, respectively]. In conclusion, the results of this study suggest that the PA athletes exhibit different thermoregulatory responses at rest and during exercise and passive recovery to those of upper body trained AB athletes. Despite this, during 90 min of arm-crank exercise in a cool environment, the PA athletes appeared to be at no greater thermal risk than the AB athletes. Accepted: 7 May 1997  相似文献   

13.
Aims: To investigate the kinetics of thermal inactivation of the bacteriocin‐like substance P34 at different pH and sodium chloride concentration. Methods and Results: Samples of bacteriocin were treated at different time–temperature combinations in the range of 0–300 min and 90–120°C and the kinetic parameters for bacteriocin inactivation were calculated. For all treatments, the thermal inactivation reaction fitted adequately to first‐order model. D‐ and k‐values were smaller and higher, respectively, for pH 4·5 than for 6·0 or 7·0, indicating that bacteriocin P34 was less thermostable at lower pH. At 120, 115 and 100°C, the addition of sodium chloride decreased thermal stability. For other temperatures, addition of NaCl increased stability of the peptide. The presence of greater amount of the salt (50 g l?1) resulted in a higher thermal stability of bacteriocin P34, suggesting that the reduction in water activity of the solution interfered on the stability of the peptide. Conclusions: Based on an isothermal experiment in the temperature range of 90–120°C, and by thermal death time models, bacteriocin P34 is less heat stable at low pH and has increased thermal stability in the presence of NaCl. Addition of NaCl improved the stability of the peptide P34 at high temperatures. Significance and Impact of the Study: Studies on kinetics of thermal inactivation of bacteriocins are essential to allow their proper utilization in the food industry.  相似文献   

14.
Muscle morphology was investigated in newly hatched barramundi Lates calcarifer larvae incubated at set temperatures (26, 29 and 31° C) prior to hatching. Three days after hatching (the start of exogenous feeding), larvae from the 26 and 31° C treatments were each divided into two groups and reared at that temperature or transferred over the period of several hours to 29° C (control temperature). Incubation temperature significantly affected muscle cellularity in the developing embryo, with larvae incubated at 26° C (mean ±s .e . 223·3 ± 7·9) having on average 14·4% more inner muscle fibres than those incubated at 31° C (195·2 ± 8·8) and 4·8% more than those incubated at 29° C (213·5 ± 4·7). Conversely, inner muscle fibre cross‐sectional area significantly increased at the warm incubation temperature in L. calcarifer, so that the total cross‐sectional muscle area was not different between treatment groups. The total cross‐sectional area of superficial muscle fibres and the proportion of superficial to total fibre cross‐sectional area in just hatched L. calcarifer were also affected by incubation temperature, with incubation at the cool temperature (26° C) increasing both the total cross‐sectional area and proportion of superficial muscle fibres. By 9 days post‐hatch, the aforementioned differences were no longer significant. Similarly, there was no difference in total superficial fibre cross‐sectional area between any treatment groups of L. calcarifer, whereas incubation temperature still significantly affected the proportion of superficial to total muscle fibre cross‐sectional area. Larvae hatched and grown at 31° C had a significantly reduced percentage of superficial muscle cross‐sectional area (mean ±s .e . 5·11 ± 0·66%) compared with those incubated and grown at 29° C (8·04 ± 0·77%) and 26° C (9·32 ± 0·56%) and those incubated at 26° C and transferred to 29° C (7·52 ± 0·53%), and incubated at 31° C and transferred to 29° C (6·28 ± 0·69%). These results indicate that changes in muscle cellularity induced by raising or lowering the incubation temperature of L. calcarifer display varying degrees of persistence over developmental time. The significance of these findings to the culture of L. calcarifer is discussed.  相似文献   

15.
When plants of Zea mays L. cv. LG11 that have been grown at optimal temperatures are transferred to chilling temperatures (0–12°C) photoinhibition of photosynthetic CO2 assimilation can occur. This study examines how growth at sub-optimal temperatures alters both photosynthetic capacity and resistance to chilling-dependent photoinhibition. Plants of Z. mays cv. LG11 were grown in controlled environments at 14, 17, 20 and 25°C. As a measure of the capacity for photosynthesis under light limiting conditions, the maximum quantum yields of CO2 assimilation (φa.c) and O2 evolution (φa.o) were determined for the laminae of the second leaves at photon fluxes of 50–150 μmol m-2s-1. To determine photosynthetic capacity at photon fluxes approaching light saturation, rates of CO2 uptake (A1500) and O2 evolution (A1500) were determined in a photon flux of 1500 μmol m-2s-1. In leaves developed at 14°C, φ and φ were 26 and 43%, respectively, of the values for leaves grown at 25°C. Leaves grown at 17°C showed intermediate reductions in φ and φ, whilst leaves developed at 20°C showed no significant differences from those grown at 25°C. Similar patterns of decrease were observed for A1500 and A1500.0 with decreasing growth temperature. Leaves developed at 25°C showed higher rates of CO2 assimilation at all light levels and measurement temperatures in comparison to leaves developed at 17 and 14°C. A greater reduction in A1500 relative to A1500.0 with decreasing growth temperature was attributed to increased stomatal limitation. Exposure of leaves to 800–1000 μmol m-2 s-1 when plant temperature was depressed to ca 6.5°C produced a photoinhibition of photosynthetic CO2 assimilation in all leaves. However, in leaves developed at 17°C the decrease in A1500 following this chilling treatment was only 25% compared to 90% in leaves developed at 25°C. Recovery following chilling was completed earlier in leaves developed at 17°C. The results suggest that growth at sub-optimal temperatures induces increased tolerance to exposure to high light at chilling temperatures. This is offset by the large loss in photosynthetic capacity imposed by leaf development at sub-optimal temperatures.  相似文献   

