Articles | Volume 17, issue 18
https://doi.org/10.5194/gmd-17-7083-2024
https://doi.org/10.5194/gmd-17-7083-2024
Model description paper
 | 
24 Sep 2024
Model description paper |  | 24 Sep 2024

PyEt v1.3.1: a Python package for the estimation of potential evapotranspiration

Matevž Vremec, Raoul A. Collenteur, and Steffen Birk
Abstract

Evapotranspiration (ET) is a crucial flux of the hydrological water balance, commonly estimated using (semi-)empirical formulas. The estimated flux may strongly depend on the formula used, adding uncertainty to the outcomes of environmental studies using ET. Climate change may cause additional uncertainty, as the ET estimated by each formula may respond differently to changes in meteorological input data. To include the effects of model uncertainty and climate change and facilitate the use of these formulas in a consistent, tested, and reproducible workflow, we present PyEt. PyEt is an open-source Python package for the estimation of daily potential evapotranspiration (PET) using available meteorological data. It allows the application of 20 different PET methods on both time series and gridded datasets. The majority of the implemented methods are benchmarked against literature values and tested with continuous integration to ensure the correctness of the implementation. This article provides an overview of PyEt's capabilities, including the estimation of PET with 20 PET methods for station and gridded data, a simple procedure for calibrating the empirical coefficients in the alternative PET methods, and estimation of PET under warming and elevated atmospheric CO2 concentration. Further discussion on the advantages of using PyEt estimates as input for hydrological models, sensitivity and uncertainty analyses, and hindcasting and forecasting studies (especially in data-scarce regions) is provided.

1 Introduction

Evaporation – the process by which water is converted from its liquid to vapor phase – is a central component of the global hydrological cycle (Katul et al.2012). Evaporation has far-reaching impacts on both human societies and ecosystems (Oki and Kanae2006; Fisher et al.2011). In the remainder of this paper, the term evapotranspiration (ET) is used to refer to the total evaporation flux from soil and waterbodies (evaporation) and vegetated surfaces (transpiration; Allen et al.1998; Dingman2015). Information about the magnitude of the ET flux is important across different geoscience disciplines: it assists in predicting irrigation demands and crop water requirements in agriculture, supports efficient water resource management, guides operational strategies in hydropower and meteorological studies, and plays a crucial role in ecological research and climate-change impact assessments. Given that climate change – through warming and elevated CO2 concentrations – is set to alter evapotranspiration, the need for its accurate estimation is of paramount importance, as it affects our understanding and assessment of past, present, and potential future impacts on ecosystem functioning (Milly and Dunne2016; Yang et al.2019; Caretta et al.2022).

Evapotranspiration can hardly be measured directly (Wang and Dickinson2012; Jensen and Allen2016) and is therefore commonly estimated using (semi-)empirical formulas from other, more easily obtained meteorological variables such as temperature, wind speed, and radiation. Over time, dozens of methods have been proposed and applied. Each of these methods generally results in slightly different estimates of evapotranspiration, depending on the methods and data used (Oudin et al.2005; McMahon et al.2013; Xu and Singh2000, 2001; Lemaitre-Basset et al.2022). Most of these formulas estimate either the reference crop evapotranspiration (ET0), which is ET from a reference surface or crop that is not short of water (Allen et al.1998), or the potential evapotranspiration (PET), which is the maximum rate of ET that would occur given a sufficient water supply (Xiang et al.2020). Potential evapotranspiration is determined by meteorological conditions, whereas water availability determines if actual evapotranspiration occurs at its potential rate (Jensen and Allen2016). Differences in the potential evapotranspiration estimate may cascade through a modeling chain and ultimately impact the results of a study. For example, Prudhomme and Williamson (2013), Lemaitre-Basset et al. (2022), and Bormann (2010) showed that the method used affects the results from hydrological climate change impact studies. Similarly, the estimation of water demand for efficient crop and irrigation management depends on potential evapotranspiration, and it may thus be impacted by the methods used (Kumar et al.2012).

To account for the structural uncertainty of the different PET models, it has been recommended to use multiple methods (Seiller and Anctil2016; Beven and Freer2001; Velázquez et al.2013). Such an approach can help improve the understanding of the effect of model uncertainty on PET estimates in, for example, historical climate studies (Zhou et al.2020; Dakhlaoui et al.2020; Yang et al.2019) and climate change impact studies (Bormann2010; Seiller and Anctil2016; Gharbia et al.2018; Shi et al.2020). Climate change impact studies often rely on climate projection data or global observational datasets, which are generally available in a gridded format (i.e., netCDF, GRIB), thus requiring tools that can efficiently process such data. Some of these datasets also contain only a limited set of observed or projected meteorological variables, requiring PET estimation methods that use fewer inputs. It may also be necessary to account for environmental variables that change over time and impact the evapotranspiration, such as vegetation changes and increases in atmospheric CO2 concentrations (Fatichi et al.2016; Ainsworth and Rogers2007; Vremec et al.2022). Studies like those mentioned above require software programs that (1) include multiple PET estimation methods, (2) are flexible in adjusting input parameters (e.g., empirical coefficients, crop data, and meteorological inputs), and (3) are applicable to both time series and gridded data given the spatial nature of many of these studies.

Existing tools for calculating evapotranspiration, such as “Evapotranspiration” in R (Guo et al.2016) and “PyETo” (Richards2019) and “pyfao56” in Python (Thorp2022), are primarily designed for station-based time series data. This limits their applicability with gridded datasets. While Peterson et al. (2020) extended “Evapotranspiration” to “AWAPer” to process gridded data, its use is limited to Australia. For the large community of geoscientists working with Python, the number of available PET methods from existing packages is limited (three for PyETo and one for pyfao56) compared to the 21 methods in the R package. This highlights a gap in the availability of a software for the estimation of multiple PET methods for both time series and gridded data, with the input parameter flexibility required for advanced studies on PET. Given the increasing need to understand and predict environmental changes accurately across the globe, the availability of such software is of paramount importance for the geoscience community.

Opportunities also exist to further align these tools with the FAIR standards of findability, accessibility, interoperability, and reusability for research software, which is of crucial importance for the credibility and reproducibility of scientific studies (Barker et al.2022). This involves improving methodological testing through continuous integration, inclusion of additional alternative PET methods, and enabling more flexibility in adjusting internal empirical coefficients. Such enhancements not only adhere to best practices in software development but also broaden the scope and applicability of these tools in diverse geoscientific contexts. The refinement and development of evapotranspiration estimation tools that fully embrace the FAIR principles are therefore crucial steps toward advancing the field, ensuring more reliable and comprehensive research outcomes in the face of evolving scientific needs (Wood et al.1998; DeJonge and Thorp2017).

In this paper we introduce PyEt, an open-source Python package for the estimation of potential evapotranspiration. The aim of PyEt is to provide researchers and practitioners with a wide variety of tested, documented, and flexible Python functions that support multiple PET methods for both station and gridded data. All methods have a common application programming interface, allowing users to easily test different PET models for their application and, if desired, address structural uncertainty and changing conditions. The majority of the implemented methods are benchmarked against literature values and tested with continuous integration to ensure the correctness of the implementation. Allowing different types of input data, PyEt is also applicable in regions with sparsely distributed measurement stations where standard meteorological data (e.g., wind, relative humidity) are often unavailable. The software is available under MIT license from the Python Package Index (PyPI) (Vremec and Collenteur2022) and developed as a community project on GitHub (http://www.github.com/pyet-org/PyEt, last access: 17 September 2024).

The remainder of this paper is structured as follows. In the next section, the software design, capabilities, and benchmarking tests are described. Section 3 introduces the software through four examples, showing potential future users how to apply PyEt in real-world applications. These examples concentrate on addressing practical problems commonly faced by geoscientists in their daily work. Section 4 discusses future potential applications of PyEt, and we detail how we think it can help the scientific community improve the estimation of potential evapotranspiration. Conclusions and future plans are outlined in Sect. 5.

2 PyEt Python package

2.1 Software design

The basic design principle for PyEt was to build a software that is intuitive and easy to use by novice users with little programming experience yet flexible enough to allow advanced users to perform more complex analyses. The software uses a modular design, with formulas shared by different PET methods implemented as a single function. This reduces the amount of code and makes it easier to maintain the software and implement new methods. All of the PET methods are intended to work with the minimum input data required by the PET models (e.g., radiation, temperature) but also allow more user input if such data are available and allowed by the PET method (e.g., humidity, surface resistance in the Penman–Monteith model). Utility functions are available to the user or are called internally to compute unavailable variables (e.g., solar radiation from latitude value). Moreover, the constants in the empirical PET formulas (e.g., the Stefan–Boltzmann constant) are function arguments with default values from the literature, which may also be changed by the user to adapt the empirical relationship to another region. Finally, the available methods should work for both station (1D) and gridded data (2D/3D).

PyEt is part of the wider Python ecosystem and depends on three widely used and well-developed Python packages from the scientific Python stack: NumPy (Harris et al.2020), Pandas (McKinney201), and Xarray (Hoyer and Hamman2017). The input and output data of PyEt are formatted as time series data in Pandas.Series or Xarray.DataArrays, which allows using all the Pandas or Xarray functions on the data (Harris et al.2020; McKinney201; Hoyer and Hamman2017). These functions include gap-filling and selection functions for interpolation, resampling, clustering, and many more. Being part of a wider ecosystem, users can leverage other Python packages for visualization (e.g., Matplotlib (Hunter2007), MetPy (May et al.2022)) and optimization and uncertainty analyses (SciPy (Virtanen et al.2020), SpotPy (Houska et al.2015)).

The software is hosted and developed on the GitHub platform and distributed under MIT license through the Python Packaging Index (PyPI). Documentation and example applications are available on a dedicated ReadTheDocs website (http://pyet.readthedocs.io, last access: 17 September 2024). The documentation for individual methods is also directly available in Python from the documentation strings. Each release of PyEt is automatically stored in the Zenodo repository and assigned a Digital Object Identifier (DOI). As such, PyEt complies with many of the recommendations for good research software development as given in, for example, Hutton et al. (2016) and the FAIR4RS (FAIR for Research Software) principles (Barker et al.2022). The scripts or the Jupyter notebooks used to apply PyEt improve the reproducibility and provide a transparent report of the entire calculation process (Kluyver et al.2016).

https://gmd.copernicus.org/articles/17/7083/2024/gmd-17-7083-2024-f01

Figure 1Scatter plots showing estimated PET with PyEt against PET values estimated with the R package “Evapotranspiration” from Guo et al. (2016) and literature values from Allen et al. (1998), McMahon et al. (2013), and Schrödter (1985).

Download

2.2 Implemented methods and benchmarking

There are 20 methods currently implemented in PyEt for the estimation of daily potential evapotranspiration. Apart from the Penman–Monteith method, which is considered the standard by the Food and Agriculture Organization (FAO) (Allen et al.1998) and the World Meteorological Organization (WMO), multiple alternative methods are also available in PyEt. An overview of these methods and the required input data is provided in Table 1. Depending on the method, different (amounts of) input data are required to compute the potential evapotranspiration. It is often also possible to provide different input data to the same method (e.g., the average or the minimum and maximum daily temperatures) or even that some input data is optional (as described in the footnotes of Table 1). In the case of optional input data, utility functions are used internally to estimate that data. In the example of the Penman–Monteith method, solar radiation does not necessarily need to be provided by the user and can be estimated from the latitude and actual duration of sunshine hours instead.

