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.
While evapotranspiration (ET) is normally measured as one hydrologic component, evaporation (E), and transpiration (T) result from different physical-biological processes. Using a two-source model, a trapezoid framework has been widely applied in recent years. The key to applying the trapezoid framework model is the determination of the dry/wet boundaries of the land surface temperature-fractional vegetation coverage trapezoid (LST-fc). Although algorithms have been developed to characterize the two boundaries, there remains a significant uncertainty near the wet boundary which scatters in a discrete and uneven manner. It is therefore difficult to precisely locate the wet boundary. To address this problem, a Wet Boundary Algorithm (WBA) was developed in this study with the algorithm applied in the region of Huang-Huai-Hai plain of China, using the Pixel Component Arranging and Comparing Algorithm (PCACA) to retrieve ET from MODerate-resolution Imaging Spectroradiometer (MODIS) Data. The eddy covariance (EC) measurements from Yucheng station was used to verify the modified model where the root mean square error (RMSE) of 17.8 W/m2, Bias of −7.2 W/m2 for latent heat flux (LE) simulation in 28 cloudless test days. The ratio of transpiration to evapotranspiration (T/ET) varied between 0.48 and 0.81 over the Huang-Huai-Hai plain. The spatial and temporal distribution of ET revealed that agriculture practices have a significant influence on the hydrological cycle, where crop growth promotes the magnitude of ET. Likewise, harvesting activities significantly reduce ET. The proposed WBA algorithm significantly reduces the uncertainty of the trapezoid ET model caused by wet edge positioning. The analysis of the impact of agricultural activities on ET provide a better understanding how human activities change the hydrological cycle at regional scales.  相似文献   

2.
An experiment on evapotranspiration from citrus trees under irrigation with saline waterwas carried out for 4 months. Two lysimeters planted with a citrus tree in the green house wereused. One lysimeter was irrigated with saline water (NaCl and CaCl2 of 2000 mg/L equivalence,EC = 3.8 dS/m, SAR = 5.9) and the other was irrigated with freshwater using drip irrigation. Theapplied irrigation water was 1.2 times that of the evapotranspiration on the previous day.Evapotranspiration was calculated as the change in lysimeter weight recorded every 30 minutes.The lysimeters were filled with soil with 95.8% sand. The results of the experiment were as follows.(i) The evapotranspiration from citrus tree was reduced after irrigation with saline water. Theevapotranspiration returns to normal after leaching. However it takes months to exhaust the saltfrom the tree. ( ii ) To estimate the impact of irrigation with saline water on the evapotranspirationfrom citrus trees, the reduction coefficient due to salt stress (Ks) was used in this experiment.Evapotranspiration under irrigation with saline water (ETs) can be calculated from evapotranspira-tion under irrigation with freshwater (ET) by the equation ETs = Ks× ET. Ks can be expressed as afunction of ECsw. (iii) The critical soil-water electrical conductivity (ECsw) is 9.5 dS/m, beyondwhich adverse effects on evapotranspiration begin to appear. If ECsw can be controlled at below9.5 dS/m, saline water can be safely used for irrigation.  相似文献   

3.
Estimation of evapotranspiration (ET) is of great significance in modeling the water and energy interactions between land and atmosphere. Negative correlation of surface temperature (Ts) versus vegetation index (VI) from remote sensing data provides diagnosis on the spatial pattern of surface soil moisture and ET. This study further examined the applicability of Ts–VI triangle method with a newly developed edges determination technique in estimating regional evaporative fraction (EF) and ET at MODIS pixel scale through comparison with large aperture scintillometer (LAS) and high‐level eddy covariance measurements collected at Changwu agro‐ecological experiment station from late June to late October, 2009. An algorithm with merely land and atmosphere products from MODIS onboard Terra satellite was used to estimate the surface net radiation (Rn) and soil heat flux. In most cases, the estimated instantaneous Rn was in good agreement with surface measurement with slight overestimation by 12 W/m2. Validation results from LAS measurement showed that the root mean square error is 0.097 for instantaneous EF, 48 W/m2 for instantaneous sensible heat flux, and 30 W/m2 for daily latent heat flux. This paper successfully presents a miniature of the overall capability of Ts–VI triangle in estimating regional EF and ET from limited number of data. For a thorough interpretation, further comprehensive investigation needs to be done with more integration of remote sensing data and in‐situ surface measurements. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

