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.
Water relations, desiccation tolerance and longevity of Taxus brevifolia (Nutt.) seeds were studied to determine the optimal stage of development and storage conditions for seeds of this species. Seeds equilibrated to a range of relative humidities (RHs) had unusually low water contents which can be accounted for by the high lipid content of gametophyte tissues (71% of the dry mass). Water relations of embryonic tissue were more typical of those reported for other seed species. The water content below which freezing transitions were not observable in the embryo was ca 0.24 g H2O (g dry weight)−1 (g g−1) for all maturity classes studied. Embryos did not achieve significant levels of desiccation tolerance (survival to water contents less than 0.5 g g−1) until the latter stages of development when dry matter was maximal. Mature embryos could be dried to 0.025 g g−1 (seed water content of 0.010 g g−1) with no loss of viability. Thus, at the latter stages of development, embryo water content could be optimized to avoid both desiccation and freezing damage. Survival of mature seeds declined over a 2-year period when seeds were stored at temperatures between 5 and 35°C and RHs between 14 and 75%, corresponding to seed water contents between 0.015 and 0.07 g g−1. The deterioration rate was slowest for seeds stored at the lowest RH and temperature. Our data indicate that seeds of Taxus brevifolia show orthodox rather than recalcitrant storage characteristics, but that the optimum water content for storage was extremely low. The results suggest that even if stored at optimal water contents and low temperatures, T. brevifolia seeds will be relatively short lived. The high quantity of lipids or reducing sugars may be contributing factors in the poor storage characteristics.  相似文献   

2.
Acquisition of desiccation tolerance in soybeans   总被引:10,自引:0,他引:10  
The entry into a desiccation-tolerant state is a major developmental component of seed maturation. Development of desiccation tolerance of embryonic axes of soybean [Glycine max (L.) Merrill cv. Chippewa 64] was studied by measuring changes in electrolyte leakage. germination and relative growth rate after axes were rapidly air-dried to various water contents. Axes acquired the full capacity for germination at 34 days after flowering (DAF). and reached physiological maturity (maximum dry weight) at 48 DAF. When dried to water content h = 0. 08 (g water g−1 dry weight). few axes germinated before 42 DAF. but more than 90% germinated after 48 DAF. However, electrolyte leakage of rehydrated axes showed a linear decline from 30 to 55 DAF. For developing axes there was a critical water content or desiccation threshold. which could be estimated by using the electrolyte leakage method. The threshold of desiccation tolerance decreased gradually from h = 1. 10 to 0. 18 as axes matured from 28 to 55 DAF. The development of desiccation tolerance continued after physiological maturity at 48 DAF. We conclude that the acquisition of desiccation tolerance of soybean axes is a gradual event, rather than an abrupt transition.  相似文献   

3.
A new terrestrial species of green alga ( Chlorella sp. strain DT) that survives under extremely dry conditions was isolated from an arid mountainous area of Taiwan. The water content of the cells dropped to less than 3% after 3 h dehydration at 42°C. The dried cells could regrow if they were rewetted. The photosynthetic activity of the cells ceased following 2 h of dehydration, but respiratory activity was still detectable after 3 h desiccation. During the rewetting process, respiration increased immediately and then decreased to a steady level after 5 days of rewetting, matching the net photosynthetic oxygen evolution. The cells had 38% neoxanthin (1520 nmol g−1 dry weight), 39% lutein (1580 nmol g−1 dry weight), and 14% violaxanthin (570 nmol g−1 dry weight) of total carotenoids, but only 5.1%β-carotene (210 nmol g−1 dry weight), 3.8%α-carotene (150 nmol−1 g dry weight) and a trace of antheraxanthin and zeaxanthin. Upon dehydration, zeaxanthin and carotene contents increased and recovered to normal levels after 8 days of rewetting. The photoprotective xanthophyll cycle was found in these cells. It appears that the energy dissipation process for preventing photodamage is perhaps one of the tolerance mechanisms in this Chlorella strain.  相似文献   