Monteith (1965)Walter et al. (2000)Allen et al. (1998)Penman (1948)Priestley and Taylor (1972)Wright (1982)Thom and Oliver (1977)Blaney and others (1952)McMahon et al. (2013)Xu and Singh (2001)Schrödter (1985)Hamon (1963)Oudin et al. (2005)Ansorge and Beran (2019)Romanenko (1961)Xu and Singh (2001)Linacre (1977)Xu and Singh (2001)Haude (1955)Schiff (1975)Turc (1961)Xu and Singh (2000)Jensen and Haise (1963)Oudin et al. (2005)Jensen and Allen (2016)McGuinness and Bordne (1972)Xu and Singh (2000)Hargreaves and Samani (1982)McMahon et al. (2013)Jensen and Allen (2016)Jensen et al. (1990)Abtew (1996)Xu and Singh (2000)Makkink (1957)McMahon et al. (2013)Oudin et al. (2005)

Table 1Data requirements for different PET or ET0 models, the corresponding PyEt function, and if benchmarking of the method was performed. The references include both the original publications of the models and the papers from which the equations were taken: (1) McMahon et al. (2013), (2) Oudin et al. (2005), (3) Xu and Singh (2001), (4) Ansorge and Beran (2019), Rosenberry et al. (2004), (5) Schrödter (1985), (6) Schiff (1975), (7) Jensen and Allen (2016), (8) Xu and Singh (2000).

a The corresponding literature for each method is provided in the Appendix B22.
b Tmax and Tmin can also be provided.
c RHmax and RHmin can also be provided.
d If actual vapor pressure is provided, RH is not needed.
e Input for radiation can be (1) net radiation, (2) solar radiation, or (3) sunshine hours. If it is (1), then latitude is not needed. If it is (1, 3) latitude and elevation are needed.
f One must provide either the atmospheric pressure or elevation.
g The PM method can be used to estimate potential crop evapotranspiration if leaf area index or crop height data are available.
h The effect of CO2 on stomatal resistance can be included using the formulation of Yang et al. (2019).
i If net radiation is provided, RH and latitude are not needed.
j If method 2 is used, u2, RHmin and sunshine hours are required.
k Additional input of Tmax and Tmin or Tdew.
l Input can be RH or actual vapor pressure.
m If method 1 is used, latitude is needed instead of Rs.
n Tmax and Tmin are also needed.

Download Print Version | Download XLSX

The PyEt project is intended to be used by a wide community, and any errors in the code may have consequences for other studies applying PyEt to obtain PET estimates. Special attention was therefore paid to benchmark the available methods to published literature values and data from well-known research and meteorological institutes (Allen et al.1998; McMahon et al.2013; Schrödter1985; Walter et al.2000). These benchmarks are also implemented in the continuous integration and tested using the unittest testing framework (unittest2022). This ensures that the benchmarks are satisfied each time the software is updated in the future. New methods added to PyEt will be required to be accompanied by the appropriate benchmark data and tests. Figure 1 shows the results for each benchmarked method, indicating that the PET estimates from all these methods are equal to the benchmark values (i.e., all values are on the 1 : 1 line). Despite our best efforts, we acknowledge here that four methods have not (yet) been benchmarked due to a lack of appropriate data.

In various sections of the paper, the performance of PET estimation methods are evaluated using three key performance metrics: model bias (mm d−1), the coefficient of determination (–), and the Kling–Gupta Efficiency (KGE, –) (Gupta et al.2009). These metrics enabled comparisons between benchmark PET values from literature and those estimated with PyEt, as well as between PET values derived from alternative models and the Penman–Monteith method, both before and after calibration. The Python implementations of the package SpotPy (Houska et al.2015) were used for each performance metric, and the formulas of these implementations are detailed in Appendix B22.

2.3 Performance

The computational efficiency of various PET methods was assessed by examining the computation times in relation to the time series length and the number of cells in a Xarray dataset. Computation time was evaluated by running all models on a benchmark configuration (with time series of varying lengths) using a 12th Gen Intel Core i7-1255U processor with 10 cores and 12 logical processors. All Xarray.DataArrays cover a period of 1 year, whereas the spatial resolution changes. This comparison highlights the trade-offs between computational complexity and data size but also demonstrates the performance of the methods.

Figure 2 demonstrates that all PyEt methods maintain computational times below 1 s for time series data with lengths ranging from 10 to 10 000 d. For multidimensional data, the computation time does not exceed 10 s for larger Xarray.DataArrays up to 100 000 cells. Moreover, the results in Fig. 2 show that the problem scales well and proportionally does not take more time for larger datasets. Notably, the Penman–Monteith and Priestley–Taylor methods exhibited the largest processing times, whereas methods like Jensen–Haise, Turc, Makkink, and Romanenko were faster. Future improvements will aim to increase this efficiency, in particular to support faster calculations in large-scale global studies.

https://gmd.copernicus.org/articles/17/7083/2024/gmd-17-7083-2024-f02

Figure 2Computational efficiency of different PET methods. This figure shows the comparative processing times of different PET estimation methods regarding the length of the time series (a) and the size of the Xarray data (b).

Download

3 Example use cases

Below, four use cases of PyEt are presented to illustrate how the software can be used. The first example shows how to efficiently compute different potential evapotranspiration estimates using 20 various methods for station data. This example also illustrates how to use PyEt in general. The second example illustrates how to provide 3D estimates of PET using three different methods and gridded Xarray data. The third example shows how to calibrate different PET methods to local conditions and use the calibrated formula for hindcasting. The fourth and final example illustrates a workflow to account for the effects of warming and elevated CO2 in climate change impact studies. The source code for these and other examples can be found in a Zenodo repository related to this paper (Vremec and Collenteur2024a).

3.1 Example 1: estimation of PET from station data

In this example, potential evapotranspiration is estimated for the town of De Bilt in The Netherlands using data provided by the Royal Netherlands Meteorological Institute (KNMI). The reference method used by the KNMI for the estimation of potential evapotranspiration is the Makkink method, which is also implemented in PyEt. The PET computed with the Makkink method is compared to the PET values from all other methods in PyEt. Several steps are taken in a Python script to estimate PET. The code implementing these steps is shown in the code example below. PyEt provides a convenient way to compute the PET with all available methods, pyet.calculate_all().

  1. Import the necessary Python packages.

    import pandas as pd
    import pyet
  2. Load the meteorological data.

    meteo = pd.read_csv("meteo.csv",
      index_col=0, parse_dates=True)
  3. Determine the necessary input data for the PET model.

    tmean, tmax, tmin, rh, rs, wind, \
    pet_knmi = (meteo[col] for col in
                meteo.columns)
    lat = 0.91  # latitude
    elev = 4  # elevation
  4. Estimate the potential evapotranspiration with all methods or the method of choice.

    pet_df = pyet.calculate_all(tmean,
      wind, rs, elev, lat, tmax, tmin, rh)
    pet_mak = pyet.makkink(tmean, rs,
      elevation=elev)
  5. Visualize and analyze the results.

    pet_df.plot()
    pet_df.boxplot()
    pet_df.cumsum().plot()
https://gmd.copernicus.org/articles/17/7083/2024/gmd-17-7083-2024-f03

Figure 3Potential evapotranspiration estimates for the year 2018 computed with all available PET methods plotted as (a) time series, (b) box plots, and (c) cumulative PET.

Download

The results from this analysis are shown in Fig. 3. From these visualizations, it is clear that the potential evapotranspiration depends on the chosen method. This can accumulate up to a 35 % deviation of the estimated annual flux from the mean in this example. Such substantial differences between the estimated fluxes motivate the use of multiple methods (ensemble modeling) (Beven and Freer2001; Krueger et al.2010; Shi et al.2020; Oudin et al.2005). This example showed how PyEt can be used efficiently for this task.

3.2 Example 2: estimate PET for gridded data

Gridded three-dimensional data (x, y, and t) obtained from satellites, radar imagery, or post-processed products are rapidly becoming widely available. More and more public datasets exist with global PET estimates at 0.1° resolution (e.g., Martens et al.2017; Xie et al.2022). PyEt also supports such gridded data, as illustrated here for the E-OBS gridded dataset (Cornes et al.2018) for Europe. The application of PyEt on gridded datasets is illustrated for the FAO-56, Makkink, and Hargreaves method. Xarray.DataArrays are used as input data instead of Pandas.Series. PyEt methods will return the same data type, a Xarray.DataArray. The workflow is comparable to that in the first example, except that now the individual PET methods are used.

https://gmd.copernicus.org/articles/17/7083/2024/gmd-17-7083-2024-f04

Figure 4Daily PET estimates for Europe from 6 to 8 June 2018 using meteorological data obtained from the E-OBS dataset (Cornes et al.2018).

The results for the three methods and three time steps are shown in Fig. 4. These again show that results may also differ spatially depending on the PET method. Looking more closely at Fig. 4, we can observe that the FAO-56 and Makkink methods do not compute PET in eastern parts of Europe. The data do not include relative humidity and solar radiation for these areas, and thus PET cannot be computed using the FAO-56 or Makkink method. If NaN (not-a-number) values are present in the required input data for a PyEt method, the method also returns a NaN value. The Hargreaves method, on the other hand, does not require solar radiation or relative humidity data. It can therefore be used to compute PET in the eastern parts of Europe. This example showed how PyEt can be applied to estimate PET using gridded data and demonstrated the benefits of using alternative PET methods when data such as radiation or relative humidity are missing.

3.3 Example 3: calibration of PET models

The available input data often does not suffice to compute potential evapotranspiration with the Penman–Monteith equation. This can be the case in data-scarce regions or time periods or when using historical data or data from climate models. In such cases, alternative PET methods can be calibrated to the estimates obtained from the Penman–Monteith equation for a period when sufficient data are available. The calibrated method can then be used to estimate PET in periods of data scarcity. As concluded by several authors (Jensen and Allen2016; Valipour2015; Yang et al.2021; Dlouhá et al.2021), calibration of alternative models is often crucial to ensure that the model fits the regional climate. In this example, it is shown how the calibration of temperature-based PET models affects the model uncertainty for studies focusing on current and past climates.

The approach is illustrated for the town of Graz, Austria, where the input data required for Penman–Monteith are only available from 2000 to 2021. Imagine, however, that for our study we also need potential evapotranspiration data for the period 1961 to 2021, but only temperature data are available (e.g., from the Spartacus temperature dataset (Hiebl and Frei2016)). Several steps are taken to calibrate the following five temperature-based methods: Oudin, Hargreaves, McGuinness–Bordne, Hamon, and Blaney–Criddle. First, the PET for the period 2000–2021 is computed using the Penman–Monteith equation. In the second step, the coefficients of the temperature-based PET equations are estimated by calibrating the estimated PET from temperature-based methods to the Penman–Monteith PET. Calibration is done by minimizing the sum of the squared residuals between these two PET estimates using SciPy's (Virtanen et al.2020) least_squares method. In the third and final step, these calibrated coefficients are used to estimate the PET for the period 1961–2021.