4.
靳铮  张雪芹 《湖泊科学》2020,32(3):877-886
源区划分和质量过滤提高湖面涡动相关通量数据可靠性的同时,却降低了通量时间序列的连续性.为此,本文基于TensorFlow机器学习框架构建了一种超宽人工神经网络(ANN)模型.在选择输入ANN模型的特征变量信息时,我们采取了尽可能获取湍流输送过程中热力、动力学同步观测背景强迫信息的原则.通过ANN模型模拟通量的插补,本文实现了通量时间序列连续性的优化,插补后的羊卓雍错湖面通量数据的时间覆盖率从不足0.40提升至超过0.98.基于10次折叠交叉验证的ANN模型通量模拟性能检验则表明,各个检验组之间ANN模型的模拟误差波动较小,这显示出了较好的稳健性.具体地讲,感热通量、潜热通量和水汽通量原始观测平均值分别约为18.8 W/m~2、81.5 W/m~2和1.84 mmol/(s·m~2),10组交叉验证的插补感热通量、潜热通量和水汽通量平均绝对误差分别为5.4 W/m~2、15.7 W/m~2和0.35 mmol/(s·m~2).这表明本文所探索的ANN建模结构和同步观测变量筛选原则可更充分地利用观测点局地同步观测信息估算通量强度,有效地优化湍流通量数据的时间连续性,从而提升通量数据的可分析性.  相似文献   

5.
An experiment on evapotranspiration from citrus trees under irrigation with saline water was carried out for 4 months. Two lysimeters planted with a citrus tree in the green house were used. One lysimeter was irrigated with saline water (NaCl and CaCl2 of 2000 mg/L equivalence,EC = 3.8 dS/m, SAR = 5.9) and the other was irrigated with freshwater using drip irrigation. The applied irrigation water was 1.2 times that of the evapotranspiration on the previous day. Evapotranspiration was calculated as the change in lysimeter weight recorded every 30 minutes. The lysimeters were filled with soil with 95.8% sand. The results of the experiment were as follows. (i) The evapotranspiration from citrus tree was reduced after irrigation with saline water. The evapotranspiration returns to normal after leaching. However it takes months to exhaust the salt from the tree. (ii) To estimate the impact of irrigation with saline water on the evapotranspiration from citrus trees, the reduction coefficient due to salt stress (Ks) was used in this experiment. Evapotranspiration under irrigation with saline water (ET s ) can be calculated from evapotranspiration under irrigation with freshwater (ET) by the equationET s =K s × ET. Ks can be expressed as a function ofEC sw . (iii) The critical soil-water electrical conductivity (EC sw ) is 9.5 dS/m, beyond which adverse effects on evapotranspiration begin to appear. IfEC sw can be controlled at below 9.5 dS/m, saline water can be safely used for irrigation.  相似文献   

6.
A large weighing lysimeter was installed at Yucheng Comprehensive Experimental Station, north China, for evapotranspiration and soil‐water–groundwater exchange studies. Features of the lysimeter include the following: (i) mass resolution equivalent to 0·016 mm of water to accurately and simultaneously determine hourly evapotranspiration, surface evaporation and groundwater recharge; (ii) a surface area of 3·14 m2 and a soil profile depth of 5·0 m to permit normal plant development, soil‐water extraction, soil‐water–groundwater exchanges, and fluctuations of groundwater level; (iii) a special supply–drainage system to simulate field conditions of groundwater within the lysimeter; (iv) a soil mass of about 30 Mg, including both unsaturated and saturated loam. The soil consists mainly of mealy sand and light loam. Monitoring the vegetated lysimeter during the growing period of winter wheat, from October 1998 through to June 1999, indicated that during the period groundwater evaporation contributed 16·6% of total evapotranspiration for a water‐table depth from 1·6 m to 2·4 m below ground surface. Too much irrigation reduced the amount of upward water flow from the groundwater table, and caused deep percolation to the groundwater. Data from neutron probe and tensiometers suggest that soil‐water‐content profiles and soil‐water‐potential profiles were strongly affected by shallow groundwater. Copyright © 2000 John Wiley & Sons, Ltd.  相似文献   

