Next Article in Journal
Decision-Making Framework for GI Layout Considering Site Suitability and Weighted Multi-Function Effectiveness: A Case Study in Beijing Sub-Center
Next Article in Special Issue
Understanding the Effect of Hydro-Climatological Parameters on Dam Seepage Using Shapley Additive Explanation (SHAP): A Case Study of Earth-Fill Tarbela Dam, Pakistan
Previous Article in Journal
Community-Scale Rural Drinking Water Supply Systems Based on Harvested Rainwater: A Case Study of Australia and Vietnam
Previous Article in Special Issue
Optimal Design of Water Treatment Contact Tanks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Estimating the Standardized Precipitation Evapotranspiration Index Using Data-Driven Techniques: A Regional Study of Bangladesh

1
Agricultural Engineering Department, Faculty of Agriculture, Mansoura University, Mansoura 35516, Egypt
2
Faculty of Maritime Studies, King Abdulaziz University, Jeddah 21589, Saudi Arabia
3
Farm Machinery and Postharvest Technology Division, Bangladesh Rice Research Institute, Gazipur 1701, Bangladesh
4
School of Civil Engineering, Faculty of Engineering, Universiti Teknologi Malaysia (UTM), Johor Bahru 81310, Malaysia
5
Irrigation and Water Management Division, Bangladesh Agricultural Research Institute, Gazipur 1701, Bangladesh
6
Agricultural Economics Division, Bangladesh Rice Research Institute, Gazipur 1701, Bangladesh
7
Tuber Crops Research Centre, Bangladesh Agricultural Research Institute, Gazipur 1701, Bangladesh
8
Irrigation and Water Management Division, Bangladesh Rice Research Institute, Gazipur 1701, Bangladesh
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Water 2022, 14(11), 1764; https://doi.org/10.3390/w14111764
Submission received: 23 April 2022 / Revised: 14 May 2022 / Accepted: 23 May 2022 / Published: 30 May 2022

Abstract

:
Drought prediction is the most effective way to mitigate drought impacts. The current study examined the ability of three renowned machine learning models, namely additive regression (AR), random subspace (RSS), and M5P tree, and their hybridized versions (AR-RSS, AR-M5P, RSS-M5P, and AR-RSS-M5P) in predicting the standardized precipitation evapotranspiration index (SPEI) in multiple time scales. The SPEIs were calculated using monthly rainfall and temperature data over 39 years (1980–2018). The best subset regression model and sensitivity analysis were used to determine the most appropriate input variables from a series of input combinations involving up to eight SPEI lags. The models were built at Rajshahi station and validated at four other sites (Mymensingh, Rangpur, Bogra, and Khulna) in drought-prone northern Bangladesh. The findings indicated that the proposed models can accurately forecast droughts at the Rajshahi station. The M5P model predicted the SPEIs better than the other models, with the lowest mean absolute error (27.89–62.92%), relative absolute error (0.39–0.67), mean absolute error (0.208–0.49), root mean square error (0.39–0.67) and highest correlation coefficient (0.75–0.98). Moreover, the M5P model could accurately forecast droughts with different time scales at validation locations. The prediction accuracy was better for droughts with longer periods.

