Next Article in Journal
The Evaluation of a Long-Term Experiment on the Relationships between Weather, Nitrogen Fertilization, Preceding Crop, and Winter Wheat Grain Yield on Cambisol
Next Article in Special Issue
Impact of Preseason Climate Factors on Vegetation Photosynthetic Phenology in Mid–High Latitudes of the Northern Hemisphere
Previous Article in Journal
Salinity Mitigates the Negative Effect of Elevated Temperatures on Photosynthesis in the C3-C4 Intermediate Species Sedobassia sedoides
Previous Article in Special Issue
Differential Water Conservation Capacity in Broadleaved and Mixed Forest Restoration in Latosol Soil-Eroded Region, Hainan Province, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Environmental Controls on Evapotranspiration and Its Components in a Qinghai Spruce Forest in the Qilian Mountains

1
College of Environment and Resource, Shanxi University, Taiyuan 030006, China
2
Academy of Water Resources Conservation Forests in Qilian Mountains of Gansu Province, Zhangye 734000, China
3
Shanxi Laboratory for Yellow River, Taiyuan 030006, China
4
Northwest Institute of Eco-Environment and Resources, Chinese Academy of Sciences, Lanzhou 730000, China
5
Gansu Qilian Mountain Forest Eco-System of the State Research Station, Zhangye 734000, China
6
China National Forestry-Grassland Development Research Center, Beijing 100714, China
*
Authors to whom correspondence should be addressed.
Plants 2024, 13(6), 801; https://doi.org/10.3390/plants13060801
Submission received: 30 January 2024 / Revised: 23 February 2024 / Accepted: 4 March 2024 / Published: 12 March 2024
(This article belongs to the Special Issue Responses of Vegetation to Global Climate Change)

Abstract

:
Qinghai spruce forests, found in the Qilian mountains, are a typical type of water conservation forest and play an important role in regulating the regional water balance and quantifying the changes and controlling factors for evapotranspiration (ET) and its components, namely, transpiration (T), evaporation (Es) and canopy interceptions (Ei), of the Qinghai spruce, which may provide rich information for improving water resource management. In this study, we partitioned ET based on the assumption that total ET equals the sum of T, Es and Ei, and then we analyzed the environmental controls on ET, T and Es. The results show that, during the main growing seasons of the Qinghai spruce (from May to September) in the Qilian mountains, the total ET values were 353.7 and 325.1 mm in 2019 and 2020, respectively. The monthly dynamics in the daily variations in T/ET and Es/ET showed that T/ET increased until July and gradually decreased afterwards, while Es/ET showed opposite trends and was mainly controlled by the amount of precipitation. Among all the ET components, T always occupied the largest part, while the contribution of Es to ET was minimal. Meanwhile, Ei must be considered when partitioning ET, as it accounts for a certain percentage (greater than one-third) of the total ET values. Combining Pearson’s correlation analysis and the boosted regression trees method, we concluded that net radiation (Rn), soil temperature (Ts) and soil water content (SWC) were the main controlling factors for ET. T was mainly determined by the radiation and soil hydrothermic factors (Rn, photosynthetic active radiation (PAR) and TS30), while Es was mostly controlled by the vapor pressure deficit (VPD), atmospheric precipitation (Pa), throughfall (Pt) and air temperature (Ta). Our study may provide further theoretical support to improve our understanding of the responses of ET and its components to surrounding environments.

1. Introduction

Evapotranspiration (ET) reflects the complex interactions of climate, vegetation, soil and terrain [1]. It is a key component of the Earth’s hydrological system and surface energy balance [2,3], with more than 60% of the annual global precipitation being returned into the atmosphere [4] through plant transpiration (T) as well as evaporation from both soil (Es) and intercepted water from wet leaves and surfaces (Ei) [5]. Among these, T is one of the fundamental ways in which ecosystems gather water and is currently the primary water flux on Earth [6]. In addition, T regulates the water transport mechanism in the soil–plant–atmosphere continuum through the stomata [7] and is also essential for hydrological processes, the carbon cycle and the energy balance of ecosystems [8,9]; it is strongly influenced by plants’ physiological characteristics. For Es, it is dominated by the physical factors that result from the diffusion of water to the soil surface and the patterns of rainfall and the structural characteristics of the vegetation stand [10]. What is more, several climatological parameters like solar radiation, air temperature (Ta, °C), atmospheric relative humidity (RH, %) and wind speed (u, m s−1) can, if not well measured or modelled, negatively affect the assessment of Es [11,12,13]. Es is often regarded as ineffective water consumption that causes waste or a low utilization rate of water resources [14]. However, Es can keep an ecosystem cool, further supporting vegetation photosynthesis and other functions, in some cases [15]. Ei has always been neglected due to its marginal proportion in the ET of vegetation in extreme arid regions. However, for forest ecosystems with dense canopies or vegetation grown in humid environmental conditions, Ei cannot be neglected, as it usually accounts for a significant proportion of ET. In such cases, the accurate partitioning of ET into T, Es and Ei is a critical step which yields both a comprehensive insight into hydrological processes and better water management [16] and is essential to improve the modeling of land–atmosphere interactions, especially by predicting the temporal response to droughts across biomes [17,18].
There are multiple established methods for partitioning ET, ranging from simple to complex, instantaneous to continuous and point scale to satellite pixel scale. These methods were reviewed by Kool et al. [14] and classified into two categories: models and measurements. The commonly used models were the Shuttleworth–Wallace model [19] and its improved structures [20,21,22,23,24,25], the clumped model [26,27], the FAO-56 dual crop coefficient model [28,29,30] and other improved dual-source models [31,32,33], while measurements were mainly eddy covariance techniques [34,35,36,37,38] and Bowen ratio systems [39,40,41] (acquiring ET), stable isotopes [42,43,44,45] (acquiring Es or T), sap flow meters [46,47,48,49] (acquiring T), microlysimeters [50,51,52] (acquiring Es) and water collection tanks [53,54] (acquiring Ei). Among these methods, the modeling approach has the advantage of its applicability over a wide range of time scales and can be applied to the spatial scale of an entire ecosystem [55,56], but these models always require complex parameterizations and still require validation. For measurements, researchers usually measure three of four parameters (ET and its components) and then calculate the last parameter based on the assumption that ET = T + Es + Ei [10,57,58,59,60]. ET partitioning is site-specific and strongly influenced by the availability of water and energy [61], so it is still urgent to study the partitioning of ET across various vegetation types under changing climatic conditions.
As ET tends to display complex spatiotemporal patterns across scales due to the complex interactions between ET and the surrounding environments [62,63], it is of critical importance to elucidate how ET responds to a variety of environmental variables across landscapes with varying land surface and climatic conditions [64,65,66]. Though a great number of studies have explored the response characteristics of ET to different environmental factors [67,68,69,70], no unified conclusion has been reached so far [71]. For each component of ET, the dominant environmental controlling factors are also different. The environmental response of T has been the focus of research in the fields of water demand and supply and the adaptation mechanism of plants to their environments [72,73,74,75]. T is generally assumed to be affected by site hydrometeorological factors [76], such as net/shortwave radiation (Rn/Rs, W m−2), vapor pressure deficit (VPD, kPa), atmospheric precipitation (Pa, mm), soil water content (SWC, %), Ta and RH across temporal scales [77,78,79,80]. Among these factors, Rn/Rs, Ta and VPD are positively correlated with T by controlling stomatal conductance over short timescales (i.e., hourly and daily), while for longer timescales (i.e., seasonally and interannually), SWC and Pa are responsible for most of the variation observed in T [73,79,81]. Consequently, high T rates always correspond to high Ta, low RH and sufficient soil water availability [82,83,84]. Es, in contrast, represents the phenomenon concerning the change in the liquid phase from water to vapor [85], which primarily depends on soil and environmental factors [16,86,87,88] and may vary between understory and bare soil circumstances. In conclusion, due to the different water consumption mechanisms, ET, T and Es are likely to have different responses to changes in environmental factors.
The Qilian mountains (QLMs), located in the upper reach of the Heihe River Basin (HRB), the second largest inland river basin in China, are an important ecological barrier in the northwest regions. As the constructive tree species in the QLMs, Qinghai spruce (QHS) forest is a typical type of water conservation forest and plays an important role in regulating the regional water balance. Given the complex topography and climatic conditions, the region’s QHS ecosystems are fragile and sensitive to climate change [89]. ET is the most important way to consume water for the QHS; therefore, identifying the characteristics and environmental responses of ET and its components is of great significance for the regular growth of QHS. In this study, we quantified ET, T and Ei by direct field measurements and calculated Es based on the assumption that total ET equals the sum of T, Es and Ei, the latter often being neglected due to its small impact on the water balance of some areas or simplifications used in calculations. Then, we analyzed the environmental controls on ET, T and Es by using the boosted regression trees (BRTs) method. The purposes of our study were to: (1) partition ET into T, Es and Ei and analyze the magnitudes and changing characteristics of each parameter; and (2) identify the dominant environmental controlling factors for ET, T and Es.

2. Materials and Methods

2.1. Experimental Site

This study was conducted in the Pailugou watershed (38°31′–38°33′ N, 100°16′–100°18′ E), located in the middle part of the QLMs (36°43′–39°36′ N, 97°25′–103°46′ E). The watershed has an area of 2.85 km2 and an elevation range between 2500 and 3800 m a.s.l. In this area, the mean total daily Rn is 110.28 kW m−2, the mean annual Ta is 1.6 °C, the mean annual Pa is 435.5 mm year−1, the mean annual RH is 60% and the mean daily u is between 0.1 and 2.8 m s−1 [90].
Our experimental site (38°33′12.15″, 100°17′8.19″, altitude 2765 m a.s.l; Figure 1) was composed of 126 pure natural QHS trees aged between 49 and 119 years. The average tree height was 10.5 m, the average diameter at breast height was 14.78 cm, the average sapwood area was 33.3 cm2 and the average crown breadth of the trees was 322 cm × 357 cm. The canopy density was 0.58, and the tree density was 1200 trees·ha−1.

2.2. Eddy Covariance (EC) and Meteorological Measurements

A uniform open-path EC system which consisted of a 3D sonic anemometer/thermometer (model CSAT3, Campbell Scientific Inc., Logan, UT, USA) and an open-path CO2/H2O gas analyzer (model LI-7500, Li-COR Inc., Lincoln, NE, USA) was installed in a tower to monitor CO2/H2O fluxes at a height of 24 m. The fetch of the EC system was approximately 1000 m in the predominant upwind direction. Signals were recorded by a datalogger (model CR3000, Campbell Scientific, Logan, UT, USA) and block-averaged over 30 min intervals.
Meteorological instruments were installed at heights of 2, 15 and 30 m in the tower to monitor all available meteorological variables. Among those, two sets of rain gauges (TE525MM, Campbell Scientific Inc., Logan, UT, USA) were installed at heights of 2 m and 30 m to measure Pa and throughfall (Pt, mm). Rn/Rs and photosynthetic active radiation (PAR; W m–2) were measured by radiometers (Li 200X, Li-COR Inc., Lincoln, NE, USA; LI-190SA, LI-COR Inc., Lincoln, NE, USA). Ta and RH were measured by instruments (HMP115A, Onset, Bourne, MA, USA; HMP45A, Campbell Scientific, USA), then VPD could be calculated from Ta [91]. Soil temperature (Ts, °C) (109SS, Campbell Scientific, Logan, UT, USA) and SWC (SMC300, Spectrum Technologies, Plainfield, IL, USA) were measured at depths of 10, 20, 30, 40, 60 and 80 cm.