7.
A typical agricultural water reservoir (AWR) of 2400 m2 area and 5 m depth, located in a semi‐arid area (southern Spain), was surveyed on a daily basis for 1 year. The annual evaporation flux was 102·7 W m?2, equivalent to an evaporated water depth of 1310 mm year?1. The heat storage rate G exhibited a clear annual cycle with a peak gain in April (G ~ 45 W m?2) and a peak loss in November (G ~ 40 W m?2), leading to a marked annual hysteretic trend when evaporation (λE) was related to net radiation (Rn). λE was strongly correlated with the available energy A, representing 91% of the annual AWR energy loss. The sensible heat flux H accounted for the remaining 9%, leading to an annual Bowen ratio in the order of 0·10. The equilibrium and advective evaporation terms of the Penman formula represented 76 and 24%, respectively, of the total evaporation, corresponding to a annual value of the Priestley–Taylor (P–T) coefficient (α) of 1·32. The P–T coefficient presented a clear seasonal pattern, with a minimum of 1·23 (July) and a maximum of 1·65 (December), indicating that, during periods of limited available energy, AWR evaporation increased above the potential evaporation as a result of the advection process. Overall, the results stressed that accurate prediction of monthly evaporation by means of the P–T formula requires accounting for both the annual cycle of storage and the advective component. Some alternative approaches to estimating Rn, G and α are proposed and discussed. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

8.
Abstract

Reliable estimation of sensible heat flux (H) is important in energy balance models for quantifying evapotranspiration (ET). This study was conducted to evaluate the value of adding the Priestley-Taylor (PT) equation to the METRIC (Mapping Evapotranspiration at high Resolution with Internalized Calibration) model. METRIC was used to estimate energy fluxes for 10 Landsat images from the 2005, 2006 and 2007 crop growing seasons in south-central Nebraska, USA, where each image owing to recent rainfall exhibited high residual moisture content even at the hot pixel. The METRIC model performed satisfactorily for net radiation (Rn ) and soil heat flux (G) estimation with a root mean square error (RMSE) of 52 and 24 W m-2, respectively. A RMSE of 122 W m-2 for H indicated the limitation of the METRIC model in estimating H for high residual moisture content of the hot pixel (Alfalfa reference ET fraction, ET r F > 0.15). The modified METRIC model (wet METRIC or wMETRIC) incorporating the PT equation was applied to calculate H at the anchor pixels (hot and cold) for high residual moisture content of the hot pixel. The α coefficient of the PT equation was locally calibrated using hourly meteorological data from an automatic weather station and Rn and G data from a Bowen ratio flux tower. The mean α coefficient value was 1.14. The wMETRIC model reduced the RMSE of H from 122 to 64 W m-2 and that of latent heat flux, LE, from 163 to 106 W m-2. The RMSE of daily ET decreased from 1.7 to 1.1 mm d-1 with wMETRIC. The results indicate that treatment of anchor pixels for high residual moisture content with the PT approach gives improved estimation of H, LE and daily ET. It is recommended that the wMETRIC model be used for estimating ET if the hot pixel has high residual moisture (i.e. reference ET fraction > 0.15).

Citation Singh, R. K. & Irmak, A. (2011) Treatment of anchor pixels in the METRIC model for improved estimation of sensible and latent heat fluxes. Hydrol. Sci. J. 56(5), 895–906.  相似文献   

