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 843 毫秒
1.
Ohne ZusammenfassungErklärung der Abbildungen a Aorta - b Blindzipfel des Staubeckens - c Colon - cb hinterer Colonscheitelbogen - ch Rückensaite - d Darmhang - e Endhang - f Lungenfuß - h Rumpfhang - H Herz - hs Hakenstrecke - i Dünndarm - k Kniestelle des Colons - l Lungenflügel - Ii links - L Leberschirm (Säuger) - Le Leberpolster (Vögel) - M Magenbecken - m Milz - n Niere - o Netz - p Hinterlappen der Leber - P Pancreas - q Querwand - r Randbogen des Colons - re rechts - rp Darmrippe - s Darmschwelle - sb Scheitelbogen des Darmwulstes - t Roden - tb Staubecken - u Urniere - v Vorderlappen der Leber - y Pförtner - z Zwerchfell - Zf Zwölffingerdarm - 1 Vorderlappen 2 Zwischenlappen 3 Hinterlappen der Leber des Meerschweinchens  相似文献   

2.
A critical appraisal of a combined stomatal-photosynthesis model for C3 plants   总被引:13,自引:13,他引:0  
Gas-exchange measurements on Eucalyptus grandis leaves and data extracted from the literature were used to test a semi-empirical model of stomatal conductance for CO2 gSc=go+a1A/(cs-I) (1+Ds/Do)] where A is the assimilation rate; Ds and cs are the humidity deficit and the CO2 concentration at the leaf surface, respectively; g0 is the conductance as A → 0 when leaf irradiance → 0; and D0 and a1 are empirical coefficients. This model is a modified version of gsc=a1A hs/cs first proposed by Ball, Woodrow & Berry (1987, in Progress in Photosynthesis Research, Martinus Mijhoff, Publ., pp. 221–224), in which hs is relative humidity. Inclusion of the CO2 compensation point, τ, improved the behaviour of the model at low values of cs, while a hyperbolic function of Ds for humidity response correctly accounted for the observed hyperbolic and linear variation of gsc and ci/cs as a function of Ds, where Ci is the intercellular CO2 concentration. In contrast, use of relative humidity as the humidity variable led to predictions of a linear decrease in gsc and a hyperbolic variation in ci/cs as a function of Ds, contrary to data from E. grandis leaves. The revised model also successfully described the response of stomata to variations in A, Ds and cs for published responses of the leaves of several other species. Coupling of the revised stomatal model with a biochemical model for photosynthesis of C3 plants synthesizes many of the observed responses of leaves to light, humidity deficit, leaf temperature and CO2 concentration. Best results are obtained for well-watered plants.  相似文献   

3.
4.
Summary A thermocouple shows a high accuracy in reading temperatures, but it does not always give a true value for the temperature of the leaf against which it is pressed on. Temperature differences between the leaf and the air moving around it cause deviations of the shown temperature from the actual leaf temperature. A method is described to calibrate thermocouples inside gas exchange cuvettes without obstructing the movement of the air around the leaf, so that the heat exchange between thermocouple and air is taken into account. The reading of the leaf temperature in a steady-state porometer was checked by this method at various temperatures of ambient air (T a ) and of the leaf (T 1 ) and was found to give an average value of T a and T 1 . The effect of incorrect estimation of the leaf temperature on computed diffusive resistances on H2O (r w ) and CO2 (r c ) and intercellular CO2 partial pressures (p i c ) is discussed.Abbreviations A net CO2 uptake - E evaporative transpiration - R correlation coefficient - T Temperature - a heat-transfer coefficient - b ratio of heat-transfer coefficients - q heat transfer - p c CO2 partial pressure - p w H2O partial pressure - r c diffusive resistance on CO2 - r w diff. resistance on H2O - a ambient - i intercellular - l leaf - p porometer - s leaf surface (=boundary layer+stomata) - t thermocouple; 1, 2: number of cuvette  相似文献   