2.3. Sap Flow and Transpiration

Three sample trees that were representative of the surrounding stand were selected (Table 1). For each tree, a thermal dissipation probe (TDP) (RR-8210, Rainroot Ltd., Beijing, China) with two needles was inserted into the sapwood, with only the upper needle continually heated and the lower one kept normal. The sensor output signal was the difference in temperature (dt) between the two probes. Before installation, a small piece of bark was removed with a length between 10 and 15 cm in diameter at breast height (DBH, cm). Afterwards, two holes were drilled (the depths of which were slightly greater than 10 mm) using a drilling plate, and the probes with lengths of 10 mm were inserted into the holes. Then, the probes were covered with foam boxes and wrapped with reflective paper to prevent the influence of the natural environment on the measurement data. Eventually, the probes were connected to a datalogger and set up to collect data every 2 s, and the data were automatically averaged and stored every 10 min [92]. The sap flow velocity (Vs, cm h–1) can be calculated according to the following formula [93]:
V s = 0.0119 d t m d t d t 1.231
where dt is the difference in temperature at one moment, in °C, and dtm is the maximum difference in temperature in a day, in °C.
Then, T can be calculated through the following formula:
T = 0.001 A i n V s i A s i N i
where Vsi is the average sap flow velocity at i diameter class, cm h–1; Asi is the total area at i diameter class, cm2; n is the total number of diameter classes; and Ni is the total number of trees at i diameter class.

2.4. Stem Flow (S) and Canopy Interception

According to the distribution of the diameter class of the QHS stand, 1~3 sample trees were chosen from each diameter class to measure S, and a total of 10 sample trees were selected in the end. When measuring, we first made grooves by cutting several polyethylene plastic pipes and fixed them at heights of 50~140 cm above the bases of the trunks by using nails and glass cements; then, the stem flow was collected by importing it into a plastic bucket with a capacity of 10 L. S could then be calculated through the following formula:
S = i = 1 n C i × N i A s t × 1000
where Ci is the volume of S at i diameter class, mL, and Ast is the area of the QHS stand, m2.
Ei can be calculated according to the following principle of water balance:
E i = P a P t S

2.5. Flux Data Processing and Gap Fillings

CO2/H2O flux and CSAT3 data obtained from the EC system were processed using the EddyPro software version 6.0.0 (LI-COR Biosciences, Inc., Lincoln, NE, USA). Then, some corrections, including time lag compensation by covariance maximization (with default), density fluctuation compensation by the use/convert feature to the mixing ratio, block average detrending and double coordinate rotation, were made. Statistical analysis of the raw time series data was processed according to [94], and quality check flagging was obtained according to the CarboEurope standard.
The gaps in the ET data were filled by using different methods according to the gap lengths. Specifically, the gaps of less than 2 h were filled through the linear interpolation method and the gaps of less than 2 days were filled through the mean diurnal variation method [95], while the gaps longer than 2 days were filled through the energy balance equation [96].

2.6. Boosted Regression Trees (BRTs)

The BRTs method developed by Elith et al. [97] was used to quantify the contribution of each environmental variable to ET, T and Es. The BRTs method combines the strengths of regression tree algorithms and the boosting technique, which can handle different data types and accommodate missing data, and its prediction performance is superior to most traditional modeling methods in fitting complex nonlinear relationships [98]. In this study, the gbm package was used to run the BRTs model in R software (version 4.3.1); the values of the learning rate, tree complex and bag fraction were 0.01, 10 and 0.5, respectively, and the family type was Gaussian. The output results represent the relative importance of a single impact factor in the form of a percentage.

2.7. Statistical Analyses

The study periods we selected were from May to September (the main growing season of QHS) in 2019 and 2020, all the measured data for which were processed to hourly values. All the statistical analyses were performed using SPSS 19.0 and Origin 8.0.

3. Results

3.1. Variations in the Main Environmental Factors

Detailed information on the key meteorological factors (Figure 2) is essential to assess their variations with respect to ET and its components. During the study periods in 2019 and 2020, Rn, PAR and Ta showed similar trends, and the daily values of all three parameters in July and August were greater than those in other months. The PAR values were 125.9 and 124.7 Wm−2, which accounted for 74.8% and 73.3% of Rn (Figure 2a). The average Ta values were both 11.1 °C (Figure 2b). The variation trends of Ts at different depths were similar, and Ts showed more significant trends that increased first and then decreased compared to Ta (Figure 2c). The VPD variations were rather consistent with Ta, as VPD was calculated from the equation based on Ta. The average VPD values were 0.57 and 0.61 kPa and ranged from 0.02 to 1.38 and 0.03 to 1.44 kPa, respectively (Figure 2b). Pt accounted for 76.8% and 89.7% of Pa, indicating that a significant proportion of precipitation had been intercepted (Figure 2d). The variation trends in SWC were consistent with the amounts of Pt and Pa, and this was because precipitation is the only source of water input for the QHS ecosystem (Figure 2e). The average values for RH were 62.8% and 60.6%, and they ranged from 23.8% to 97.1% and 16.1% to 96.7%, respectively. u remained relatively stable (there were no significant seasonal differences), and its variation was fairly similar each year (its average values were 0.58 and 0.59 m s−1 and ranged from 0.05 to 0.99 and 0.03 to 0.93 m s−1, respectively) (Figure 2f).

3.2. Variations in ET, T and Es

Monthly mean diurnal variations in ET and T are shown in Figure 3. Both parameters increased rapidly from sunrise and reached a maximum at approximately midday and then decreased (Figure 3a,b), and the maximum values of T were greater than 0.3 mm h−1. We can also see the much lower values of T between sunset and sunrise on the next day (Figure 3b), which confirmed the existence of nocturnal transpiration. Es and Ei values were not analyzed at the monthly mean diurnal scale, as Ei was not measured hourly and Es was calculated by the difference between ET and T and Ei.
The monthly dynamics of daily variations in ET, T and Es during the main growing season of the QHS are shown in Figure 4. The peak values for ET (4.4 and 4.3 mm) and T (2.8 and 2.7 mm) occurred at the end of June or in July. Both parameters increased first and then decreased during the whole study period in each year, especially for T, which showed even more significant trends, as the period (between July and August) was the flourishing growing stage of the QHS (Figure 4a,b). Total values between July and August were 108.3 and 104.0 mm for T, which accounted for 57.2% and 56.0% of the total amounts in 2019 and 2020, respectively. There were no similar variation trends for Es compared to ET and T (Figure 4c). This is because Es was mainly controlled by the hydrothermic condition of the soil and was not correlated with plant growth.
Monthly cumulative values for ET, T and Es during the study periods in each year are shown in Figure 5. The total amounts of ET were 353.7 mm in 2019 and 325.1 mm in 2020, and, among all the components, T was always the largest part (189.5 mm in 2019 and 185.7 mm in 2020). ET values in each month in both years were not significantly different, except for July, during which the difference was 24.6 mm (Figure 5a), and this was mainly due to the significant difference in P (51.1 mm) during the period (Figure 2d), which influenced Es rather than T (Figure 5b,c).

3.3. Variations in T/ET, Es/ET and Ei/ET

Monthly dynamics of daily variations in T/ET and Es/ET are shown in Figure 6. T/ET increased until July and gradually decreased afterwards, while Es/ET showed opposite trends. The daily maximum T/ET and Es/ET values can both reach up to as high as 90%. In each month, in both years, T/ET was nearly always the greatest among all the components, except for June in 2019, when the T/ET (46.8%) value was slightly lower than that for Ei/ET (48.5%), and this was mainly due to the significant amount of P (104.6 mm) in the period. Overall, T accounted for at least one-third of the ET and Es/ET accounted for as little as 3%, while Ei/ET was uncertain, as it was mainly controlled by the amount of P (Figure 7). From the total amounts of T, Es and Ei, the T/ET values were always the greatest at 53.6% in 2019 and 57.1% in 2020. Ei accounted for a certain percentage of ET (32.9% in 2019 and 31.6% in 2020), which indicated the important role of Ei in the water balance of the Qinghai spruce forest ecosystem. The contribution of Es to ET was minimal among all the components, which reflects the excellent water conservation function of the QHS ecosystem (Figure 8).

3.4. Dominant Environmental Controlling Factors for ET, T and Es

Firstly, we selected the environmental factors that were significantly related to ET, T and Es by using Pearson’s correlation analysis. Then, we used the BRTs model to identify the environmental factors that greatly contributed to ET, T and Es. Pearson’s correlation analysis revealed that ET was significantly positively correlated with Rn and PAR (the correlation coefficients were 0.75 and 0.76, respectively) and then u2 and u15 (the correlation coefficients were 0.51 and 0.57, respectively). T was significantly positively correlated with Ta, Rn, PAR, Ts and SWC, with correlation coefficients all greater than 0.5, and Es was significantly positively correlated with VPD (for which the correlation coefficient was 0.51). In addition, ET, T and Es were all negatively correlated with RH (Figure 9).
Then, we conducted a multicollinearity assessment of the environmental variables and selected those which passed the multicollinearity test (the absolute value of the related coefficient was lower than 0.8, and the variance inflation factor was lower than 10). Eventually, some of the many environmental factors were selected as the predictor variables. The BRTs analysis showed that, among the selected predictor variables for ET, T and Es, ET was mainly controlled by Rn, Ts30 and SWC40, with contributions at 51.0%, 15.7% and 8.5%, respectively. T was mainly controlled by Rn and PAR, with contributions at 34.9% and 34.2%, respectively, and then Ts30 (the contribution was 13.2%). For Es, it was mainly controlled by VPD, Pt, Ta and Pa, with contributions at 28.5%, 11.8%, 10.9% and 9.9%, respectively (Figure 10).

4. Discussion

4.1. Partitioning Methods for ET

Accurately partitioning ET into T, Es and Ei can not only help us better understand the water and energy exchanges between the surface and atmosphere, but also determine the water demand and further improve water resource management [15]. Thus, partitioning ET is crucial for understanding water resources in the context of global climate change [99]. ET can be partitioned mainly through the methods of modeling and measuring. For modeling, the SW model and the clumped model and their improved forms have been widely used due to their suitable physical mechanisms. Researchers nowadays mainly focus on optimizing the structures of the dual- or multi-source models [20,22,100,101,102], but these models incorporate a large number of parameters, some of which are difficult to parameterize [31]. The FAO-56 dual crop coefficient model is an indirect method for calculating ET from reference crop ET and crop coefficients. Based on the basal crop coefficient and the soil evaporation coefficient recommended by the FAO, the FAO-56 dual crop coefficient model has been widely used in estimating crop ET and its components. However, crop coefficients for many types of vegetations have not been provided by the FAO. In general, though great efforts have been made and modeling accuracies have also been improved, there are limitations to the use of these models. Moreover, modeling can never represent the true amounts of ET and its components, and the applicability of the models for different ecosystems still requires extensive verification. In contrast, measuring methods can accurately quantify ET and its components. The commonly used measuring methods include stable isotope [103,104,105,106], sap flow and microlysimeters [10,22,107,108]; the EC high-frequency correlation approach [109,110,111,112]; and large-aperture scintillometers with Bowen ratio systems [113,114,115,116]. Studies based on these instruments and the assumption that ET equals the sum of all the components have mostly focused on partitioning ET into T and Es [60,117,118,119,120,121,122,123]; in fact, Ei should also be considered when it accounts for a significant percentage.