Figure 5 shows the computed PET with the default (row 1) and calibrated coefficients (row 2). The model bias (mm d−1) and the Kling–Gupta efficiency between simulated and observed (Penman–Monteith) PET show an improved model fit for all methods after calibration. The use of calibrated methods reduces the model bias, which is visually illustrated by the annual PET flux (composed of daily values) in the last column of Fig. 5. Using the Spartacus temperature dataset (Hiebl and Frei2016), PET can now be estimated back to 1961 using the calibrated alternative PET methods.

https://gmd.copernicus.org/articles/17/7083/2024/gmd-17-7083-2024-f05

Figure 5Density scatter plots comparing simulated and observed (FAO-56) PET for uncalibrated (row 1) and calibrated models (row 2). The performance of the calibrated models is evaluated using the model bias in mm d−1 (Bias), the coefficient of determination (R2), and the Kling–Gupta efficiency (KGE). The last column shows the annual PET sums for the period 1961–2021 using uncalibrated (row 1) and calibrated models (row 2).

Download

3.4 Example 4: the effect of CO2 on future PET estimates

In this example, it is shown how to account for changing environmental conditions affecting the PET flux when modeling the effects of climate change. Under a warmer and CO2-richer future (Caretta et al.2022), potential evapotranspiration tends to increase with increasing temperature (and vapor pressure deficit). A reduction in PET is expected under elevated CO2 due to an increased stomatal resistance (Field et al.1995; Ainsworth and Rogers2007). The increase in CO2 is still commonly ignored in PET models employed for climate change studies, although excluding its stomatal effect may lead to an overestimation of PET (Kingston et al.2009; Milly and Dunne2016; Vremec et al.2022; Riedel et al.2023). The effect of temperature increases on PET can be easily modeled with all available PET methods, as temperature is an input for all methods. The CO2 stomatal effect, however, can only be directly accounted for with the Penman–Monteith method (Liu et al.2022). Using a CO2-dependent stomatal resistance model implemented in PyEt (Yang et al.2019), the effect of elevated CO2 on stomatal resistance can be considered (see Eq. B2). When calculating PET with alternative methods, Kruijt et al. (2008) and Trnka et al. (2014) argued that an adjustment factor for the atmospheric CO2 concentration (fCO2) can be used to account for the effect of elevated CO2 concentrations on PET. The scaling factor can be obtained from literature values (Kruijt et al.2008; Trnka et al.2014). Alternatively, the factor can be calibrated using the Penman–Monteith equation together with the CO2-dependent stomatal resistance model (Eq. B1) to match the local climate and vegetation:

(1) PET CO 2 = f CO 2 PET 300 = ( 1 + S PET CO 2 ( CO 2 - 300 ) ) PET 300 ,

where SPETCO2 is the relative sensitivity of PET to CO2, PET300 is the computed Penman–Monteith estimate at 300 ppm [CO2] (preindustrial concentration), and PETCO2 is the computed Penman–Monteith estimate under elevated CO2 concentration (Yang et al.2019). Such relationships can be easily implemented in PyEt, and fCO2 can be obtained by calculating PET300 and PETCO2 with the Penman–Monteith equation (Eq. B1) at ambient and elevated CO2 concentration, respectively.

Building on the previous example, the Graz study area served as a practical example to demonstrate the application of the calibrated models in assessing the impact of warming and elevated CO2 concentration on PET based on the projected increase in temperature and CO2 concentration from the representative concentration pathways (RCPs) (Van Vuuren et al.2011). Daily PET was calculated for each RCP scenario (2.6, 4.5, 6.0, and 8.5) by adding the projected increase in temperature and CO2 concentration to the existing data for 2020–2021. Figure 6 shows the increase in the average annual PET (aggregated from daily values) under warming and elevated CO2 concentrations according to the RCP scenarios. In Fig. 6c, the effects of elevated CO2 concentration on PET were neglected, and only increases in temperature were considered. Similar to Milly and Dunne (2016), Yang et al. (2019), and Vremec et al. (2022), this example shows that neglecting the effect of elevated CO2 on PET (Fig. 6c) can lead to overestimation of PET under future conditions.

https://gmd.copernicus.org/articles/17/7083/2024/gmd-17-7083-2024-f06

Figure 6Projected increase in temperature (a) and atmospheric CO2 concentration (b) under the RCP scenarios, calculated increase in the average annual PET with warming (c), and PET with warming and elevated CO2 concentration (d). The uncertainty bounds represent the 5th–95th percentile of the PET model ensemble.

Download

4 Discussion

4.1 Improved handling of PET in scientific studies

Evapotranspiration data from lysimeter or eddy correlation measurements (Pastorello et al.2020) are rare and, if available at all, only locally available for relatively short time periods. Thus, there is a widespread need to estimate evapotranspiration from more readily available meteorological data using (semi-)empirical approaches. In general, these approaches follow three steps, as outlined for example by Allen et al. (1998). Firstly, the potential evapotranspiration of a reference surface (hence reference evapotranspiration) is estimated using meteorological data. Secondly, a crop coefficient may be applied to transform the reference evapotranspiration into the potential crop evapotranspiration. Thirdly, a soil–water balance approach is used to account for reduced actual evapotranspiration if the soil–water storage is depleted. PyEt is designed to perform the first two steps. It can be easily complemented by soil–water balance approaches to calculate actual evapotranspiration. Hydrological models, however, often use PET directly as input.

Rainfall–runoff models represent one type of hydrological model where PET is commonly used as an input, either as gridded data in distributive models or as a spatially aggregated values in lumped-parameter models. Some studies (e.g., Andréassian et al.2004; Oudin et al.2005; Sperna Weiland et al.2012) found that PET had little impact on the performance of such models, and they thus advocated for the use of simplistic PET models. However, Jayathilake and Smith (2021) found that model performance was clearly sensitive to PET at sites with water-limited evapotranspiration. More importantly, the choice of the PET model has been shown to affect the results of hydrological projections in climate change impact assessments (Kay and Davies2008; Seiller and Anctil2016; Dallaire et al.2021; Lemaitre-Basset et al.2022). PET is expected to be even more influential in the assessment of groundwater recharge (e.g., Bakundukize et al.2011) and crop water demands (e.g., Webber et al.2016), which – compared to runoff – are more directly linked to evapotranspiration. Thus, the selection of appropriate PET models needs to account for the research context and variable of interest.

Bormann (2010) found that PET models that are based on the same or similar climate variables exhibit different sensitivity to observed climate change. This finding suggests that appropriate PET models need to be specifically selected for the given region of interest. Guo et al. (2017) provides pointers to examining which variables are likely to be the most important for a particular location. For more detailed insights into the application of individual PET methods across various climates, refer to Table A2 and the studies by Allen et al. (1998), McMahon et al. (2013), Jensen and Allen (2016), Yang et al. (2021), and Pimentel et al. (2023). The comparison of PET estimates for Europe shown in Fig. 4 illustrates the spatial variability of differences in PET estimates obtained from different methods; as can be seen, the magnitude and pattern of PET estimates are similar in some regions (e.g., Scandinavia) but differ more strongly in others (e.g., southeastern Europe).

As indicated above, the performance of PET models may vary depending on the region considered. Approaches that were found to be applicable in one region may perform less well in other regions. In this case, PET models can be calibrated to a reference dataset by adjustment of the coefficients in the model equation, as shown in the third example. The reference dataset can either be observed evapotranspiration (e.g., from lysimeters) or PET obtained from a model considered to be reliable. This has been illustrated by Example 3, where the coefficients of temperature-based models were adjusted to achieve the best fit to the Penman–Monteith model. This approach can also be used to obtain consistent spatial distributions of PET. As shown in Example 2 (Fig. 4), the limited data availability for eastern Europe did not allow the application of the FAO-56 or Makkink method, while sufficient data were available for the Hargreaves method. Thus, one may consider calibrating the latter to one of the former methods where these are applicable and only then applying it to obtain estimates for the entire region. For a more advanced calibration procedure, see, for example, Haslinger and Bartsch (2016).

Often the range of PET models that can potentially be employed is pre-determined by data availability. This may be the case if historical records of climate data are to be used for the PET estimation, for example, as many weather stations do not measure all climate variables included in the Penman–Monteith equation. However, this is also often the case in assessments of hydrological impacts of climate change if projected climate variables have high uncertainty. Lai et al. (2022), for example, concluded that the high uncertainty of wind speed projected in complex terrain may increase the uncertainty in PET, whereas air temperature and solar radiation have low uncertainty and thus should be the parameters preferred in the PET model. Given the climate variables for which data is available, Table 1 can be used to identify the PET models that come into consideration. However, it is advised to evaluate the assumptions and limitations of the individual methods regarding their applicability in the given case. Please refer to the comments in Table A2 and the references in Table 1 for this purpose. We generally recommend applying all models that have been identified as suitable (PET model ensemble), but the purpose and specific implementation of such a multi-model approach will depend on the research context. Example 4 (Sect. 3.4) illustrated how PET model ensembles can be used to include model uncertainties in PET projections under warming and elevated atmospheric CO2 concentration. Since the latter effect is frequently excluded in hydrological projections, Milly and Dunne (2016) and Yang et al. (2019) advocate the inclusion of the effect of elevated CO2 on stomatal resistance when estimating PET under warming and elevated atmospheric CO2 concentrations.

To evaluate the performance of the estimated PET against observed values or other PET methods, performance metrics are employed. In this paper, the model bias and coefficient of determination (R2) are used to provide readers and practitioners with a clear and concise assessment of the overall deviation in flux and the percentage of the variability explained by the model (Onyutha2024). Additionally, the Kling–Gupta Efficiency (KGE) (Gupta et al.2009) is used, offering a comprehensive view by incorporating correlation, variability, and bias into a single statistic. KGE is particularly valued in hydrological modeling, as it addresses some limitations of the R2 by providing a differentiated perspective on model performance. The choice of the most appropriate performance metric is important to reduce model uncertainty, as it affects the judgment of the performance of a method. For further guidance on the evaluation of various performance metrics, readers are referred to Krause et al. (2005) and Onyutha (2024).

To improve reliability and efficiency in estimating PET, it is crucial to use a reproducible workflow. Scripts provide an efficient way to document the modeling process and are an important step towards full reproducibility. As shown in the examples, Jupyter Notebooks (Kluyver et al.2016) provide a solution for publishing code, results, and explanations in a single document. As such, the presented package and its application in this paper are in line with the steps suggested by Hutton et al. (2016) to improve reproducibility in hydrological studies. To speed up adaptation of the methods and allow a faster transfer between research teams, formal procedures such as benchmarking (e.g., Maxwell et al.2014) can help to ensure confidence in key complex codes.