4.
The effects of water hardness (9 and 220 mgl−1 as CaCO3) upon zinc exchange in brown trout exposed to 0.77 μmol Zn 1−1 have been investigated using artificial soft water (<49.9 μmol Ca l-1, <40.1 μmol Mg 1−1) and mains hard water (1671.7 μmol Ca 1−1, 493.6 μmol Mg 1−1) of known composition. Both hard and soft water-adapted fish exhibited a bimodal pattern of net zinc influx. Net zinc influxes during both fast and slow uptake phases were significantly greater ( P <0.001) in soft (82.9 and 6.2 μmol Zn 100 g−1 h−1) than in hard water (46.3 and 2.4 μmol Zn 100 g h−1). Zinc efflux (- 0.2 μmol Zn 100 g−1 h−1) was enhanced only in hard water during the slow net influx phase.
Brown trout exposed to zinc in hard water and placed in metal-free media exhibited a greater net efflux (- 25.6 μmol Zn 100 g−1 h−1) of the metal than did fish in soft water (-4.2 μmol Zn 100 g−1 h−1) treated in the same manner. Tissue 65Zn activities reflected both the differences in uptake and excretion rates of the metal between hard and soft water fish. During zinc exposure (0.77 μmol Zn 1−1) high water hardness reduced tissue burdens of the metal by reducing net branchial influx, and enhancing efflux of the metal in hard water fish.  相似文献   

5.
Desiccation tolerance is initiated in wheat (Triticum aestivum L.) embryos in planta at 22 to 24 d after anthesis, at the time that the embryo water content has decreased from about 73% fresh weight (2.7 g water/g dry weight) to about 65% fresh weight (1.8 g water/g dry weight). To determine if desiccation tolerance is fully induced by the loss of a relatively small amount of water, detached wheat grains were treated to reduce the embryo water content by just a small amount to approximately 69% (2.2 g water/g dry weight). After 24 h of such incipient water loss, subsequently excised embryos were able to withstand severe desiccation, whereas those embryos that had not previously lost water could not. Therefore, a relatively small decrease in water content for only 24 h acts as the signal for the development of desiccation tolerance. Embryos that were induced into tolerance by a 24-h water loss had no detectable raffinose. The oligosaccharide accumulated at later times even in embryos of detached grains that had not become desiccation tolerant, although tolerant embryos (i.e. those that previously had lost some water) contained larger amounts of the carbohydrate. It is concluded that desiccation tolerance and the occurrence of raffinose are not correlated. Immunodetected dehydrins accumulated in embryos in planta as desiccation tolerance developed. Detachment of grains induced the appearance of dehydrins at an earlier age, even in embryos that had not been made desiccation tolerant by incipient drying. It is concluded that a small reduction in water content induces desiccation tolerance by initiating changes in which dehydrins might participate but not by their interaction with raffinose.  相似文献   

6.
Hydration, protons and onset of physiological activities in maize seeds   总被引:1,自引:0,他引:1  
Dry maize ( Zea mays L.) seed components, namely, embryo and endosperm, provide model materials for studies on water-dependent mechanisms in cellular function. We explored the thermodynamics of hydration for both tissues, along with their dielectric behavior, as a function of water content. In addition, we evaluated the direct current (DC) conductivity due to water protons. Our data on embryo tissue show large enthalpic and entropic peaks at water content [h, in g H2O (g dry sampie)−1] around 0.08 g g−1, indicating very tight binding and ordering of water molecules. With increasing water content both enthalpy and entropy decrease, and the completion of primary hydration requires h ∼ 0.26 g g−1. Data for endosperm tissue show the absence of such an enthalpic peak and a reduced degree of ordering for h < 0.10 g g−1. The DC protonic conductivity shows explosive growth above a threshold hydration level hc= 0.082 g g−1 and hc= 0.12 g g−1, for embryo and endosperm, respectively. Protonic conduction can be considered within the framework of a percolation modell characterized by a hydration threshold and by a power law increase in conductivity with further hydration. The critical exponent of the power law is in agreement with theory for a two-dimensional percolative process. This percolative water-assisted behavior reflects the presence of an extended network of water molecules adsorbed on the surface of proteins and/or membranes inside cells. We consider this percolative protonic conduction as being a prerequisite to respiration processes.  相似文献   