4.2. Proportions of T and Ei on ET

Understanding the annual variation in T/ET remains a challenge and is essential to garner a thorough understanding of plant responses to the changing environment [124]. A number of studies carried out with different methodologies have identified that there exists a large variability (appropriately 20–90%) in T/ET across biomes or even at the global scale [125]. For a typical inland river basin, among the various environmental conditions, the T/ET for vegetations in arid and semi-arid climate zones in lower reaches with sufficient water supply was between 20 and 70% [126,127,128,129], while for vegetations in high-altitude mountains in upper reaches, the mean values of this ratio were slightly greater. For the QHS in the QLMs in the upper reaches of the Heihe River in our study, the T/ET values were 53.6% and 57.1% in 2019 and 2020, respectively, which were close to the results for the QHS in the QLMs obtained by other researchers [130,131], while being lower than those obtained for subtropical coniferous plantations, e.g., 63~68% for Wei et al. [132], 69~85% for Zhu et al. [133] and 77.4% for Ren et al. [134]. Monthly variations in T/ET showed that T contributed the most to the ET, especially in summer (the T/ET reached up to 75% on July in 2020), indicating that T is the most important factor for ET consumption and that it changes with the growing stages of the QHS.
Ei is an essential component of the forest hydrological cycles. Studies on Ei processes have been conducted on various plants, all of which clearly indicated that it cannot be ignored as an important component of water balance [130,135,136,137]. In our study, the Ei/ET values for the QHS in the QLMs were 32.9% and 31.6% in 2019 and 2020, respectively, which further emphasized that Ei plays an undeniable role in ET. The amount and percentage of Ei vary between climatic regions [138], as they depend on factors such as vegetation characteristics [139,140], meteorological conditions [141] and precipitation characteristics [142,143]. According to previous studies of the QHS in the QLMs, Ei is mainly controlled by P. Generally speaking, a low amount or intensity of P leads to small impacts on the branches and leaves, and the rainwater can be intercepted with great probability. Furthermore, the high values for the leaf area index (the maximum value was 3.96) of the QHS in the QLMs also contributed to the greater Ei/ET.
Proportions of T, Es and Ei on the ET of the QHS and other coniferous forests are summarized in Table 2.

4.3. Environmental Controls for ET, T and Es

The role of environmental variables in controlling ET and its components is an important but not well-understood aspect in arid areas in northwestern China. In our study, we concluded that Rn, Ts and SWC were the main controlling factors for ET, and numerous previous studies reached the same conclusions [150,151,152,153,154,155,156]. Among these, Rn is the primary driver of ET, as radiation is the energy source and most of the radiant energy absorbed by the leaves is used for T [157]. SWC can not only affect the development of the leaf area index, but also directly and positively affect surface conductance by decreasing soil surface resistance and stomatal resistance and then increasing soil evaporation and transpiration [146]. The effect of Ts on ET can be explained, as rising Ts can accelerate processes of Es and T directly by promoting the roots’ absorption of water from the soil, and the water conservation function of the QHS ensures the presence of an adequate water resource in the soil. For T and Es, the main environmental controlling factors were different. T was mainly determined by the radiation and soil hydrothermic factors (Rn, PAR and TS), which was consistent with the work by Du et al. [158] and Yang et al. [152], who also reached the same conclusion for the QHS in the QLMs, and also consistent with the work by Chen et al. [159,160]. VPD contributed less than the radiation and soil hydrothermic factors in our study, which was also consistent with the work by Du et al. [158], and the reason for this is that VPD is not the most important controlling factor, but the result of the comprehensive effect of meteorological factors to some extent [158]. Es was mostly controlled by VPD, Pa, Pt and Ta, which was consistent with the work by Wang et al. [158], who also reached the same conclusion for the QHS in the QLMs. The climate of the QLMs is typically continental, and in this area, especially after spring, the Mongolia high press reduces and the subtropical high pressure expands gradually, which results in Ta and u increasing rapidly, thus promoting the Es process. Consequently, it is of great significance to keep the soil moisture needed for the regular growth of the QHS and prevent spring drought in the QLMs [161].

5. Conclusions

Identifying ET partitioning results, as well as the variations and controlling factors for ET and its components, is essential for ensuring the regular growth of the QHS in the QLMs. In this study, we analyzed all these aspects, and the primary results are as follows: (1) ET increased first and then decreased during the main growing season of the QHS, with total values of 353.7 and 325.1 mm in 2019 and 2020, respectively. (2) The monthly dynamics of the daily variations in T/ET and Es/ET showed that T/ET increased until July and gradually decreased afterwards, while Es/ET showed opposite trends. (3) T always accounted for the largest part of ET and was significantly greater than Es; meanwhile, Ei cannot be neglected, as it accounts for a certain percentage. (4) The main environmental controlling factors for ET and its components were different. ET was mainly controlled by Rn, Ts and SWC; T was mainly determined by the radiation and soil hydrothermic factors (Rn, PAR and TS30); and Es was mostly affected by VPD, Pa, Pt and Ta.

Author Contributions

W.J. and X.R. collected the experimental data; Y.H., X.G., J.S., B.W., Y.W. (Yujing Wen) and Y.W. (Yin Wang) conducted the analyses; G.G. and E.X wrote the manuscript. The other co-authors participated equally in the conceptualization, data analysis and manuscript preparation and review. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Post-Doctoral Special Program of Gansu Province (23JRRG0026), the Youth Fund of the National Natural Science Foundation of China (42301030), the National Natural Science Foundation of China (U21A20468), the Innovation Base and Talent Planning Project of Gansu Province (23JDKG0001), the Fundamental Research Program of Shanxi Province (202203021221035), the Scientific and Technological Funds for Young Scientists of Gansu Province (20JR10RG824) and the Postgraduate Education Innovation Projects of Shanxi Province (2022Y107, 2023KY110, 2023KY111).

Data Availability Statement

Data are contained within the article.

Acknowledgments

We wish to express our thanks for the support received from the Gansu Qilian Mountain Forest Eco-System of the State Research Station, Academy of Water Resources Conservation Forests in Qilian Mountains of Gansu Province. The authors also offer their sincere appreciation for the helpful and constructive comments of the reviewers of the draft manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

DescriptionsSymbolsUnits
The Qilian mountainsQLMs
Heihe River BasinHRB
Qinghai spruceQHS
EvapotranspirationETmm
TranspirationTmm
EvaporationEsmm
Canopy interceptionEimm
Net radiationRnW m−2
Shortwave radiationRsW m−2
Photosynthetic active radiationPARW m−2
Air temperatureTa°C
Soil temperatureTs°C
Vapor pressure deficitVPDkPa
Atmospheric precipitationPamm
ThroughfallPtmm
Atmospheric relative humidityRH%
Soil water contentSWC%
Wind speedum s−1
Stem flowSmm
Diameter at breast heightDBHmm
Sapwood areaAscm2
Boosted regression treesBRTs