5.
This study tested a multiplicative model of stomatal response to environment for drought‐affected trees of Eucalyptus globulus Labill. growing in southern Australia. The model incorporates a feed‐forward response to vapour pressure deficit of ambient air (δea) and performed well if evaluated using reduced major axis regression and log‐transformed data. There was strong evidence from gas‐exchange data, leaf water potentials and sapflow measurements of the feed‐forward response by stomata to leaf‐to‐air vapour pressure deficit (δel). The response of stomata to δel was irreversible. Stomatal conductance and the rate of net photosynthesis were highly correlated and declined, together with the rate of transpiration, throughout the afternoon as δea increased despite increasing leaf water potentials. The concentration of CO2 inside leaves (ci) increased as stomatal conductance declined indicating increasing non‐stomatal limitations to photosynthesis. The stomatal response to δel of E. globulus in the field is best described as an ‘apparent feed‐forward response’ that probably results from both slowly reversible depression of net photosynthesis and abscisic acid accumulation in guard cells. We suggest that the stomatal response to ci may strengthen the link between photosynthetic capacity and stomatal conductance during leaf drying as a result of either drought or large δ el.  相似文献   

6.
Claisen-Schmidt condensation of 3-(1,2,3,6-tetrahydro-1-methylpyridin-4-yl)-2,4,5- trimethoxybenzaldehyde 3 and various aromatic, heterocyclic and alicyclic amides of 3- aminoacetophenone 6(a–s) afforded novel curcumin mimics. All the synthesized compounds were characterized by IR, 1H NMR, Mass spectroscopy and evaluated for antioxidant, cytotoxicity and antimicrobial activity. Out of the 20 compounds screened, compounds 7i, 7l, 7q, and 7n have shown excellent radical scavenging activity, compounds 7o, 7t, 7f, and 7r have shown significant xanthine oxidase inhibition, and compounds 7a, 7k and 7l were found to be potent inhibitors of selected cancer cell lines. Compounds 7h, 7t, 7l, 7i, and 7e have shown good antibacterial activity, whereas compounds 7j, 7f, 7o, 7h, and 7t exhibited significant antifungal activity.  相似文献   

7.
Proceeding from three previously derived expressions for the intensity of nitrification in soil as a function of time (logΣN=K.logt+q), as a function of incubation moisture (logΣN=A.pF i+B), as a function of initial moisture (logΣN=C.pF v+D), it was shown that the nitrification intensity as a function of time and of moisture can be expressed by the bilinear function log ΣN=a.pF i.logT+b.pF i+c.logt+d; as a function of time and of initial moisture by the bilinear function logΣ=N=a.pF v.logt+b.pF v+c.logt+d; as a function of initial and incubation moisture by the bilinear function log ΣN=a.pF ipF v+b.pF i+c.pF v+d. The intensity of nitrification as a function of time, incubation moisture and initial moisture may be expressed by the multilinear function log ΣN=a.pF i.pF v.logt+b.pF i.pF v+c.pF i.logt+d.pF v.logt+e .pF i+f.pF v=g.logt+h. This function is valid for all the incubation moistures lying between pF i 3.0 and 4.0 and for all initial moistures between 3.5 and 5.9 provided that the incubation temperature remains constant.  相似文献   

8.
Summary Well watered plants of Vigna unguiculata (L.) Walp cv. California Blackeye No. 5 had maximum photosynthetic rates of 16 mol m-2 s-1 (at ambient CO2 concentration and environmental parameters optimal for high CO2 uptake). Leaf conductance declined with increasing water vapour concentration difference between leaf and air (w), but it increased with increasing leaf temperature at a constant small w. When light was varied, CO2 assimilation and leaf conductance were correlated linearly. We tested the hypothesis that g was controlled by photosynthesis via intercellular CO2 concentration (c i). No unique relationship between (1) c i, (2) the difference between ambient CO2 concentration (c a) and c i, namely c a-c i, or (3) the c i/c a ratio and g was found. g and A appeared to respond to environmental factors fairly independently of each other. The effects of different rates of soil drying on leaf gas exchange were studied. At unchanged air humidity, different rates of soil drying were produced by using (a) different soils, (b) different irrigation schemes and (c) different soil volumes per plant. Although the soil dried to wilting point the relative leaf water content was little affected. Different soil drying rates always resulted in the same response of photosynthetic capacity (A max) and corresponding leaf conductance (g(Amax)) when plotted against percent relative plant-extractable soil water content (W e %) but the relationship with relative soil water content (W e ) was less clear. Above a range of W e of 15%–25%, A max and g(Amax) were both high and responded little to decreasing W e . As soon as W e fell below this range, A max and g(Amax) declined. The data suggest root-to-leaf communication not mediated via relative leaf water content. However, g(Amax) was initially more affected than A max.List of abbreviations A CO2 assimilation - A max photosynthetic capacity at favourable ambient conditions - c a CO2 concentration of the air in the leaf chamber - c i intercellular - CO2 concentration - E transpiration - g leaf conductance - g(Amax) leaf conductance corresponding to photosynthetic capacity - I photon flux rate - T l leaf temperature - W e relative plant-extractable soil water content - W e absolute plant-extractable soil water content - W l relative leaf water content - W s relative soil water content - w difference in water vapour mole fraction between leaf and air - leaf water potential  相似文献   