9.
Summary The mean annual cycle of the net energy flux through the sea surface and of the heat storage are investigated in detail using observations of the Light Vessel LV Elbe 1 for the period 1962-1986 in the German Bight and at Ocean Weather Ship OWS Famita for the period 1965-1978 in the central North Sea. The investigation confirms the general geographical picture of the heat budget of the German Bight that shows a net loss to the atmosphere by a long-term mean of -15 W m-2. The radiative surface input of 113 W m-2 is balanced by -62 W m-2 net back radiation, -51 W m-2 of latent heat flux and -15 W m-2 of sensible heat flux. The heat advection calculated as the residual of the heat storage rate and surface energy balance is 16 W m-2. The mean annual cycles of heat storage and surface energy balance are nearly equal, and the temperature variations are mainly driven by local heat input. The small differences build up the annual advection cycle. Warm water advection occurs from October to April and cold water advection in summer from May to September. The seasonal advection variability is extreme in winter and summer, and the ranges slow down in spring and autumn, when the sign of the heat balance changes. The OWS Famita is situated also in an area of net energy loss to the atmosphere, showing a long-term annual mean loss of -16 W m-2. The surface radiation input of 105 W m-2 is mainly balanced by outgoing long wave back radiation of -60 W m-2 and a latent heat flux of -49 W m-2. A minor contribution to the heat balance is the sensible heat flux of -12 W m-2. Warm water advection occurs in winter and spring. Variability is greater than at LV Elbe 1. Calculated monthly fluxes show the dominance of the energy gain of incoming solar radiation. Net long-wave radiation is nearly constant with time. The sensible heat flux serves as a heat source only at LV Elbe 1 from May to June. The latent heat flux is a loss term all the year. The heat storage cycle is nearly equal to the surface energy balance at LV Elbe 1 ; the differences are more irregular at OWS Famita. The temperature variations are mainly driven by local heat input. The simplified one-dimensional balance holds generally for the heating period in both regions, although for some months the magnitude of the advection is up to a third of the net surface fluxes or the storage rate. At LV Elbe 1 from April to December, the heat budget is dominated by local dynamics. At OWS Famita the advective contribution is less than 30% of net surface heat input from May to October. The dominance of solar radiation in determining the surface heat fluxes, and the annual cycles of the storage rate in phase justify the use of one-dimensional models as a first approximation of the temperature dynamics in these regions. Comparisons of simulations of the temperature cycle at both sites with observations give sufficient precision during most parts of the seasonal cycle. Suitable data sets to drive and validate these models are now available and different models should be tested.
Advektive beitr?ge zur w?rmebilanz der deutschen bucht (feuerschiff elbe 1) und zur w?rmebilanz der zentralen nordsee (Wetterschiff Famita)
Zusammenfassung Untersucht wurde der mittlere Jahresgang vom W?rmeeintrag durch die Meeresoberfl?che und vom W?rmeinhalt der Wassers?ule. Dazu wurden Messungen aus der Deutschen Bucht vom Feuerschiff Elbe 1 für die Jahre 1962-1986 und Messungen in der zentralen Nordsee vom Wetterschiff Famita für die Jahre 1965-1978 verwendet. Die Untersuchung best?tigt das generelle Bild einer W?rmeabgabe an die Atmosph?re von -15 W m-2 im langj?hrigen Mittel für die Deutsche Bucht. Die kurzwellige Einstrahlung von 113 W-2 wird durch -62 W m-2 langwellige Ausstrahlung, -51 W m-2 latenten W?rmefluβ und -15 W m-2 sensiblen W?rmefluβ nahezu balanciert. Die berechnete W?rmeadvektion als Residuum aus W?rmeinhalt und Nettow?rmefluβ an der Meeresoberfl?che betr?gt 16 W m-2 Der Jahresgang des W?rmeinhaltes und der Jahresgang des Nettow?rmeflusses an der Oberfl?che sind fast gleich, so daβ der Temperaturjahresgang haupts?chlich durch den lokalen W?rmeeintrag gesteuert wird. Kleine Abweichungen hiervon bestimmen den Jahresgang der W?rmeadvektion. Warmwasseradvektion tritt von Oktober bis April auf. Kaltwasseradvektion liegt im Sommer von Mai bis September vor. Die Variabilit?t der W?rmeadvektion ist im Winter und Sommer am gr?βten, w?hrend die Variabilit?t im Frühjahr und Herbst geringer ist, wenn sich das Vorzeichen der W?rmebilanz ?ndert. Das Wetterschiff Famita befindet sich ebenfalls in einer Region, in der W?rme an die Atmosph?re abgegeben wird. Die W?rmeabgabe betr?gt -16 W m-2 im langzeitlichen Mittel. Die kurzwellige Einstrahlung von 105 W m-2 wird haupts?chlich durch -60 W m-2 langwellige Ausstrahlung, -49 W m-2 latenten W?rmefluβ und -12 W m-2 sensiblen W?rmefluβ balanciert. Warmwasseradvektion tritt im Winter und Frühjahr auf. Die Variabilit?t der W?rmeadvektion ist gr?βer als bei Feuerschiff Elbe 1. Die berechneten monatlichen Energieflüsse zeigen, daβ die solare Einstrahlung den Jahresgang der W?rmebilanz dominiert. Die effektive Ausstrahlung ist nahezu konstant. Die sensible W?rme wirkt nur bei Feuerschiff Elbe 1 von Mai bis Juni als W?rmequelle. Der latente W?rmefluβ ist w?hrend des gesamten Jahres negativ. Für Feuerschiff Elbe 1 ist der W?rmeinhalt der Wassers?ule mit dem Energieeintrag an der Oberfl?che in Phase, w?hrend bei Wetterschiff Famita Differenzen auftreten. Die Temperaturvariationen sind haupts?chlich durch den lokalen W?rmeeintrag bestimmt. Diese vereinfachten Verh?ltnisse gelten für beide Regionen, obwohl für einige Monate die W?rmeadvektion bis zu einem Drittel des Nettow?rmeflusses an der Oberfl?che betragen kann. Bei Feuerschiff Elbe 1 wird die W?rmebilanz von April bis Dezember durch die lokale Dynamik bestimmt. Bei Wetterschiff Famita ist die W?rmeadvektion von Mai bis Oktober kleiner als 30% vom Oberfl?cheneintrag. Die Dominanz der solaren Einstrahlung für die W?rmebilanz an der Oberfl?che und der phasengleiche Jahresgang des W?rmeinhaltes rechtfertigen es, eindimensionale Wassers?ulenmodelle für die Region zu verwenden, um die Dynamik der Temperatur zu berechnen. So zeigt der Vergleich von simulierten und gemessenen Temperaturjahresg?ngen an beiden Positionen eine ausreichende Genauigkeit über weite Teile des Jahres. Damit stehen neben der gezeigten W?rmebilanzabsch?tzung zwei Datens?tze zur Verfügung, um Modelle zu betreiben, zu validieren und verschiedenartige Modelle zu vergleichen.
  相似文献   