1. Introduction

Drought is one of the most complicated recurring natural disasters, defined by a deficiency of precipitation, causing prolonged water scarcity. Failure to manage drought risk effectively has the potential to have dire consequences for people, livelihoods, the economy, and ecosystems [1,2,3,4,5,6]. Like the rest of South Asia, Bangladesh is plagued by periodic droughts. Due to the adverse impact on agricultural productivity and the environment, it is one of Bangladesh’s most expensive natural disasters [7,8]. Drought-related damage to agriculture is more prevalent than any other natural disaster in the country. Pre- and post-monsoon droughts (March–May and October–November) are the most common in Bangladesh, and pre-monsoon droughts can last into the monsoon season, delaying the arrival of rain. Drought susceptibility, on the other hand, varies according to region. In recent years, annual rainfall has increased across Bangladesh, including in the drought-prone northern region, while the western region has seen a decrease [2,9]. Extreme climate events such as droughts have increased in frequency over the past two decades, which has put food security in some regions at risk [10,11]. The frequency of extreme droughts will nearly double in the coming decades [12]. Thus, drought forecasting, and early warning are critical for agricultural resilience to climate change.
Due to its complexity and spatial–temporal development, drought forecasting is one of the most difficult issues for climate scientists and hydrologists [13]. Statistical, dynamical [14,15], and hybrid models [16] are typically utilized to forecast drought occurrences. For drought forecasting, statistical models employ correlation relationships between climate variables and drought indicators [17]. As opposed to statistical models, dynamical models are founded on the physical connections between the earth, ocean, and climate. These interactions are theoretically defined and resolved in dynamic models to develop drought forecasts and predictions [18]. A hybrid model, in contrast, is a combination of statistical and dynamical techniques [19,20]. For instance, to create an ensemble prediction, multiple dynamical model predictions can be combined using a statistical framework that assigns weights to the various dynamical model predictions [21]. Due to their simplicity [22] and low processing costs, statistical models are widely used to predict droughts [23,24,25]. Recently, machine learning (ML) techniques have also been used to predict drought in many global regions.
ML strategies involve a set of commands that enable systems to learn and improve without extensive programming [26,27,28,29]. In different climatological applications, such as rainfall prediction, ML algorithms have been used to develop models that can reproduce the empirical relationships between the variables [30], drought prediction [31], forecasting heat waves [32], and runoff simulation [17]. There are many critical issues to consider when developing a drought forecasting model. Systematic monitoring and early warning of impending droughts are essential for effective drought management. Various statistical models are widely used to predict droughts, such as ARIMA, multilinear regression (MLR), and the Markov chain [33]. As a result, hydrological research typically employs non-linear time series models. When it comes to predicting drought and climate, ML applications have performed exceptionally well in recent years [34,35]. SVR, RF, ELM, M5T, ANFIS, LGP, ERT, LSSVR, and MVR have all been successfully applied to drought forecasting [22,36,37,38,39,40,41,42,43,44,45,46,47,48,49]. Several recent studies have shown that hybrid ML models perform exceptionally well and accurately [22,42,50,51,52,53]. However, it is not always the case. Thus, evaluation of the models is essential to predict drought in a particular region.
Additionally, models for drought forecasting exist in different parts of the world, but a universal or ideal model cannot be developed for all climates. Incorrect model structure variables can lead to erroneous predictions. Thus, it is necessary to evaluate the models at the regional scale. Only a few studies have been performed using ML algorithms to predict drought in Bangladesh. SPI has been used in previous studies to predict droughts [44,54]. Despite this, no research has been conducted in Bangladesh on the use of ML and hybrid methods to predict SPEI. This study’s primary goal was to fill in this knowledge gap.
Compared to other parts of the country, the Barind tract and the Teesta floodplain regions of the northern and northwestern parts (known as North Bengal) are highly impacted by drought due to high poverty rates, dependency on agriculture, low adaptive capacity and high variability of annual and seasonal rainfall [55,56]. Drought is a recurrent event in these regions [1,57]. Over the years, drought severity, frequency, and variability have increased in North Bengal [1,2,7,8,58,59,60]. Several studies indicated that droughts have significantly affected agricultural production and the natural environment in the country in recent years [7,61]. Although there have been tremendous improvements in irrigation systems in Bangladesh in recent decades, agricultural activities remain dependent on seasonal rainfall [62]. A study conducted by Islam [63] indicated that North Bengal has experienced significant increases in rainfall variability, long seasonal-scale dry spells and numerous instances of below-normal rainfall in recent decades, significantly hampering crop growth. In addition, variability in temperature has a substantial effect on crop yields (such as rice and wheat) in North Bengal [64]. However, there is a dearth of literature regarding the evolution of drought in eastern Bangladesh.
Drought analysis and prediction need reliable rainfall and temperature data recorded for a longer period. However, an uninterrupted climate record for a longer period is not available in most of the meteorological stations of Bangladesh. This limits drought prediction using conventional statistical and numerical models. ML algorithms could be one of the solutions for bridging the climate data gap. Furthermore, to the best of our knowledge, no research has been published in the literature that predicts the SPEI in the study region using ML approaches applied. Hence, this work aimed to assess the performance of standalone models and novel hybrid ML algorithms in SPEI estimation. Additive regression, random subspace, and M5P tree models and their hybridized versions were used to predict SPEI. Agricultural drought is measured in Bangladesh using a 6-month SPEI, while the water scarcities, river flow declines, and hydrological droughts are measured using 9, 12, and 24-month SPEIs. As a result, forecasting models were developed to predict SPEI for those four timescales over 38 years (1980–2018). The models were developed for a single station in northwestern Bangladesh (Rajshahi), which is extremely prone to droughts, and tested at four locations, including Bogra, Rangpur, Mymensingh and Khulna.
The organization of the remaining parts of this article is as follows. Section 2 explains the study area and the data used in the study. Additionally, the theories of the ML algorithms used and the construction of prediction models using ML are discussed in this section. Section 3 explains the results obtained in this study. Section 4 provides a discussion of the study’s findings. Finally, conclusions drawn from the results and discussion are provided in Section 5.