References

  1. Feng, J.J.; Wang, W.Z.; Che, T.; Xu, F.N. Performance of the improved two-source energy balance model for estimating evapotranspiration over the heterogeneous surface. Agric. Water Manag. 2023, 278, 108159. [Google Scholar] [CrossRef]
  2. Jung, M.; Reichstein, M.; Ciais, P.; Seneviratne, S.I.; Sheffield, J.; Goulden, M.L.; Bonan, G.; Cescatti, A.; Chen, J.; Jeu, R.; et al. Recent decline in the global land evapotranspiration trend due to limited moisture supply. Nature 2010, 467, 951–954. [Google Scholar] [CrossRef]
  3. Wang, K.; Dickinson, R.E. A review of global terrestrial evapotranspiration: Observation, modeling, climatology, and climatic variability. Rev. Geophys. 2012, 50, 1–54. [Google Scholar] [CrossRef]
  4. Burba, G.G.; Verma, S.B. Seasonal and interannual variability in evapotranspiration of native tallgrass prairie and cultivated wheat ecosystems. Agric. For. Meteorol. 2005, 135, 190–201. [Google Scholar] [CrossRef]
  5. Chen, X.; Yu, L.Y.; Cui, N.B.; Cai, H.J.; Jiang, X.L.; Liu, C.W.; Shu, Z.K.; Wu, Z.J. Modeling maize evapotranspiration using three types of canopy resistance models coupled with single-source and dual-source hypotheses—A comparative study in a semi-humid and drought-prone region. J. Hydrol. 2022, 614, 128638. [Google Scholar] [CrossRef]
  6. Jasechko, S.; Sharp, Z.D.; Gibson, J.J.; Birks, S.J.; Yi, Y.; Fawcett, P.J. Terrestrial water fluxes dominated by transpiration. Nature 2013, 496, 347–350. [Google Scholar] [CrossRef]
  7. Faralli, M.; Bontempo, L.; Bianchedi, P.L.; Moser, C.; Bertamini, M.; Lawson, T.; Camin, F.; Stefanini, M.; Varotto, C. Natural variation in stomatal dynamics drives divergence in heat stress tolerance and contributes to seasonal intrinsic water use efficiency in Vitis vinifera (subsp. sativa and sylvestris). J. Exp. Bot. 2022, 73, 3238–3250. [Google Scholar] [CrossRef] [PubMed]
  8. Jiao, L.; Lu, N.; Fang, W.W.; Li, Z.S.; Wang, J.; Jin, Z. Determining the independent impact of soil water on forest transpiration: A case study of a black locust plantation in the loess plateau, China. J. Hydrol. 2019, 572, 671–681. [Google Scholar] [CrossRef]
  9. Wang, H.L.; Guan, H.D.; Liu, N.; Soulsby, C.; Tetzlaff, D.; Zhang, X.P. Improving the jarvis-type model with modified temperature and radiation functions for sap flow simulations. J. Hydrol. 2020, 587, 124981. [Google Scholar] [CrossRef]
  10. Qubaja, R.; Amer, M.; Tatarinov, F.; Rotenberg, E.; Preisler, Y.; Sprintsin, M.; Yakir, D. Partitioning evapotranspiration and its long-term evolution in a dry pine forest using measurement-based estimates of soil evaporation. Agric. For. Meteorol. 2020, 281, 107831. [Google Scholar] [CrossRef]
  11. Droogers, P.; Allen, R.G. Estimating reference evapotranspiration under inaccurate data conditions. Irrig. Drain. Syst. 2002, 16, 33–45. [Google Scholar] [CrossRef]
  12. Mobilia, M.; Longobardi, A. Prediction of potential and actual evapotranspiration fluxes using six meteorological data-based approaches for a range of climate and land cover types. ISPRS Int. J. Geo-Inf. 2021, 10, 192. [Google Scholar] [CrossRef]
  13. Allen, R.G.; Pereira, L.S.; Howell, T.A.; Jensen, M.E. Evapotranspiration information reporting: I. Factors governing measurement accuracy. Agric. Water Manag. 2011, 98, 899–920. [Google Scholar] [CrossRef]
  14. Kool, D.; Agam, N.; Lazarovitch, N.; Heitman, J.L.; Sauer, T.J.; Ben-Gal, A. A review of approaches for evapotranspiration partitioning. Agric. For. Meteorol. 2014, 184, 56–70. [Google Scholar] [CrossRef]
  15. Xu, Z.W.; Zhu, Z.L.; Liu, S.M.; Song, L.S.; Wang, X.C.; Zhou, S.; Yang, X.F.; Xu, T.R. Evapotranspiration partitioning for multiple ecosystems within a dryland watershed: Seasonal variations and controlling factors. J. Hydrol. 2021, 598, 126483. [Google Scholar] [CrossRef]
  16. Nguyen, M.N.; Hao, Y.F.; Baik, J.; Choi, M. Partitioning evapotranspiration based on the total ecosystem conductance fractions of soil, interception, and canopy in different biomes. J. Hydrol. 2021, 603, 126970. [Google Scholar] [CrossRef]
  17. Williams, I.N.; Lu, Y.; Kueppers, L.M.; Riley, W.J.; Biraud, S.C.; Bagley, J.E.; Torn, M.S. Land-atmosphere coupling and climate prediction over the U.S. Southern great plains. J. Geophys. Res. 2016, 121, 125–144. [Google Scholar] [CrossRef]
  18. Konings, A.G.; Williams, A.P.; Gentine, P. Sensitivity of grassland productivity to aridity controlled by stomatal and xylem regulation. Nat. Geosci. 2017, 10, 284–288. [Google Scholar] [CrossRef]
  19. Shuttleworth, W.J.; Wallace, J.S. Evaporation from sparse crops-an energy combination theory. Q. J. R. Met. Soc. 1985, 111, 839–855. [Google Scholar] [CrossRef]
  20. Bao, Y.Z.; Liu, T.X.; Duan, L.M.; Tong, X.; Zhang, L.; Singh, V.P.; Lei, H.M.; Wang, G.Q. Comparison of an improved Penman-Monteith model and SWH model for estimating evapotranspiration in a meadow wetland in a semiarid region. Sci. Total Environ. 2021, 795, 148736. [Google Scholar] [CrossRef]
  21. Bao, Y.Z.; Duan, L.M.; Liu, T.X.; Tong, X.; Wang, G.Q.; Lei, H.M.; Zhang, L.; Singh, V.P. Simulation of evapotranspiration and its components for the mobile dune using an improved dual-source model in semi-arid regions. J. Hydrol. 2021, 592, 125796. [Google Scholar] [CrossRef]
  22. Qin, S.J.; Li, S.E.; Cheng, L.; Zhang, L.; Qiu, R.J.; Liu, P.; Xi, H.Y. Partitioning evapotranspiration in partially mulched interplanted croplands by improving the Shuttleworth-Wallace model. Agric. Water Manag. 2023, 276, 108040. [Google Scholar] [CrossRef]
  23. Jiang, Z.Y.; Yang, Z.G.; Zhang, S.Y.; Liao, C.M.; Hu, Z.M.; Cao, R.C.; Wu, H.W. Revealing the spatio-temporal variability of evapotranspiration and its components based on an improved Shuttleworth-Wallace model in the Yellow River Basin. J. Environ. Manag. 2020, 262, 110310. [Google Scholar] [CrossRef]
  24. Chen, H.; Jiang, A.Z.; Huang, J.H.J.; Li, H.; Mcbean, E.; Sing, V.P.; Zhang, J.W.; Lan, Z.Q.; Gao, J.J.; Zhou, Z.Q. An enhanced shuttleworth-wallace model for simulation of evapotranspiration and its components. Agric. For. Meteorol. 2022, 313, 108769. [Google Scholar] [CrossRef]
  25. Chen, H.; Huang, J.H.J.; Mcbean, E. Partitioning of daily evapotranspiration using a modified shuttleworth-wallace model, random Forest and support vector regression, for a cabbage farmland. Agric. Water Manag. 2020, 228, 105923. [Google Scholar] [CrossRef]
  26. Villagarcía, L.; Were, A.; García, M.; Domingo, F. Sensitivity of a clumped model of evapotranspiration to surface resistance parameterisations: Application in a semi-arid environment. Agric. For. Meteorol. 2010, 150, 1065–1078. [Google Scholar] [CrossRef]
  27. Zhang, B.Z.; Kang, S.Z.; Li, F.S.; Zhang, L. Comparison of three evapotranspiration models to Bowen ration-energy balance method for a vineyard in an arid desert region of northwest China. Agric. For. Meteorol. 2008, 148, 1629–1640. [Google Scholar] [CrossRef]
  28. Anapalli, S.S.; Fisher, D.K.; Pinnamaneni, S.R.; Reddy, K.N. Quantifying evapotranspiration and crop coefficients for cotton (Gossypium hirsutum L.) using an eddy covariance approach. Agric. Water Manag. 2020, 233, 106091. [Google Scholar] [CrossRef]
  29. Ding, R.; Kang, S.; Zhang, Y.; Hao, X.; Tong, L.; Du, T. Partitioning evapotranspiration into soil evaporation and transpiration using a modified dual crop coefficient model in irrigated maize field with ground-mulching. Agric. Water Manag. 2013, 127, 85–96. [Google Scholar] [CrossRef]
  30. Liu, C.; Cui, N.; Gong, D.; Hu, X.; Feng, Y. Evaluation of seasonal evapotranspiration of winter wheat in humid region of East China using large weighted lysimeter and three models. J. Hydrol. 2020, 590, 125388. [Google Scholar] [CrossRef]
  31. Gao, G.L.; Feng, Q.; Liu, X.D.; Zhao, Y.H. Measuring and modeling evapotranspiration of a Populus euphratica forest in northwestern China. J. For. Res. 2021, 32, 1963–1977. [Google Scholar] [CrossRef]
  32. Gao, G.L.; Zhang, X.Y.; Yu, T.F.; Liu, B. Comparison of three evapotranspiration models with eddy covariance measurements for a Populus euphratica Oliv. forest in an arid region of northwestern China. J. Arid Environ. 2016, 8, 146–156. [Google Scholar] [CrossRef]
  33. Dzikiti, S.; Volschenk, T.; Midgley, S.J.E.; Lötze, E.; Taylor, N.J.; Gush, M.B.; Ntshidi, Z.; Zirebwa, S.F.; Doko, Q.; Schmeisser, M.; et al. Estimating the water requirements of high yielding and young apple orchards in the winter rainfall areas of South Africa using a dual source evapotranspiration model. Agric. Water Manag. 2018, 208, 152–162. [Google Scholar] [CrossRef]
  34. Wagle, P.; Skaggs, T.H.; Gowda, P.H.; Northup, B.K.; Neel, J.P.S. Flux variance similarity-based partitioning of evapotranspiration over a rainfed alfalfa field using high frequency eddy covariance data. Agric. For. Meteorol. 2020, 285–286, 107907. [Google Scholar] [CrossRef]
  35. Paul-Limoges, E.; Wolf, S.; Schneider, F.D.; Longo, M.; Moorcroft, P.; Gharun, M.; Damm, A. Partitioning evapotranspiration with concurrent eddy covariance measurements in a mixed forest. Agric. For. Meteorol. 2020, 280, 107786. [Google Scholar] [CrossRef]
  36. Aouade, G.; Ezzahar, J.; Amenzou, N.; Er-Raki, S.; Benkaddour, A.; Khabba, S.; Jarlan, L. Combining stable isotopes, Eddy Covariance system and meteorological measurements for partitioning evapotranspiration, of winter wheat, into soil evaporation and plant transpiration in a semi-arid region. Agric. Water Manag. 2016, 177, 181–192. [Google Scholar] [CrossRef]
  37. Zhang, W.; Jung, M.; Migliavacca, M.; Poyatos, R.; Miralles, D.G.; El-Madany, T.S.; Galvagno, M.; Carrara, A.; Arriga, N.; Ibrom, A. The effect of relative humidity on eddy covariance latent heat flux measurements and its implication for partitioning into transpiration and evaporation. Agric. For. Meteorol. 2023, 330, 109305. [Google Scholar] [CrossRef]
  38. Rafi, Z.; Merlin, O.; Dantec, V.L.; Khabba, S.; Mordelet, P.; Er-Raki, S.; Amazirh, A.; Olivera-Guerra, L.; Hssaine, B.A.; Simonneaux, V.; et al. Partitioning evapotranspiration of a drip-irrigated wheat crop: Inter comparing eddy covariance-, sap flow-, lysimeter- and FAO-based methods. Agric. For. Meteorol. 2019, 265, 310–326. [Google Scholar] [CrossRef]