10.
A comparison between half‐hourly and daily measured and computed evapotranspiration (ET) using three models of different complexity, namely, the Priestley–Taylor (P‐T), the reference Penman–Monteith (P‐M) and the Common Land Model (CLM), was conducted using three AmeriFlux sites under different land cover and climate conditions (i.e. arid grassland, temperate forest and subhumid cropland). Using the reference P‐M model with a semiempirical soil moisture function to adjust for water‐limiting conditions yielded ET estimates in reasonable agreement with the observations [root mean square error (RMSE) of 64–87 W m?2 for half‐hourly and RMSE of 0.5–1.9 mm day?1 for daily] and similar to the complex Common Land Model (RMSE of 60–94 W m?2 for half‐hourly and RMSE of 0.4–2.1 mm day?1 for daily) at the grassland and cropland sites. However, the semiempirical soil moisture function was not applicable particularly for the P‐T model at the forest site, suggesting that adjustments to key model variables may be required when applied to diverse land covers. On the other hand, under certain land cover/environmental conditions, the use of microwave‐derived soil moisture information was found to be a reliable metric of regional moisture conditions to adjust simple ET models for water‐limited cases. Further studies are needed to evaluate the utility of the simplified methods for different landscapes. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

11.
With the complex nature of land surfaces, more attention should be paid to the performance of remotely sensed models to estimate evapotranspiration from moderate and low spatial resolution data. Taking into account the characteristic of a stable evaporative fraction (EF) in the daytime, this paper uses the surface energy balance system (SEBS) to estimate the EF from MODIS data for a subtropical evergreen coniferous plantation in southern China and evaluates the stability of the SEBS model in estimating the EF under complex surface conditions. The results show that the SEBS‐estimated EF is larger than the measured EF partly because of the serious lack of energy‐balance closure. This difference can be largely reduced by the residual energy correction method. More evaporative land cover within the MODIS pixel is a main reason for the overestimated EF. SEBS underestimates sensible heat flux, and the underestimation of surface available energy also contributes to the overestimation of the EF. The EF estimated from MODIS/Terra data is in agreement with that from MODIS/Aqua data with a coefficient of determination (R2) of 0.552, a mean bias error (BIAS) of 0.028, and a root mean square error (RMSE) of 0.079, which is consistent with the result from in situ measurements. In addition, the estimation of surface available energy from remotely sensed data is evaluated on this complex underlying surface. Compared with in situ measurements, the available energy is underestimated by 28 W m?2 with an RMSE = 50 W m?2 and an R2 = 0.87. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