9.
The effects of the substrate conditions on the volumetric productivity of Lactobacillus helveticus at different cell densities up to 60 g l−1 in a continuous stirred-tank reactor with microfiltration to retain the biomass were investigated. At low dilution rates, D, the steady-state volumetric productivity, r p, gradually increased to a maximum at D = 1.2–1.5 h−1, because of reduced product inhibition. At higher D values, r p unexpectedly decreased, although the substrate conditions further improved. The maxima of r p at different cell densities coincided with a critical specific substrate utilization rate beyond which the cell metabolism seems to be controlled through a catabolic modulator factor, and r p decreases. Received: 8 September 1997 / Received last revision: 31 December 1997 / Accepted: 2 January 1998  相似文献   

10.
I. Nijs  I. Impens  T. Behaeghe 《Planta》1989,177(3):312-320
The relationship between leaf photosynthetic capacity (p n, max), net canopy CO2- and H2O-exchange rate (NCER and E t, respectively) and canopy dry-matter production was examined in Lollium perenne L. cv. Vigor in ambient (363±30 l· l-1) and elevated (631±43 l·l-1) CO2 concentrations. An open system for continuous and simultaneous regulation of atmospheric CO2 concentration and NCER and E t measurement was designed and used over an entire growth cycle to calculate a carbon and a water balance. While NCERmax of full-grown canopies was 49% higher at elevated CO2 level, stimulation of p n, max was only 46% (in spite of a 50% rise in one-sided stomatal resistance for water-vapour diffusion), clearly indicating the effect of a higher leaf-area index under high CO2 (approx. 10% in one growing period examined). A larger amount of CO2-deficient leaves resulted in higher canopy dark-respiration rates and higher canopy light compensation points. The structural component of the high-CO2 effect was therefore a disadvantage at low irradiance, but a far greater benefit at high irradiance. Higher canopy darkrespiration rates under elevated CO2 level and low irradiance during the growing period are the primary causes for the increase in dry-matter production (19%) being much lower than expected merely based on the NCERmax difference. While total water use was the same under high and low CO2 levels, water-use efficiency increased 25% on the canopy level and 87% on a leaf basis. In the course of canopy development, allocation towards the root system became greater, while stimulation of shoot dry-matter accumulation was inversely affected. Over an entire growing season the root/shoot production ratio was 22% higher under high CO2 concentration.Abbreviations and symbols C350 ambient CO2, 363±30 l·l-1 - C600 high CO2, 631±43 l·l-1 - c a atmospheric CO2 level - c i CO2 concentration in the intracellular spaces of the leaf - Et canopy evapotranspiration - I o canopy light compensation point - NCER canopy CO2-exchange rate - p n leaf photosynthetic rate - PPFD photosynthetic photon flux density - r a leaf boundary-layer resistance - RD canopy dark-respiration rate - r s stomatal resistance - WUE water use efficiency  相似文献   