2. Materials and Methods

2.1. Study Area

Bangladesh covers a land area of 147,570 km2 and is situated between latitudes 20.34° N and 26.38° N and longitudes 88.01° E to 92.41° E. It is surrounded by the deltas of several rivers that originate in the Himalayas (Figure 1). A humid tropical climate characterizes most of the country. The temperature is below 12.8 °C in January and above 31.1 °C in May. The temperature’s spatial variability is extremely low because of the country’s extremely flat topography. Different seasons have different temperature gradient orientations. As a result, the yearly average temperature in different regions is very similar. Rainfall in Bangladesh varies between northwest and northeastern Bangladesh by 1600 and 4400 mm, respectively. Seasonal and yearly variations in rainfall are significant. Nearly 70% of the year’s rainfall occurs between May and September, with only 3% occurring in the winter months between December and February. In large parts of the country, the annual monsoon rainfall variability coefficient is greater than 30%. Since the country experiences a wide range of rainfall variability, droughts are common. Following its independence in 1973, Bangladesh was hit by droughts in 1978, 1979, 1981, 1982, 1989, 1992, 1994, and 1995. In recent memory, the worst droughts were in 1973, 1979, and 1994–1995, when the northwestern region lost 3.5 million tons of rice alone. A 25–30% decrease in average crop production was recorded in this region during the drought of 2006 [65,66].

2.2. Data

Monthly temperature and precipitation data were collected from five meteorological stations (Table 1) from the Bangladesh Agricultural Research Council (BARC) for a 38-year period (1980–2018). The BARC team checked all the datasets for errors and inconsistencies. The rainfall and temperature data were seasonally divided using the R function na.seasplit [67] to examine missing data in the monthly climate variables. No missing data were found at these sites during the research period.

2.3. Methodology

Data were collected in the first phase to estimate the SPEI from 1980 to 2018. The best subset regression model for selecting the optimal combination of climatic variables was then used to create seven machine learning models (standalone and hybrid) to predict the multiscale SPEI. The validation stations were used to forecast SPEIs using the best prediction model.

2.3.1. SPEI Calculation

Standardized precipitation evapotranspiration indexes (SPEIs) were derived from precipitation and temperature data as part of a simple water balance to account for surface evaporation changes, which is more sensitive to drought reactions caused by rising global temperatures.
The difference in the water balance was normalized as a log-logistic probability distribution in order to estimate SPEI’s value. In mathematical terms, the probability density function is represented by the following equation:
f x = β α x λ α 1 + x λ α 2
Scale, shape and origin were all represented by the aforementioned parameters α, β, and γ, respectively. The probability distribution function can, therefore, be expressed as follows.
F x = 1 + α x γ β 1
The SPEI may be simply calculated using the standardized values of F(x)
SPEI = W C 0 + C 1 W + C 2 W 2 1 + d 1 W + d 2 W 2 + d 3 W 3
when P ≤ 0.5, W = 2 l n P , and when P > 0.5, W = 2 l n 1 P , C0 = 2.5155, C1 = 0.8028, C2 = 0.0203, d1 = 1.4327, d2 = 0.1892, and d3 = 0.0013.
The drought index was calculated using the SPEI package in the R environment [29]. Negative SPEI values, accompanied by a decrease in rainfall, signal drought, whereas positive SPEI values signal wetter conditions.

2.3.2. Machine Learning Algorithms

For the perdition multiscale SPEI, this study considered three tree-based algorithms, random subspace (RSS), additive regression (AR), and M5 pruned (M5P), and their stacking hybrid forms, AR-RSS, AR-M5P, RSS-M5P, and AR-RSS-M5P. The basic concept of tree-based algorithms is fitting decision trees to different sub-samples of calibration datasets and then integrating the prediction of each tree to provide the final output. Fitting models to different sub-samples allows data decomposition in detail for better identification of complex input–output relations. The advantages of these non-parametric methods are accurate prediction, little data pre-processing such as normalization or scaling, and no assumptions on space distributions. A detailed description of the standalone model can be found in the study by Elbeltagi et al. [68]. Wolpert [69] proposed the use of a hybrid stacking algorithm. This method facilitates the use of ensemble algorithms, which combine two or more algorithms throughout the training period. The idea behind stacking hybrid generalization is to use first-level learners to train and predict training datasets. For the meta learner, a new training dataset was created by combining the predicted outcomes from first-level learners. Sikora et al. [70] provided comprehensive information on stacked hybrid generalization.
All modelings were carried out using the WEKA (Version3.8.4, Waikato University, Hamilton, New Zealand) software (https://www.cs.waikato.ac.nz/ml/weka, accessed on 20 May 2021). Table 2 contains the model parameters. Additionally, a flowchart illustrating the methodology used is shown in Figure 2. The WEKA software is a collection of ML techniques related to data mining. It can manage and process various data types using multiple tools, i.e., regression, clustering, and visualization. The program also features a graphical user interface (GUI), which facilitates the program’s operation. There are additional alternatives, such as MATLAB, Python, and R, but preparing and implementing computer codes using those programs requires considerable time. However, the implementation of models in different environments does not affect the obtained results. Therefore, WEKA was chosen considering its simple implementation procedure compared to other software.
A grid search optimization algorithm was used to select optimum ML model parameters. It randomly searches the optimum parameters within a discrete grid space. The value within the range that provides the most accurate prediction is selected. Table 2 shows the selected optimum ML model hyperparameter values.