  39. Holland, S.; Heitman, J.L.; Howard, A.; Sauer, T.J.; Giese, W.; Ben-Gal, A.; Agam, N.; Kool, D.; Havlin, J. Micro-Bowen ratio system for measuring evapotranspiration in a vineyard interrow. Agric. For. Meteorol. 2013, 177, 93–100. [Google Scholar] [CrossRef]
  40. Uddin, J.; Hancock, N.H.; Smith, R.J.; Foley, J.P. Measurement of evapotranspiration during sprinkler irrigation using a precision energy budget (Bowen ratio, eddy covariance) methodology. Agric. Water Manag. 2013, 116, 89–100. [Google Scholar] [CrossRef]
  41. Maruyama, T.; Ito, K.; Takimoto, H. Abnormal data rejection range in the Bowen ratio and inverse analysis methods for estimating evapotranspiration. Agric. For. Meteorol. 2019, 269–270, 323–334. [Google Scholar] [CrossRef]
  42. Guo, X.Y.; Tian, L.D.; Wang, L.; Yu, W.S.; Qu, D.M. River recharge sources and the partitioning of catchment evapotranspiration fluxes as revealed by stable isotope signals in a typical high-elevation arid catchment. J. Hydrol. 2017, 549, 616–630. [Google Scholar] [CrossRef]
  43. Yuan, Y.; Wang, L.; Eang, H.; Lin, W.; Jiao, W.; Du, T. A modified isotope-based method for potential high-frequency evapotranspiration partitioning. Adv. Water Resour. 2022, 160, 104103. [Google Scholar] [CrossRef]
  44. Gillefalk, M.; Tetzlaff, D.; Marx, C.; Smith, A.; Meier, F.; Hinkelmann, R.; Soulsby, C. Estimates of water partitioning in complex urban landscapes with isotope-aided ecohydrological modelling. Hydrol. Process. 2022, 36, 14532. [Google Scholar] [CrossRef]
  45. Iraheta, A.; Birkel, C.; Benegas, L.; Rios, N.; Beyer, M. A preliminary isotope-based evapotranspiration partitioning approach for tropical Costa Rica. Ecohydrology 2021, 14, e2297. [Google Scholar] [CrossRef]
  46. Uddin, J.; Foley, J.P.; Smith, R.J.; Hancock, N.H. A new approach to estimate canopy evaporation and canopy interception capacity from evapotranspiration and sap flow measurements during and following wetting. Hydrol. Process. 2016, 30, 1757–1767. [Google Scholar] [CrossRef]
  47. Jiang, X.L.; Kang, S.Z.; Li, F.S.; Du, T.S.; Tong, L.; Comas, L. Evapotranspiration partitioning and variation of sap flow in female and male parents of maize for hybrid seed production in arid region. Agric. Water Manag. 2016, 176, 132–141. [Google Scholar] [CrossRef]
  48. Paço, T.A.; Pôças, I.; Cunha, M.; Silvestre, J.C.; Santos, F.L.; Paredes, P.; Pereira, L.S. Evapotranspiration and crop coefficients for a super intensive olive orchard. An application of SIMDualKc and METRIC models using ground and satellite observations. J. Hydrol. 2014, 519, 2067–2080. [Google Scholar] [CrossRef]
  49. Gong, D.; Kang, S.; Yao, L.; Zhang, L. Estimation of evapotranspiration and its components from an apple orchard in northwest China using sap flow and water balance methods. Hydrol. Process. 2007, 21, 931–938. [Google Scholar] [CrossRef]
  50. Liu, M.H.; Shi, H.B.; Paredes, P.; Ramos, T.B.; Dai, L.P.; Feng, Z.Z.; Pereira, L.S. Estimating and partitioning maize evapotranspiration as affected by salinity using weighing lysimeters and the SIMDualKc model. Agric. Water Manag. 2022, 261, 107362. [Google Scholar] [CrossRef]
  51. Liu, M.H.; Paredes, P.; Shi, H.B.; Ramos, T.B.; Dou, X.; Dai, L.P.; Pereira, L.S. Impacts of a shallow saline water table on maize evapotranspiration and groundwater contribution using static water table lysimeters and the dual Kc water balance model SIMDualKc. Agric. Water Manag. 2022, 273, 107887. [Google Scholar] [CrossRef]
  52. Schrader, F.; Durner, W.; Fank, J.; Gebler, S.; Pütz, T.; Hannes, M.; Wollschläger, U. Estimating precipitation and actual evapotranspiration from precision lysimeter measurements. Procedia Environ. Sci. 2013, 19, 543–552. [Google Scholar] [CrossRef]
  53. Zhu, Z.R.; Li, J.S.; Zhu, D.L. Influence of biotic and abiotic factors and water partitioning on the kinetic energy of sprinkler irrigation on a maize canopy. Agric. Water Manag. 2024, 293, 108700. [Google Scholar] [CrossRef]
  54. Zhu, Z.R.; Li, J.S.; Zhu, D.L.; Gao, Z. The impact of maize canopy on splash erosion risk on soils with different textures under sprinkler irrigation. Catena 2024, 234, 107608. [Google Scholar] [CrossRef]
  55. Xing, L.W.; Cui, N.B.; Liu, C.W.; Guo, L.; Zhao, L.; Wu, Z.J.; Jiang, X.L.; Wen, S.L.; Zhao, L.; Gong, D.Z. Estimating daily kiwifruit evapotranspiration under regulated deficit irrigation strategy using optimized surface resistance based model. Agric. Water Manag. 2024, 295, 108745. [Google Scholar] [CrossRef]
  56. Liu, X.; Xu, J.; Liu, B.; Wang, W.; Li, Y. A novel model of water-heat coupling for water-saving irrigated rice fields based on water and energy balance: Model formulation and verification. Agric. Water Manag. 2019, 223, 105705. [Google Scholar] [CrossRef]
  57. Paredes, P.; Rodrigues, G.J.; Petry, M.T.; Severo, P.O.; Carlesso, R.; Pereira, L.S. Evapotranspiration partition and crop coefficients of Tifton 85 Bermudagrass as affected by the frequency of cuttings. Application of the FAO56 Dual Kc Model. Water 2018, 10, 558. [Google Scholar] [CrossRef]
  58. Fisher, J.B.; Tu, K.P.; Baldocchi, D.D. Global estimates of the land-atmosphere water flux based on monthly AVHRR and ISLSCP-II data, validated at 16 FLUXNET sites. Remote Sens. Environ. 2008, 112, 901–919. [Google Scholar] [CrossRef]
  59. Yao, Y.; Liang, S.; Cheng, J.; Liu, S.; Fisher, J.B.; Zhang, X.; Li, Y.; Zhao, S. MODIS-driven estimation of terrestrial latent heat flux in China based on a modified Priestley-Taylor algorithm. Agric. For. Meteorol. 2013, 171–172, 187–202. [Google Scholar] [CrossRef]
  60. Yu, T.F.; Feng, Q.; Si, J.H.; Xi, H.Y.; O’Grady, A.P.; Pinkard, E.A. Responses of riparian forests to flood irrigation in the hyper-arid zone of NW China. Sci. Total Environ. 2019, 648, 1421–1430. [Google Scholar] [CrossRef]
  61. Lai, J.B.; Liu, T.G.; Luo, Y. Evapotranspiration partitioning for winter wheat with shallow groundwater in the lower reach of the Yellow River Basin. Agric. Water Manag. 2022, 266, 107561. [Google Scholar] [CrossRef]
  62. Vivoni, E.R.; Rodríguez, J.C.; Watts, C.J. On the spatiotemporal variability of soil moisture and evapotranspiration in a mountainous basin within the north American monsoon region. Water Resour. Res. 2010, 46, w02509. [Google Scholar] [CrossRef]
  63. Han, Q.; Wang, T.; Wang, L.; Smettem, K.; Mai, M.; Chen, X. Comparison of nighttime with daytime evapotranspiration responses to environmental controls across temporal scales along a climate gradient. Water Resour. Res. 2021, 57, e2021WR029638. [Google Scholar] [CrossRef]
  64. Amatya, D.M.; Irmak, S.; Gowda, P.; Sun, G.; Nettles, J.E.; Douglas-Mankin, K.R. Ecosystem evapotranspiration: Challenges in measurements, estimates, and modeling. Trans. ASABE 2016, 59, 555–560. [Google Scholar]
  65. Fisher, J.B.; Melton, F.; Middleton, E.; Hain, C.; Anderson, M.; Allen, R.; McCabe, M.F.; Hook, S.; Baldocchi, D.; Townsend, P.A.; et al. The future of evapotranspiration: Global requirements for ecosystem functioning, carbon and climate feedbacks, agricultural management, and water resources. Water Resour. Res. 2017, 53, 2618–2626. [Google Scholar] [CrossRef]
  66. Liu, Z.; Cheng, L.; Zhou, G.; Chen, X.; Lin, K.; Zhang, W.F.; Chen, X.Z.; Zhou, P. Global response of evapotranspiration ratio to climate conditions and watershed characteristics in a changing environment. J. Geophys. Res.-Atmos. 2020, 125, e2020JD032371. [Google Scholar] [CrossRef]
  67. Chang, L.L.; Dwivedi, R.; Knowles, J.F.; Fang, Y.H.; Niu, G.Y.; Pelletier, J.D.; Rasmussen, C.; Durcik, M.; Barron-Gafford, G.A.; Meixner, T. Why do large-scale land surface models produce a low ratio of transpiration to evapotranspiration? J. Geophys. Res.-Atmos. 2018, 123, 9109–9130. [Google Scholar] [CrossRef]
  68. Sun, X.M.; Wilcox, B.P.; Zou, C.B. Evapotranspiration partitioning in dryland ecosystems: A global meta-analysis of in situ studies. J. Hydrol. 2019, 576, 123–136. [Google Scholar] [CrossRef]
  69. Zhu, S.H.; Zhang, C.; Fang, X.; Cao, L.Z. Interactive and individual effects of multifactor controls on water use efficiency in Central Asian ecosystems. Environ. Res. Lett. 2020, 15, 8. [Google Scholar] [CrossRef]
  70. Gao, G.Y.; Wang, D.; Zha, T.S.; Wang, L.X.; Fu, B.J. A global synthesis of transpiration rate and evapotranspiration partitioning in the shrub ecosystems. J. Hydrol. 2022, 606, 127417. [Google Scholar] [CrossRef]
  71. Zhu, S.H.; Fang, X.; Cao, L.Z.; Hang, X.; Xie, X.P.; Sun, L.X.; Li, Y.C. Multivariate drives and their interactive effects on the ratio of transpiration to evapotranspiration over Central Asia ecosystems. Ecol. Model. 2023, 478, 110294. [Google Scholar] [CrossRef]
  72. Bazie, H.R.; Sanou, J.; Bayala, J.; Bargues-Tobella, A.; Zombre, G.; Ilstedt, U. Temporal variations in transpiration of Vitellaria paradoxa in West African agroforestry parklands. Agrofor. Syst. 2018, 92, 1673–1686. [Google Scholar] [CrossRef]
  73. Iqbal, S.; Zha, T.S.; Jia, X.; Hayat, M.; Qian, D.; Bourque, C.P.A.; Tian, Y.; Bai, Y.J.; Liu, P.; Yang, R.Z.; et al. Interannual variation in sap flow response in three xeric shrub species to periodic drought. Agric. For. Meteorol. 2021, 297, 108276. [Google Scholar] [CrossRef]
  74. Zhang, R.F.; Xu, X.L.; Liu, M.X.; Zhang, Y.H.; Xu, C.H.; Yi, R.Z.; Luo, W. Comparing ET-VPD hysteresis in three agroforestry ecosystems in a subtropical humid karst area. Agric. Water Manag. 2018, 208, 454–464. [Google Scholar] [CrossRef]
  75. Zhang, X.; Yu, X.X.; Ding, B.B.; Liu, Z.H.; Jia, G.D. Water storage and use by Platycladus orientalis under different rainfall conditions in the rocky mountainous area of Northern China. Forests 2022, 13, 1761. [Google Scholar] [CrossRef]
  76. Hayat, M.; Xiang, J.; Yan, C.H.; Xiong, B.W.; Wang, B.; Qin, L.J.; Saeed, S.; Hussain, S.; Zou, Z.D.; Qiu, G.Y. Environmental control on transpiration and its cooling effect of Ficus concinna in a subtropical city Shenzhen, southern China. Agric. For. Meteorol. 2022, 312, 108715. [Google Scholar] [CrossRef]