12.
Evapotranspiration (ET) from riparian vegetation can be difficult to estimate due to relatively abundant water supply, spatial vegetation heterogeneity, and interactions with anthropogenic influences such as shallower groundwater tables, increased salinity, and nonpoint source pollution induced by irrigation. In semiarid south-eastern Colorado, reliable ET estimates are scarce for the riparian corridor that borders the Arkansas River. This work investigates relationships between the riparian ecosystem along the Arkansas River and an underlying alluvial aquifer using ET estimates from remotely sensed data and modelled water table depths. Results from a calibrated, finite-difference groundwater model are used to estimate weekly water table fluctuations in the riparian ecosystem from 1999 to 2009, and estimates of ET are calculated using the Operational Simplified Surface Energy Balance (SSEBop) model with over 200 Landsat scenes covering over 30 km2 of riparian ecosystem along a 70-km stretch of the river. Comparison of calculated monthly SSEBop ET to estimated alfalfa reference ET from local micrometeorological station data indicated statistically significant high linear correspondence (R2 = .87). Daily calculated SSEBop ET showed statistically significant moderate linear correspondence with data from a local weighing lysimeter (R2 = .59). Simulated monthly SSEBop ET values were larger in drier years compared with wetter years, and ET variability was also larger in drier years. Peak ET most commonly occurred during the month of June for all 11 years of analysis. Relationships between ET and water table depth showed that peak monthly ET was highest when groundwater depths were less than about 3 m, and ET values were significantly lower for groundwater depths greater than 3 m. Negative sample Spearman correlation highlighted riparian corridor locations where ET increased as a result of decreased groundwater depths across years with different hydroclimatic conditions. This study shows how a combination of remotely sensed riparian ET estimates and a regional groundwater model can improve our understanding of linkages between riparian consumptive use and near-river groundwater conditions influenced by irrigation return flow and different climatic drivers.  相似文献   

13.
The backward‐averaged iterative two‐source surface temperature and energy balance solution (BAITSSS) model was developed to calculate evapotranspiration (ET) at point to regional scales. The BAITSSS model is driven by micrometeorological data and vegetation indices and simulates the water and energy balance of the soil and canopy sources separately, using the Jarvis model to calculate canopy resistance. The BAITSSS model has undergone limited testing in Idaho, United States. We conducted a blind test of the BAITSSS model without prior calibration for ET against weighing lysimeter measurements, net radiation, and surface temperature of drought‐tolerant corn (Zea mays L. cv. PIO 1151) in a semiarid, advective climate (Bushland, Texas, United States) in 2016. Later in the season (20 days), BAITSSS consistently overestimated ET by up to 3 mm d?1. For the entire growing season (127 days), simulated versus measured ET resulted in a 7% error in cumulative ET, RMSE = 0.13 mm h?1, and 1.70 mm d?1; r2 = 0.66 (daily) and r2 = 0.84 (hourly); MAE = 0.08 mm h?1 and 1.24 mm d?1; and MBE = 0.02 mm h?1 and 0.58 mm d?1. The results were comparable with thermally driven instantaneous ET models that required some calibration. Next, the initial soil water boundary condition was reduced, and model revisions were made to resistance terms related to incomplete cover and assumption of canopy senescence. The revisions reduced discrepancies between measured and modelled ET resulting in <1% error in cumulative ET, RMSE = 0.1 mm h?1, and 1.09 mm d?1; r2 = 0.86 (daily) and r2 = 0.90 (hourly); MAE = 0.06 mm h?1 and 0.79 mm d?1; and MBE = 0.0 mm h?1 and 0.17 mm d?1 and generally mitigated the previous overestimation. The advancement in ET modelling with BAITSSS assists to minimize uncertainties in crop ET modelling in a time series.  相似文献   

14.
Eddy covariance (EC) and micro‐meteorological data were collected from May 2010 to January 2013 from urban, non‐irrigated bahiagrass (Paspalum notatum) in subtropical south Florida. The objectives were to determine monthly crop coefficients (Kc) for non‐irrigated bahiagrass by using EC evapotranspiration (ET) data and the Food and Agriculture Organization 56 Penman–Monteith reference evapotranspiration equation; compare crop ET (ETc) calculated with new Kc values to ETc obtained using Kc values available in the literature; and compare results and methodologies for statistical differences. New Kc values ranged from 0.62 to 0.92 and were different from Kc values found in the scientific literature for bahiagrass. Resulting ETc calculated using literature Kc values were significantly different from EC ET data, whereas ETc using the new Kc values was not. Specifically, literature Kc values were temporally biased to miscalculate the timing of convergence between potential and actual ET, assuming that our new Kc values calculated with EC methods were most accurate. As a consequence, ETc calculated using the literature Kc values was either too large or too small. However, one set of literature Kc values from a similar climate and water table depth were closer to our new Kc values, indicating that climate should be considered when selecting urban non‐irrigated Kc from the literature to estimate ET. Results also indicated that more than 1 year of EC ET data was needed when establishing monthly Kc values because of annual variability in factors controlling ET, such as water availability. The new Kc values reported herein could be used as an estimate for urban non‐irrigated bahiagrass within similar climates. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