4.2 Building the PyEt community and outlook

As a community project, the success of PyEt depends on the uptake from and interaction with the community. This, in turn, depends on the ease of use and the trust in the project. Emphasis was put on designing user-friendly and well-documented software, including various user examples, and extensive benchmark testing using continuous integration. Since the initial launch of PyEt, the package has already seen a good community uptake. Apart from applications of the software in projects related to the authors, which include estimating PET under conditions of warming and elevated CO2 concentrations, assessing potential crop evapotranspiration, and providing inputs for hydrological models (e.g., Vremec et al.2022; Forstner et al.2022; Collenteur et al.2023; Jemeljanova et al.2023), PyEt has also been independently used by other researchers. These studies have used PyEt for various purposes: integrating PET estimates with machine learning models for enhanced analytical capabilities (Vaz et al.2022; Kajári et al.2024) and combining them with software for computing groundwater recharge and the water balance (Hassanzadeh et al.2024). The software has also played a role in evaluating hydroclimatic changes and generating regional and global PET estimates (e.g., Tercini and Mello Júnior2023; Aguayo et al.2024; Ha et al.2024). The quick uptake of the software by the community confirms the need for this software.

The primary channel for communication with the PyEt community is GitHub, which provides several options for discussions, tracking code issues, and code development. Users are encouraged to ask questions in GitHub discussions and to report potential issues, suggest improvements, and feature requests via the GitHub issue tracker. As a community project, we plan to continue to improve the existing code and develop new capabilities based on feedback and with help from the community. An example of developments that are currently underway is the adaptation of the current methods to also work for hourly data, allowing the estimation of hourly PET. Other future work will focus on improvements in usability and the inclusion of other alternative methods.

5 Conclusions

This paper introduced PyEt, a Python package for the estimation of daily potential evapotranspiration (PET). The package enables the inclusion of model uncertainty and climate change in the PET estimation in a consistent, tested, and reproducible environment. With PyEt, PET can be estimated using 20 different methods with just a few lines of Python code. Unlike existing tools for PET calculation, which are designed for station-based time series, PyEt can also be applied to gridded (3D) datasets. This is of great practical relevance, particularly in climate impact studies, where gridded datasets are often used. The examples described in this paper illustrate how PyEt can be used in geoscientific studies to (1) facilitate the characterization of model uncertainty using a multimodel approach (model ensembles), (2) calibrate PET models and apply them in data-scarce regions and time periods, and (3) include the effects of warming and elevated atmospheric CO2 concentrations. The use of Python scripts and Jupyter Notebooks ensures reproducibility and provides a transparent report of the PET computation process. We believe that PyEt will help improve the handling of PET and allow a more sophisticated and comprehensive consideration of PET in environmental studies, particularly those related to climate change.

Appendix A: PET methods – climate suitability and applications

Table A1List of variables and symbols used in the paper and appendix.

Download Print Version | Download XLSX

Table A2Overview of PET methods: climate suitability, applications, and limitations. ET0 refers to reference crop or surface ET, PETC refers to potential crop or surface ET, PETOW refers to potential ET for open water, PETSL refers to potential ET for shallow lakes, PETRR refers to potential ET for rainfall–runoff modeling. Based on Allen et al. (1998), McMahon et al. (2013), Jensen and Allen (2016), Yang et al. (2021), and Pimentel et al. (2023).

Download Print Version | Download XLSX

Appendix B: PET methods – equations

B1 Penman–Monteith (pm)

Through the introduction of the Penman–Monteith equation by Monteith (1965), a broad applicability of PET estimation to different surfaces and vegetation types was achieved (Jensen and Allen2016). This was done by implementing the plant aerodynamic resistance (ra) and the surface resistance (rs) in the PET formula:

(B1) PET = 1 λ ρ w Δ ( R n - G ) + ρ a c p K u ( e s - e a ) / r a Δ + γ 1 + r s r a .

Users of PyEt can include leaf or canopy cover measurements (leaf area index, LAI) to calculate surface resistance (rs), thereby accounting for the effects of crop management and phenology on PET. A modified stomatal resistance model also allows for the inclusion of the sensitivity of the stomatal resistance (rl) to the atmospheric CO2 concentration (e.g., Yang et al.2019; Vremec et al.2022):

(B2) r s = r l ( CO 2 ) 0.5 LAI = r r l - 300 1 + S r l - CO 2 ( CO 2 - 300 ) 0.5 LAI ,

where Srl-[CO2] [ppm−1] is the relative sensitivity of rl to Δ [CO2] and rrl-300 [s m−1] is the reference stomatal resistance when atmospheric CO2 concentration is 300 ppm. The relative sensitivity of rl to Δ [CO2] represents the change in rl per ppm increase in CO2 concentration.

If measurements of crop height exist, these data can be used to calculate the aerodynamic resistance to vapor and heat transfer (ra) to represent the effects of crop phenology on PET:

(B3) r a = ln z m - d z om ln z h - d z oh k 2 u z ,

where zm is the reference level at which the wind speed is measured; zh is the height of the temperature and humidity measurements; k is the von Karman constant (=0.41); uz is the measured wind speed (Allen et al.1998); d is the zero plane displacement height, taken as 0.67hc [m]; zom is the roughness parameter for momentum (=0.123hc) [m]; and zoh is the roughness parameter for heat and water vapor (=0.1zom) [m] (Jensen and Allen2016).

Free parameters in the Penman–Monteith equation, available for calibration, include the bulk stomatal rl=100 (ranging between 40–150) and surface resistance (ranging between 50–200), with values for specific surfaces or crops found in Jensen and Allen (2016). Additionally, one can adjust the surface albedo (α = 0.23, ranges between 0.04 for water surfaces and 0.9 for snow) (Jensen and Allen2016) to estimate net shortwave radiation (Eq. 38 in Allen et al.1998), the empirical coefficients for net longwave radiation a=1.35 and b=-0.35 (Eq. 39 in Allen et al.1998), or the empirical coefficients for clear-sky radiation as1=0.25 and bs1=0.5 (Eq. 36 in Allen et al.1998). Optional meteorological inputs include G, Tmax, Tmin, RHmax, RHmin, p, and N.

B2 ASCE-PM (pm_asce)

The ASCE Penman–Monteith equation for is computed after Walter et al. (2000) (Eq. 1):

(B4) PET = 0.408 Δ ( R n - G ) + γ C n T + 273 u 2 ( e s - e a ) Δ + γ ( C d u 2 ) ,

where Cn=900 and Cd=0.34 are empirical coefficients for short reference vegetation (grass), while for tall reference vegetation (alfalfa) Cn=1600 and Cd=0.38 can be specified. The free parameters a, b, as1, bs1, and α are consistent with those specified for the Penman–Monteith method, while the optional meteorological inputs remain the same.

B3 FAO-56 (pm_fao56)

The FAO-56 Penman–Monteith equation for reference crop evapotranspiration is computed after Allen et al. (1998) (Eq. 6):

(B5) PET = 0.408 Δ ( R n - G ) + γ 900 T + 273 u 2 ( e s - e a ) Δ + γ ( 1 + 0.34 u 2 ) .

The free parameters a, b, as1, bs1, and α are consistent with those specified for the Penman–Monteith method, while the optional meteorological inputs remain the same.

B4 Penman (penman)

The Penman's PET formulation is computed after Penman (1948):

(B6) PET = 1 λ ρ w Δ ( R n - G ) + γ ( e s - e a ) ( a w + b w u 2 ) Δ + γ ,

where free parameters for the Penman's wind function include aw=1 and bw=0.537 (Valiantzas2006), while Penman (1948) suggested values of aw=2.626 and bw=1.381. The free parameters a, b, as1, bs1, and α are consistent with those specified for the Penman–Monteith method, while the optional meteorological inputs remain the same.

B5 Priestley–Taylor (priestley_taylor)

Priestley–Taylor's PET formulation is computed after Priestley and Taylor (1972):

(B7) PET = α L Δ ( R n - G ) λ ρ w ( Δ + γ ) ,

where αL=1.26 is an empirical coefficient. The free parameters a, b, as1, bs1, and α are consistent with those specified for the Penman–Monteith method, while the optional meteorological inputs remain the same.

B6 Kimberly–Penman (kimberly_penman)

The Kimberly–Penman equation (Wright1982) is computed after Oudin et al. (2005):

(B8) PET = Δ ( R n - G ) + γ ( e s - e a ) w λ ρ w ( Δ + γ ) ,

where w=u20.4+0.14exp-(j-173)582+0.605+0.345exp-j-243802.

The free parameters a, b, as1, bs1, and α are consistent with those specified for the Penman–Monteith method, while the optional meteorological inputs remain the same.

B7 Thom–Oliver (thom_oliver)

Thom–Oliver's PET formulation is computed (Thom and Oliver1977) the same way as in Oudin et al. (2005)

(B9) PET = Δ ( R n - G ) + 2.5 γ ( e s - e a ) a w ( 1 + b w u 2 ) λ ρ w ( Δ + γ ( 1 + r s r a ) ) ,

where aw=2.6 and bw=0.536. The free parameters a, b, as1, bs1, α, rl, rs, and Srl-[CO2] are consistent with those specified for the Penman–Monteith method, while the optional meteorological inputs remain the same.

B8 Blaney–Criddle (blaney_criddle)

Three different approaches can be taken to estimate the Blaney–Criddle PET, depending on the selected method.

(B10) PET = a + b P y ( 0.46 T + 8.13 ) if  Method = 0  (after  ), k P y ( 0.457 T + 8.128 ) if  Method = 1  (after  , Eq. 6), k 1 + b var P y ( 0.46 T + 8.13 ) if  Method = 2  (after  , Eqs. S9.7 and S9.8) ,

Here, k=0.65, a=-1.55, and b=0.96 are empirical coefficients, while k1=0.0043RHmin-nN-1.41 and bvar=0.81917-0.0040922RHmin+1.0705nN+0.065649u2-0.0059684RHminnN-0.0005967RHminu2.

B9 Hamon (hamon)

The PET formulation after Hamon (1963), as used in Oudin et al. (2005), is as follows:

(B11) PET = k N 12 2 exp T 16 ,

where k=1 is a calibration coefficient.

B10 Romanenko (romanenko)

Romanenko's PET formulation (Romanenko1961), as used in Oudin et al. (2005), is as follows:

(B12) PET = k 1 + T 25 2 1 - e a e s ,

where k=4.5 is an empirical coefficient (Oudin et al.2005). Optional meteorological inputs include Tmax, Tmin, RHmax, and RHmin.

B11 Linacre (linacre)

Linacre's PET formula (Linacre1977), as used in Oudin et al. (2005), is as follows:

(B13) PET = 500 T m 100 - lat deg + 15 ( T - T dew ) 80 - T ,

where Tm=T+0.006Elev.

B12 Haude (haude)

Haude's PET formulation (Haude1955), as used in Schiff (1975), is computed as follows:

(B14) PET = k F K ( e 0 - e a ) 10 ,

where k=1 is a calibration coefficient and FK represents Haude's monthly coefficients, as adapted by Schiff (1975).

B13 Turc (turc)

The PET formula, as derived from Turc (1961) and used in McMahon et al. (2013) (Eqs. S9.10 and S9.11), is computed as follows:

(B15) PET = k c T T + 15 ( 23.88 R s + 50 ) ,

where k=0.013 is an empirical coefficient and c, dependent on the relative humidity (RH), is defined as follows:

(B16) c = 1 + 50 - rh 70 if  RH < 50 , 1 otherwise .

B14 Jensen–Haise (jensen_haise)

The PET according to the Jensen–Haise model (Jensen and Haise1963), varies depending on the chosen method.