  77. Tognetti, R.; Giovannelli, A.; Lavini, A.; Morelli, G.; Fragnito, F.; d’Andria, R. Assessing environmental controls over conductances through the soil–plant–atmosphere continuum in an experimental olive tree plantation of southern Italy. Agric. For. Meteorol. 2009, 149, 1229–1243. [Google Scholar] [CrossRef]
  78. Rahman, M.A.; Moser, A.; Rötzer, T.; Pauleit, S. Microclimatic differences and their influence on transpirational cooling of Tilia cordata in two contrasting street canyons in Munich, Germany. Agric. For. Meteorol. 2017, 232, 443–456. [Google Scholar] [CrossRef]
  79. Hayat, M.; Zha, T.; Jia, X.; Iqbal, S.; Qian, D.; Bourque, C.P.A.; Yang, R. A multiple-temporal scale analysis of biophysical control of sap flow in Salix psammophila growing in a semiarid shrubland ecosystem of northwest China. Agric. For. Meteorol. 2020, 288, 107985. [Google Scholar] [CrossRef]
  80. Hayat, M.; Iqbal, S.; Zha, T.; Jia, X.; Qian, D.; Bourque, C.P.A.; Yang, R. Biophysical control on nighttime sap flow in Salix psammophila in a semiarid shrubland ecosystem. Agric. For. Meteorol. 2021, 300, 108329. [Google Scholar] [CrossRef]
  81. Chen, X.; Zhao, P.; Hu, Y.; Ouyang, L.; Zhu, L.; Ni, G. Canopy transpiration and its cooling effect of three urban tree species in a subtropical city-Guangzhou, China. Urban For. Urban Green. 2019, 43, 126368. [Google Scholar] [CrossRef]
  82. Hiraoka, H. An investigation of the effect of environmental factors on the budgets of heat, water vapor, and carbon dioxide within a tree. Energy 2005, 30, 281–298. [Google Scholar] [CrossRef]
  83. Zhu, L.W.; Zhao, P. Temporal variation in sap-flux-scaled transpiration and cooling effect of a subtropical Schima superba plantation in the urban area of Guangzhou. J. Integr. Agr. 2013, 12, 1350–1356. [Google Scholar] [CrossRef]
  84. Zha, T.; Qian, D.; Jia, X.; Bai, Y.; Tian, Y.; Bourque, C.P.A.; Peltola, H. Soil moisture control of sap-flow response to biophysical factors in a desert-shrub species, Artemisia ordosica. Biogeosciences 2017, 14, 4533–4544. [Google Scholar] [CrossRef]
  85. Rosa Ferraz Jardim, A.M.; Morais, J.E.F.; Souza, L.S.B.; Souza, C.A.A.; Nascimento Araújo Júnior, G.; Alves, C.P.; Silva, G.Í.N.; Renan, M.C.L.; Moura, M.S.B.; Lima, J.L.M.P.; et al. Monitoring energy balance, turbulent flux partitioning, evapotranspiration and biophysical parameters of Nopalea cochenillifera (Cactaceae) in the Brazilian semi-arid environment. Plants 2023, 12, 2562. [Google Scholar] [CrossRef] [PubMed]
  86. Li, X.; Gentine, P.; Lin, C.J.; Zhou, S.; Sun, Z.; Zheng, Y.; Liu, J.; Zheng, C.M. A simple and objective method to partition evapotranspiration into transpiration and evaporation at eddy-covariance sites. Agric. For. Meteorol. 2019, 265, 171–182. [Google Scholar] [CrossRef]
  87. Wang, F.; Liang, W.; Fu, B.; Jin, Z.; Yan, J.; Zhang, W.; Fu, S.; Yan, N. Changes of cropland evapotranspiration and its driving factors on the loess plateau of China. Sci. Total Environ. 2020, 728, 138582. [Google Scholar] [CrossRef]
  88. Zhang, J.W.; Guan, K.Y.; Peng, B.; Pan, M.; Zhou, W.; Jiang, C.Y.; Kimm, H.; Franz, T.E.; Grant, R.F.; Yang, Y.; et al. Sustainable irrigation based on co-regulation of soil water supply and atmospheric evaporative demand. Nat. Commun. 2021, 12, 5549. [Google Scholar] [CrossRef]
  89. Yang, L.S.; Feng, Q.; Adamowski, J.F.; Alizadeh, M.R.; Yin, Z.L.; Wen, X.H.; Zhu, M. The role of climate change and vegetation greening on the variation of terrestrial evapotranspiration in northwest China’s Qilian Mountains. Sci. Total Environ. 2021, 759, 143532. [Google Scholar] [CrossRef]
  90. Wan, Y.F.; Yu, P.T.; Li, X.Q.; Wang, Y.H.; Wang, B.; Yu, Y.P.; Zhang, L.; Liu, X.D.; Wang, S.L. Seasonal pattern of stem diameter growth of Qinghai spruce in the Qilian mountains, northwestern China. Forests 2020, 11, 494. [Google Scholar] [CrossRef]
  91. Song, L.N.; Zhu, J.J.; Zhang, X.; Wang, K.; Lü, L.Y.; Zhang, X.L.; Hao, G.Y. Transpiration and canopy conductance dynamics of Pinus sylvestris var. mongolica in its natural range and in an introduced region in the sandy plains of Northern China. Agric. For. Meteorol. 2020, 281, 107830. [Google Scholar] [CrossRef]
  92. Gao, G.L.; Feng, Q.; Liu, X.D.; Yu, T.F.; Wang, R.X. The photosynthesis of Populus euphratica Oliv. is not limited by drought stress in the hyper-arid zone of Northwest China. Forests 2022, 13, 2096. [Google Scholar] [CrossRef]
  93. Clearwater, M.J.; Meinzer, F.C.; Andrade, J.L.; Goldstein, G.; Holbrook, N.M. Potential errors in measurement of non-uniform sap flow using heat dissipation probes. Tree Physiol. 1999, 19, 681–687. [Google Scholar] [CrossRef]
  94. Vickers, D.; Mahrt, L. Quality control and flux sampling problems for tower and aircraft data. J. Atmos. Ocean. Technol. 1997, 14, 512–526. [Google Scholar] [CrossRef]
  95. Falge, E.; Baldocchi, D.; Olson, R.; Anthoni, P.; Aubinet, M.; Bernhofer, C.; Burba, G.; Ceulemans, R.; Clement, R.; Dolman, H.; et al. Gap filling strategies for defensible annual sums of net ecosystem exchange. Agric. For. Meteorol. 2001, 107, 43–69. [Google Scholar] [CrossRef]
  96. Gao, G.L.; Zhang, X.Y.; Yu, T.F. Evapotranspiration of a Populus euphratica forest during the growing season in an extremely arid region of northwest China using the Shuttleworth–Wallace model. J. For. Res. 2016, 27, 879–887. [Google Scholar] [CrossRef]
  97. Elith, J.; Leathwick, J.R.; Hastie, T. A working guide to boosted regression trees. J. Anim. Ecol. 2008, 77, 802–813. [Google Scholar] [CrossRef]
  98. Li, C.; Li, Z.Z.; Zhang, F.M.; Lu, Y.Y.; Duan, C.F.; Xu, Y. Seasonal dynamics of carbon dioxide and water fluxes in a rice-wheat rotation system in the Yangtze-Huaihe region of China. Agric. Water Manag. 2023, 275, 107992–108000. [Google Scholar] [CrossRef]
  99. Sutanto, S.J.; van den Hurk, B.; Dirmeyer, P.A.; Seneviratne, S.I.; Rckmann, T.; Trenberth, K.E.; Blyth, E.M.; Wenninger, J.; Hoffmann, G. HESS opinions: A perspective on isotope versus non-isotope approaches to determine the contribution of transpiration to total evaporation. Hydrol. Earth Syst. Sci. 2014, 18, 2815–2827. [Google Scholar] [CrossRef]
  100. Cao, R.C.; Huang, H.; Wu, G.N.; Han, D.R.; Jiang, Z.Y.; Di, K.; Hu, Z.M. Spatiotemporal variations in the ratio of transpiration to evapotranspiration and its controlling factors across terrestrial biomes. Agric. For. Meteorol. 2022, 321, 108984. [Google Scholar] [CrossRef]
  101. Gao, X.; Mei, X.R.; Zhang, J.S.; Cai, J.F.; Gu, F.X.; Hao, W.P.; Gong, D.Z. Comparison of three modified models in evapotranspiration and its components over a rainfed spring maize cropland on the Loess Plateau, China. Agric. For. Meteorol. 2023, 330, 109322. [Google Scholar] [CrossRef]
  102. Meng, W.; Sun, X.H.; Ma, J.J.; Guo, X.H.; Lei, T.; Li, R.F. Measurement and simulation of the water storage pit irrigation trees evapotranspiration in the Loess Plateau. Agric. Water Manag. 2019, 226, 105804. [Google Scholar] [CrossRef]
  103. Good, S.P.; Soderberg, K.; Guan, K.; King, E.G.; Scanlon, T.M.; Caylor, K.K. δ2H isotopic flux partitioning of evapotranspiration over a grass field following a water pulse and subsequent dry down. Water Resour. Res. 2014, 50, 1410–1432. [Google Scholar] [CrossRef]
  104. Lu, X.; Liang, L.L.; Wang, L.; Jenerette, G.D.; McCabe, M.F.; Grantz, D.A. Partitioning of evapotranspiration using a stable isotope technique in an arid and high temperature agricultural production system. Agric. Water Manag. 2017, 179, 103–109. [Google Scholar] [CrossRef]
  105. Tarin, T.; Yepez, E.A.; Garatuza-Payan, J.; Rodriguez, J.C.; Méndez-Barroso, L.A.; Watts, C.J.; Vivoni, E.R. Evapotranspiration flux partitioning at a multi-species shrubland with stable isotopes of soil, plant, and atmosphere water pools. Atmosfera 2020, 33, 319–335. [Google Scholar] [CrossRef]
  106. Ma, Y.; Song, X. Applying stable isotopes to determine seasonal variability in evapotranspiration partitioning of winter wheat for optimizing agricultural management practices. Sci. Total Environ. 2019, 654, 633–642. [Google Scholar] [CrossRef] [PubMed]
  107. Tie, Q.; Hu, H.; Tian, F.; Holbrook, N.M. Comparing different methods for determining forest evapotranspiration and its components at multiple temporal scales. Sci. Total Environ. 2018, 633, 12–29. [Google Scholar] [CrossRef] [PubMed]
  108. Zhang, B.Z.; Xu, D.; Liu, Y.; Li, F.; Cai, J.; Du, L. Multi-scale evapotranspiration of summer maize and the controlling meteorological factors in north China. Agric. For. Meteorol. 2016, 216, 1–12. [Google Scholar] [CrossRef]
  109. Zhou, S.; Yu, B.; Zhang, Y.; Huang, Y.; Wang, G. Partitioning evapotranspiration based on the concept of underlying water use efficiency. Water Resour. Res. 2016, 52, 1160–1175. [Google Scholar] [CrossRef]
  110. Scott, R.L.; Biederman, J.A. Partitioning evapotranspiration using long-term carbon dioxide and water vapor fluxes. Geophys. Res. Lett. 2017, 44, 6833–6840. [Google Scholar] [CrossRef]
  111. Scanlon, T.M.; Kustas, W.P. Partitioning carbon dioxide and water vapor fluxes using correlation analysis. Agric. For. Meteorol. 2010, 150, 89–99. [Google Scholar] [CrossRef]
  112. Jiang, S.Z.; Liang, C.; Zhao, L.; Gong, D.Z.; Huang, Y.W.; Xing, L.W.; Zhu, S.D.; Feng, Y.; Guo, L.; Cui, N.B. Energy and evapotranspiration partitioning over a humid region orchard: Field measurements and partitioning model comparisons. J. Hydrol. 2022, 610, 127890. [Google Scholar] [CrossRef]
  113. Wilson, K.B.; Hanson, P.J.; Mulholland, P.J.; Baldocchi, D.D.; Wullschleger, S.D. A comparison of methods for determining forest evapotranspiration and its components: Sap-flow, soil water budget, eddy covariance and catchment water balance. Agric. For. Meteorol. 2001, 106, 153–168. [Google Scholar] [CrossRef]
  114. Lopez-Olivari, R.; Ortega-Farias, S.; Poblete-Echeverria, C. Partitioning of net radiation and evapotranspiration over a superintensive drip-irrigated olive orchard. Irrig. Sci. 2016, 34, 17–31. [Google Scholar] [CrossRef]