15.
We have used stable water isotopes (δ18O, δ2H) in combination with lumped-parameter modeling for characterizing unsaturated flow in two lysimeters vegetated with maize. The lysimeters contained undisturbed soil cores dominated by sandy gravel (Ly1) and clayey sandy silt (Ly2). Stable water isotopes were analyzed in precipitation and lysimeter outflow water over about 3 years. The mean transit time of water T and dispersion parameter PD, obtained from modeling, were higher for the silt soil in Ly2 than for the gravel soil in Ly1 (T of 362 vs. 129 d, PD of 0.7 vs. 0.12). The consideration of preferential flow (PF) paths could substantially improve the model curve fits, with 13 and 11% contribution of PF for Ly1 and Ly2 as best estimates. Different assumptions were compared to estimate the input function, that is, stable water isotope content in the recharging water. Using the isotopic composition of precipitation as input (no modification) resulted in reasonable model estimations. Best model fits for the entire observation were obtained by weighting the recharging isotopes according to average precipitation within periods of 3 and 6 months, in correspondence to changing vegetation phases and seasonal influences. Input functions that consider actual evapotranspiration could significantly improve modeling at some periods, however, this led to deviations between modeled and observed δ18O at other periods. This may indicate the influence of variable flow, so that dividing the whole observation period into hydraulically characteristic sub-periods for lumped-parameter modeling (which implements steady-state flow) is recommended for possible further improvement.  相似文献   

16.
Moringa oleifera (MO) seed extract coupled with electrocoagulation (EC) was used to remove fluoride from water. Different MO extract volumes (5.0, 12.5, and 25.0 mL of MO extract per water liter) were coupled with EC, using aluminum electrodes at different current density values (J = 0.7, 2.0, and 3.3 mA/cm2) and different electrode separations (1.0, 2.0, and 4.0 cm), tested in batch and recirculation experiments. Control experiments using MO extract and EC alone achieved 5% and 54% water defluoridation, respectively. Best experimental batch conditions were achieved using 12.5 mL of MO extract followed by EC (3.3 mA/cm2) with a 1.0 cm electrode separation, producing >90% fluoride removal. Recirculation experiments with the EC reactor were performed with DI water and tap water using 1.0 cm electrode separation, 12.5 mL of MO extract and different current densities. More than 90% fluoride removal was achieved with the EC/MO process, using 3.3 mA/cm2, in both DI and tap water after 30 and 60 min, respectively. An energy consumption index (ECI) was developed, which showed that 1.51 and 0.67 W/h/mg were achieved for batch experiments of EC alone and EC/MO extract, respectively. For EC/MO extract, recirculation experiments with tap and DI water resulted in 0.35 and 0.22 W/h/mg, respectively. A cost analysis showed that $0.18 will be needed to treat one cubic meter of water.  相似文献   

17.
In this study, Surface Energy Balance Algorithm for Land (SEBAL) was evaluated for its ability to derive aerodynamic components and surface energy fluxes from very high resolution airborne remote sensing data acquired during the Bushland Evapotranspiration and Agricultural Remote Sensing Experiment 2008 (BEAREX08) in Texas, USA. Issues related to hot and cold pixel selection and the underlying assumptions of difference between air and surface temperature (dT) being linearly related to the surface temperature were also addressed. Estimated instantaneous evapotranspiration (ET) and other components of the surface energy balance were compared with measured data from four large precision weighing lysimeter fields, two each managed under irrigation and dryland conditions. Instantaneous ET was estimated with overall mean bias error and root mean square error (RMSE) of 0.13 and 0.15 mm h−1 (23.8 and 28.2%) respectively, where relatively large RMSE was contributed by dryland field. Sensitivity analysis of the hot and cold pixel selection indicated that up to 20% of the variability in ET estimates could be attributed to differences in the surface energy balance and roughness properties of the anchor pixels. Adoption of an excess resistance to heat transfer parameter model into SEBAL significantly improved the instantaneous ET estimates.  相似文献   