(B17) PET = k R s λ ρ w ( T - T x ) if  Method = 0  (after  ) , k R a T λ ρ w if  Method = 1  (after  ) ,

Here, cr=0.025 is an empirical coefficient, while Tx=-3, as used in Jensen and Allen (2016).

B15 McGuinness–Bordne (mcguinness_bordne)

McGuinness–Bordne's PET equation (McGuinness and Bordne1972), as used in Oudin et al. (2005), is as follows:

(B18) PET = k R a ( T + 5 ) λ ρ w ,

where k=0.0147 is an empirical coefficient as suggested by Xu and Singh (2000).

B16 Hargreaves (hargreaves)

The Hargreaves PET equation (Hargreaves and Samani1982) is computed as follows:

(B19) PET = k c HS R a λ ρ w T max - T min ( T + 17.8 ) .

B17 FAO-24 (fao_24)

The FAO-24 PET equation from Doorenbos (1977); Jensen et al. (1990), as used in Xu and Singh (2000) (Eqs. 11 and 12), is as follows:

(B20) PET = a + Δ Δ + γ R s λ ρ w ( 1 - α ) b ,

where a=-0.3 and b=1.066-0.13RH100+0.045u2-0.02RH100u2-0.315(RH100)2-0.0011u2. Free parameters include the surface albedo α, consistent with those used in the Penman–Monteith method.

B18 Abtew (abtew)

Abtew's PET equation (Abtew1996), as used in Xu and Singh (2000) (Eq. 14), is as follows:

(B21) PET = k R s λ ρ w ,

where k=0.53 is an empirical coefficient as suggested by Xu and Singh (2000).

B19 Makkink (makink)

Makkink's PET equation (Makkink1957) is as follows:

(B22) PET = Δ Δ + γ R s λ ρ w ,

where k=0.65 is the empirical coefficient recommended by Hiemstra and Sluiter (2011) and ranging between (0.61–0.77) (Jensen and Allen2016).

B20 Makkink–KNMI (makink_knmi)

The Royal Netherlands Meteorological Institute (KNMI) employs a slightly modified version of the Makkink equation, tailored specifically for conditions in the Netherlands, as described by Hiemstra and Sluiter (2011):

(B23) PET = k Δ Δ + γ R s λ ρ w ,

where k=0.65. The calculations for s, es, γ, and λ are specified as follows:

  • Δ=7.5237.3(237.3+T)2ln(10)es,

  • es=0.6107107.5T237.3+T,

  • γ=0.0646+0.00006T,

  • λ=(2501-2.375T)1000.

B21 Oudin (oudin)

According to Oudin et al. (2005) (Eq. 2), the potential evapotranspiration (PET) can be expressed as follows:

(B24) PET = R a ( T + k 2 ) λ ρ w k 1 if  T + k 2 > 0 0 otherwise,

where k2=5 and k1=100 (ranging between 75–100) are empirical coefficients recommended by Oudin et al. (2005).

B22 Performance metrics

He we provide mathematical definitions of the performance metrics used to evaluate the PET models discussed in the main text. The model bias (mm d−1) is calculated as the average difference between the PET values estimated using PyEt (y^i) and the reference PET values, which include (i) literature values presented in Fig. 1 and (ii) PET values computed using the Penman–Monteith method in Example 3 (yi) over n time steps:

(B25) Bias = 1 n i = 1 n ( y ^ i - y i ) .

The coefficient of determination (R2) was computed as follows:

(B26) R 2 = 1 - i = 1 n ( y i - y ^ i ) 2 i = 1 n ( y i - y ) 2 ,

where y is the mean of the reference data.

The Kling–Gupta efficiency (KGE) is defined as follows:

(B27) KGE = 1 - ( r - 1 ) 2 + ( α - 1 ) 2 + ( β - 1 ) 2 ,

where r is the correlation coefficient between the reference and estimated data, α is the ratio of the standard deviation of estimated data to that of reference data, and β is the ratio of the mean of estimated data to that of reference data (bias ratio).

Code and data availability

The Jupyter Notebook and data used in this study are available in the “examples” folder of the GitHub repository (http://www.github.com/pyet-org/pyet, last access: 17 September 2024) and are also available on Zenodo (version v.1.3.1; DOI: https://doi.org/10.5281/zenodo.5896799; Vremec and Collenteur2024a). The authors welcome code contributions, bug reports, and feedback from the community to further improve the software. PyEt is free and open-source software available under the MIT license. The source code is available at the project's home page on GitHub (http://www.github.com/pyet-org/pyet). The full documentation is available on https://PyEt.readthedocs.io (Vremec and Collenteur2024b). PyEt is meant as a community project, and contributions and feedback to continue to improve and develop the project are welcome.

Author contributions

Conceptualization: MV, SB, and RAC; software: MV and RAC; investigation: MV; writing – original draft preparation: MV and RAC; writing – review and editing: SB and RAC; supervision, SB. All authors have read and agreed to the published version of the manuscript.

Competing interests

The contact author has declared that none of the authors has any competing interests.

Disclaimer

Publisher's note: Copernicus Publications remains neutral with regard to jurisdictional claims made in the text, published maps, institutional affiliations, or any other geographical representation in this paper. While Copernicus Publications makes every effort to include appropriate place names, the final responsibility lies with the authors.

Acknowledgements

We acknowledge the financial support provided by the University of Graz. We acknowledge the ZAMG dataset (https://data.hub.zamg.ac.at, last access: 17 September 2024) and the KNMI dataset (https://www.knmi.nl/home, last access: 17 September 2024). We also acknowledge the E-OBS dataset from the EU-FP6 project UERRA (http://www.uerra.eu, last access: 17 September 2024), the Copernicus Climate Change Service, and the data providers of the ECA&D project (https://www.ecad.eu, last access: 17 September 2024). We appreciate the insightful comments and suggestions from the two anonymous reviewers and the handling editor that greatly improved this paper. We also thank the contributors from the preprint discussions in Hydrology and Earth System Sciences and the PyEt community for their valuable input.

Financial support

This research has been supported by the Österreichischen Akademie der Wissenschaften (ClimGrassHydro).

Review statement

This paper was edited by Charles Onyutha and reviewed by two anonymous referees.

References

Abtew, W.: Evapotranspiration measurements and modeling for three wetland systems in South Florida 1, JAWRA Journal of the American Water Resources Association, 32, 465–473, 1996. a, b

Aguayo, R., León-Muñoz, J., Aguayo, M., Baez-Villanueva, O. M., Zambrano-Bigiarini, M., Fernández, A., and Jacques-Coper, M.: PatagoniaMet: A multi-source hydrometeorological dataset for Western Patagonia, Sci. Data, 11, 6, https://doi.org/10.1038/s41597-023-02828-2, 2024. a

Ainsworth, E. A. and Rogers, A.: The response of photosynthesis and stomatal conductance to rising [CO2]: mechanisms and environmental interactions, Plant Cell Environ., 30, 258–270, https://doi.org/10.1111/j.1365-3040.2007.01641.x, 2007. a, b

Allen, R. G., Pereira, L. S., Raes, D., and Smith, M.: Crop evapotranspiration-Guidelines for computing crop water requirements-FAO Irrigation and drainage paper 56, Fao, Rome, 300, D05109, ISBN 92-5-104219-5, 1998. a, b, c, d, e, f, g, h, i, j, k, l, m, n

Andréassian, V., Perrin, C., and Michel, C.: Impact of imperfect potential evapotranspiration knowledge on the efficiency and parameters of watershed models, J. Hydrol., 286, 19–35, https://doi.org/10.1016/j.jhydrol.2003.09.030, 2004. a

Ansorge, L. and Beran, A.: Performance of simple temperature-based evaporation methods compared with a time series of pan evaporation measures from a standard 20 m2 tank, J. Water Land Develop., 41, 1–11, https://doi.org/10.2478/jwld-2019-0021, 2019. a, b

Bakundukize, C., Van Camp, M., and Walraevens, K.: Estimation of groundwater recharge in Bugesera region (Burundi) using soil moisture budget approach, GEOLOGICA BELGICA, 14, 85–102, http://hdl.handle.net/1854/LU-1204652 (last access: 17 September 2024), 2011. a

Barker, M., Chue Hong, N., Katz, D. S., Lamprecht, A.-L., Martinez Ortiz, C., Psomopoulos, F., Harrow, J., Castro, L., Gruenpeter, M., Martinez, P., and Honeyman, T.: Introducing the FAIR Principles for research software, Sci. Data, 9, 622, https://doi.org/10.1038/s41597-022-01710-x, 2022. a, b

Beven, K. and Freer, J.: J. Hydrol., 249, 11–29, https://doi.org/10.1016/S0022-1694(01)00421-8, 2001. a, b

Blaney, H. F. and others: Determining water requirements in irrigated areas from climatological and irrigation data, Tech. rep., United States Department Of Agriculture, 1952. a

Bormann, H.: Sensitivity analysis of 18 different potential evapotranspiration models to observed climatic change at German climate stations, Clim. Change, 104, 729–753, https://doi.org/10.1007/s10584-010-9869-7, 2010. a, b, c

Caretta, M. A., Mukherji, A., Arfanuzzaman, M., Betts, R. A., Gelfan, A., Hirabayashi, Y., Lissner, T. K., Liu, J., Gunn, E. L., Morgan, R., Mwanga, S., and Supratid, S.: Water, in: Climate Change 2022: Impacts, Adaptation, and Vulnerability, Contribution of Working Group II to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change, edited by: Pörtner, H.-O., Roberts, D. C., Tignor, M., Poloczanska, E. S., Mintenbeck, K., Alegría, A., Craig, M., Langsdorf, S., Löschke, S., Möller, V., Okem, A., and Rama, B., Cambridge University Press, https://doi.org/10.1017/9781009325844.006, 2022. a, b

Collenteur, R. A., Moeck, C., Schirmer, M., and Birk, S.: Analysis of nationwide groundwater monitoring networks using lumped-parameter models, J. Hydrol., 626, 130120, https://doi.org/10.1016/j.jhydrol.2023.130120, 2023. a

Cornes, R. C., van der Schrier, G., van den Besselaar, E. J. M., and Jones, P. D.: An Ensemble Version of the E-OBS Temperature and Precipitation Data Sets, J. Geophys. Res.-Atmos., 123, 9391–9409, https://doi.org/10.1029/2017JD028200, 2018. a, b

Dakhlaoui, H., Seibert, J., and Hakala Assendelft, K.: Sensitivity of discharge projections to potential evapotranspiration estimation in Northern Tunisia, Reg. Environ. Change, 20, 34, https://doi.org/10.1007/s10113-020-01615-8, 2020. a

Dallaire, G., Poulin, A., Arsenault, R., and Brissette, F.: Uncertainty of potential evapotranspiration modelling in climate change impact studies on low flows in North America, Hydrol. Sci. J., 66, 689–702, https://doi.org/10.1080/02626667.2021.1888955, 2021. a

DeJonge, K. C. and Thorp, K. R.: Implementing Standardized Reference Evapotranspiration and Dual Crop Coefficient Approach in the DSSAT Cropping System Model, Transactions of the ASABE, 60, 1965–1981, https://doi.org/10.13031/trans.12321, 2017. a

Dingman, S. L.: Physical hydrology, Waveland Press, Inc, Long Grove, Illinois, 3rd edn., ISBN 9781478611189, 2015. a