2.4. Constructing and Evaluating Models

Choosing suitable input and output variables for non-linear hydrologic systems can be time-consuming. SPEI 6, 9, 12, and 24 were calculated using precipitation and temperature data from Rajshahi station. The most important inputs (at different time lags) for the target variable (output) were selected by using subset regression and sensitivity analyses. When there are many input variables, one of the most critical stages in the soft computing model is feature selection. Numerous procedures, such as best subset regression, forward stepwise selection, and mutual information, can be used to determine the best combinations of all possible inputs. The current study employed best subset regression to determine the best possible input combinations for the SPI 6, 9-, 12-, and 24-month models. The optimal input combination was determined using six statistical criteria (MSE, determination coefficients (R2), adjusted R2, Akaike’s AIC, Mallows’ Cp, Akaike’s AIC, and Amemiya’s PC).
Multiscale SPEI data were predicted using standalone ML models and their stacking hybrid forms at the Rajshahi station. The models were developed using 30-year monthly data (1980–2009), and 9-year data (2010–2018) were used to test the model at Rajshahi station. Evaluation of model outputs was carried out using statistical indices and visual interpretation, including Taylor diagram, scatter, and boxplots. The models with the lowest RMSE, MAE, RAE, RRSE, and greater R during testing were considered superior for drought prediction. An evaluation was performed on the generalizability of the most successful model developed at Rajshahi station to predict SPEI at other stations (Bogra, Mymensingh, Khulna and Rangpur).

2.5. Statistical Performance Measurement

Several performance metrics were calculated to assess the model performance, including the root mean square error (RMSE), coefficient of determination (R), mean absolute error (MAE), relative absolute error (RAE), and root relative squared error (RRSE). Definitions for all parameters are as follows:
R M S E = i = 1 N SPEI O b s SPEI p r e 2 N
M A E = i = 1 N S P E I O b s SPEI p r e N
R A E = SPEI O b s SPEI p r e SPEI p r e × 100
R R S E = i = 1 N SPEI p r e SPEI O b s 2 i = 1 N SPEI O b s SPEI p r e 2
R = i = 1 N SPEI O b s SPEI O b s ¯   SPEI p r e SPEI p r e ¯   i = 1 N SPEI O b s SPEI O b s ¯   2 i = 1 N SPEI p r e SPEI p r e ¯   2
The actual and predicted SPEIs are represented by SPEIobs and SPEIpre, respectively, with SPEI ¯ representing the average values of the actual SPEI index, and N representing the number of observations.

3. Results

3.1. Model Input Selection

Table 3 summarizes the statistical indices of model performance for various input combinations. Best results are indicated by bolded numbers in Table 3. The results indicate that three lag values, 1, 6, and 7, were the most accurate predictors of SPI-6. Similarly, inputs 1, 2, 3, 4, 6, 7, and 8 provided the best forecasts for SPEI-9 and SPEI-12, but only lag 1 did so for SPEI-24.
The input variables were analyzed for sensitivity at a 5% significance level. The results from the regression analysis for SPEI-6, SPEI-9, SPEI-12, and SPEI-24 are depicted in Figure 3. Spikes with lags of 1, 6, 7, 8 for SPEI-6; 1, 2, 3, 4, 6, 7, 8 for SPEI-9; 1, 3, 4, 5, 6, 7, 8 for SPEI-12, and 1 for SPEI-24 were statistically significant, according to the findings. Sensitivity analysis verified the results of the statistical metrics presented in Table 3. In this way, the SPEI of the corresponding scale was predicted using those lags. The inputs used to develop the SPEI for various temporal scales at the Rajshahi station are summarized in Table 4.