18.
Accurate estimation of evapotranspiration (ET) is essential in water resources management and hydrological practices. Estimation of ET in areas, where adequate meteorological data are not available, is one of the challenges faced by water resource managers. Hence, a simplified approach, which is less data intensive, is crucial. The FAO‐56 Penman–Monteith (FAO‐56 PM) is a sole global standard method, but it requires numerous weather data for the estimation of reference ET. A new simple temperature method is developed, which uses only maximum temperature data to estimate ET. Ten class I weather stations data were collected from the National Meteorological Agency of Ethiopia. This method was compared with the global standard PM method, the observed Piche evaporimeter data, and the well‐known Hargreaves (HAR) temperature method. The coefficient of determination (R2) of the new method was as high as 0.74, 0.75, and 0.91, when compared with that of PM reference evapotranspiration (ETo), Piche evaporimeter data, and HAR methods, respectively. The annual average R2 over the ten stations when compared with PM, Piche, and HAR methods were 0.65, 0.67, and 0.84, respectively. The Nash–Sutcliff efficiency of the new method compared with that of PM was as high as 0.67. The method was able to estimate daily ET with an average root mean square error and an average absolute mean error of 0.59 and 0.47 mm, respectively, from the PM ETo method. The method was also tested in dry and wet seasons and found to perform well in both seasons. The average R2 of the new method with the HAR method was 0.82 and 0.84 in dry and wet seasons, respectively. During validation, the average R2 and Nash–Sutcliff values when compared with Piche evaporation were 0.67 and 0.51, respectively. The method could be used for the estimation of daily ETo where there are insufficient data. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

19.
This study explores surface-layer stability effects on the Penman-Monteith (P-M) method, one of the most widely used methods to estimate evapotranspiration (ET). Stability correction is applied to the original (neutral stability) formula by using atmospheric exchange coefficients developed by Louiset al. (1981). First, the effects of stability on the P-M formula are explored theoretically. ET is then computed from field data using both P-M formulas and the values are compared to measured ET from Bowen-ratio and lysimeter data. Theoretical investigation of the P-M formula is performed by varying stability conditions for a range of vapor pressure deficit and canopy conductance. Infinitely large canopy conductance (zero canopy resistance) gives potential ET with results similar to Mahrt and Ek (1984): the stability-dependent formula gives larger (smaller) potential ET than the original formula under unstable (stable) conditions. For finite canopy conductance, ET behaves differently from potential ET as a result of coupling between canopy and atmospheric conductances. Stability-dependent ET values become smaller than the original formula values under extreme unstable conditions (when atmospheric conductance becomes large compared to canopy conductance) because stability-dependent sensible heat flux is considerably larger than its neutral counterpart under same net input energy. Field data collected during both wet and dry growing seasons indicate that both P-M formulas track the Bowen-ET estimates and lysimeter measurements quite closely.  相似文献   

20.
The dissipation method, the method preferred for estimating scalar surface fluxes over open water has not traditionally been used by agronomists, whereas the surface renewal (SR) theory in conjunction with the analysis of the scalar time trace offers tremendous advantages for estimating fluxes over agronomic crops. For a steady and horizontally homogeneous flow, it is shown that the dissipation method and SR analysis are closely related. As a consequence, a new dissipation–SR analysis expression for estimating scalar surface fluxes was derived. The new equation requires no calibration, and the scalar time trace measured at a frequency capable of identifying canopy‐scale coherent structures (typically 4–10 Hz in agriculture) is the only input required. Sensible and latent heat flux estimates obtained from 10 Hz air temperature and water vapour concentration measurements in the inertial sub‐layer (2 m height) over short, homogeneous rangeland grass at a site where similarity does not hold gave similar results to those measured with the eddy covariance (EC) method. For unstable cases, the new equation provided a root mean square error of 57 W m?2 for the surface energy‐balance closure. For stable cases, the performance was difficult to evaluate because the EC fluxes were similar in magnitude to the sensor error. It is concluded that the proposed method can contribute to a better understanding of hydrological processes and water requirements by providing an accurate, less costly, alternative method to indirectly estimate evapotranspiration as the residual of the energy balance equation. Copyright © 2009 John Wiley & Sons, Ltd.  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