11.
 The rates of convection and evaporation at the interface between the human body and the surrounding air are expressed by the parameters convective heat transfer coefficient h c, in W m–2°C–1 and evaporative heat transfer coefficient h e, W m–2 hPa–1. These parameters are determined by heat transfer equations, which also depend on the velocity of the airstream around the body, that is still air (free convection) and moving air (forced convection). The altitude dependence of the parameters is represented as an exponential function of the atmospheric pressure p: h cp n and h ep 1–n, where n is the exponent in the heat transfer equation. The numerical values of n are related to airspeed: n=0.5 for free convection, n=0.618 when airspeed is below 2.0 ms–1 and n=0.805 when airspeed is above 2.0 ms–1. This study considers the coefficients h c and h e with respect to the similarity of the two processes, convection and evaporation. A framework to explain the basis of established relationships is proposed. It is shown that the thickness of the boundary layer over the body surface increases with altitude. As a medium of the transfer processes, the boundary layer is assumed to be a layer of still air with fixed insulation which causes a reduction in the intensity of heat and mass flux propagating from the human body surface to its surroundings. The degree of reduction is more significant at a higher altitude because of the greater thickness of the boundary layer there. The rate of convective and evaporative heat losses from the human body surface at various altitudes in otherwise identical conditions depends on the following factors: (1) during convection – the thickness of the boundary layer, plus the decrease in air density, (2) during evaporation (mass transfer) – the thickness of the boundary layer, plus the increase with altitude in the diffusion coefficient of water vapour in the air. The warming rate of the air volume due to convection and evaporation is also considered. Expressions for the calculation of altitude dependences h c (p) and h e (p) are suggested. Received: 23 June 1998 / Accepted: 10 February 1999  相似文献   

12.
The lugworm Arenicola marina is a typical inhabitant of intertidal flats. In its L-shaped burrow the animal is exposed to varying concentrations of O2 and toxic sulfide depending on the tides. The lugworm is able to detoxify sulfide through its oxidation to thiosulfate. When exposed to declining O2 tensions Arenicola marina reacted as an oxyconformer. In the presence of 25 μmol · l−1 sulfide the respiration was not affected. In contrast, the lugworm consumed significantly less O2 at any Po2 in the presence of 200 μmol · l−1 sulfide. Without sulfide anaerobic metabolism started at a Po2 of approximatedly 10 kPa. Even at high O2 tensions animals exposed to sulfide produced significantly more anaerobic metabolites compared with the controls. Accordingly the critical value PcM, the ambient Po2 below which anaerobic metabolism starts, was shifted towards normoxia. Since O2 supply was sufficient for aerobic metabolism, anaerobiosis was induced by sulfide. An influx of sulfide was observed at 25 as well as at 200 μmol · l−1 sulfide. The main product of sulfide detoxification in the lugworm was thiosulfate. Its synthesis increased with ambient Po2 and depended on the sulfide concentration. Sulfide and thiosulfate were detected in the coelomic fluid, the blood, and the body wall of Arenicola marina. Only about 2% of the ambient O2 was used for sulfide detoxification at 25 μmol · l−1 sulfide and about 50% at 200 μmol · l−1 sulfide, respectively. Even at the low sulfide concentration Arenicola marina's capacity to detoxify sulfide was too low to maintain a complete aerobic metabolism. Accepted: 19 February 1997  相似文献   

13.
To assess muscle metabolism and inorganic phosphate (Pi) peak splitting during exercise, 31-phosphorus nuclear magnetic resonance spectroscopy was performed during ramp incremental and submaximal step exercise with and without circulatory occlusion. Seven healthy men performed calf flexion in a superconducting magnet. There was no Pi splitting during ramp incremental exercise with the circulation present and phosphocreatine (PCr) decreased linearly by 0.07 (SEM 0.01) mmol · l−1 · s−1, while exercise with the circulation occluded caused the Pi peak to split into a high and a low pH peak. The rate of PCr decrease during exercise with the circulation occluded was 0.15 (SEM 0.03) mmol · l−1 · s−1 which with the efficiency of the adenosine 5′-triphosphate (ATP) hydrolysis reaction corresponded well to the mechanical energy. Both with and without occlusion of the circulation PCr decreased with some time lag which may reflect the consumption of residual oxygen. In submaximal step exercise PCr decreased exponentially at the onset of exercise with the circulation open whereas it decreased linearly by 0.15␣mmol · l−1 · s−1 when the circulation was occluded. After exercise, occlusion of the circulation was maintained for 1 min more and there was no PCr resynthesis. It is suggested that ATP synthesis was limited by the availability of oxygen. Accepted: 14 August 1996  相似文献   