3.2. Prediction of Droughts Using Machine Learning Techniques

To forecast multiscale SPEI data at Rajshahi station, individual and hybrid model versions were developed. It took 30 years of data from 1980 to 2009 to create models and data from 2010 to 2018 for testing at Rajshahi station. Various statistical indices and visual interpretations were used to evaluate the models’ performance, including Taylor diagrams, scatter diagrams, and boxplots. As a general rule, models with a lower RMSE, MAE, RAE, RRSE, and a larger R during testing were deemed more accurate for drought prediction. According to statistical metrics, the models’ performance at Rajshahi station is shown in Table 5. The best results are shown in the table in bold type.
Based on the testing data, the best-performing model was selected. Table 5 shows that M5P was the best predictor of SPEI on all time scales, with the highest accuracy. M5P had the lowest RRSE, RAE, MAE and RMSE, and the highest R of any other model when predicting SPEI over various time scales, with R ranging from 0.750 to 0.98, MAE ranging from 0.208 to 0.49, and the RMSE from 0.389 to 0.67; RAE was between 16.60% and 55.26%, and RRSE was between 27.89% and 62.92%. There was a positive correlation between increasing SPEI time scales and improved model accuracy. Consequently, SPEI-24 had the best prediction accuracy.
Taylor diagrams (TD) were used to examine the geographic configuration of predicted and calculated (observed) multiscale SPEI values based on various ML models during testing. Model performance can be evaluated visually using Taylor [64], which provides a polar plot to show how models reproduce observed values while emphasizing their accuracy and precision. TD shows three statistical matrices, including standard deviation (SD), root mean square error (RMSE), and correlation coefficient (R). There was a good correlation between the Taylor diagram (Figure 4) and the derived performance indicators in Table 4. SPEI has the best M5P prediction–observation agreement (orange triangle) of any timeframe except SPEI9, as shown by the Taylor diagram. The lowest RMSE (0.424) and highest correlations (0.869) were for the M5P model for SPEI9, while the variation in RSS-M5P (0.968) was the most in line with observations.
During the testing period (2010–2018), model performance at Rajshahi station was evaluated using scatter plots (Figure 5). Predicted and observed values showed good correlation. The ideal line (a 45° line) was aligned with the majority of the predicted points, indicating that all models had a high degree of accuracy in their predictions. For all timeframes, M5P had the best correlation coefficient performance (R = 0.76–0.93), while AR-RSS had the worst (R = 0.64–0.87). On average, the scatter plots indicated that the M5P outperformed the other models.
SPEI data were used to create boxplots showing the 25%, 50%, and 75% quantiles of the observed and predicted SPEI values (Figure 6). All prediction models performed admirably in predicting SPEI quantiles at various scales, especially at higher orders. In simulating SPEI quantiles at all scales, the M5P model showed the best overall accuracy, while the RSS and AR models performed the worst.

3.3. The Best Predictive Model Is Applied in Various Regions

The best predictive model (M5P) was used to predict multi-scaler SPEI values at four different stations located throughout the drought-prone northern part of Bangladesh: Rangpur, Borga, Mymensingh, and Khulna. Model performance was evaluated at the test locations for the whole study period (1980–2018). According to various statistical indices, the performance evaluation results are summarized in Table 6. The model performed satisfactorily in terms of the statistical indices computed. The M5P predicted the SPEI-6, 9, 12, and 24 with R values ranging from 0.787 to 0.802, 0.850 to 0.882, 0.899 to 0.938, and 0.927 to 0.966, respectively. The MAE, RMSE, RAE, and RRSE were all low in terms of prediction on all time scales. In terms of accuracy, SPEI-24 was predicted the best by the M5P model, followed by SPEI-9 and SPEI-6. According to the findings, drought forecasting in western Bangladesh could benefit from the M5P model developed in this study.