Dlouhá, D., Dubovský, V., and Pospíšil, L.: Optimal Calibration of Evaporation Models against Penman–Monteith Equation, Water, 13, 1484, https://doi.org/10.3390/w13111484, 2021. a

Doorenbos, J.: Guidelines for predicting crop water requirements, FAO irrigation and drainage paper, 24, 1–179, 1977. a

Fatichi, S., Leuzinger, S., Paschalis, A., Langley, J. A., Barraclough, A. D., and Hovenden, M. J.: Partitioning direct and indirect effects reveals the response of water-limited ecosystems to elevated CO2, P. Natl. Acad. Sci. USA, 113, 12757–12762, https://doi.org/10.1073/pnas.1605036113, 2016. a

Field, C. B., Jackson, R. B., and Mooney, H. A.: Stomatal responses to increased CO2: implications from the plant to the global scale, Plant Cell Environ., 18, 1214–1225, https://doi.org/10.1111/j.1365-3040.1995.tb00630.x, 1995. a

Fisher, J. B., Whittaker, R. J., and Malhi, Y.: ET come home: potential evapotranspiration in geographical ecology, Global Ecol. Biogeogr., 20, 1–18, https://doi.org/10.1111/j.1466-8238.2010.00578.x, 2011. a

Forstner, V., Vremec, M., Herndl, M., and Birk, S.: Effects of dry spells on soil moisture and yield anomalies at a montane managed grassland site: A lysimeter climate experiment, Ecohydrology, 16, e2518, https://doi.org/10.1002/eco.2518, 2022. a

Gharbia, S. S., Smullen, T., Gill, L., Johnston, P., and Pilla, F.: Spatially distributed potential evapotranspiration modeling and climate projections, Sci. Total Environ., 633, 571–592, https://doi.org/10.1016/j.scitotenv.2018.03.208, 2018. a

Guo, D., Westra, S., and Maier, H. R.: An R package for modelling actual, potential and reference evapotranspiration, Environ. Model. Softw., 78, 216–224, https://doi.org/10.1016/j.envsoft.2015.12.019, 2016. a, b

Guo, D., Westra, S., and Maier, H. R.: Sensitivity of potential evapotranspiration to changes in climate variables for different Australian climatic zones, Hydrol. Earth Syst. Sci., 21, 2107–2126, https://doi.org/10.5194/hess-21-2107-2017, 2017. a

Gupta, H. V., Kling, H., Yilmaz, K. K., and Martinez, G. F.: Decomposition of the mean squared error and NSE performance criteria: Implications for improving hydrological modelling, J. Hydrol., 377, 80–91, https://doi.org/10.1016/j.jhydrol.2009.08.003, 2009. a, b

Ha, T. V., Uereyen, S., and Kuenzer, C.: Spatiotemporal analysis of tropical vegetation ecosystems and their responses to multifaceted droughts in Mainland Southeast Asia using satellite-based time series, GIScience & Remote Sensing, 61, 2387385, https://doi.org/10.1080/15481603.2024.2387385, 2024. a

Hamon, W. R.: Estimating potential evapotranspiration, T. Am. Soc. Civ. Eng., 128, 324–338, 1963. a, b

Hargreaves, G. H. and Samani, Z. A.: Estimating potential evapotranspiration, J. Irr. Drain. Div.-ASCE, 108, 225–230, 1982. a, b

Harris, C. R., Millman, K. J., Van der Walt, S. J., Gommers, R., Virtanen, P., Cournapeau, D., Wieser, E., Taylor, J., Berg, S., Smith, N. J., Kern, R., Picus, M., Hoyer, S., van Kerkwijk, M. H., Brett, M., Haldane, A., Fernández del Río, J., Wiebe, M., Peterson, P., Gérard-Marchant, P., Sheppard, K., Reddy, T., Weckesser, W., Abbasi, H., Gohlke, C., and Oliphant, T. E.: Array programming with NumPy, Nature, 585, 357–362, https://doi.org/10.1038/s41586-020-2649-2, 2020. a, b

Haslinger, K. and Bartsch, A.: Creating long-term gridded fields of reference evapotranspiration in Alpine terrain based on a recalibrated Hargreaves method, Hydrol. Earth Syst. Sci., 20, 1211–1223, https://doi.org/10.5194/hess-20-1211-2016, 2016. a

Hassanzadeh, A., Vázquez-Suñé, E., Valdivielso, S., and Corbella, M.: WaterpyBal: A comprehensive open-source python library for groundwater recharge assessment and water balance modeling, Environ. Model. Softw., 172, 105934, https://doi.org/10.1016/j.envsoft.2023.105934, 2024. a

Haude, W.: Determination of evapotranspiration by an approach as simple as possible, Mitt. Dt. Wetterdienst, 2, 1955. a, b

Hiebl, J. and Frei, C.: Daily temperature grids for Austria since 1961 – concept, creation and applicability, Theor. Appl. Climatol., 124, 161–178, https://doi.org/10.1007/s00704-015-1411-4, 2016. a, b

Hiemstra, P. and Sluiter: Interpolation of Makkink evaporation in the Netherlands, Royal Netherlands Meteorological Instituute (KNMI), 2011. a, b

Houska, T., Kraft, P., Chamorro-Chavez, A., and Breuer, L.: SPOTting Model Parameters Using a Ready-Made Python Package, PLOS ONE, 10, 1–22, https://doi.org/10.1371/journal.pone.0145180, 2015. a, b

Hoyer, S. and Hamman, J.: xarray: ND labeled arrays and datasets in Python, J. Open Res. Softw., 5, 261–272, https://doi.org/10.1038/s41592-019-0686-2, 2017. a, b

Hunter, J. D.: Matplotlib: A 2D graphics environment, Comput. Sci. Eng., 9, 90–95, https://doi.org/10.1109/MCSE.2007.55, 2007. a

Hutton, C., Wagener, T., Freer, J., Han, D., Duffy, C., and Arheimer, B.: Most computational hydrology is not reproducible, so is it really science?, Water Resour. Res., 52, 7548–7555, https://doi.org/10.1002/2016WR019285, 2016. a, b

Jayathilake, D. I. and Smith, T.: Assessing the impact of PET estimation methods on hydrologic model performance, Hydrol. Res., 52, 373–388, https://doi.org/10.2166/nh.2020.066, 2021. a

Jemeljanova, M., Collenteur, R. A., Kmoch, A., Bikše, J., Popovs, K., and Kalvāns, A.: Modeling hydraulic heads with impulse response functions in different environmental settings of the Baltic countries, J. Hydrol.-Regional Studies, 47, 101416, https://doi.org/10.1016/j.ejrh.2023.101416, 2023. a

Jensen, M. E. and Allen, R. G.: Evaporation, Evapotranspiration, and Irrigation Water Requirements, American Society of Civil Engineers, 2nd edn., https://doi.org/10.1061/9780784414057, 2016. a, b, c, d, e, f, g, h, i, j, k, l, m, n

Jensen, M. E. and Haise, H. R.: Estimating evapotranspiration from solar radiation, J. Irr. Drain. Div.-ASCE, 89, 15–41, https://doi.org/10.1061/JRCEA4.0000287, 1963. a, b

Jensen, M. E., Burman, R. D., and Allen, R. G.: Evapotranspiration and irrigation water requirements, ASCE, New York, 1990. a, b

Kajári, B., Tobak, Z., Túri, N., Bozán, C., and Van Leeuwen, B.: Prediction of Inland Excess Water Inundations Using Machine Learning Algorithms, Water, 16, 1267, https://doi.org/10.3390/w16091267, 2024. a

Katul, G. G., Oren, R., Manzoni, S., Higgins, C., and Parlange, M. B.: Evapotranspiration: A process driving mass transport and energy exchange in the soil-plant-atmosphere-climate system, Rev. Geophys., 50, RG3002, https://doi.org/10.1029/2011RG000366, 2012. a

Kay, A. and Davies, H.: Calculating potential evaporation from climate model data: A source of uncertainty for hydrological climate change impacts, J. Hydrol., 358, 221–239, https://doi.org/10.1016/j.jhydrol.2008.06.005, 2008. a

Kingston, D. G., Todd, M. C., Taylor, R. G., Thompson, J. R., and Arnell, N. W.: Uncertainty in the estimation of potential evapotranspiration under climate change, Geophys. Res. Lett., 36, L20403, https://doi.org/10.1029/2009GL040267, 2009. a

Kluyver, T., Ragan-Kelley, B., Pérez, F., Granger, B., Bussonnier, M., Frederic, J., Kelley, K., Hamrick, J., Grout, J., Corlay, S., Ivanov, P., Avila, D., Abdalla, S., Willing, C., and Team, J. D.: Jupyter Notebooks – a publishing format for reproducible computational workflows, Positioning and Power in Academic Publishing: Players, Agents and Agendas, 87–90, https://doi.org/10.3233/978-1-61499-649-1-87, 2016. a, b

Krause, P., Boyle, D. P., and Bäse, F.: Comparison of different efficiency criteria for hydrological model assessment, Adv. Geosci., 5, 89–97, https://doi.org/10.5194/adgeo-5-89-2005, 2005. a

Krueger, T., Freer, J., Quinton, J. N., Macleod, C. J. A., Bilotta, G. S., Brazier, R. E., Butler, P., and Haygarth, P. M.: Ensemble evaluation of hydrological model hypotheses, Water Resour. Res., 46, W07516, https://doi.org/10.1029/2009WR007845, 2010. a

Kruijt, B., Witte, J.-P. M., Jacobs, C. M. J., and Kroon, T.: Effects of rising atmospheric CO2 on evapotranspiration and soil moisture: A practical approach for the Netherlands, J. Hydrol., 349, 257–267, https://doi.org/10.1016/j.jhydrol.2007.10.052, 2008. a, b

Kumar, R., Jat, M. K., and Shankar, V.: Methods to estimate irrigated reference crop evapotranspiration – a review, Water Sci. Technol., 66, 525–535, https://doi.org/10.2166/wst.2012.191, 2012. a

Lai, C., Chen, X., Zhong, R., and Wang, Z.: Implication of climate variable selections on the uncertainty of reference crop evapotranspiration projections propagated from climate variables projections under climate change, Agric. Water Manage., 259, 107273, https://doi.org/10.1016/j.agwat.2021.107273, 2022. a

Lemaitre-Basset, T., Oudin, L., Thirel, G., and Collet, L.: Unraveling the contribution of potential evaporation formulation to uncertainty under climate change, Hydrol. Earth Syst. Sci., 26, 2147–2159, https://doi.org/10.5194/hess-26-2147-2022, 2022. a, b, c

Linacre, E. T.: A simple formula for estimating evaporation rates in various climates, using temperature data alone, Agric. Meteorol., 18, 409–424, https://doi.org/10.1016/0002-1571(77)90007-3, 1977. a, b

Liu, Z., Han, J., and Yang, H.: Assessing the ability of potential evaporation models to capture the sensitivity to temperature, Agric. Forest Meteorol., 317, 108886, https://doi.org/10.1016/j.agrformet.2022.108886, 2022. a

Makkink, G. F.: Testing the Penman formula by means of lysimeters, Journal of the Institution of Water Engineers, 11, 277–288, 1957. a, b

Martens, B., Miralles, D. G., Lievens, H., van der Schalie, R., de Jeu, R. A. M., Fernández-Prieto, D., Beck, H. E., Dorigo, W. A., and Verhoest, N. E. C.: GLEAM v3: satellite-based land evaporation and root-zone soil moisture, Geosci. Model Dev., 10, 1903–1925, https://doi.org/10.5194/gmd-10-1903-2017, 2017. a

Maxwell, R. M., Putti, M., Meyerhoff, S., Delfs, J., Ferguson, I. M., Ivanov, V., Kim, J., Kolditz, O., Kollet, S. J., Kumar, M., Lopez, S., Niu, J., Paniconi, C., Park, Y., Phanikumar, M. S., Shen, C., Sudicky, E. A., and Sulis, M.: Surface–subsurface model intercomparison: A first set of benchmark results to diagnose integrated hydrology and feedbacks, Water Resour. Res., 50, 1531–1549, https://doi.org/10.1002/2013WR013725, 2014. a

May, R. M., Goebbert, K. H., Thielen, J. E., Leeman, J. R., Camron, M. D., Bruick, Z., Bruning, E. C., Manser, R. P., Arms, S. C., and Marsh, P. T.: MetPy: A meteorological Python library for data analysis and visualization, B. Am. Meteorol. Soc., 103, E2273–E2284, https://doi.org/10.1175/BAMS-D-21-0125.1, 2022. a

McGuinness, J. and Bordne, E.: A comparison of lysimeter derived potential evapotranspiration with computed values, Tech. Bull., 1452, Agric. Res. Serv., US Dep. of Agric., Washington, DC, https://doi.org/10.22004/ag.econ.171893, 1972. a, b

McKinney, W.: Data Structures for Statistical Computing in Python, in: Proceedings of the 9th Python in Science Conference, edited by: Walt, S. V. D. and Millman, J., 56–61, https://doi.org/10.25080/Majora-92bf1922-00a, 201. a, b

McMahon, T. A., Peel, M. C., Lowe, L., Srikanthan, R., and McVicar, T. R.: Estimating actual, potential, reference crop and pan evaporation using standard meteorological data: a pragmatic synthesis, Hydrol. Earth Syst. Sci., 17, 1331–1363, https://doi.org/10.5194/hess-17-1331-2013, 2013. a, b, c, d, e, f, g, h, i, j

Milly, P. and Dunne, K.: Potential evapotranspiration and continental drying, Nat. Clim. Change, 6, 946–949, https://doi.org/10.1038/nclimate3046, 2016. a, b, c, d

Monteith, J. L.: Evaporation and environment, in: Symposia of the society for experimental biology, vol. 19, pp. 205–234, Cambridge University Press (CUP) Cambridge, 1965. a, b

Oki, T. and Kanae, S.: Global Hydrological Cycles and World Water Resources, Science, 313, 1068–1072, https://doi.org/10.1126/science.1128845, 2006. a

Onyutha, C.: Pros and cons of various efficiency criteria for hydrological model performance evaluation, Proc. IAHS, 385, 181–187, https://doi.org/10.5194/piahs-385-181-2024, 2024. a, b

Oudin, L., Michel, C., and Anctil, F.: Which potential evapotranspiration input for a lumped rainfall-runoff model?, J. Hydrol., 303, 275–289, https://doi.org/10.1016/j.jhydrol.2004.08.025, 2005. a, b, c, d, e, f, g, h, i, j, k, l, m, n, o, p

Pastorello, G., Trotta, C., Canfora, E., Chu, H., Christianson, D., Cheah, Y.-W., Poindexter, C., Chen, J., Elbashandy, A., Humphrey, M., Isaac, P., Polidori, D., Reichstein, M., Ribeca, A., van Ingen, C., Vuichard, N., Zhang, L., Amiro, B., Ammann, C., Arain, M. A., Ardö, J., Arkebauer, T., Arndt, S. K., Arriga, N., Aubinet, M., Aurela, M., Baldocchi, D., Barr, A., Beamesderfer, E., Marchesini, L. B., Bergeron, O., Beringer, J., Bernhofer, C., Berveiller, D., Billesbach, D., Black, T. A., Blanken, P. D., Bohrer, G., Boike, J., Bolstad, P. V., Bonal, D., Bonnefond, J.-M., Bowling, D. R., Bracho, R., Brodeur, J., Brümmer, C., Buchmann, N., Burban, B., Burns, S. P., Buysse, P., Cale, P., Cavagna, M., Cellier, P., Chen, S., Chini, I., Christensen, T. R., Cleverly, J., Collalti, A., Consalvo, C., Cook, B. D., Cook, D., Coursolle, C., Cremonese, E., Curtis, P. S., D'Andrea, E., da Rocha, H., Dai, X., Davis, K. J., Cinti, B. D., Grandcourt, A. D., Ligne, A. D., De Oliveira, R. C., Delpierre, N., Desai, A. R., Di Bella, C. M., Tommasi, P. d., Dolman, H., Domingo, F., Dong, G., Dore, S., Duce, P., Dufrêne, E., Dunn, A., Dušek, J., Eamus, D., Eichelmann, U., ElKhidir, H. A. M., Eugster, W., Ewenz, C. M., Ewers, B., Famulari, D., Fares, S., Feigenwinter, I., Feitz, A., Fensholt, R., Filippa, G., Fischer, M., Frank, J., Galvagno, M., Gharun, M., Gianelle, D., Gielen, B., Gioli, B., Gitelson, A., Goded, I., Goeckede, M., Goldstein, A. H., Gough, C. M., Goulden, M. L., Graf, A., Griebel, A., Gruening, C., Grünwald, T., Hammerle, A., Han, S., Han, X., Hansen, B. U., Hanson, C., Hatakka, J., He, Y., Hehn, M., Heinesch, B., Hinko-Najera, N., Hörtnagl, L., Hutley, L., Ibrom, A., Ikawa, H., Jackowicz-Korczynski, M., Janouš, D., Jans, W., Jassal, R., Jiang, S., Kato, T., Khomik, M., Klatt, J., Knohl, A., Knox, S., Kobayashi, H., Koerber, G., Kolle, O., Kosugi, Y., Kotani, A., Kowalski, A., Kruijt, B., Kurbatova, J., Kutsch, W. L., Kwon, H., Launiainen, S., Laurila, T., Law, B., Leuning, R., Li, Y., Liddell, M., Limousin, J.-M., Lion, M., Liska, A. J., Lohila, A., López-Ballesteros, A., López-Blanco, E., Loubet, B., Loustau, D., Lucas-Moffat, A., Lüers, J., Ma, S., Macfarlane, C., Magliulo, V., Maier, R., Mammarella, I., Manca, G., Marcolla, B., Margolis, H. A., Marras, S., Massman, W., Mastepanov, M., Matamala, R., Matthes, J. H., Mazzenga, F., McCaughey, H., McHugh, I., McMillan, A. M. S., Merbold, L., Meyer, W., Meyers, T., Miller, S. D., Minerbi, S., Moderow, U., Monson, R. K., Montagnani, L., Moore, C. E., Moors, E., Moreaux, V., Moureaux, C., Munger, J. W., Nakai, T., Neirynck, J., Nesic, Z., Nicolini, G., Noormets, A., Northwood, M., Nosetto, M., Nouvellon, Y., Novick, K., Oechel, W., Olesen, J. E., Ourcival, J.-M., Papuga, S. A., Parmentier, F.-J., Paul-Limoges, E., Pavelka, M., Peichl, M., Pendall, E., Phillips, R. P., Pilegaard, K., Pirk, N., Posse, G., Powell, T., Prasse, H., Prober, S. M., Rambal, S., Rannik, Ã., Raz-Yaseef, N., Rebmann, C., Reed, D., Dios, V. R. d., Restrepo-Coupe, N., Reverter, B. R., Roland, M., Sabbatini, S., Sachs, T., Saleska, S. R., Sánchez-Cañete, E. P., Sanchez-Mejia, Z. M., Schmid, H. P., Schmidt, M., Schneider, K., Schrader, F., Schroder, I., Scott, R. L., Sedlák, P., Serrano-Ortíz, P., Shao, C., Shi, P., Shironya, I., Siebicke, L., Šigut, L., Silberstein, R., Sirca, C., Spano, D., Steinbrecher, R., Stevens, R. M., Sturtevant, C., Suyker, A., Tagesson, T., Takanashi, S., Tang, Y., Tapper, N., Thom, J., Tomassucci, M., Tuovinen, J.-P., Urbanski, S., Valentini, R., van der Molen, M., van Gorsel, E., van Huissteden, K., Varlagin, A., Verfaillie, J., Vesala, T., Vincke, C., Vitale, D., Vygodskaya, N., Walker, J. P., Walter-Shea, E., Wang, H., Weber, R., Westermann, S., Wille, C., Wofsy, S., Wohlfahrt, G., Wolf, S., Woodgate, W., Li, Y., Zampedri, R., Zhang, J., Zhou, G., Zona, D., Agarwal, D., Biraud, S., Torn, M., and Papale, D.: The FLUXNET2015 dataset and the ONEFlux processing pipeline for eddy covariance data, Sci. Data, 7, 225, https://doi.org/10.1038/s41597-020-0534-3, 2020. a

Penman, H. L.: Natural evaporation from open water, bare soil and grass, P. Roy. Soc. Lond. A, 193, 120–145, publisher: The Royal Society London, 1948. a, b, c

Peterson, T. J., Wasko, C., Saft, M., and Peel, M. C.: AWAPer: An R package for area weighted catchment daily meteorological data anywhere within Australia, Hydrol. Process., 34, 1301–1306, https://doi.org/10.1002/hyp.13637, 2020. a

Pimentel, R., Arheimer, B., Crochemore, L., Andersson, J. C. M., Pechlivanidis, I. G., and Gustafsson, D.: Which Potential Evapotranspiration Formula to Use in Hydrological Modeling World-Wide?, Water Resour. Res., 59, e2022WR033447, https://doi.org/10.1029/2022WR033447, 2023. a, b

Priestley, C. H. B. and Taylor, R. J.: On the assessment of surface heat flux and evaporation using large-scale parameters, Mon. Weather Rev., 100, 81–92, 1972. a, b

Prudhomme, C. and Williamson, J.: Derivation of RCM-driven potential evapotranspiration for hydrological climate change impact analysis in Great Britain: a comparison of methods and associated uncertainty in future projections, Hydrol. Earth Syst. Sci., 17, 1365–1377, https://doi.org/10.5194/hess-17-1365-2013, 2013. a

Richards, M.: PyETo, https://github.com/woodcrafty/PyETo (last access: 17 September 2024), 2019. a

Riedel, T., Weber, T. K. D., and Bergmann, A.: Near constant groundwater recharge efficiency under global change in a central European catchment, Hydrol. Process., 37, e14805, https://doi.org/10.1002/hyp.14805, 2023. a

Romanenko, V.: Computation of the autumn soil moisture using a universal relationship for a large area, Proc. of Ukrainian Hydrometeorological Research Institute, 3, 12–25, 1961. a, b

Rosenberry, D. O., Stannard, D. I., Winter, T. C., and Martinez, M. L.: Comparison of 13 equations for determining evapotranspiration from a prairie wetland, Cottonwood Lake Area, North Dakota, USA, Wetlands, 24, 483–497, https://doi.org/10.1672/0277-5212(2004)024[0483:COEFDE]2.0.CO;2, 2004. a

Schiff, H.: Berechnung der potentiellen Verdunstung und deren Vergleich mit aktuellen Verdunstungswerten von Lysimetern, Archiv für Meteorologie, Geophysik und Bioklimatologie, Serie B, 23, 331–342, https://doi.org/10.1007/BF02242689, 1975. a, b, c, d

Schrödter, H.: Hinweise Für den Einsatz Anwendungsorientierter Bestimmungsverfahren, in: Verdunstung: Anwendungsorientierte Meßverfahren und Bestimmungsmethoden, Springer Berlin Heidelberg, Berlin, Heidelberg, ISBN 978-3-642-70434-5, https://doi.org/10.1007/978-3-642-70434-5_8, 1985. a, b, c, d

Seiller, G. and Anctil, F.: How do potential evapotranspiration formulas influence hydrological projections?, Hydrol. Sci. J., 61, 2249–2266, https://doi.org/10.1080/02626667.2015.1100302, 2016. a, b, c

Shi, L., Feng, P., Wang, B., Liu, D. L., Cleverly, J., Fang, Q., and Yu, Q.: Projecting potential evapotranspiration change and quantifying its uncertainty under future climate scenarios: A case study in southeastern Australia, J. Hydrol., 584, 124756, https://doi.org/10.1016/j.jhydrol.2020.124756, 2020. a, b

Sperna Weiland, F. C., Tisseuil, C., Dürr, H. H., Vrac, M., and van Beek, L. P. H.: Selecting the optimal method to calculate daily global reference potential evaporation from CFSR reanalysis data for application in a hydrological model study, Hydrol. Earth Syst. Sci., 16, 983–1000, https://doi.org/10.5194/hess-16-983-2012, 2012. a

Tercini, J. R. B. and Mello Júnior, A. V.: Impact of Hydroclimatic Changes on Water Security in the Cantareira Water Production System, Brazil, Atmosphere, 14, 1836, https://doi.org/10.3390/atmos14121836, 2023. a

Thom, A. and Oliver, H.: On Penman's equation for estimating regional evaporation, Q. J. Roy. Meteor. Soc., 103, 345–357, 1977. a, b

Thorp, K. R.: pyfao56: FAO-56 evapotranspiration in Python, SoftwareX, 19, 101208, https://doi.org/10.1016/j.softx.2022.101208, 2022. a

Trnka, M., Rötter, R. P., Ruiz-Ramos, M., Kersebaum, K. C., Olesen, J. E., Žalud, Z., and Semenov, M. A.: Adverse weather conditions for European wheat production will become more frequent with climate change, Nat. Clim. Change, 4, 637–643, https://doi.org/10.1038/nclimate2242, 2014. a, b

Turc, L.: Estimation of irrigation water requirements, potential evapotranspiration: a simple climatic formula evolved up to date, Ann. Agron., 12, 13–49, 1961. a, b

unittest: unittest, https://docs.python.org/3/library/unittest.html (last access: 17 September 2024), 2022. a

Valiantzas, J. D.: Simplified versions for the Penman evaporation equation using routine weather data, J. Hydrol., 331, 690–702, https://doi.org/10.1016/j.jhydrol.2006.06.012, 2006. a

Valipour, M.: Evaluation of radiation methods to study potential evapotranspiration of 31 provinces, Meteorol. Atmos. Phys., 127, 289–303, https://doi.org/10.1007/s00703-014-0351-3, 2015. a

Van Vuuren, D. P., Edmonds, J., Kainuma, M., Riahi, K., Thomson, A., Hibbard, K., Hurtt, G. C., Kram, T., Krey, V., Lamarque, J.-F., Masui, T., Meinshausen, M., Nakicenovic, N., Smith, S. J., and Rose, S. K.: The representative concentration pathways: an overview, Clim. Change, 109, 5–31, https://doi.org/10.1007/s10584-011-0148-z, 2011. a

Vaz, P. J., Schütz, G., Guerrero, C., and Cardoso, P. J. S.: A Study on the Prediction of Evapotranspiration Using Freely Available Meteorological Data, in: Computational Science – ICCS 2022, edited by: Groen, D., de Mulatier, C., Paszynski, M., Krzhizhanovskaya, V. V., Dongarra, J. J., and Sloot, P. M. A., 436–450, Springer International Publishing, Cham, ISBN 978-3-031-08760-8, 2022. a

Velázquez, J. A., Schmid, J., Ricard, S., Muerth, M. J., Gauvin St-Denis, B., Minville, M., Chaumont, D., Caya, D., Ludwig, R., and Turcotte, R.: An ensemble approach to assess hydrological models' contribution to uncertainties in the analysis of climate change impact on water resources, Hydrol. Earth Syst. Sci., 17, 565–578, https://doi.org/10.5194/hess-17-565-2013, 2013. a

Virtanen, P., Gommers, R., Oliphant, T. E., Haberland, M., Reddy, T., Cournapeau, D., Burovski, E., Peterson, P., Weckesser, W., Bright, J., van der Walt, S. J., Brett, M., Wilson, J., Millman, K. J., Mayorov, N., Nelson, A. R. J., Jones, E., Kern, R., Larson, E., Carey, C. J., Polat, Ä., Feng, Y., Moore, E. W., VanderPlas, J., Laxalde, D., Perktold, J., Cimrman, R., Henriksen, I., Quintero, E. A., Harris, C. R., Archibald, A. M., Ribeiro, A. H., Pedregosa, F., van Mulbregt, P., and SciPy 1.0 Contributors: SciPy 1.0: Fundamental Algorithms for Scientific Computing in Python, Nature Methods, 17, 261–272, https://doi.org/10.1038/s41592-019-0686-2, 2020. a, b

Vremec, M. and Collenteur, R.: pyet-pypi, https://pypi.org/project/pyet/ (last access: 17 September 2024), 2022. a

Vremec, M. and Collenteur, R.: PyEt: A Python package for estimating potential evapotranspiration, Zenodo [code], https://doi.org/10.5281/ZENODO.5896799, 2024a. a, b

Vremec, M. and Collenteur, R.: pyet – Estimation of Potential Evapotranspiration, PyData [code and data set], https://PyEt.readthedocs.io (last access: 17 September 2024), 2024b. a

Vremec, M., Forstner, V., Herndl, M., Collenteur, R., Schaumberger, A., and Birk, S.: Sensitivity of evapotranspiration and seepage to elevated atmospheric C O 2 from lysimeter experiments in a montane grassland, J. Hydrol., 617, 128875, https://doi.org/10.1016/j.jhydrol.2022.128875, 2022. a, b, c, d, e

Walter, I. A., Allen, R. G., Elliott, R., Jensen, M., Itenfisu, D., Mecham, B., Howell, T., Snyder, R., Brown, P., Echings, S., Spofford, T., Hattendorf, M., Martin, D., Cuenca, R. H., and Wright, L.: ASCE's standardized reference evapotranspiration equation, in: Watershed management and operations management 2000, pp. 1–11, 2000. a, b, c

Wang, K. and Dickinson, R. E.: A review of global terrestrial evapotranspiration: Observation, modeling, climatology, and climatic variability, Rev. Geophys., 50, https://doi.org/10.1029/2011RG000373, 2012. a

Webber, H., Gaiser, T., Oomen, R., Teixeira, E., Zhao, G., Wallach, D., Zimmermann, A., and Ewert, F.: Uncertainty in future irrigation water demand and risk of crop failure for maize in Europe, Environ. Res. Lett., 11, 074007, https://doi.org/10.1088/1748-9326/11/7/074007, 2016. a

Wood, E. F., Lettenmaier, D. P., Liang, X., Lohmann, D., Boone, A., Chang, S., Chen, F., Dai, Y., Dickinson, R. E., Duan, Q., Ek, M., Gusev, Y. M., Habets, F., Irannejad, P., Koster, R., Mitchel, K. E., Nasonova, O. N., Noilhan, J., Schaake, J., Schlosser, A., Shao, Y., Shmakin, A. B., Verseghy, D., Warrach, K., Wetzel, P., Xue, Y., Yang, Z.-L., and Zeng, Q.-C.: The Project for Intercomparison of Land-surface Parameterization Schemes (PILPS) Phase 2(c) Red–Arkansas River basin experiment, Global Planet. Change, 19, 115–135, https://doi.org/10.1016/S0921-8181(98)00044-7, 1998. a

Wright, J. L.: New evapotranspiration crop coefficients, Proceedings of the American Society of Civil Engineers, J. Irr. Drain. Div.-ASCE, 108, 57–74, 1982. a, b

Xiang, K., Li, Y., Horton, R., and Feng, H.: Similarity and difference of potential evapotranspiration and reference crop evapotranspiration – a review, Agric. Water Manage., 232, 106043, https://doi.org/10.1016/j.agwat.2020.106043, 2020. a

Xie, Z., Yao, Y., Zhang, X., Liang, S., Fisher, J. B., Chen, J., Jia, K., Shang, K., Yang, J., Yu, R., Guo, X., Liu, L., Ning, J., and Zhang, L.: The Global LAnd Surface Satellite (GLASS) evapotranspiration product Version 5.0: Algorithm development and preliminary validation, J. Hydrol., 610, 127990, https://doi.org/10.1016/j.jhydrol.2022.127990, 2022. a

Xu, C.-Y. and Singh, V. P.: Evaluation and generalization of radiation-based methods for calculating evaporation, Hydrol. Process., 14, 339–349, 2000. a, b, c, d, e, f, g, h, i

Xu, C.-Y. and Singh, V. P.: Evaluation and generalization of temperature-based methods for calculating evaporation, Hydrol. Process., 15, 305–319, https://doi.org/10.1002/hyp.119, 2001. a, b, c, d, e

Yang, Y., Roderick, M., Zhang, S., McVicar, T., and Donohue, R.: Hydrologic implications of vegetation response to elevated CO2 in climate projections, Nat. Clim. Change, 9, 44–48, https://doi.org/10.1038/s41558-018-0361-0, 2019. a, b, c, d, e, f, g, h

Yang, Y., Chen, R., Han, C., and Liu, Z.: Evaluation of 18 models for calculating potential evapotranspiration in different climatic zones of China, Agric. Water Manage., 244, 106545, https://doi.org/10.1016/j.agwat.2020.106545, 2021.  a, b, c

Zhou, J., Wang, Y., Su, B., Wang, A., Tao, H., Zhai, J., Kundzewicz, Z. W., and Jiang, T.: Choice of potential evapotranspiration formulas influences drought assessment: A case study in China, Atmos. Res., 242, 104979, https://doi.org/10.1016/j.atmosres.2020.104979, 2020. a

Download
Short summary
Geoscientists commonly use various potential evapotranpiration (PET) formulas for environmental studies, which can be prone to errors and sensitive to climate change. PyEt, a tested and open-source Python package, simplifies the application of 20 PET methods for both time series and gridded data, ensuring accurate and consistent PET estimations suitable for a wide range of environmental applications.