  115. Khatun, R.; Baten, M.A.; Hossen, M.S.; Khan, M.B.; Miyata, A. Estimation of Aman rice evapotranspiration using Bowen ratio energy balance method and its comparison with some other methods under Bangladesh condition. Res. J. Recent Sci. 2016, 5, 28–33. [Google Scholar]
  116. Zeggaf, A.T.; Takeuchi, S.; Dehghanisanij, H.; Anyoji, H.; Yano, T. A Bowen ratio technique for partitioning energy fluxes between maize transpiration and soil surface evaporation. Agron. J. 2008, 100, 988–996. [Google Scholar] [CrossRef]
  117. Iida, S.i.; Shimizu, T.; Tamai, K.; Kabeya, N.; Shimizu, A.; Ito, E.; Ohnuki, Y.; Chann, S.; Levia, D.F. Evapotranspiration from the understory of a tropical dry deciduous forest in Cambodia. Agric. For. Meteorol. 2020, 295, 108170. [Google Scholar] [CrossRef]
  118. Shimizu, T.; Kumagai, T.o.; Kobayashi, M.; Tamai, K.; Iida, S.i.; Kabeya, N.; Ikawa, R.; Tateishi, M.; Miyazawa, Y.; Shimizu, A. Estimation of annual forest evapotranspiration from a coniferous plantation watershed in Japan (2): Comparison of eddy covariance, water budget and sap-flow plus interception loss. J. Hydrol. 2015, 522, 250–264. [Google Scholar] [CrossRef]
  119. Wen, X.; Yang, B.; Sun, X.; Lee, X. Evapotranspiration partitioning through in-situ oxygen isotope measurements in an oasis cropland. Agric. For. Meteorol. 2016, 230–231, 89–96. [Google Scholar] [CrossRef]
  120. Zhang, Y.; Kang, S.; Ward, E.J.; Ding, R.; Zhang, X.; Zheng, R. Evapotranspiration components determined by sap flow and microlysimetry techniques of a vineyard in northwest China: Dynamics and influential factors. Agric. Water Manag. 2011, 98, 1207–1214. [Google Scholar] [CrossRef]
  121. Ma, L.; Li, Y.; Wu, P.; Zhao, X.; Chen, X.; Gao, X. Coupling evapotranspiration partitioning with water migration to identify the water consumption characteristics of wheat and maize in an intercropping system. Agric. For. Meteorol. 2020, 290, 108034. [Google Scholar] [CrossRef]
  122. Scott, R.L.; Knowles, J.F.; Nelson, J.A.; Gentine, P.; Li, X.; Barron-Gafford, G.; Bryant, R.; Biederman, J.A. Water availability impacts on evapotranspiration partitioning. Agric. For. Meteorol. 2021, 297, 108251. [Google Scholar] [CrossRef]
  123. Wang, Y.Y.; Horton, R.; Xue, X.; Ren, T. Partitioning evapotranspiration by measuring soil water evaporation with heat-pulse sensors and plant transpiration with sap flow gauges. Agric. Water Manag. 2021, 252, 106883. [Google Scholar] [CrossRef]
  124. Kang, E.S.; Cheng, G.D.; Song, K.C.; Jin, B.W.; Liu, X.D.; Wang, J.Y. Simulation of energy and water balance in Soil-Vegetation-Atmosphere Transfer system in the mountain area of Heihe River Basin at Hexi Corridor of northwest China. Sci. China Earth Sci. 2005, 48, 538–548. [Google Scholar] [CrossRef]
  125. Fatichi, S.; Pappas, C. Constrained variability of modeled T:ET ratio across biomes. Geophys. Res. Lett. 2017, 44, 6795–6803. [Google Scholar] [CrossRef]
  126. Wilcox, B.P.; Breshears, D.D.; Seyfried, M.S. Rangelands, water balance on. In Encyclopedia of Water Science; Stewart, B.A., Howell, T.A., Eds.; Marcel Dekker: New York, NY, USA, 2003; pp. 791–794. [Google Scholar]
  127. Tian, X.F.; Zhao, C.Y.; Feng, Z.D. Simulating evapotranspiration of Qinghai spruce (Picea crassifolia) forest in the Qilian Mountains, northwestern China. J. Arid Environ. 2011, 75, 648–655. [Google Scholar] [CrossRef]
  128. Wang, P.; Grinevsky, S.O.; Pozdniakov, S.P.; Yu, J.J.; Dautova, D.S.; Min, L.L.; Du, C.Y.; Zhang, Y.C. Application of the water table fluctuation method for estimating evapotranspiration at two phreatophyte-dominated sites under hyper-arid environments. J. Hydrol. 2014, 519, 2289–2300. [Google Scholar] [CrossRef]
  129. Lang, P.; Ahlborn, J.; Schäfer, P.; Wommelsdorf, T.; Jeschke, M.; Zhang, X.M.; Thomas, F.M. Growth and water use of Populus euphratica trees and stands with different water supply along the Tarim River, NW China. For. Ecol. Manag. 2016, 380, 139–148. [Google Scholar] [CrossRef]
  130. Si, J.H.; Feng, Q.; Zhang, X.Y.; Chang, Z.Q.; Su, Y.H.; Xi, H.Y. Sap flow of Populus euphratica in a desert riparian forest in an extreme arid region during the growing season. J. Integr. Plant Biol. 2007, 49, 425–436. [Google Scholar] [CrossRef]
  131. Peng, W.L.; Zhao, L.J.; Xie, C.; Huang, X.Y.; Liu, Q.Y.; Li, R.F.; Pan, Z.Y. Evapotranspiration partitioning of the Picea crassifolia forest ecosystem in the upper reaches of the Heihe River. J. Glaciol. Geocryol. 2020, 42, 629–640. (In Chinese) [Google Scholar]
  132. Wei, H.Q.; He, H.L.; Liu, M.; Zhang, L.; Yu, G.R.; Min, C.C.; Wang, H.M.; Liu, Y. Modeling evapotranspiration and its components in Qianyanzhou Planation based on remote sensing data. J. Nat. Resour. 2012, 27, 778–789. (In Chinese) [Google Scholar]
  133. Zhu, X.J.; Yu, G.R.; Hu, Z.M.; Wang, Q.F.; He, H.L.; Yan, J.H.; Wang, H.M.; Zhang, J.H. Spatiotemporal variations of T/ET (the ratio of transpiration to evapotranspiration) in three forests of Eastern China. Ecol. Indic. 2015, 52, 411–421. [Google Scholar] [CrossRef]
  134. Ren, X.L.; He, H.L.; Moore, D.J.P.; Zhang, L.; Liu, M.; Li, F.; Yu, G.R.; Wang, H.M. Uncertainty analysis of modeled carbon and water fluxes in a subtropical coniferous plantation. J. Geophys. Res. 2013, 118, 1674–1688. [Google Scholar] [CrossRef]
  135. Yang, J.J.; He, Z.B.; Feng, J.M.; Lin, P.F.; Du, J.; Guo, L.X.; Liu, Y.F.; Yan, J.L. Rainfall interception measurements and modeling in a semiarid evergreen spruce (Picea crassifolia) forest. Agric. For. Meteorol. 2022, 328, 109257. [Google Scholar] [CrossRef]
  136. Peng, H.H.; Zhao, C.Y.; Shen, W.H.; Xu, Z.L.; Feng, Z.D. Modeling rainfall canopy interception of Picea crassifolia forest in Qilian mountains: A case of Pailugou catchent. Arid Reg. Geogr. 2010, 33, 600–606. (In Chinese) [Google Scholar]
  137. Yuan, R.Q.; Chang, L.L.; Niu, G.Y. Annual variations of T/ET in a semi-arid region: Implications of plant water use strategies. J. Hydrol. 2021, 603, 126884. [Google Scholar] [CrossRef]
  138. Zhang, R.; Seke, K.; Wang, L. Quantifying the contribution of meteorological factors and plant traits to canopy interception under maize cropland. Agric. Water Manag. 2023, 279, 108195. [Google Scholar] [CrossRef]
  139. Wang, D.; Wang, L. Canopy interception of apple orchards should not be ignored when assessing evapotranspiration partitioning on the Loess Plateau in China. Hydrol. Process. 2019, 33, 372–382. [Google Scholar] [CrossRef]
  140. Yan, T.; Wang, Z.H.; Liao, C.G.; Xu, W.Y.; Wan, L. Effects of the morphological characteristics of plants on rainfall interception and kinetic energy. J. Hydrol. 2020, 592, 125807. [Google Scholar] [CrossRef]
  141. Kozak, J.A.; Ahuja, L.R.; Green, T.R.; Ma, L.W. Modelling crop canopy and residue rainfall interception effects on soil hydrological components for semi-arid agriculture. Hydrol. Process. 2007, 21, 229–241. [Google Scholar] [CrossRef]
  142. Zhang, Z.S.; Zhao, Y.; Li, X.R.; Huang, L.; Tan, H.J. Gross rainfall amount and maximum rainfall intensity in 60-minute influence on interception loss of shrubs: A 10-year observation in the Tengger Desert. Sci. Rep. 2016, 6, 26030. [Google Scholar] [CrossRef] [PubMed]
  143. Grunicke, S.; Queck, R.; Bernhofer, C. Long-term investigation of forest canopy rainfall interception for a spruce stand. Agric. For. Meteorol. 2020, 292–293, 108125. [Google Scholar] [CrossRef]
  144. Dong, X.H. Simulation on Hydrological Effects of Forest Vegetation in Drainage Gully Watershed of Qilian. Ph.D. Dissertation, Chinese Academy of Forestry, Beijing, China, 2007. (In Chinese). [Google Scholar]
  145. Wang, L. Study on Water Consumption Characteristics of Typical Forests in Loess Alpine Region, in Datong County, Qinghai Province. Master’s Dissertation, Beijing Forestry University, Beijing, China, 2020. (In Chinese). [Google Scholar]
  146. Lu, Q.Q.; He, H.L.; Zhu, X.J.; Yu, G.R.; Wang, H.M.; Zhang, J.H.; Yan, J.H. Study on the variations of forest evapotranspiration and its components in eastern China. J. Nat. Resour. 2015, 30, 1436–1448. (In Chinese) [Google Scholar]
  147. Cao, G.X.; Wang, Y.N.; Wang, Y.H.; Xu, L.H.; Ji, M. The characteristics of evapotranspiration and its components of three coniferous forests in Liupan mountains of Ningxia. Sci. Technol. Eng. 2018, 18, 16–22. (In Chinese) [Google Scholar]
  148. Han, J.J.; Tian, L.D.; Cai, Z.Y.; Ren, W.; Liu, W.W.; Li, J.; Tai, J.R. Season-specific evapotranspiration partitioning using dual water isotopes in a Pinus yunnanensis ecosystem, southwest China. J. Hydrol. 2022, 608, 127672. [Google Scholar] [CrossRef]
  149. Sun, J.Y.; Sun, X.Y.; Hu, Z.Y.; Wang, G.X. Exploring the influence of environmental factors in partitioning evapotranspiration along an elevation gradient on mount Gongga, eastern edge of the Qinghai-Tibet Platea, China. J. Mt. Sci. 2020, 17, 384–396. [Google Scholar] [CrossRef]
  150. Wang, Y.F.; Zou, Y.F.; Cai, H.J.; Zeng, Y.J.; He, J.Q.; Yu, L.Y.; Zhang, C.; Saddique, Q.; Peng, X.B.; Siddique, K.H.M.; et al. Seasonal variation and controlling factors of evapotranspiration over dry semi-humid cropland in Guanzhong Plain, China. Agric. Water Manag. 2022, 259, 107242. [Google Scholar] [CrossRef]
  151. Pei, W.W.; Wang, X.; Wang, Y.Y.; Du, Y.G. Water and heat flux characteristics of Qinghai spruce forests in the Qilian mountains during growing season. J. Arid Land Resour. Environ. 2022, 36, 144–150. (In Chinese) [Google Scholar]
  152. Yang, W.J. Spatial Distribution, Structural Characteristics and Evapotranspiration of Qinghai Spruce Forests in the Qilian Mountains, Northwest China. Ph.D. Dissertation, Chinese Academy of Forestry, Beijing, China, 2018. [Google Scholar]
  153. Zhu, G.F.; Lu, L.; Su, Y.H.; Wang, X.F.; Cui, X.; Ma, J.Z.; He, J.H.; Zhang, K.; Li, C.B. Energy flux partitioning and evapotranspiration in a sub-alpine spruce forest ecosystem. Hydrol. Process. 2014, 28, 5093–5104. [Google Scholar]
  154. Ma, J.Y. Water Vapour, Heat Fluxes and Water Use Efficiency over a Young Pinus tabuliformis plantation in Beijing. Ph.D. Dissertation, Beijing Forestry University, Beijing, China, 2019. [Google Scholar]