4. Discussion

Droughts frequently affect agriculture and the livelihoods of farmers in northern Bangladesh. Reliable forecasting of droughts is important to reduce drought impacts in the region. However, drought analysis and prediction need reliable rainfall and temperature data recorded for longer periods, which most of the meteorological stations in Bangladesh do not have. ML algorithms were used in this study to overcome the limitations of climate data. The present study revealed that the SPEI, the most widely used DI, can be accurately forecasted using ML models for a multi-month horizon (i.e., 6, 9, 12 and 24). SPEI6, SPEI9, SPEI12, and SPEI24-month models were optimized using the best subset regression analysis. Predictions of SPEI using M5P were the most accurate across all time scales. The temporal variability of SPEIs could be replicated in different parts of western Bangladesh. Water professionals and policymakers may find this model useful for making intelligent decisions.
The studied models better predicted SPEIs over longer periods than shorter periods. This is expected as the smoothness and randomness of SPEI time series increase with increasing time scale. More linear data improve the machine learning models’ performance, as was found in the current study. M5P was still able to predict lower-scale SPEI series with an R = 0.76. A highly non-linear process can be captured using the M5P model.
A multivariate linear algorithm is how the M5P model works. The leaves on the tree represent different linear regression models. Segmenting data and fitting it with an appropriate regression model are made easier with this method [71]. Due to its decomposition capability, it can fit different models to various non-linear datasets. M5P’s ability to simulate each data point in a data series helped it to better predict linear model phenomena than the other models tested in this study. M5P’s ability to learn efficiently and model high-dimensional data also improved as a result of this change.
Several drought indicators have been predicted using algorithms inspired by nature and a stochastic (time-series) model. The M5P results were compared to those of these models (DIs). The AR and non-linear bi-linear (BL) models based on global climate indicators and delayed SPI data were used to predict the meteorological drought in Ankara, Turkey [45]. The predictive power of the BL and AR models was lower than the prediction accuracy obtained in this study. The least-squares support vector machine (LSSVM), multivariate adaptive regression splines (MARS), and M5 tree models were used to predict droughts in eastern Australia [42]. Prediction accuracy was reported to be higher for the M5 tree approach. Özger et al. [72] used ANN and SVM and their hybrids with wavelet decomposition for forecasting droughts in the Antalya region of Turkey. They showed that hybridization improved model performance significantly, which agrees with the present study’s findings. Nguyen et al. [73] used different ML models, including ANFIS, M5, M11 and M13 models, to predict SPEI in the Cai River basin of Vietnam. They reported M5 as the best performing model, followed by M11 and M13. Adarsh and Janga Reddy [74] employed stepwise linear regression, genetic programming, and M5 methods to predict standardized precipitation indices for different regions of India. They reported superior performance from M5 in predicting droughts in all regions. Shamshirband et al. [75] predicted SPEI using support vector regression, gene expression programming (GEP), and M5 models and reported higher performance with the M5 model. Barzkar et al. [76] used three ML models, GEP, M5, and multivariate adaptive regression spline (MARS), to predict SPEIs for different climatic conditions. They showed that the M5 model performed better in all cases.
The literature mentioned above clearly indicates the ML models’ capability to predict droughts in different meteorological settings. The present study revealed that the ML model, specifically M5P, was more capable of forecasting meteorological droughts over a wide range of timescales. The study revealed that a hybrid ML model can significantly improve the standalone ML models’ performance. Longer-period droughts are more predictable than shorter-period droughts. This may be due to the higher variability of shorter-period droughts. Droughts have a devastating impact on society and the economy. The findings of this study indicate that drought forecasting models in drought-prone eastern regions of Bangladesh have the potential to be installed as an early warning system to mitigate the effects of drought.

5. Conclusions

Newly developed machine learning models for forecasting SPEI in Bangladesh were evaluated in this study. In Rajshahi, Bangladesh’s most drought-prone location, the models were developed to predict SPEI for the period from 1980 to 2018. The model was validated at four stations distributed over the country’s western region. SPEI was accurately predicted in Rajshahi using the M5P model. R values for SPEI-6, 9, 12, and 24 at the validation stations were 0.787–0.802, 0.850–0.882, 0.899–0.938 and 0.927–0.966, respectively. The M5P model can forecast droughts on multiple timescales according to correlation and low errors. As a result of climate change, the model’s output could help predict when droughts will occur in the country and help mitigate their growing negative effects. This study considered only three tree-based algorithms for drought prediction. Bangladesh’s droughts can be predicted using other ML models to see how well they work. Optimization algorithms can be combined with machine learning models to improve predictability.

Author Contributions

Each author contributed to the composition of the paper. S.S., F.A. and M.K. supervised and coordinated this study; A.E. performed the calculations and modeling; D.K.R., P.K.K., L.D., M.M.I. and M.M.R. assisted in reviewing the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