16.
Mucosal acidification (from pH 8.1 to 6.0) reversibly inhibited the hydroosmotic responses to oxytocin, cyclic AMP and 8-bromo-cyclic AMP in frog urinary bladder. These inhibitory effects were only observed in the presence of a permeant buffer in the apical medium and could also be elicited by CO2 bubbling, even when the mucosal pH was clamped at 8.1. Acid pH reduced the oxytocin-induced net water flux faster than norepinephrine or oxytocin removal and the difference was especially important at low temperature. The time course of recovery from acid pH inhibition was, at 20°C, similar to that of the hormonal action, but when the medium temperature was reduced to 6–7°C, the recovery from acid pH inhibition paradoxically became faster while the oxytocin action was markedly slowed down (t12 of changes in net water fluxes (expressed in min): oxytocin addition at 20°C, 6.2 ± 0.9; at 6°C, 24 ± 3; oxytocin removal at 20°C, 4.7 ± 0.8; at 6°C, 22 ± 3; pH inhibition at 20°C, 2.6 ± 0.2; at 6°C 2.5 ± 0.2; recovery from pH 6 at 20°C, 6.5 ± 0.9; at 6°C, 2.7 ± 0.3). These results can be explained by accepting two main loci sensitive to medium acidification: (1) the cyclase system and (2) an intracellular, temperature-independent, post-cyclic AMP site. The fact that the intramembranous particle aggregates associated with the oxytocin-induced water permeability increase did not disappear after the flow inhibition by acid pH at low temperature suggests that the second effect could be located at the water channel itself.  相似文献   

17.
Diapause in a predaceous mite, Metaseiulus occidentalis, from a Californian vineyard population is a photoperiodically induced, facultative, adult reproductive diapause in females. The laboratory-determined critical photophase at 19°C was estimated at 11·2 hr. At 16°C, the critical photophase under laboratory conditions was approximately 11·6 hr. Temperature influenced the photoresponse of M. occidentalis so that diapause was entirely averted at temperatures of 22, 25, and 30°C. Aestival diapause at higher temperatures and long photophases was lacking. Development was continuous under constant darkness at all the temperatures tested. Diapause termination in laboratory-reared mites occurred spontaneously under the inductive conditions. Under constant 19°C temperatures, females responded to photophases so that diapause was terminated most rapidly under a 16 hr photophase (in 18·6 days); the 12 and 8 hr photophases, at this temperature, were next in their effectiveness, with 27·9 and 73·0 days, respectively, required for termination.  相似文献   

18.
Spore survival and moisture content are two important properties of biopesticides, and both are related to field biocontrol efficacy and storage shelf life. In this study, Paenibacillus polymyxa (HY96-2) was spray-dried on both pilot plant and production scales, and the effects of inlet and outlet temperatures on spore survival and moisture content were investigated. The results showed that inlet temperatures ranging from 170 to 230 °C (at an outlet temperature of 80 °C) had no obvious effect on the two properties during pilot scale processing, although an inlet temperature of 230 °C resulted in higher feed speed. When the outlet temperature on the pilot scale was reduced from 100 to 80 °C, no obvious variations in spore survival and moisture content were found, while a further reduction from 80 to 65 °C resulted in a decline in spore survival from 81.0 to 67.0% and an increase in moisture content from 2.3 to 31.7%. These results indicate that both outlet temperature and moisture content have an effect on spore survival. Optimum inlet and outlet temperatures for P. polymyxa processing were 230 °C and 85–90 °C on a production scale. Under these conditions, spore survival and moisture content were 83.5–86.6% and 2.73––4.12%, respectively.  相似文献   

19.
Honeybees (Apis mellifera) and three species of blowflies (Calliphora vomitoria, Lucilia caesar and L. sericata) were observed on mass-pollinated populations of onions (Allium cepa) in 4·4 × 3·6 m × 2-2·5 m cages in June-August 1977. Pollination activity at temperatures from 14° to 28°C was compared on the basis of mean time per flower touched on each umbel visit. Honeybees did not forage below 16°C. Above 16°C their mean time per flower was short (1·4 s) and varied little with temperature. For blowflies, it decreased markedly from 12·1 s at 14–15·5°C to 2·7 s at 26°C and above, largely because at low temperatures flies spent long periods quiescent or grooming rather than actively feeding. When flowering was not completely synchronous between cultivars, honeybees were more selective than blowflies, but where flowering was synchronous, both types of pollinator visited the two cultivars at random. For the size of cage used, neither type of insect had a distinct advantage as a pollinator of onions, despite their different behaviour patterns.  相似文献   

20.
Samples of three seed lots of each of three cultivars of carrot, celery and onion were primed in polyethylene glycol solution for 2 weeks at 15 °C. Priming reduced the mean germination times (recorded at 15 °C) of all seed lots (compared to the untreated control) by 3–4 days in carrot, 6–10 days in celery and 3–5 days in onion. The largest reductions in mean germination time occurred in the slowest-germinating seed lots. There were highly significant interactions between priming and cultivars, and between priming and seed lots within cultivars for each species. Drying back the primed seeds at 15 °C increased the mean germination times (compared to primed seed which had not been dried) by 0·6 day in carrot and 1·4 days in celery, and there was no interaction with cultivars or seed lots. The corresponding increase for onion was either 1·0 or 1·8 days, according to the cultivar, but this variation was largely attributable to differences in time taken for the dried seeds to re-imbibe. Seeds dried back at 30 °C germinated 0·2·0·7 day (depending on the species) later than those dried at 15 °C. Percentage germination was not affected by either priming or drying back. Priming reduced the spread of germination times in all cultivars. For primed and dried-back seed, the spread of germination times was larger than that of primed seed in certain cultivars, but was always smaller than that of untreated seeds.  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