7.
Water content of Atlantic salmon parr fell from about 84% at emergence (late May) to just under 79% in September but rose again towards March. Na+ content consequently rose from 3·3 mg g−1 dry wt at the beginning of June to 6·2 mg g−1 in early July. It then fell to 4·4 mg g−1 in September, rising again towards March. K+ content rose to a maximum in July to stabilize at 16·6 mg g−1 dry wt in September. The resultant Na+/K+ ratio peaked at 0·43: 1 in mid-June, falling to a minimum in mid-August but rising again in March reflecting changes in the relative proportions of intra and extracellular water. The changes in whole-body chemistry suggest a period of nutritional stress immediately after emergence and during the winter. In streams at higher altitude and of lower nutrient status, nutritional stress during the winter appears to be more severe.  相似文献   

8.
Depending on the environmental conditions, imbibed seeds survive subzero temperatures either by supercooling or by tolerating freezing-induced desiccation. We investigated what the predominant survival mechanism is in freezing canola ( Brassica napus cv. Quest) and concluded that it depends on the cooling rate. Seeds cooled at 3°C h−1 or faster supercooled, whereas seeds cooled over a 4-day period to −12°C and then cooled at 3°C h−1 to−40°C did not display low temperature exotherms. Both differential thermal analysis and nuclear magnetic resonance (NMR) spectroscopy confirmed that imbibed canola seeds undergo freezing-induced desiccation at slow cooling rates. The freezing tolerance of imbibed canola seed (LT50) was determined by slowly cooling to −12°C for 48 h, followed with cooling at 3°C h−1 to −40°C, or by holding at a constant −6°C (LD50). For both tests, the loss in freezing tolerance of imbibed seeds was a function of time and temperature of imbibition. Freezing tolerance was rapidly lost after radicle emergence. Seeds imbibed in 100 μ M abscisic acid (ABA), particularly at 2°C, lost freezing tolerance at a slower rate compared with water-imbibed seeds. Seeds imbibed in water either at 23°C for 16 h, or 8°C for 6 days, or 2°C for 6 days were not germinable after storage at −6°C for 10 days. Seeds imbibed in ABA at 23°C for 24 h, or 8°C for 8 days, or 2°C for 15 days were highly germinable after 40 days at a constant −6°C. Desiccation injury induced at a high temperature (60°C), as with injury induced by freezing, was found to be a function of imbibition temperature and time.  相似文献   

9.
The ability of seeds to withstand desiccation develops during embryogenesis and differs considerably among species. Paddy rice (Oryza sativa L.) grains readily survive dehydration to as low as 2% water content, whereas North American wild rice (Zizania palustris var interior [Fasset] Dore) grains are not tolerant of water contents below 6% and are sensitive to drying and imbibition conditions. During embryogenesis, dehydrin proteins, abscisic acid (ABA), and saccharides are synthesized, and all have been implicated in the development of desiccation tolerance. We examined the accumulation patterns of dehydrin protein, ABA, and soluble saccharides (sucrose and oligosaccharides) of rice embryos and wild rice axes in relation to the development of desiccation tolerance during embryogenesis. Dehydrin protein was detected immunologically with an antibody raised against a conserved dehydrin amino acid sequence. Both rice and wild rice embryos accumulated a 21-kD dehydrin protein during development, and an immunologically related 38-kD protein accumulated similarly in rice. Dehydrin protein synthesis was detected before desiccation tolerance had developed in both rice embryos and wild rice axes. However, the major accumulation of dehydrin occurred after most seeds of both species had become desiccation tolerant. ABA accumulated in wild rice axes to about twice the amount present in rice embryos. There were no obvious relationships between ABA and the temporal expression patterns of dehydrin protein in either rice or wild rice. Wild rice axes accumulated about twice as much sucrose as rice embryos. Oligosaccharides were present at only about one-tenth of the maximum sucrose concentrations in both rice and wild rice. We conclude that the desiccation sensitivity displayed by wild rice grains is not due to an inability to synthesize dehydrin proteins, ABA, or soluble carbohydrates.  相似文献   

10.
Ethylene production from an embryogenic culture of Norway spruce ( Picea abies L.) was generally low. ca 2.5 nl g−1 h−1, whereas 1-aminocyclopropane-1 -carboxylic acid (ACC) concentration was high, fluctuating between 50 and 500 nmol g−1 during the 11-day incubation period. Hypoxia (2.5 and 5 kPa O2) rapidly inhibited ethylene production without subsequent accumulation of ACC. Exogenous ACC (1, 10 and 100 μ M ) did not increase ethylene production, but the highest concentrations inhibited tissue growth. Ethylene (7 μl I−1) did not inhibit growth either when supplied as ethephon in the medium or in a continuous flow system. Benzyladenine (BA) had little effect on ethylene production, although it was necessary for sustaining the ACC level. Omission of 2.4-dichloro-phenoxyacetic acid (2.4-D) from the medium caused ethylene production to increase from about 2.5 to 7 nl g−1 h−1 within the 11-day incubation period. Although 2.4-D did not specifically alter the endogenous level of ACC, the lowest ACC level, 33 nmol g−1, was observed in tissue treated with 2.4-D (22.5 μ M ) and no BA for 11 days. Data from this treatment were used to estimate the kinetic constants for ACC oxidase, the apparent Km was 50 μ M and Vmax 2.7 nl g−1 h−1. Growth of the tissue was strongly inhibited by 2.4-D in the absence of BA, but weakly in the presence of BA (4.4 μ M ). The results suggest that ethylene or ACC may be involved in the induction of embryogenic tissue and in the early stages of embryo maturation.  相似文献   

11.
ABSTRACT Arboreal and terrestrial ants were exposed to 0, 25, 50, 75 and 100 (control)% r.h., at 30oC. Desiccation resistance increased with body size (as dry weight0.55), but not as quickly as expected from the consequences of the surface area and volume relationship (as dry weight0.67). Arboreal ants took 8 times longer to die than terrestrial ants of comparable size. Even after size effects were removed, desiccation resistance differed between various terrestrial species and showed a correlation with foraging patterns.
Arboreal and terrestrial ants whose waterproofing epicuticular lipids were removed by chloroform: methanol extraction had equally high water loss rates at 0% r.h. Unextracted arboreal ants had water loss rates half those of unextracted terrestrial ants, suggesting that differences between them were based on differences in epicuticular lipids. The lower water loss rates of arboreal ants contributed significantly to their longer survival under desiccation. Arboreal ants also had greater total rectal pad area than terrestrial ants, suggesting that they may be able to reclaim faecal water more effectively. There were no differences in the minimum viable water content between the two groups of ants. Both had water loss tolerances comparable with those of arthropods adapted to xeric environments. Initial water loss rates could not account for all of the differences in desiccation resistance between arboreal and terrestrial ants. Other adaptations to desiccation stress by arboreal ants are likely.
Comparisons of water loss rates and desiccation resistance between arboreal and desert ants suggest that the arboreal habitat is at least as stressful as the desert habitat.  相似文献   

12.
The energy density ( E D) of anchovy Engraulis encrasicolus in the Bay of Biscay was determined by direct calorimetry and its evolution with size, age and season was investigated. The water content and energy density varied seasonally following opposite trends. The E D g−1 of wet mass ( M W) was highest at the end of the feeding season (autumn: c . 8 kJ g−1 M W) and lowest in late winter ( c . 6 kJ g−1 M W). In winter, the fish lost mass, which was partially replaced by water, and the energy density decreased. These variations in water content and organic matter content may have implications on the buoyancy of the fish. The water content was the major driver of the energy density variations for a M W basis. A significant linear relationship was established between E D g−1 ( y ) and the per cent dry mass ( M D; x ): y =−4·937 + 0·411 x . In the light of the current literature, this relationship seemed to be not only species specific but also ecosystem specific. Calibration and validation of fish bioenergetics models require energy content measurements on fish samples collected at sea. The present study provides a first reference for the energetics of E. encrasicolus in the Bay of Biscay.  相似文献   

13.
Abstract: Substrate utilization of microbial cells extracted from soil with a 0.85% aqueous sodium chloride solution, was determined to estimate effects on soil microorganisms at the community level with microtiter plates (Biolog GN®) containing 95 different sources of organic carbon. A consistent pattern of utilized substrates was obtained after 24 h of microtiter plate incubation at 28°C. The absorbance values (OD590) obtained from a microtiter plate reader after background correction were transformed by using the average absorbance values of oxidized substrates as a threshold to distinguish between well utilized and poorly or non-utilized substrates and thereby reduce variances between replicates. Doubling times of the extracted soil microorganisms in the microtiter plates were tested with 12 substrates and ranged from 1.96 h to 3.23 h, depending on the carbon source. The carbon source utilization assay was used to assess the effects of soil inoculation with Corynebacterium glutamicum with and without a genetically engineered plasmid (pUN1; 6.3 kb), which encoded for the synthesis of the mammalian protease inhibiting peptide, aprotinin. Additionally, aprotinin itself was added at two concentrations to soil samples. An identical decrease in the number of carbon sources utilized, especially carbohydrates, occurred upon soil inoculation with both C. glutamicum strains after inoculation with 106 cells g−1 soil. This effect was only detectable during the first three weeks of incubation, as long as cell numbers of C. glutamicum (pUN1) were above 105 cfu g−1. Soil amendment with aprotinin resulted in utilization of additional substrates, most of them carbohydrates. With 0.1 mg aprotinin g−1 soil this stimulation lasted 2 days and with 10 mg g−1 it lasted for 7 days.  相似文献   

14.
Abstract. Primary myogenic cell cultures derived from 12-day embryos of genetically fast-growing chickens (fast cultures) and slow-growing chickens (slow cultures) were grown under identical conditions to examine differences in growth and differentiation at the cellular level. The two types of cultures exhibited significant ( P <0.01) differences in proliferation, protein accumulation, response to the addition of insulin to the culture medium and the amount of insulin bound per nucleus. The fast cultures exhibited a larger number of both total nuclei and fused nuclei at 48, 72 and 96 h in culture, accumulated more protein per nucleus at 24, 48 and 72 h in culture and demonstrated a greater response to the addition of insulin to the culture medium, as reflected by increased fusion rate and protein accumulation at 24 h in culture. Maximal response to insulin in both types of cultures was obtained at 24 h to added insulin concentrations of 10−10−10−9 M . Slow cultures bound more [125I]-insulin than fast cultures at 24 h in culture. These experiments suggest that different muscle growth potentials in animals of the same species are at least partly due to intrinsic cellular differences in the myogenic cells that give rise to adult muscle tissue.  相似文献   

15.
Measurements of dry weight (wt), carbon (C), nitrogen (N) and calories were made on walleye pollock eggs (0.24 mg, 35.3% C, 8.3% N, and 4.6 kcal g−1 dry wt), larvae (0.16 g, 42.9% C, 11.1% N and 5.1 kcal g−1 dry wt) and juveniles (22.4 g, 47.2% C, 9.0% N and 5.6 kcal g−1 dry wt). For juvenile fish (9–360 g wet wt) the measured values were related to dry weight and Fulton's condition factor index (CFI) by regression models. The CFI was a better predictor of body composition than dry weight. As CFI improved from a minimum starvation level of 0.42 to a maximum of 1.16, body caloric content, percentage C, and the C/N ratio increased (kcal g−1 dry wt = 4.4 CFI + 1.7, percentage carbon = 49.7 CFI0.5, C/N ratio = 5.0 CFI + 0.9), while percentage N and percentage ash decreased (percentage N =−3.5 CFI + 12.1; percentage ash = 9.1 CFI−1.4). The results of this study suggest that seasonal C, N and caloric content of young pollock can be estimated from measurements of Fulton's condition factor index.  相似文献   

16.
Wet mass and water content of four lots of whole eggs did not change throughout embryonic development of rainbow trout Oncorhynchus mykiss. Eggs in all four lots accumulated Na+. Eggs in lots 2 and 4 also accumulated Ca2+ and Cl-, whereas eggs in lot 1 showed no significant change in Ca2+ or Cl- and eggs in lot 3 showed no change in Cl-and a small loss of Ca2+. Although the Na+ content of embryonic tissues increases in the later stages of development, the yolk sac content remained constant, indicating uptake of Na+ from the environment. Na+ uptake by whole eggs was non-saturable, consistent with diffusion of Na+ across the chorion into the perivitelline fluid. Na+ uptake in dechorionated embryos was saturable, as was Ca2+ uptake by both whole eggs and dechorionated embryos, consistent with active uptake or facilitated diffusion mechanisms at the surface of embryos. Very low Ca2+ uptake rates in dechorionated embryos suggest that the Ca2+ uptake mechanism is not fully developed until after hatching.  相似文献   

17.
18.
Ascorbic acid (AA) in the leaf apoplast has the potential to limit ozone injury by participating in reactions that detoxify ozone and reactive oxygen intermediates and thus prevent plasma membrane damage. Genotypes of snap bean ( Phaseolus vulgaris L) were compared in controlled environments and in open-top field chambers to assess the relationship between extracellular AA content and ozone tolerance. Vacuum infiltration methods were employed to separate leaf AA into extracellular and intracellular fractions. For plants grown in controlled environments at low ozone concentration (4 nmol mol−1 ozone), leaf apoplast AA was significantly higher in tolerant genotypes (300–400 nmol g−1 FW) compared with sensitive genotypes (approximately 50 nmol g−1 FW), evidence that ozone tolerance is associated with elevated extracellular AA. For the open top chamber study, plants were grown in pots under charcoal-filtered air (CF) conditions and then either maintained under CF conditions (29 nmol mol−1 ozone) or exposed to elevated ozone (67 nmol mol−1 ozone). Following an 8-day treatment period, leaf apoplast AA was in the range of 100–190 nmol g−1 FW for all genotypes, but no relationship was observed between apoplast AA content and ozone tolerance. The contrasting results in the two studies demonstrated a potential limitation in the interpretation of extracellular AA data. Apoplast AA levels presumably reflect the steady-state condition between supply from the cytoplasm and utilization within the cell wall. The capacity to detoxify ozone in the extracellular space may be underestimated under elevated ozone conditions where the dynamics of AA supply and utilization are not adequately represented by a steady-state measurement.  相似文献   

19.
Germination of Archontophoenix alexandrae seeds and embryos were studied under gradient water content treatments throughout the seed development phases of maturation in 2005 to investigate seed desiccation tolerance and storage characteristics. During the maturation process, seed water content decreased gradually from55 DAF (days after flowering) to 70 DAF, and seeds reached the maximum dry-weight at 90 DAF. Seed germinability appeared after 60 DAF. Seeds germinated with a temperature range from15℃- 40℃ under alternating photoperiod (14 h light, 10 h dark, 12μmol m- 2s - 1 ), while the best germination percentage was obtained between 30℃- 35℃. A maximum germination capacity reached at 70 DAF. However, seed germination was greatly inhibited by light. Desiccation tolerance of seeds and embryos increasedgradually from 55 DAF to 90 DAF and reached the maximum at 90 DAF with a semilethal water content of 0.18 g/g ( seed) and 0.3 g/g ( embryo) respectively. Rapid dehydration maintained higher seed germination percentage than thatof slow dehydration when drying to the same water content. Seeds with without water content treatments failed to germinate after 1 month storage under - 18℃, whereas appropriate desiccation treatment prolonged seed longevity under 4℃, 10℃ and 15℃ storage temperatures. It revealed obviously the recalcitrant characteristics of Archontophoenix alexandrae seeds torage behaviour which are tolerant toward neither deep desiccation nor low temperatures.  相似文献   

20.
Abstract— To establish compartments involved in depolarization-induced release of γ-aminobutyric acid (GABA) in rat brain slices, the amount of exogenous labeled and endogenous GABA released and retained was followed during 48 min exposure to 50 m m -K+ or to 50 μ m -veratridine. Endogenous GABA was measured with high performance liquid chromatography. The presence of 10 μ m -aminooxyacetic acid throughout prevented both the metabolism of GABA and the formation of endogenous GABA due to depolarization. During super-fusion with 50 m m -K+ and 2.6 m m -Ca2+ the efflux of labeled and endogenous GABA after an initial large increase declined to 10% of the highest value with constant and identical rates. Kinetic analysis of efflux showed that 10% of endogenous and 25% of labeled GABA present is available for release by high K+ and Ca2+. In the absence of Ca2+, release by high K+ of both labeled and endogenous GABA was nearly suppressed. Veratridine, unlike high K+, caused an efflux which declined with an initial fast and late very slow phase. The slow efflux by veratridine was doubled in the absence of Ca2+. Exposure to veratridine in the absence of Ca2+ during 120 min released nearly 70% of labeled and endogenous GABA present. Results suggest that only about 0.25 μmol g−1 endogenous GABA is the source of physiological Ca2+-dependent release, while much of the remaining GABA present is released only under unphysiological conditions.  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