There was no external funding for this research.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kamruzzaman, M.; Hwang, S.; Cho, J.; Jang, M.W.; Jeong, H. Evaluating the Spatiotemporal Characteristics of Agricultural Drought in Bangladesh Using Effective Drought Index. Water 2019, 11, 2437. [Google Scholar] [CrossRef] [Green Version]
  2. Kamruzzaman, M.; Jang, M.-W.W.; Cho, J.; Hwang, S. Future Changes in Precipitation and Drought Characteristics over Bangladesh under CMIP5 Climatological Projections. Water 2019, 11, 2219. [Google Scholar] [CrossRef] [Green Version]
  3. Lorenzo-Lacruz, J.; Vicente-Serrano, S.M.; López-Moreno, J.I.; Beguería, S.; García-Ruiz, J.M.; Cuadrat, J.M. The Impact of Droughts and Water Management on Various Hydrological Systems in the Headwaters of the Tagus River (Central Spain). J. Hydrol. 2010, 386, 13–26. [Google Scholar] [CrossRef] [Green Version]
  4. Potop, V.; Možný, M.; Soukup, J. Drought Evolution at Various Time Scales in the Lowland Regions and Their Impact on Vegetable Crops in the Czech Republic. Agric. For. Meteorol. 2012, 156, 121–133. [Google Scholar] [CrossRef]
  5. Qin, Y.; Yang, D.; Lei, H.; Xu, K.; Xu, X. Comparative Analysis of Drought Based on Precipitation and Soil Moisture Indices in Haihe Basin of North China during the Period of 1960–2010. J. Hydrol. 2015, 526, 55–67. [Google Scholar] [CrossRef]
  6. Yang, X.; Zheng, W.; Lin, C.; Ren, L.; Wang, Y.; Zhang, M.; Yuan, F.; Jiang, S. Prediction of Drought in the Yellow River Based on Statistical Downscale Study and SPI. J. Hohai Univ. 2017, 45, 377. [Google Scholar] [CrossRef]
  7. Alamgir, M.; Shahid, S.; Hazarika, M.K.; Nashrrullah, S.; Harun, S.B.; Shamsudin, S. Analysis of Meteorological Drought Pattern during Different Climatic and Cropping Seasons in Bangladesh. J. Am. Water Resour. Assoc. 2015, 51, 794–806. [Google Scholar] [CrossRef]
  8. Islam, A.R.M.T.; Salam, R.; Yeasmin, N.; Kamruzzaman, M.; Shahid, S.; Fattah, M.A.; Uddin, A.S.; Shahariar, M.H.; Mondol, M.A.H.; Jhajharia, D.; et al. Spatiotemporal Distribution of Drought and Its Possible Associations with ENSO Indices in Bangladesh. Arab. J. Geosci. 2021, 14, 2681. [Google Scholar] [CrossRef]
  9. Fahad, M.G.R.; Saiful Islam, A.K.M.; Nazari, R.; Alfi Hasan, M.; Tarekul Islam, G.M.; Bala, S.K. Regional Changes of Precipitation and Temperature over Bangladesh Using Bias-Corrected Multi-Model Ensemble Projections Considering High-Emission Pathways. Int. J. Climatol. 2018, 38, 1634–1648. [Google Scholar] [CrossRef]
  10. Dai, A. Characteristics and Trends in Various Forms of the Palmer Drought Severity Index during 1900–2008. J. Geophys. Res. Atmos. 2011, 116. [Google Scholar] [CrossRef] [Green Version]
  11. Dai, A. Increasing Drought under Global Warming in Observations and Models. Nat. Clim. Chang. 2013, 3, 52–58. [Google Scholar] [CrossRef]
  12. Aadhar, S.; Mishra, V. On the Occurrence of the Worst Drought in South Asia in the Observed and Future Climate. Environ. Res. Lett. 2021, 16, 024050. [Google Scholar] [CrossRef]
  13. Hao, Z.; Singh, V.P.; Xia, Y. Seasonal Drought Prediction: Advances, Challenges, and Future Prospects. Rev. Geophys. 2018, 56, 108–141. [Google Scholar] [CrossRef] [Green Version]
  14. Mishra, A.K.; Singh, V.P. Drought Modeling—A Review. J. Hydrol. 2011, 403, 157–175. [Google Scholar] [CrossRef]
  15. Shahid, S. Rainfall Variability and the Trends of Wet and Dry Periods in Bangladesh. Int. J. Climatol. 2010, 30, 2299–2313. [Google Scholar] [CrossRef]
  16. Pozzi, W.; Sheffield, J.; Stefanski, R.; Cripe, D.; Pulwarty, R.; Vogt, J.V.; Heim, R.R.; Brewer, M.J.; Svoboda, M.; Westerhoff, R.; et al. Toward Global Drought Early Warning Capability: Expanding International Cooperation for the Development of a Framework for Monitoring and Forecasting. Bull. Am. Meteorol. Soc. 2013, 94, 776–785. [Google Scholar] [CrossRef]
  17. Yaseen, Z.M.; El-shafie, A.; Jaafar, O.; Afan, H.A.; Sayl, K.N. Artificial Intelligence Based Models for Stream-Flow Forecasting: 2000–2015. J. Hydrol. 2015, 530, 829–844. [Google Scholar] [CrossRef]
  18. Turco, M.; Ceglar, A.; Prodhomme, C.; Soret, A.; Toreti, A.; Doblas-Reyes Francisco, J. Summer Drought Predictability over Europe: Empirical versus Dynamical Forecasts. Environ. Res. Lett. 2017, 12, 084006. [Google Scholar] [CrossRef] [Green Version]