  155. Nakai, T.; Kim, Y.; Busey, R.C.; Suzuki, R.; Nagai, S.; Kobayashi, H.; Park, H.; Sugiura, K.; Ito, A. Characteristics of evapotranspiration from a permafrost black spruce forest in interior Alaska. Polar Sci. 2013, 7, 136–148. [Google Scholar] [CrossRef]
  156. Iwata, H.; Harazono, Y.; Ueyama, M. The role of permafrost in water exchange of a black spruce forest in interior Alaska. Agric. For. Meteorol. 2012, 161, 107–115. [Google Scholar] [CrossRef]
  157. Zhu, Y.; Cheng, Z.F.; Feng, K.; Chen, Z.; Cao, C.X.; Huang, J.; Ye, H.; Gao, Y.F. Influencing factors for transpiration rate: A numerical simulation of an individual leaf system. Therm. Sci. Eng. Prog. 2022, 27, 101110. [Google Scholar] [CrossRef]
  158. Du, M.G. Study on the Characteristics of Transpiration and Water Consumption in Spruce Forest of Qilian Mountains. Master’s Thesis, Northwest A&F University, Yangling, China, 2019. [Google Scholar]
  159. Chen, S.N.; Chen, Z.S.N.; Feng, Z.Y.; Kong, Z.; Xu, H.; Zhang, Z.Q. Species difference of transpiration in three urban coniferous forests in a semiarid region of China. J. Hydrol. 2023, 617, 129098. [Google Scholar] [CrossRef]
  160. Chen, S.N.; Zhang, Z.Q.; Chen, Z.S.N.; Xu, H.; Li, J.L. Responses of canopy transpiration and conductance to different drought levels in Mongolian pine plantations in a semiarid urban environment of China. Agric. For. Meteorol. 2024, 347, 109897. [Google Scholar] [CrossRef]
  161. Wang, J.; Niu, Y.; Jing, W.M.; Ma, J. Comparative study on woodland and meadow of meteorological factors, soil properties and its evaporation in composite basin of Qilian mountains. J. Cent. South Univ. For. Technol. 2014, 34, 90–94. (In Chinese) [Google Scholar]
Figure 1. Location of the study area.
Figure 1. Location of the study area.
Plants 13 00801 g001
Figure 2. Diurnal variations in environmental factors, including (a) net radiation (Rn) and photosynthetic active radiation (PAR), (b) air temperature (Ta) and vapor pressure deficit (VPD), (c) soil temperature (Ts) at different depths, (d) precipitation (Pa) and throughfall (Pt), (e) soil water content (SWC) at different depths, and (f) relative humidity (RH) and wind speed (u), in 2019 and 2020 at the study site.
Figure 2. Diurnal variations in environmental factors, including (a) net radiation (Rn) and photosynthetic active radiation (PAR), (b) air temperature (Ta) and vapor pressure deficit (VPD), (c) soil temperature (Ts) at different depths, (d) precipitation (Pa) and throughfall (Pt), (e) soil water content (SWC) at different depths, and (f) relative humidity (RH) and wind speed (u), in 2019 and 2020 at the study site.
Plants 13 00801 g002
Figure 3. The dynamics of monthly mean diurnal variations in (a) evapotranspiration (ET) and (b) transpiration (T) of the Qinghai spruce in the Qilian mountains in 2019 and 2020.
Figure 3. The dynamics of monthly mean diurnal variations in (a) evapotranspiration (ET) and (b) transpiration (T) of the Qinghai spruce in the Qilian mountains in 2019 and 2020.
Plants 13 00801 g003
Figure 4. The monthly dynamics of daily variations in (a) evapotranspiration (ET), (b) transpiration (T) and (c) evaporation (Es) during the main growing season of the Qinghai spruce in the Qilian mountains in 2019 and 2020.
Figure 4. The monthly dynamics of daily variations in (a) evapotranspiration (ET), (b) transpiration (T) and (c) evaporation (Es) during the main growing season of the Qinghai spruce in the Qilian mountains in 2019 and 2020.
Plants 13 00801 g004
Figure 5. Monthly cumulative values of (a) evapotranspiration (ET), (b) transpiration (T) and (c) evaporation (Es) for the Qinghai spruce in the Qilian mountains during the study periods in 2019 and 2020.
Figure 5. Monthly cumulative values of (a) evapotranspiration (ET), (b) transpiration (T) and (c) evaporation (Es) for the Qinghai spruce in the Qilian mountains during the study periods in 2019 and 2020.
Plants 13 00801 g005
Figure 6. Monthly dynamics of daily variations in the proportions of (a) transpiration to evapotranspiration (T/ET) and (b) evaporation to evapotranspiration (Es/ET) of the Qinghai spruce in the Qilian mountains in 2019 and 2020.
Figure 6. Monthly dynamics of daily variations in the proportions of (a) transpiration to evapotranspiration (T/ET) and (b) evaporation to evapotranspiration (Es/ET) of the Qinghai spruce in the Qilian mountains in 2019 and 2020.
Plants 13 00801 g006
Figure 7. Proportions of transpiration to evapotranspiration (T/ET), evaporation to evapotranspiration (Es/ET) and canopy interception to evapotranspiration (Ei/ET) of the Qinghai spruce in each month in 2019 and 2020.
Figure 7. Proportions of transpiration to evapotranspiration (T/ET), evaporation to evapotranspiration (Es/ET) and canopy interception to evapotranspiration (Ei/ET) of the Qinghai spruce in each month in 2019 and 2020.
Plants 13 00801 g007
Figure 8. Proportions of transpiration to evapotranspiration (T/ET), evaporation to evapotranspiration (Es/ET) and canopy interception to evapotranspiration (Ei/ET) of the Qinghai spruce during the study periods in 2019 and 2020.
Figure 8. Proportions of transpiration to evapotranspiration (T/ET), evaporation to evapotranspiration (Es/ET) and canopy interception to evapotranspiration (Ei/ET) of the Qinghai spruce during the study periods in 2019 and 2020.
Plants 13 00801 g008
Figure 9. Pearson’s correlation analysis between (a) evapotranspiration (ET), (b) transpiration (T), (c) evaporation (Es) and the environmental factors, including air temperature (Ta), relative humidity (RH), vapor pressure deficit (VPD), net radiation (Rn), photosynthetic active radiation (PAR), wind speed (u), precipitation (Pa) and throughfall (Pt). ** means significant correlation.
Figure 9. Pearson’s correlation analysis between (a) evapotranspiration (ET), (b) transpiration (T), (c) evaporation (Es) and the environmental factors, including air temperature (Ta), relative humidity (RH), vapor pressure deficit (VPD), net radiation (Rn), photosynthetic active radiation (PAR), wind speed (u), precipitation (Pa) and throughfall (Pt). ** means significant correlation.
Plants 13 00801 g009
Figure 10. Contributions of environmental factors to (a) evapotranspiration (ET), (b) transpiration (T) and (c) evaporation (Es) based on the boosted regression trees method. The environmental factors included net radiation (Rn); soil temperature at depths of 30 cm (Ts30) and 80 cm (Ts80); soil water content at depths of 10 cm (SWC10), 40 cm (SWC40), 60 cm (SWC60) and 80 cm (SWC80); vapor pressure deficit (VPD); relative humidity (RH); wind speed (u); precipitation (Pa); and throughfall (Pt).
Figure 10. Contributions of environmental factors to (a) evapotranspiration (ET), (b) transpiration (T) and (c) evaporation (Es) based on the boosted regression trees method. The environmental factors included net radiation (Rn); soil temperature at depths of 30 cm (Ts30) and 80 cm (Ts80); soil water content at depths of 10 cm (SWC10), 40 cm (SWC40), 60 cm (SWC60) and 80 cm (SWC80); vapor pressure deficit (VPD); relative humidity (RH); wind speed (u); precipitation (Pa); and throughfall (Pt).
Plants 13 00801 g010
Table 1. Summary of biological parameters for the three selected Qinghai spruce trees.
Table 1. Summary of biological parameters for the three selected Qinghai spruce trees.
Tree NumberHeight (m)DBH (cm)As (cm2)
111.014.9111.1
213.521.0186.5
315.537.4525.7
Sig. (two-tailed)0.3310.7360.912
Significant differences were tested by the Student’s t-test at a significance level of p = 0.05. DBH, diameter at breast height; As, sapwood area.
Table 2. Summary of transpiration (T), evaporation (Es) and canopy interception (Ei) on evapotranspiration (ET) of Qinghai spruce and other coniferous forests.
Table 2. Summary of transpiration (T), evaporation (Es) and canopy interception (Ei) on evapotranspiration (ET) of Qinghai spruce and other coniferous forests.
Tree SpeciesCoordinatesAltitude (m)Study PeriodsT/ETEs/ETEi/ETReferences
Qinghai spruce38°31′–38°33′ N
100°16′–100°18′ E
27002019–202053.6%
57.1%
32.9%
31.6%
13.5%
11.3%
This study
Qinghai spruce38°32′ N, 100°15′ E2835200851.3%16.5%32.2%Tian et al. [127]
Qinghai spruce38°31′–38°33′ N
100°17′–100°18′ E
2700
2900
201152.2–88.4%11.6–47.8% Peng et al. [131]
Qinghai spruce 26052001–200260.9%1.1%38.0%Dong [144]
Qinghai spruce36°43′–37°23′ N
100°51′–101°56′ E
2519201982.6%17.4% Wang [145]
Coniferous and broad-leaved mixed forest42°24′ N, 128°05′ E7382003–200865.7%19.3%15.0%Lu et al. [146]
Evergreen coniferous forest26°44′ N, 115°03′ E1022003–200861.4%10.8%27.8%
Artificial coniferous forests26°44’ N, 115°03’ E110.82003–200865.0%23.0%12.0%Wei et al. [132]
Larch forest35°15′–35°41′ N
106°09′–106°30′ E
2264201241.6%27.2%31.19%Cao et al. [147]
Pine forest35°15′–35°41′ N
106°09′–106°30′ E
2264201247.2%17.2%35.5%
Pinus tabuliformis forest35°15′–35°41′ N
106°09′–106°30′ E
2264201242.4%27.1%30.5%
Pinus yunnanensis forest27°00′ N, 100°10′ E3250201959.0–81.0% Han et al. [148]
Coniferous broad-leaved mixed forest29°20′–30°20′ N
101°30′–102°15′ E
7556201648.0% Sun et al. [149]
Evergreen coniferous forest29°20′–30°20′ N
101°30′–102°15′ E
7556201650.0%
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gao, G.; Guo, X.; Feng, Q.; Xu, E.; Hao, Y.; Wang, R.; Jing, W.; Ren, X.; Liu, S.; Shi, J.; et al. Environmental Controls on Evapotranspiration and Its Components in a Qinghai Spruce Forest in the Qilian Mountains. Plants 2024, 13, 801. https://doi.org/10.3390/plants13060801

AMA Style

Gao G, Guo X, Feng Q, Xu E, Hao Y, Wang R, Jing W, Ren X, Liu S, Shi J, et al. Environmental Controls on Evapotranspiration and Its Components in a Qinghai Spruce Forest in the Qilian Mountains. Plants. 2024; 13(6):801. https://doi.org/10.3390/plants13060801

Chicago/Turabian Style

Gao, Guanlong, Xiaoyun Guo, Qi Feng, Erwen Xu, Yulian Hao, Rongxin Wang, Wenmao Jing, Xiaofeng Ren, Simin Liu, Junxi Shi, and et al. 2024. "Environmental Controls on Evapotranspiration and Its Components in a Qinghai Spruce Forest in the Qilian Mountains" Plants 13, no. 6: 801. https://doi.org/10.3390/plants13060801

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop