Next Article in Journal
Direct and Activated Chlorine Dioxide Oxidation for Micropollutant Abatement: A Review on Kinetics, Reactive Sites, and Degradation Pathway
Previous Article in Journal
In Search of Periodicity in the Annual Precipitation in Europe (1881–2020)
Previous Article in Special Issue
Estimating the Soil Erosion Response to Land-Use Land-Cover Change Using GIS-Based RUSLE and Remote Sensing: A Case Study of Miyun Reservoir, North China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Artificial Neural Networks for the Prediction of the Reference Evapotranspiration of the Peloponnese Peninsula, Greece

Department of Geology, University of Patras, 26504 Patras, Greece
*
Author to whom correspondence should be addressed.
Water 2022, 14(13), 2027; https://doi.org/10.3390/w14132027
Submission received: 25 May 2022 / Revised: 19 June 2022 / Accepted: 21 June 2022 / Published: 24 June 2022
(This article belongs to the Special Issue Remote Sensing Application on Soil Moisture)

Abstract

:
The aim of the study was to investigate the utility of artificial neural networks (ANNs) for the estimation of reference evapotranspiration (ETo) on the Peloponnese Peninsula in Greece for two representative months of wintertime and summertime during 2016–2019 and to test if using fewer inputs could lead to satisfactory predictions. Datasets from sixty-two meteorological stations were employed. The available inputs were mean temperature (Tmean), sunshine (N), solar radiation (Rs), net radiation (Rn), vapour pressure deficit (es-ea), wind speed (u2) and altitude (Z). Nineteen Multi-layer Perceptron (MLP) and Radial Basis Function (RBF) models were tested and compared against the corresponding FAO-56 Penman Monteith (FAO PM) estimates of a previous study, via statistical indices. The MLP1 7-2 model with all the variables as inputs outperformed the rest of the models (RMSE = 0.290 mm d−1, R2 = 98%). The results indicate that even ANNs with simple architecture can be very good predictive models of ETo for the Peloponnese, based on the literature standards. The MLP1 model determined Tmean, followed by u2, as the two most influential factors for ETo. Moreover, when one input was used (Tmean, Rn), RBFs slightly outperformed MLPs (RMSE < 0.385 mm d−1, R2 ≥ 96%), which means that even a sole-input ANN resulted in satisfactory predictions of ETo.

Graphical Abstract

1. Introduction

Reference Evapotranspiration (ETo) is a key climate parameter investigated in the frame of climate crisis and water resources management [1,2,3,4]. Moreover, ETo has an effect on the productive sector of agriculture [5]. In particular, the precision irrigation techniques and the decision-making irrigation systems demand accurate ETo values [6]. Therefore, the importance of effective estimation and prediction methods of ETo is crucial [7,8].
The necessity of acquiring ETo values led to the development of several methods of estimation, ranging from simple empirical or physically based models [9,10] to complex neuro-fuzzy and machine learning algorithms [11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47]. The methods incorporate data from meteorological stations or, due to the scarcity of the former, remotely sensed data [48,49,50,51,52,53,54,55,56,57,58]. The FAO-56 Penman Monteith (FAO PM) equation requires numerous meteorological variables for effective application [59,60]. Even the more refined empirical methods of ETo, such as Valiantzas’ equations [61], share a common denominator; the more meteorological inputs, the higher the accuracy of ETo [62]. On the other hand, there are empirical methods requiring limited climatic variables (e.g., the Hargreaves’-Samani equation). However, the accuracy of these methods, compared to FAO PM, is limited and performance in some cases has been reported to be season-dependent [63]. The unavailability of input data is a global issue, due to the high cost of equipping and running meteorological stations, especially for developing countries. Thus, reducing the number of inputs for the predicting models to only trivial parameters is highly recommended [64]. Those parameters can be air temperature (minimum, maximum and mean values) plus parameters that can be either measured or easily computed based on latitude and Julian day (e.g., sunshine, solar radiation, net radiation, vapor pressure deficit and wind speed) [59]. Computing the missing parameters is a commonly used practice in ETo research [61,65]. However, even if input parameters are trivial, their inner relationships are non-linear and complex. Among different methods, Artificial Neural Networks (ANNs) stand out for their capability to model complex non-linear relationships. ANNs demand short real-time control and computational effort [66]. Hence, they are considered as a suitable approach to model ETo.
Artificial neural networks are utilized in a broad spectrum of applications. ANN architectures consist of one input layer (including several input variables), multiple hidden layers and one output layer. The hidden layers consist of several neurons (or nodes) and each neuron can have multiple inputs and one output. The output of each neuron is connected to a neuron of the successive layer via a synapse. Each synapse is a link which has a particular weight to represent the strength of the connection between two successive neurons. After computing the sum of all weighted inputs, neurons transfer this sum into an output via an activation function. This output is fed to the next layer. The output of the network is computed by the last neuron of the network [67]. Multi-layer Perceptron (MLP) and Radial Basis Function (RBF) are common ANNs with capabilities in ETo forecasting [68,69,70,71,72].
Trajkovic (2009) [68], based on Trajkovic et al. (2003) [69], demonstrated that an RBF model with only two input variables is able to predict ETo with greater accuracy than FAO PM, amongst other equations. Kumar et al. (2020) [70] tested two models with five and two input variables, respectively, for the Himalayan region, employing ground-based data from forty-seven stations. They found that MLP could satisfactorily predict ETo even with limited inputs. Tabari et al. (2012) [71] investigated four MLP models with four, three and two input variables, for a semi-arid region in Iran. They concluded that the model with the most inputs resulted in the most accurate predictions of ETo and that tanh was the best activation function in all cases. Sattari et al. (2021) tried ten combinations of input variables (from eight to one inputs) [72]. The best performance yielded an R2 = 0.978. Shamshirband et al. (2016) [73] used data from twelve meteorological stations in Serbia for optimizing an ANN with very low error levels, but an R2 below 0.97. Gavili et al. (2017) [74] tested soft computing methods versus empirical ones and deduced that the former exhibited better performance in modeling ETo. Amongst these methods, the ANN had the best performance. As far as Greece is concerned, Diamantopoulou et al. (2011) [75] tested two ANNs to estimate daily ETo with limited meteorological data, separately for four meteorological stations in the Northern Greece. The best results obtained an RMSE = 0.545 mm d−1 and an R2 = 95.7%. Antonopoulos et al. (2016) [76] applied ANNs also in the Northern Greece (Lake Vegoritis) to predict ETo. For years 2003 and 2004, they obtained: RMSE = 0.69 mm d−1, R2 = 79.2% and RMSE = 1.09 mm d−1, R2 = 82.8%, respectively. Antonopoulos and Antonopoulos (2017) [77], in consecutive research in the same study area, trained ANNs with datasets from one meteorological station (Amyntaio) over the period 2009–2013 to estimate ETo. Their best results yielded an RMSE = 0.574 mm d−1 and an R2 around 97.2%, while the indices’ values were slightly deteriorated for lesser (i.e., three and two) inputs.
Although ANN is considered a suitable method to model ETo, the literature lacks in studies of areas with interchangeable land use/land cover (LU/LC) and relief characteristics over short distances. In those areas the performance of a model predicting ETo would be challenging in terms of accuracy and seasonal consistency. Moreover, the number of meteorological stations employed in the literature is usually low compared to the corresponding study area, which leads to a coarser spatial representation of the climate variables. Lastly, as shown above, there are only a few applications of ANNs for ETo prediction in Greece in the literature.
The aim of the study was to investigate whether ANNs could predict ETo satisfactorily close to the reference method (FAO PM) in the Peloponnese, Greece, for the selected period. The study also aimed to determine the model that best fits the data, based on the literature standards, and to indicate the most influential factors on ETo. This study examines nineteen MLP and RGB (ANN) models utilizing sixty-two meteorological stations in the Peloponnese peninsula in Southwestern Greece (Table A1). The Peloponnese peninsula is a study area with distinguished differences over short distances. Empirical methods applied across the area in a previous study [63] showed inferior performance in terms of accuracy, especially for August (summertime). A further objective was to examine whether fewer input variables would produce satisfactory predictions of ETo in terms of accuracy. Therefore, the number of inputs of the models were consecutively reduced to examine whether fewer than the seven available inputs, or even a sole input, could predict ETo acceptably. The novelty of the study is twofold: it is the first time that ANNs are used to predict the ETo across a large area of Southern Greece, which includes elevation, relief and LU/LC interchanges that are challenging in terms of the consistency of model performance. Even more challenging is the usage of continuous climate datasets over a recent four-year period, instead of datasets for individual days or short intervals. In addition, the exploitation of a dense meteorological station network, instead of a couple of stations, enhances the significance of the results. A good performance from ANNs, despite the aforementioned difficulties, would prove the flexibility and the potential of ANNs in ETo modeling. The latter would be useful in cases where there is a shortage of climate data. Moreover, it could save time and decrease computational load for interdisciplinary research purposes.

2. Materials and Methods

2.1. The Study Area

The Peloponnese is a peninsula in Southwestern Greece that occupies about 1/6th of the Greek territory (21,439 km2), with a population of 1086.935 (census 2011; https://www.statistics.gr/el/statistics/-/publication/SAM03/2011 (accessed on 10 March 2022). A large part is covered by high hills and mountains, running NW to SE, with an elevation up to 2407 m. Lithology, tectonic activity and climate conditions have resulted in the relief formation of the study area. A well-developed hydrographic network, though with few large rivers, has formed [78]. Based on the latest Copernicus LU/LC classification, the widest urban area is located at the northmost edge [79]. In addition to urban areas, the main LU/LC types are forest and transitional vegetation, as well as crop plots covering the plains (Figure 1). The broadest plain lies over the western coastal part. According to Köppen-Geiger’s classification, the climate of the Peloponnese is Mediterranean warm temperate with dry summers and mild winters (Csa) [80]. The annual normal measurements (1971–2000) of air temperature, precipitation and sunshine range between 8–20 °C, 400 to over 2000 mm and 1900–3100 h, respectively (http://climatlas.hnms.gr/sdi/?lang=EN (accessed on 27 April 2022).

2.2. Methods

Meteorological datasets of daily measurements from sixty-two stations under the National Observatory of Athens, for the months August and December of 2016–2019, were utilized (Table A1). These months and years were selected for the application of ANNs in methodological consistency with our previous study [63], where ETo was computed for the Peloponnese by FAO PM, our reference method. In that study, August and December had been selected as typical months of summer and winter, respectively. FAO PM has been widely used as a reference method to estimate ETo in studies investigating ANNs, either as the only reference method [10,70,81] or combined with direct methods [82]. Multi-Layer Perceptron (MLP) and Radial Basis Function (RBF) ANN models were examined in this study. Nineteen combinations of input variables were tested (from seven inputs to one input) with both MLP and RBF models. The inputs included climate and non-climate (such as altitude) variables that affect ETo [83], namely, solar radiation (Rs), net radiation (Rn), sunshine hours (N), mean air temperature (Tmean), vapour pressure deficit (es-ea), wind speed at 2 m from the surface (u2) and altitude (Z). The first three were previously calculated as functions of the station latitude and the Julian day, as in Zanetti et al. (2008) [36]. Based on Rahimikhoob (2020) [84], the input combinations were subsequently limited, aiming to explore the possibility of producing acceptably accurate predictions of ETo with fewer variables than the available seven or even with a sole input variable (such as Tmean or Rn). About three fifths of the sample data were used for training and the other two fifths were used for testing and validation (about one fifth each) [40]. The tested architecture was based on the trial-and-error method and the ANNs were trained using the Levenberg-Marquardt algorithm [40,41]. Hyperbolic tangent (Equation (1); [85]) was the utilized activation function, based on the literature [71,86]. The hyperbolic tangent, along with the sigmoid function, is a non-linear function widely used as an activation function in ANNs [85]. The hyperbolic tangent exhibits the advantage of giving higher enhancement to the negative values. The output of the former spans in [−1, 1] while the sigmoid outputs are only half of the previous ([0, 1]) [87].
T a n h ( x ) = e x e x e x + e x
The values predicted by the ANN models were then compared against the values by FAO PM via statistical indices that were computed during the training, testing and validation phases (SSE, RE) (Table 1). Furthermore, the error levels between predicted and reference values have been computed via measures such as Root Mean Square Error (RMSE, mm d1), Normalized Root Mean Square Error (NRMSE, %), Mean Absolute Error (MAE, mm d1), Mean Bias (MB, mm d1) and Sum of Squares Error (SSE, mm2 d1) (Table 1). The error values depict the magnitude of the computed ETo values, except from NRMSE, which is expressed in *100%. Mean Bias is a signed measure so that, in addition to producing the Mean Bias Error (MBE = |MB|), it provides some extra information; the minus sign depicts that the reference ETo value (by FAO PM) is greater than the model predicted value and vice versa. Moreover, three measures that express correlation, strength of fit or agreement were used, namely, Pearson correlation coefficient (Pearson’s r), coefficient of determination (R2) and Index of Agreement (IoA) (Table 1).
In Table 1, pi stands for the ith predicted value by an ANN, oi stands for the ith observed value, which in this study is the ith reference value estimated by FAO PM, o is the mean of the observed (reference) values and n = 290 is the sample size.

3. Results

In total, nineteen models with different input combinations were tested. The results are presented in Table 2.
As presented in Table 2, MLP1 7-2, with all the available parameters as inputs, appears to provide the best performance, expressed via the minimum error values in the testing and validation data and the maximum correlation and agreement indices (R2, Pearson’s r and IoA) between prediction and reference values. The architecture is simple, consisting of one hidden layer with two neurons (Figure 2).
The parameter-estimates of the MLP1 model are displayed in Table 3. As presented, the values of both the nodes and the hidden layers are not negligible values, a fact that constitutes one of the desirable characteristics of an ANN model.
As presented in Table 2, MLP9 6-4-3 follows closely the model with the superior performance (MLP1), except for the testing SSE value, which is almost double the corresponding MLP1 value. It is interesting that the MLP10 4-3-2 model, with only four inputs and two hidden layers, exhibits a performance very close to that of MLP1, with a couple of values (MBE and SSE) even better than those of MLP1. The added value is that it produces almost the same values with only four basic parameters as inputs, specifically, Rs, es-ea, u2 and Tmean. The data of the aforementioned parameters are usually available either from meteorological stations or via remote sensing. Moreover, in the event of any missing data, the parameters can be easily computed [59]. It is obvious that the RBF network displays inferior performance to MLP for our data. RBF3 6-9 is the best among the RBF networks and fourth in the total rank of model performance. The relative errors of MLPs in the validation phase were between 2.2–4.2% for all trials, whereas for RBFs the relative errors lay between 2.8–10%. For the majority of the models, the holdout RE values were greater than the testing RE values, despite the satisfactory error levels (2.8–4.5%). This indicates that those models were overtrained towards the testing data.
The cases where only a sole basic parameter (Tmean or Rn) was set as input seem noteworthy. In those cases, the RBF models exhibited better performance than the MLP ones. RBF9 (Rn as input) obtained RMSE = 0.383 mm d1 and R2 = 96.5% with RE = 3.3%. In the case of Tmean as the sole input, the RMSE was around 0.360 mm d1 and the R2 was above 96.7%. However, as previously commented, there was evidence of overtraining of some models (Table 2). Therefore, those models are not recommended. The influence of each climatic variable on ETo presented is considered important in respect of the ANN that best fits the data and bears interesting determination [14]. As shown in Figure 3, the most influential factor is Tmean followed by u2. The third factor in the rank is vapour pressure deficit (es-ea). Those results are aligned with the ETo estimates by FAO PM of the same period for the Peloponnese [63].

4. Discussion

The determination of ETo is crucial for climate crisis research, hydrological cycle, water resources management and irrigation precision techniques. The available methods of estimation require the knowledge of numerous climate parameters, which means an increased cost. Since ETo includes complex and nonlinear relationships, ANNs have been proven to be a suitable modeling choice. In this study, ETo has been estimated via nineteen MLP and RBF ANN architectures, with different combination and number of inputs (from seven to one).
This period is interesting in the frame of the climate crisis, since it is recent and includes the first (2016) and second (2019) warmest years since the preindustrial era, which challenges the performance of ANNs.
For the majority of the trials, the RMSE was below 0.4 mm d−1 and the relative error of the testing phase was below 4%. MLPs performed generally better than RBFs for multiple inputs, whereas RBFs performed slightly better when only one input (Rn or Tmean) was set. The model that best fits the reference values was that with the most input parameters and only one hidden layer (MLP1 7-2), bearing RMSE = 0.290 mm d−1, R2 = 98% and RE of testing and validation phases equal to 2.7% and 2.2%, respectively. The different runs with the same input combination, but different model architecture, showed that any increase in the number of hidden layers and the number of neurons in the hidden layer exhibited negligible improvement in prediction accuracy. Those conclusions are in line with the findings by Tabari et al. (2012) [71]. The data of the seven used inputs can be easily derived, either by meteorological stations or via remote sensing, or can be easily computed as missing data by FAO guidelines [59]. However, during the trials, the number of the inputs was gradually limited in order to examine whether ANNs can provide satisfactory estimates of ETo, when incorporating only the most commonly available climate data. It is interesting that the MLP10 with four climate parameters as inputs exhibited results very close to the best model (MLP1). Moreover, models with two basic parameters as inputs, exhibited RMSE up to 0.352 mm d−1, testing RE below 3% and R2 at least equal to 97% (MLP6, MLP13). When only one input was used (i.e., Tmean or Rn) in RBF models, the RMSE was below 0.385 mm d−1, testing RE was below 3.6% and R2 was at least equal to 96%. According to the literature, those values are considered very good to excellent. For example, Rahimikhoob (2010) recommended an ANN for the coastal area of the Caspian Sea in Northern Iran, which used only air temperature as an input, with an RMSE = 0.41 mm d−1 and R2 = 95% [84], whereas in this study MLP14 with Tmean as an input has better accuracy (RMSE = 0.360 mm d−1 and R2 = 96.9%). In the same vein, Zanetti et al. (2008) used MLP ANN with only temperature and radiation inputs to predict ETo for Campos dos Goytacazes, Brazil [66], while Ravindran et al. (2021) deduced that Rs was the most influential parameter to ETo and used it as a sole input in ANNs for California (R2 up to 95.4%) [14]. This proves that ANNs with simple architecture can be good predictive models of ETo over the Peloponnese for the examined period. In addition, based on the best ANN model (MLP1), we found that Tmean and u2 were the two most influential factors on ETo, out of the seven examined. This is in line with the findings of a previous study of the same period for the Peloponnese, where ETo had been computed by FAO PM [63]. Tmean, as a proxy of the energy state of the system, is one of the most influential factors on ETo. This depicts the altitude and land cover difference over short distances across the Peloponnese. Probably, due to the fact that the Peloponnese has a very low variance in latitude, the radiation factors were not that influential for the determination of the ETo. Regarding the second most influential factor (u2), in contrast with December (winter), increased values of u2 are not frequent in Greece during August (summer). Therefore, where increased u2 values occurred in August, they affected the determination of the local ETo values. In conclusion, ANNs resulted in predictions very close to FAO PM, which is the most established reference method, for the examined period for the Peloponnese. Therefore, ANNs present the potential of general usage in modeling ETo across Greece, after further investigation.
This study proves that ANNs can be useful alternatives for predicting ETo, requiring limited climate data as input. The former is of considerable usefulness in cases where climate data are in shortage or the cost of the meteorological stations is not affordable, such as in developing countries. Despite the cost, ANNs provide a time saving and a computational effort-decreasing alternative to the complex algorithms currently used, which is applicable for interdisciplinary research purposes.

5. Conclusions

Among the tested ANNs, MLPs performed generally better than RBFs for multiple inputs, whereas RBFs performed slightly better when one sole input (Rn or Tmean) was set. The results revealed better performance of the MLP1 7-2 model, with all the available variables as inputs and only one hidden layer, bearing an RMSE = 0.290 mm d−1, R2 = 98% and RE values of testing and validation phases equal to 2.7% and 2.2%, respectively. The former proves that even simple ANN architectures can constitute very satisfactory predicting models. Models with only two parameters as inputs exhibited RMSE values up to 0.352 mm d−1 and R2 values at least equal to 97% (MLP6, MLP13). When one sole input was used (Tmean or Rn) in RBF models, RMSE was below 0.385 mm d−1 and R2 was at least equal to 96%. The results in both cases are very satisfactory. The MLP1, which outperformed the rest of the ANNs, determined the order of importance of parameters that affect ETo. The first two most influential parameters were Tmean and u2. Tmean is commonly a parameter to which ETo variances are attributed, as it depicts the overall energy state of the system. For the Peloponnese, where the variance of the latitude (and consequently of solar radiation) values is minor, Tmean variances occur mostly due to distinguished differences in relief, LU/LC types and proximity to the coast over short distances. Wind speed (u2) plays a substantial role, especially in August, when any increased u2 values directly affect the ETo values, since those are not frequent in summertime. Future research could test the MLP1 performance for a larger period and across different areas of Greece that differentiate in micro-climatic conditions and regimes. Moreover, direct measurements such as pan evaporation measurements employing ANNs could be investigated on a local scale. Another interesting idea, based on the satisfactory results of this study regarding a sole input variable, would be to explore the potential of multilinear regression analysis, which is a simpler method and comprehensible by a wider interdisciplinary audience.

Author Contributions

Conceptualization, S.D. and K.G.N.; methodology, S.D. and K.G.N.; software, S.D.; validation S.D.; formal analysis, S.D.; investigation, S.D., data curation, S.D., writing—original draft preparation, S.D., writing and editing, S.D. and K.G.N.; supervision, K.G.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are private, available only for the review process of the present article.

Acknowledgments

The authors acknowledge the National Observatory of Athens (https://meteosearch.meteo.gr (accessed on 15 April 2022)) for ground-based data availability of sixty-two meteorological stations.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. Meteorological stations (62) used (source: https://meteosearch.meteo.gr (accessed on 15 April 2022)).
Table A1. Meteorological stations (62) used (source: https://meteosearch.meteo.gr (accessed on 15 April 2022)).
IDStationXYElevation (m)MunicipalityIDStationXYElevation (m)Municipality
Meteorological Stations for the 3 Empirical Methods (ETo)Meteorological Stations for the 3 Empirical Methods (ETo)
1Kalavrita334349.94210128781Achaia32Oleni282783.4417787261Ilia
2Kato Vlassia317683.44208558773Achaia33Pineia285425.34191240184Ilia
3Lappa273550421892815Achaia34Pirgos273886.9417189122Ilia
4Olenia288845.1422165434Achaia35Vartholomio253773.8419312715Ilia
5Panachaiko313491.442358001588Achaia36Zacharo290302.641508065Ilia
6Panagopoula318709.5424384215Achaia37Amoni Sofikou424227.5418689855Korinthia
7Panepistimio305972.3423928966Achaia38Derveni362057.142217375Korinthia
8Patra301697.842366946Achaia39Isthmos408645.442004996Korinthia
9Rio305898.142421772Achaia40Kiato389163.5420772215Korinthia
10Romanos313476.14235744228Achaia41Krioneri378491.94203310887Korinthia
11Sageika280638.4421957526Achaia42Loutraki410248.7420263630Korinthia
12Argos386329.1416505938Argolida43Nemea381197.94188976290Korinthia
13Didima426936.94146702175Argolida44Perigiali397303.1419934438Korinthia
14Kranidi424615.74137411110Argolida45Trikala Korinthias365493.742068351077Korinthia
15Lagadia326139.94172057970Arkadia46Agioi Theodoroi423533.6419839537Korinthia
16Levidi349386.54171330853Arkadia47Apidia392819.74082655230Lakonia
17Lykochia337772.64151113870Arkadia48Asteri386527.140767578Lakonia
18Magouliana334497.741712751256Arkadia49Geraki384706.64094508330Lakonia
19Megalopoli3353634140782432Arkadia50Krokees371576.24082640241Lakonia
20Stemnitsa330377.841579671094Arkadia51Molaoi397984.64072957128Lakonia
21Tripoli359989.34152250650Arkadia52Monemvasia413811.4405905117Lakonia
22Vytina339989.841704091013Arkadia53Sparti360929.94101670204Lakonia
23Spetses424919.541246623Attiki54Alagonia343840.94107863765Messinia
24Taktikoupoli Troizinias443373.2415237415Attiki55Arfara326299.4411366696Messinia
25Ydra452645.841337272Attiki56Filiatra285439.9411517565Messinia
26Amaliada264604.9418692326Ilia57Kalamata33112740989745Messinia
27Andritsaina314220.34152125731Ilia58Kalamata Dytika329347.3410000110Messinia
28Archaia Olympia287981.3416385645Ilia59Kardamili347857.7407465113Messinia
29Foloi297082.74174732600Ilia60Kopanaki306288.64128741184Messinia
30Katakolo263537.241693272Ilia61Kyparissia291691412358436Messinia
31Lampeia306840.34192041840Ilia62Pylos294556.840875905Messinia

References

  1. Malamos, N.; Tegos, A. Advances in Evaporation and Evaporative Demand. Hydrology 2022, 9, 78. [Google Scholar] [CrossRef]
  2. Shamshirband, S.; Hashemi, S.; Salimi, H.; Samadianfard, S.; Asadi, E.; Shadkani, S.; Kargar, K.; Mosavi, A.; Nabipour, N.; Chau, K.W. Predicting Standardized Streamflow index for hydrological drought using machine learning models. Eng. Appl. Comput. Fluid Mech. 2020, 14, 339–350. [Google Scholar] [CrossRef]
  3. Xu, S.; Yu, Z.; Yang, C.; Ji, X.; Zhang, K. Trends in evapotranspiration and their responses to climate change and vegetation greening over the upper reaches of the Yellow River Basin. Agric. For. Meteorol. 2018, 263, 118–129. [Google Scholar] [CrossRef]
  4. Dimitriadou, S.; Nikolakopoulos, K.G. Evapotranspiration Trends and Interactions in Light of the Anthropogenic Footprint and the Climate Crisis: A Review. Hydrology 2021, 8, 163. [Google Scholar] [CrossRef]
  5. Pereira, L.S.; Perrier, A.; Allen, R.G.; Alves, I. Evapotranspiration: Concepts and Future Trends. J. Irrig. Drain. Eng. 1999, 125, 45–51. [Google Scholar] [CrossRef]
  6. Malamos, N.; Tsirogiannis, I.L.; Tegos, A.; Efstratiadis, A.; Koutsoyiannis, D. Spatial interpolation of potential evapotranspiration for precision irrigation purposes. Eur. Water 2017, 59, 303–309. [Google Scholar]
  7. Saggi, M.K.; Jain, S. Reference evapotranspiration estimation and modeling of the Punjab Northern India using deep learning. Comput. Electron. Agric. 2019, 156, 387–398. [Google Scholar] [CrossRef]
  8. Zhang, Y.; Wei, Z.; Zhang, L.; Du, J. Applicability evaluation of different algorithms for daily reference evapotranspiration model in KBE system. Int. J. Comput. Sci. Eng. 2019, 18, 361–374. [Google Scholar] [CrossRef]
  9. Li, M.; Chu, R.; Islam, A.R.M.T.; Shen, S. Reference Evapotranspiration Variation Analysis and Its Approaches Evaluation of 13 Empirical Models in Sub-Humid and Humid Regions: A Case Study of the Huai River Basin, Eastern China. Water 2018, 10, 493. [Google Scholar] [CrossRef] [Green Version]
  10. Tabari, H.; Grismer, M.E.; Trajkovic, S. Comparative analysis of 31 reference evapotranspiration methods under humid conditions. Irrig. Sci. 2013, 31, 107–117. [Google Scholar] [CrossRef]
  11. Roy, D.K.; Sarkar, T.K.; Alam Kamar, S.S.; Goswami, T.; Muktadir, A.; Al-Ghobari, H.M.; Alataway, A.; Dewidar, A.Z.; El-Shafei, A.A.; Mattar, M.A. Daily Prediction and Multi-Step Forward Forecasting of Reference Evapotranspiration Using LSTM and Bi-LSTM Models. Agronomy 2022, 12, 594. [Google Scholar] [CrossRef]
  12. Ferreira, L.B.; da Cunha, F.F. Multi-step ahead forecasting of daily reference evapotranspiration using deep learning. Comput. Electron. Agric. 2020, 178, 105728. [Google Scholar] [CrossRef]
  13. Ravindran, S.M.; Bhaskaran, S.K.M.; Ambat, S.K.N. A Deep Neural Network Architecture to Model Reference Evapotranspiration Using a Single Input Meteorological Parameter. Environ. Process. 2021, 8, 1567–1599. [Google Scholar] [CrossRef]
  14. Granata, F.; Di Nunno, F. Forecasting evapotranspiration in different climates using ensembles of recurrent neural networks. Agric. Water Manag. 2021, 255, 107040. [Google Scholar] [CrossRef]
  15. Mosavi, A.; Edalatifar, M. A Hybrid Neuro-Fuzzy Algorithm for Prediction of Reference Evapotranspiration. In Recent Advances in Technology Research and Education; Laukaitis, G., Ed.; INTER-ACADEMIA 2018; Lecture Notes in Networks and Systems; Springer: Cham, Switzerland, 2019; Volume 53. [Google Scholar] [CrossRef]
  16. Andreu, A.; Kustas, W.P.; Polo, M.J.; Carrara, A.; González-Dugo, M.P. Modeling Surface Energy Fluxes over a Dehesa (Oak Savanna) Ecosystem Using a Thermal Based Two Source Energy Balance Model (TSEB) II—Integration of Remote Sensing Medium and Low Spatial Resolution Satellite Images. Remote Sens. 2018, 10, 558. [Google Scholar] [CrossRef] [Green Version]
  17. Silva, A.M.; da Silva, R.M.; Santos, C.A.G. Automated surface energy balance algorithm for land (ASEBAL) based on automating endmember pixel selection for evapotranspiration calculation in MODIS orbital images. Int. J. Appl. Earth Obs. Geoinf. 2019, 79, 1–11. [Google Scholar] [CrossRef]
  18. Zakeri, M.S.; Mousavi, S.F.; Farzin, S.; Sanikhani, H. Modeling of Reference Crop Evapotranspiration in Wet and Dry Climates Using Data-Mining Methods and Empirical Equations. J. Soft Comput. Civ. Eng. 2022, 6, 1–28. [Google Scholar] [CrossRef]
  19. Achite, M.; Jehanzaib, M.; Sattari, M.T.; Toubal, A.K.; Elshaboury, N.; Wałęga, A.; Krakauer, N.; Yoo, J.-Y.; Kim, T.-W. Modern Techniques to Modeling Reference Evapotranspiration in a Semiarid Area Based on ANN and GEP Models. Water 2022, 14, 1210. [Google Scholar] [CrossRef]
  20. Elbeltagi, A.; Kumar, N.; Chandel, A.; Arshad, A.; Pande, C.B.; Islam, A.R.M.T. Modelling the reference crop evapotranspiration in the Beas-Sutlej basin (India): An artificial neural network approach based on different combinations of meteorological data. Environ. Monit. Assess. 2022, 194, 141. [Google Scholar] [CrossRef]
  21. Elbeltagi, A.; Kushwaha, N.L.; Rajput, J.; Vishwakarma, D.K.; Kulimushi, L.C.; Kumar, M.; Zhang, J.; Pande, C.B.; Choudhari, P.; Meshram, S.G.; et al. Modelling daily reference evapotranspiration based on stacking hybridization of ANN with meta-heuristic algorithms under diverse agro-climatic conditions. Stoch. Hydrol. Hydraul. 2022, 1–24. [Google Scholar] [CrossRef]
  22. Long, X.; Wang, J.; Gong, S.; Li, G.; Ju, H. Reference evapotranspiration estimation using long short-term memory network and wavelet-coupled long short-term memory network. Irrig. Drain. 2022. [CrossRef]
  23. Chia, M.Y.; Huang, Y.F.; Koo, C.H. ANN-Based Reference Evapotranspiration Estimation: Effects of Data Normalization and Parameters Selection. In Proceedings of International Conference on Emerging Technologies and Intelligent Systems; Al-Emran, M., Al-Sharafi, M.A., Al-Kabi, M.N., Shaalan, K., Eds.; ICETIS 2021; Lecture Notes in Networks and Systems; Springer: Cham, Switzerland, 2022; pp. 3–12. [Google Scholar] [CrossRef]
  24. Ferreira, L.B.; da Cunha, F.F.; Filho, E.I.F. Exploring machine learning and multi-task learning to estimate meteorological data and reference evapotranspiration across Brazil. Agric. Water Manag. 2022, 259, 107281. [Google Scholar] [CrossRef]
  25. Manikumari, N.; Vinodhini, G.; Murugappan, A. Modelling of Reference Evapotransipration using Climatic Parameters for Irrigation Scheduling using Machine learning. ISH J. Hydraul. Eng. 2022, 28, 272–281. [Google Scholar] [CrossRef]
  26. Krishnashetty, P.H.; Balasangameshwara, J.; Sreeman, S.; Desai, S.; Kantharaju, A.B. Cognitive computing models for estimation of reference evapotranspiration: A review. Cogn. Syst. Res. 2021, 70, 109–116. [Google Scholar] [CrossRef]
  27. Algretawee, H.; Alshama, G. Modeling of Evapotranspiration (ETo) in a Medium Urban Park within a Megacity by Using Artificial Neural Network (ANN) Model. Period. Polytech. Civ. Eng. 2021, 65, 1260–1268. [Google Scholar] [CrossRef]
  28. Gao, L.; Gong, D.; Cui, N.; Lv, M.; Feng, Y. Evaluation of bio-inspired optimization algorithms hybrid with artificial neural network for reference crop evapotranspiration estimation. Comput. Electron. Agric. 2021, 190, 106466. [Google Scholar] [CrossRef]
  29. Yurtseven, I.; Serengil, Y. Comparison of different empirical methods and data-driven models for estimating reference evapotranspiration in semi-arid Central Anatolian Region of Turkey. Arab. J. Geosci. 2021, 14, 2033. [Google Scholar] [CrossRef]
  30. He, Y.; Lin, E.S.; Tan, C.L.; Tan, P.Y.; Wong, N.H. Quantitative evaluation of plant evapotranspiration effect for green roof in tropical area: A case study in Singapore. Energy Build. 2021, 241, 110973. [Google Scholar] [CrossRef]
  31. Gocić, M.; Amiri, M.A. Reference Evapotranspiration Prediction Using Neural Networks and Optimum Time Lags. Water Resour. Manag. 2021, 35, 1913–1926. [Google Scholar] [CrossRef]
  32. del Cerro, R.T.G.; Subathra, M.; Kumar, N.M.; Verrastro, S.; George, S.T. Modelling the daily reference evapotranspiration in semi-arid region of South India: A case study comparing ANFIS and empirical models. Inf. Process. Agric. 2020, 8, 173–184. [Google Scholar] [CrossRef]
  33. Khedkar, D.D.; Singh, P.K. Comparison of neural network models for estimation of reference crop evapotranspiration. Agric. Res. J. 2021, 58, 60–67. [Google Scholar] [CrossRef]
  34. Abdullahi, J.; Rotimi, A.; Malami, S.I.; Jibrin, H.B.; Tahsin, A.; Abba, S. Feasibility of artificial intelligence and CROPWAT models in the estimation of uncertain combined variable using nonlinear sensitivity analysis. In Proceedings of the 2021 1st International Conference on Multidisciplinary Engineering and Applied Science (ICMEAS), Abuja, Nigeria, 15–16 July 2021; pp. 1–7. [Google Scholar] [CrossRef]
  35. Reddy, K.C. Machine Intelligence-Based Reference Evapotranspiration Modelling: An application of Neural Networks. In Proceedings of the 2021 International Conference on Artificial Intelligence and Machine Vision (AIMV), Gandhinagar, India, 24–26 September 2021; pp. 1–5. [Google Scholar] [CrossRef]
  36. Nawandar, N.K.; Cheggoju, N.; Satpute, V. ANN-Based Model to Predict Reference Evapotranspiration for Irrigation Estimation. In Advances in Intelligent Systems and Computing: Proceedings of International Conference on Recent Trends in Machine Learning, IoT, Smart Cities and Applications; Gunjan, V.K., Zurada, J.M., Eds.; Springer: Singapore, 2021; Volume 1245, pp. 671–679. [Google Scholar] [CrossRef]
  37. Yamaç, S.S. Reference evapotranspiration estimation with k-Nearest Neighbour and Artificial neural network models using different climate input variables in the semi-arid environment. J. Agric. Sci. 2021, 27, 129–137. [Google Scholar] [CrossRef]
  38. Ferreira, L.B.; da Cunha, F.F.; da Silva, G.H.; Campos, F.B.; Dias, S.H.B.; Santos, J.E.O. Generalizability of machine learning models and empirical equations for the estimation of reference evapotranspiration from temperature in a semiarid region. Anais da Academia Brasileira de Ciências 2021, 93, e20200304. [Google Scholar] [CrossRef]
  39. Yu, H.; Wen, X.; Li, B.; Yang, Z.; Wu, M.; Ma, Y. Uncertainty analysis of artificial intelligence modeling daily reference evapotranspiration in the northwest end of China. Comput. Electron. Agric. 2020, 176, 105653. [Google Scholar] [CrossRef]
  40. Maroufpoor, S.; Bozorg-Haddad, O.; Maroufpoor, E. Reference evapotranspiration estimating based on optimal input combination and hybrid artificial intelligent model: Hybridization of artificial neural network with grey wolf optimizer algorithm. J. Hydrol. 2020, 588, 125060. [Google Scholar] [CrossRef]
  41. Alizamir, M.; Kisi, O.; Adnan, R.M.; Kuriqi, A. Modelling reference evapotranspiration by combining neuro-fuzzy and evolutionary strategies. Acta Geophys. 2020, 68, 1113–1126. [Google Scholar] [CrossRef]
  42. Tikhamarine, Y.; Malik, A.; Souag-Gamane, D.; Kisi, O. Artificial intelligence models versus empirical equations for modeling monthly reference evapotranspiration. Environ. Sci. Pollut. Res. 2020, 27, 30001–30019. [Google Scholar] [CrossRef]
  43. Seifi, A.; Riahi, H. Estimating daily reference evapotranspiration using hybrid gamma test-least square support vector machine, gamma test-ANN, and gamma test-ANFIS models in an arid area of Iran. J. Water Clim. Chang. 2020, 11, 217–240. [Google Scholar] [CrossRef]
  44. Nourani, V.; Elkiran, G.; Abdullahi, J. Multi-step ahead modeling of reference evapotranspiration using a multi-model approach. J. Hydrol. 2019, 581, 124434. [Google Scholar] [CrossRef]
  45. Zhu, B.; Feng, Y.; Gong, D.; Jiang, S.; Zhao, L.; Cui, N. Hybrid particle swarm optimization with extreme learning machine for daily reference evapotranspiration prediction from limited climatic data. Comput. Electron. Agric. 2020, 173, 105430. [Google Scholar] [CrossRef]
  46. Reis, M.M.; Silva, A.; Junior, J.Z.; Santos, L.T.; Azevedo, A.M.; Lopes, M.G. Empirical and learning machine approaches to estimating reference evapotranspiration based on temperature data. Comput. Electron. Agric. 2019, 165, 104937. [Google Scholar] [CrossRef]
  47. Ferreira, L.B.; da Cunha, F.F.; de Oliveira, R.A.; Filho, E.I.F. Estimation of reference evapotranspiration in Brazil with limited meteorological data using ANN and SVM—A new approach. J. Hydrol. 2019, 572, 556–570. [Google Scholar] [CrossRef]
  48. Sidiropoulos, P.; Dalezios, N.R.; Loukas, A.; Mylopoulos, N.; Spiliotopoulos, M.; Faraslis, I.N.; Alpanakis, N.; Sakellariou, S. Quantitative Classification of Desertification Severity for Degraded Aquifer Based on Remotely Sensed Drought Assessment. Hydrology 2021, 8, 47. [Google Scholar] [CrossRef]
  49. Raoufi, R.; Beighley, E. Estimating Daily Global Evapotranspiration Using Penman–Monteith Equation and Remotely Sensed Land Surface Temperature. Remote Sens. 2017, 9, 1138. [Google Scholar] [CrossRef] [Green Version]
  50. Allen, R.G.; Tasumi, M.; Trezza, R. Satellite-Based Energy Balance for Mapping Evapotranspiration with Internalized Calibration (METRIC)—Model. J. Irrig. Drain. Eng. 2007, 133, 380–394. [Google Scholar] [CrossRef]
  51. Dimitriadou, S.; Nikolakopoulos, K.G. Remote sensing methods to estimate evapotranspiration incorporating MODIS de-rived data and applications over Greece: A review. In Proceedings of the SPIE 11524, Eighth International Conference on Remote Sensing and Geoinformation of the Environment (RSCy2020), Paphos, Cyprus, 26 August 2020. [Google Scholar] [CrossRef]
  52. Liu, Y.; Zhang, S.; Zhang, J.; Tang, L.; Bai, Y. Assessment and Comparison of Six Machine Learning Models in Estimating Evapotranspiration over Croplands Using Remote Sensing and Meteorological Factors. Remote Sens. 2021, 13, 3838. [Google Scholar] [CrossRef]
  53. Proias, G.; Gravalos, I.; Papageorgiou, E.; Poczęta, K.; Sakellariou-Makrantonaki, M. Forecasting Reference Evapotranspiration Using Time Lagged Recurrent Neural Network. WSEAS Trans. Environ. Dev. 2020, 16, 699–707. [Google Scholar] [CrossRef]
  54. Kitsara, G.; Papaioannou, G.; Retalis, A.; Paronis, D.; Kerkides, P. Estimation of air temperature and reference evapotranspiration using MODIS land surface temperature over Greece. Int. J. Remote Sens. 2018, 39, 924–948. [Google Scholar] [CrossRef]
  55. Falalakis, G.; Gemitzi, A. A simple method for water balance estimation based on the empirical method and remotely sensed evapotranspiration estimates. J. Hydroinform. 2020, 22, 440–451. [Google Scholar] [CrossRef]
  56. Tsouni, A.; Kontoes, C.; Koutsoyiannis, D.; Elias, P.; Mamassis, N. Estimation of Actual Evapotranspiration by Remote Sensing: Application in Thessaly Plain, Greece. Sensors 2008, 8, 3586–3600. [Google Scholar] [CrossRef]
  57. Vasiliades, L.; Spiliotopoulos, M.; Tzabiras, J.; Loukas, A.; Mylopoulos, N. Estimation of crop water requirements using remote sensing for operational water resources management. In Proceedings of the Third International Conference on Remote Sensing and Geoinformation of the Environment (RSCy2015), Paphos, Cyprus, 19 June 2015. [Google Scholar] [CrossRef]
  58. Dimitriadou, S.; Nikolakopoulos, K.G. Development of GIS models via optical programming and python scripts to implement four empirical methods of reference and actual evapotranspiration (ETo, ETa) incorporating MODIS LST inputs. In Proceedings of the SPIE 11856, Remote Sensing for Agriculture, Ecosystems, and Hydrology XXIII, Madrid, Spain, 12 September 2021. [Google Scholar] [CrossRef]
  59. Allen, R.; Pereira, L.; Raes, D.; Smith, M. Crop Evapotranspiration-Guidelines for Computing Crop Water Requirements-FAO Irrigation and Drainage Paper 56; FAO: Rome, Italy, 1998; Volume 300, p. D05109. [Google Scholar]
  60. Pereira, L.S.; Allen, R.G.; Smith, M.; Raes, D. Crop evapotranspiration estimation with FAO56: Past and future. Agric. Water Manag. 2015, 147, 4–20. [Google Scholar] [CrossRef]
  61. Valiantzas, J.D. Simplified Reference Evapotranspiration Formula Using an Empirical Impact Factor for Penman’s Aerodynamic Term. J. Hydrol. Eng. 2013, 18, 108–114. [Google Scholar] [CrossRef]
  62. Djaman, K.; Rudnick, D.; Mel, V.C.; Mutiibwa, D.; Diop, L.; Sall, M.; Kabenge, I.; Bodian, A.; Tabari, H.; Irmak, S. Evaluation of Valiantzas’ Simplified Forms of the FAO-56 Penman-Monteith Reference Evapotranspiration Model in a Humid Climate. J. Irrig. Drain. Eng. 2017, 143, 06017005. [Google Scholar] [CrossRef]
  63. Dimitriadou, S.; Nikolakopoulos, K.G. Reference Evapotranspiration (ETo) Methods Implemented as ArcMap Models with Remote-Sensed and Ground-Based Inputs, Examined along with MODIS ET, for Peloponnese, Greece. ISPRS Int. J. Geo-Inf. 2021, 10, 390. [Google Scholar] [CrossRef]
  64. Tegos, A.; Efstratiadis, A.; Malamos, N.; Mamassis, N.; Koutsoyiannis, D. Evaluation of a Parametric Approach for Estimating Potential Evapotranspiration Across Different Climates. Agric. Agric. Sci. Procedia 2015, 4, 2–9. [Google Scholar] [CrossRef] [Green Version]
  65. Zanetti, S.S.; Sousa, E.F.; Oliveira, V.P.S.; Almeida, F.T.; Bernardo, S. Estimating Evapotranspiration Using Artificial Neural Network and Minimum Climatological Data. J. Irrig. Drain. Eng. 2007, 133, 83–89. [Google Scholar] [CrossRef]
  66. ASCE Task Committee on Application of Artificial Neural Networks in Hydrology. Artificial neural networks in hydrology. I: Preliminary concepts. J. Hydrol. Eng. 2000, 5, 115–123. [Google Scholar] [CrossRef]
  67. Shakiba, F.M.; Zhou, M. Novel Analog Implementation of a Hyperbolic Tangent Neuron in Artificial Neural Networks. IEEE Trans. Ind. Electron. 2020, 68, 10856–10867. [Google Scholar] [CrossRef]
  68. Huss, M.; Farinotti, D.; Bauder, A.; Funk, M. Modelling runoff from highly glacierized alpine drainage basins in a changing climate. Hydrol Process. 2009, 2274, 2267–2274. [Google Scholar] [CrossRef]
  69. Trajkovic, S.; Todorovic, B.; Stankovic, M. Forecasting of Reference Evapotranspiration by Artificial Neural Networks. J. Irrig. Drain. Eng. 2003, 129, 454–457. [Google Scholar] [CrossRef]
  70. Kumar, N.; Maharshi, S.; Poddar, A.; Shankar, V. Evaluation of Artificial Neural Networks for Estimating Reference Evapotranspiration in Western Himalayan Region. In Proceedings of the 2020 International Conference on Computational Performance Evaluation (ComPE), Shillong, India, 2–4 July 2020; pp. 163–167. [Google Scholar] [CrossRef]
  71. Tabari, H.; Talaee, P.H. Multilayer perceptron for reference evapotranspiration estimation in a semiarid region. Neural Comput. Appl. 2012, 23, 341–348. [Google Scholar] [CrossRef]
  72. Shamshirband, S.; Amirmojahedi, M.; Gocić, M.; Akib, S.; Petković, D.; Piri, J.; Trajkovic, S. Estimation of Reference Evapotranspiration Using Neural Networks and Cuckoo Search Algorithm. J. Irrig. Drain. Eng. 2016, 142, 04015044. [Google Scholar] [CrossRef]
  73. Sattari, M.T.; Apaydin, H.; Band, S.S.; Mosavi, A.; Prasad, R. Comparative analysis of kernel-based versus ANN and deep learning methods in monthly reference evapotranspiration estimation. Hydrol. Earth Syst. Sci. 2021, 25, 603–618. [Google Scholar] [CrossRef]
  74. Gavili, S.; Sanikhani, H.; Kisi, O.; Mahmoudi, M.H. Evaluation of several soft computing methods in monthly evapotranspiration modelling. Meteorol. Appl. 2017, 25, 128–138. [Google Scholar] [CrossRef] [Green Version]
  75. Diamantopoulou, M.J.; Georgiou, P.E.; Papamichail, D.M. Performance evaluation of artificial neural networks in estimating reference evapotranspiration with minimal meteorological data. Glob. NEST J. 2011, 13, 18–27. [Google Scholar]
  76. Antonopoulos, V.Z.; Gianniou, S.K.; Antonopoulos, A.V. Artificial neural networks and empirical equations to estimate daily evaporation: Application to Lake Vegoritis, Greece. Hydrol. Sci. J. 2016, 61, 2590–2599. [Google Scholar] [CrossRef] [Green Version]
  77. Antonopoulos, V.Z.; Antonopoulos, A.V. Daily reference evapotranspiration estimates by artificial neural networks technique and empirical equations using limited input climate variables. Comput. Electron. Agric. 2017, 132, 86–96. [Google Scholar] [CrossRef]
  78. Dimitriadou, S.; Katsanou, K.; Charalabopoulos, S.; Lambrakis, N. Interpretation of the Factors Defining Groundwater Quality of the Site Subjected to the Wildfire of 2007 in Ilia Prefecture, South-Western Greece. Geosciences 2018, 8, 108. [Google Scholar] [CrossRef] [Green Version]
  79. Copernicus Land Monitoring Service. CLC 2018. ©European Union, Copernicus Land Monitoring Service 2018, European Environment Agency (EEA). Available online: https://land.copernicus.eu/pan-european/corine-land-cover/clc2018 (accessed on 11 April 2022).
  80. Kottek, M.; Grieser, J.; Beck, C.; Rudolf, B.; Rubel, F. World map of the Köppen-Geiger climate classification updated. Meteorol. Z. 2006, 15, 259–263. [Google Scholar] [CrossRef]
  81. Elbeltagi, A.; Nagy, A.; Mohammed, S.; Pande, C.B.; Kumar, M.; Bhat, S.A.; Zsembeli, J.; Huzsvai, L.; Tamás, J.; Kovács, E.; et al. Combination of Limited Meteorological Data for Predicting Reference Crop Evapotranspiration Using Artificial Neural Network Method. Agronomy 2022, 12, 516. [Google Scholar] [CrossRef]
  82. Abrishami, N.; Sepaskhah, A.R.; Shahrokhnia, M.H. Estimating wheat and maize daily evapotranspiration using artificial neural network. Arch. Meteorol. Geophys. Bioclimatol. Ser. B 2019, 135, 945–958. [Google Scholar] [CrossRef]
  83. Kisi, O.; Sanikhani, H.; Zounemat-Kermani, M.; Niazi, F. Long-term monthly evapotranspiration modeling by several data-driven methods without climatic data. Comput. Electron. Agric. 2015, 115, 66–77. [Google Scholar] [CrossRef]
  84. Rahimikhoob, A. Estimation of evapotranspiration based on only air temperature data using artificial neural networks for a subtropical climate in Iran. Arch. Meteorol. Geophys. Bioclimatol. Ser. B 2010, 101, 83–91. [Google Scholar] [CrossRef]
  85. Gomar, S.; Mirhassani, M.; Ahmadi, M. Precise digital implementations of hyperbolic tanh and sigmoid function. In Proceedings of the 2016 50th Asilomar Conference on Signals, Systems and Computers, Pacific Grove, CA, USA, 6–9 November 2016; pp. 1586–1589. [Google Scholar] [CrossRef]
  86. Genaidy, M.A. Estimating of evapotranspiration using artificial neural network. Misr. J. Agric. Eng. 2020, 37, 81–94. Available online: https://mjae.journals.ekb.eg/article_94971_382909fe6190f50b2edcab1dc6d3c8b9.pdf (accessed on 15 May 2022). [CrossRef]
  87. Cococcioni, M.; Rossi, F.; Ruffaldi, E.; Saponara, S. A Fast Approximation of the Hyperbolic Tangent When Using Posit Numbers and Its Application to Deep Neural Networks. In Applications in Electronics Pervading Industry, Environment and Society; Saponara, S., De Gloria, A., Eds.; ApplePies. Lecture Notes in Electrical Engineering; Springer: Cham, Switzerland, 2019; Volume 627, pp. 213–221. [Google Scholar] [CrossRef]
Figure 1. Land use/land cover map of the Peloponnese (Adapted with permission from Ref. [79] 2018, © European Union).
Figure 1. Land use/land cover map of the Peloponnese (Adapted with permission from Ref. [79] 2018, © European Union).
Water 14 02027 g001
Figure 2. MLP1 model with the best performance on ETo predictions for the Peloponnese (August and December of 2016–2020).
Figure 2. MLP1 model with the best performance on ETo predictions for the Peloponnese (August and December of 2016–2020).
Water 14 02027 g002
Figure 3. Normalized importance of the input variables of the model with the best performance (MLP1) in ETo prediction of the Peloponnese.
Figure 3. Normalized importance of the input variables of the model with the best performance (MLP1) in ETo prediction of the Peloponnese.
Water 14 02027 g003
Table 1. Formulae of the indices utilized to evaluate the performance of the examined ANN models.
Table 1. Formulae of the indices utilized to evaluate the performance of the examined ANN models.
Performance Evaluating Indices
RMSE i = 1 n ( p i o i ) 2 n NRMSE = i = 1 n ( p i o i ) 2 n o Ι o Α = 1 i = 1 n ( p i o i ) 2 i = 1 n ( | p i o | | o i o | 2
MAE = i = 1 n | p i o i | n MB = i = 1 n ( p i o i ) n r = i = 1 n ( p i p ) ( o i o ) i = 1 n ( p i p ) 2 i = 1 n ( o i o ) 2
SSE = i = 1 n ( p i o i ) 2 RE = | p i o i o i | R 2 = ( n i = 1 n p i   o i i = 1 n p i   i = 1 n o i n i = 1 n p i   2 ( i = 1 n p i   ) 2 n i = 1 n o i   2 ( i = 1 n p i   ) 2 ) 2
Table 2. The examined ANN models (19) and the corresponding indices of performance.
Table 2. The examined ANN models (19) and the corresponding indices of performance.
ANNArchitectureRMSENRMSEMAEMBER2Pearson’s rIoASSE TestingRE TestingRE Holdout
7 inputs: N, Rs, Rn, u2, es-ea, Tmean, Z
MLP17-20.2900.0860.2170.0170.9800.9900.9950.6950.0270.022
MLP27-4-30.3050.0900.2300.0240.9780.9890.9940.5730.0210.029
RBF17-50.4230.1250.2670.0040.9570.9780.9890.9480.0270.100
RBF27-90.3330.0980.246−0.0320.9740.9870.9931.5470.0450.028
6 inputs: Rs, Rn, u2, es-ea, Tmean, Z
MLP86-40.3110.0920.2400.0050.9770.9880.9940.8270.0240.025
MLP96-4-30.2960.0880.218−0.0080.9790.9890.9951.1880.0350.024
RBF36-90.3180.0940.2320.0160.9760.9880.9940.6230.0260.022
4 inputs: Rs, u2, es-ea, Tmean and 4’ inputs: Rn, u2, es-ea, Tmean
MLP74-30.3090.0910.233−0.0020.9770.9890.9940.5430.0200.027
MLP104-3-20.3000.0890.2210.0010.9780.9890.9950.5410.0200.028
RBF44-100.4060.1200.263−0.0420.9610.9800.9901.0440.0310.092
MLP34’-10.3190.0940.245−0.0250.9760.9880.9940.4590.1600.300
3 inputs: Rn, u2, Tmean
MLP53-10.3140.0930.244−0.0190.9760.9880.9941.1010.0320.030
2 inputs: Rn, Tmean; 2’ inputs: Rn, u2; 2’’ inputs: u2, Tmean
MLP42-10.3770.1110.268−0.0230.9660.9830.9911.4430.0420.034
MLP62’-10.3430.1010.2680.0260.9720.9860.0930.9610.0370.028
MLP132’’-10.3520.1040.267−0.0080.9700.9850.9920.8190.0290.029
1 input: Tmean; 1’ input: Rn
MLP141-10.3600.1060.2580.0140.9690.9840.9920.7470.0270.041
MLP161’-10.4040.1190.313−0.0140.9610.9800.9900.5940.0300.042
RBF61-10.3630.1070.2590.0060.9680.9840.9920.7100.0280.033
RBF91’-10.3830.1130.298−0.0050.9650.9820.9911.1090.0360.033
Note: The values in italics (RE holdout column) indicate overtraining of the corresponding model. The models in bold exhibit performance that is discussed in detail. RMSE, MAE, MB are in mm d1 and SSE in mm2 d1.
Table 3. Parameter estimates of the model with the best performance (MLP1).
Table 3. Parameter estimates of the model with the best performance (MLP1).
PredictorPredicted
Hidden Layer 1Output Layer
H (1:1)H (1:2)ETο
Input Layer(Bias)−0.616−0.638
Rs−0.010−0.355
Rn0.3970.173
es-ea0.224−0.181
u20.214−0.049
N−0.100−0.520
Tmean0.461−0.394
Z0.059−0.040
Hidden Layer 1(Bias) −0.009
H (1:1) 0.578
H (1:2) −0.867
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dimitriadou, S.; Nikolakopoulos, K.G. Artificial Neural Networks for the Prediction of the Reference Evapotranspiration of the Peloponnese Peninsula, Greece. Water 2022, 14, 2027. https://doi.org/10.3390/w14132027

AMA Style

Dimitriadou S, Nikolakopoulos KG. Artificial Neural Networks for the Prediction of the Reference Evapotranspiration of the Peloponnese Peninsula, Greece. Water. 2022; 14(13):2027. https://doi.org/10.3390/w14132027

Chicago/Turabian Style

Dimitriadou, Stavroula, and Konstantinos G. Nikolakopoulos. 2022. "Artificial Neural Networks for the Prediction of the Reference Evapotranspiration of the Peloponnese Peninsula, Greece" Water 14, no. 13: 2027. https://doi.org/10.3390/w14132027

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