14.
A mixed microbial culture was immobilized by entrapment into silica gel (SG) and entrapment/ adsorption on polyurethane foam (PU) and ceramic foam. The phenol degradation performance of the SG biocatalyst was studied in a packed-bed reactor (PBR), packed-bed reactor with ceramic foam (PBRC) and fluidized-bed reactor (FBR). In continuous experiments the maximum degradation rate of phenol (q s max) decreased in the order: PBRC (598 mg l−1 h−1) > PBR (PU, 471 mg l−1 h−1) > PBR (SG, 394 mg l−1 h−1) > FBR (PU, 161 mg l−1 h−1) > FBR (SG, 91 mg l−1 h−1). The long-term use of the SG biocatalyst in continuous phenol degradation resulted in the formation of a 100–200 μm thick layer with a high cell density on the surface of the gel particles. The abrasion of the surface layer in the FBR contributed to the poor degradation performance of this reactor configuration. Coating the ceramic foam with a layer of cells immobilized in colloidal SiO2 enhanced the phenol degradation efficiency during the first 3 days of the PBRC operation, in comparison with untreated ceramic packing. Received: 2 December 1999 / Revision received: 2 February 2000 / Accepted: 4 February 2000  相似文献   

15.
Rising atmospheric [CO2], ca, is expected to affect stomatal regulation of leaf gas‐exchange of woody plants, thus influencing energy fluxes as well as carbon (C), water, and nutrient cycling of forests. Researchers have proposed various strategies for stomatal regulation of leaf gas‐exchange that include maintaining a constant leaf internal [CO2], ci, a constant drawdown in CO2 (ca ? ci), and a constant ci/ca. These strategies can result in drastically different consequences for leaf gas‐exchange. The accuracy of Earth systems models depends in part on assumptions about generalizable patterns in leaf gas‐exchange responses to varying ca. The concept of optimal stomatal behavior, exemplified by woody plants shifting along a continuum of these strategies, provides a unifying framework for understanding leaf gas‐exchange responses to ca. To assess leaf gas‐exchange regulation strategies, we analyzed patterns in ci inferred from studies reporting C stable isotope ratios (δ13C) or photosynthetic discrimination (?) in woody angiosperms and gymnosperms that grew across a range of ca spanning at least 100 ppm. Our results suggest that much of the ca‐induced changes in ci/ca occurred across ca spanning 200 to 400 ppm. These patterns imply that ca ? ci will eventually approach a constant level at high ca because assimilation rates will reach a maximum and stomatal conductance of each species should be constrained to some minimum level. These analyses are not consistent with canalization toward any single strategy, particularly maintaining a constant ci. Rather, the results are consistent with the existence of a broadly conserved pattern of stomatal optimization in woody angiosperms and gymnosperms. This results in trees being profligate water users at low ca, when additional water loss is small for each unit of C gain, and increasingly water‐conservative at high ca, when photosystems are saturated and water loss is large for each unit C gain.  相似文献   

16.
Consider the model yijk=u ± ai ± bi ± cij ± eijk i=1, 2,…, t; j=1, 2,…b; k=1, 2,…,nij where μ is a constant and ai, bi, cij are distributed independently and normally with zero means and variances Δ2 Δ2/bdij and δ2 respectively. It is assumed that di's, and dij's are known (positive) constants (for all i and j). In this paper procedures for estimating the variance components (Δ2, Δ2b and Δ2a) and for testing the hypothesis Hoc2c2 = y3 and Hoa2b2 = y4 (where y2, y3, and y4, are specified constants) are presented. A generalization for the mixed model case is discussed in the last section.  相似文献   

17.
 The effects of propagation microclimate and foliar area on the rooting of Cordia alliodora (Ruiz & Pavon) Oken cuttings were investigated using non-mist propagators with and without shade. Photosynthetic rates (P n ), stomatal conductance (g s ) and chlorophyll fluorescence ratio (Fv/Fm) of the cuttings were assessed during propagation. Pronounced differences in microclimate were recorded between treatments, with lower temperatures and vapour pressure deficit (VPD) under shade. During the first 8 days after insertion, P n varied between 2.21 and 4.96 and 0.47 – 2.54 μmol CO2 m –  2s –  1 in the shaded and unshaded propagators, respectively. In the unshaded propagator, Fv/Fm decreased to a minimum of 0.72 2 days after insertion, recovering thereafter. In two separate rooting experiments, rooting percentage was reduced by high irradiance in the 20 and 30 cm2 leaf area treatments, but not in the 10 cm2 treatment. P n decreased with an increase in leaf area in both shaded and unshaded propagators. Fv/Fm also declined with increasing leaf area in the high irradiance treatment. PAR and P n were positively correlated under shade (r 2 = 0.51) but negatively correlated in the unshaded treatment (r 2 = 0.49); maximum P n values were recorded at a PAR of 400 μmol m –  2 s –  1. No significant differences in g s were found between treatments, values ranging between 130 and 194 mmol H2O m –  2 s –  1. Positive correlations were found between rooting percentage and mean Fv/Fm. These results indicate that rooting of C. alliodora cuttings is related to photosynthetic activity during propagation, which is itself influenced both by propagator microclimate and cutting leaf area. Received: 7 May 1996 / Accepted: 17 December 1996  相似文献   

18.
The experiments and simulations reported in this paper show that, for stomata sensitive to both CO2 and water vapour concentrations, responses of stomatal conductance (gws) to boundary layer thickness have two components, one resulting from changes in intercellular CO2 concentration (χci) and another from changes in leaf surface water vapour saturation deficit (Dws). The experiments and simulations also show that the boundary layer conductance (gwb) can significantly alter the apparent response of gws to ambient air CO2 mole fraction (χca) and water vapour mole fraction (χwa). Because of the feedback loop involved the responses of gws for χca and χwa each include responses to both χci and Dws. The boundary layer alters the state of the variables sensed by the guard cells—i.e. χci and Dws—and so it is a source of feedback. Thus, when scaling up from responses of stomata to the response of gws for a whole leaf, the effect of the boundary layer must be considered. The results indicate that, for given responses of gws to χci and Dws, the apparent responses of gws to Dwa and χca depend on the size of the leaf and wind speed, showing that this effect of the boundary layer should be considered when comparing data measured under different conditions, or with different methods.  相似文献   

19.
The objective of this study was to compare the effects of different farming methods on bird communities. In particular, the study surveyed the frequency and number of bird species, the number of individuals and bird diversity (Shannon index) found in organic (o, non use of synthetic pesticides), integrated (i, reduced use of pesticides on the basis of the economic threshold) and conventional (c, conventional use of pesticides) orchards. A simplified version of the mapping method was applied, from May to June 1998, to 60 fruit farms (15o, 19 i, 26c), over a total of 483 ha. Twenty-six bird species were observed (19o, 17i, 18c). Granivorous species were the most abundant and unaffected by pesticide management. Insectivorous species were less abundant in general, but more frequent on o and i farms (chi-square test, p<0.05). Bird diversity was greater in o and i farm management than in c (Tukey test, respectively, p = 0.001 and 0.004). There was a non-significant trend for bird density to be higher in o/i than c orchards (ANOVA, F 2.57 = 2.76, p = 0.07). The results confirm the positive effects of (o) and (i) agriculture on bird communities, including some species of special interest for nature conservation. These effects could be attributed mostly to the different pest management and secondarily to some aspects of environmental differentiation (type of orchard, type of farm, age and height of trees, and increased presence of hedgerows and woodlots). Surprisingly no correlation was observed between hedgerows and woodlots and bird diversity (Pearson test, R = 0.0019).  相似文献   

20.
 Stand structure and leaf area distribution of a laurel forest in the Agua García mountains of Tenerife are described. The site is situated at 820 m a.s.l., faces NNE, and has a humid mediterranean climate. Summer droughts are mitigated by relatively high air humidity and clouds. The natural mixed hardwood forest is composed of six major tree species: Laurus azorica (Seub.) Franco, Persea indica (L.) Spreng, Myrica faya Ait., Erica arborea L. and two species of Ilex (I. platyphylla Webb & Berth. and I. canariensis Poivet.). The experimental stand had a density of 1693 trees ha – 1, a basal area of 33.7 m2ha – 1, and a cumulated volume of above-ground parts of trees of 231 m3 ha – 1 with a corresponding dry mass of 204 ton ha – 1. Diameters at breast height ranged from 6 to 46 cm. Mean concentration of plant dry mass per volume was 1.17 kg m – 3. The vertical pattern of leaf area distribution in individual trees for all tree species was characterized by a Gaussian-like curve. Stand leaf area index was 7.8. These evergreen, broad-leaved (laurisilva or lucidophyllous) forests represent a relic forest that was widespread in the Mediterranean region some 20 million years ago. Our data illustrate some of the structural characteristics of this historically widespread forest type. Received: 2 December 1994 / Accepted: 6 November 1995  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