  19. Murakami, H.; Villarini, G.; Vecchi, G.A.; Zhang, W.; Gudgel, R. Statistical-Dynamical Seasonal Forecast of North Atlantic and U.S. Landfalling Tropical Cyclones Using the High-Resolution GFDL FLOR Coupled Model. Mon. Weather Rev. 2016, 144, 2101–2123. [Google Scholar] [CrossRef]
  20. Strazzo, S.; Collins, D.C.; Schepen, A.; Wang, Q.J.; Becker, E.; Jia, L. Application of a Hybrid Statistical-Dynamical System to Seasonal Prediction of North American Temperature and Precipitation. Mon. Weather Rev. 2019, 147, 607–625. [Google Scholar] [CrossRef]
  21. Madadgar, S.; AghaKouchak, A.; Shukla, S.; Wood, A.W.; Cheng, L.; Hsu, K.L.; Svoboda, M. A Hybrid Statistical-Dynamical Framework for Meteorological Drought Prediction: Application to the Southwestern United States. Water Resour. Res. 2016, 52, 5095–5110. [Google Scholar] [CrossRef] [Green Version]
  22. Belayneh, A.; Adamowski, J.; Khalil, B.; Quilty, J. Coupling Machine Learning Methods with Wavelet Transforms and the Bootstrap and Boosting Ensemble Approaches for Drought Prediction. Atmos. Res. 2016, 172–173, 37–47. [Google Scholar] [CrossRef]
  23. Ganguli, P.; Janga Reddy, M. Ensemble Prediction of Regional Droughts Using Climate Inputs and the SVM-Copula Approach. Hydrol. Process. 2014, 28, 4989–5009. [Google Scholar] [CrossRef]
  24. Mariotti, A.; Schubert, S.; Mo, K.; Peters-Lidard, C.; Wood, A.; Pulwarty, R.; Huang, J.; Barrie, D. Advancing Drought Understanding, Monitoring, and Prediction. Bull. Am. Meteorol. Soc. 2013, 94, ES186–ES188. [Google Scholar] [CrossRef]
  25. Xu, L.; Chen, N.; Zhang, X.; Chen, Z. An Evaluation of Statistical, NMME and Hybrid Models for Drought Prediction in China. J. Hydrol. 2018, 566, 235–249. [Google Scholar] [CrossRef] [Green Version]
  26. Chatterjee, I. Patenting Machine-Learning: Review and Discussions. Int. J. Mod. Res. 2021, 1, 15–21. [Google Scholar]
  27. Kumar Vaishnav, P.; Sharma, S.; Sharma, P. Analytical Review Analysis for Screening COVID-19 Disease. Int. J. Mod. Res. 2021, 1, 22–29. [Google Scholar]
  28. Kumar, R.; Dhiman, G. A Comparative Study of Fuzzy Optimization through Fuzzy Number. Int. J. Mod. Res. 2021, 1, 1–14. [Google Scholar]
  29. Sachindra, D.A.; Kanae, S. Machine Learning for Downscaling: The Use of Parallel Multiple Populations in Genetic Programming. Stoch. Environ. Res. Risk Assess. 2019, 33, 1497–1533. [Google Scholar] [CrossRef] [Green Version]
  30. Parmar, A.; Mistree, K.; Sompura, M. Machine Learning Techniques For Rainfall Prediction: A Review. In Proceedings of the 2017 International Conference on Innovations in information Embedded and Communication Systems (ICIIECS), Coimbatore, India, 17–18 March 2017. [Google Scholar]
  31. Tian, Y.; Xu, Y.P.; Wang, G. Agricultural Drought Prediction Using Climate Indices Based on Support Vector Regression in Xiangjiang River Basin. Sci. Total Environ. 2018, 622–623, 710–720. [Google Scholar] [CrossRef]
  32. Khan, N.; Shahid, S.; Juneng, L.; Ahmed, K.; Ismail, T.; Nawaz, N. Prediction of Heat Waves in Pakistan Using Quantile Regression Forests. Atmos. Res. 2019, 221, 1–11. [Google Scholar] [CrossRef]
  33. Fung, K.F.; Huang, Y.F.; Koo, C.H.; Soh, Y.W. Drought Forecasting: A Review of Modelling Approaches 2007–2017. J. Water Clim. Change 2020, 11, 771–799. [Google Scholar] [CrossRef]
  34. Malik, A.; Kumar, A.; Salih, S.Q.; Kim, S.; Kim, N.W.; Yaseen, Z.M.; Singh, V.P. Drought Index Prediction Using Advanced Fuzzy Logic Model: Regional Case Study over Kumaon in India. PLoS ONE 2020, 15, e0233280. [Google Scholar] [CrossRef] [PubMed]
  35. Kisi, O.; Choubin, B.; Deo, R.C.; Yaseen, Z.M. Incorporating Synoptic-Scale Climate Signals for Streamflow Modelling over the Mediterranean Region Using Machine Learning Models. Hydrol. Sci. J. 2019, 64, 1240–1252. [Google Scholar] [CrossRef]
  36. Rhee, J.; Im, J. Meteorological Drought Forecasting for Ungauged Areas Based on Machine Learning: Using Long-Range Climate Forecast and Remote Sensing Data. Agric. For. Meteorol. 2017, 237–238, 105–122. [Google Scholar] [CrossRef]
  37. Mouatadid, S.; Deo, R.C.; Adamowski, J.F. Prediction of SPEI Using MLR and ANN: A Case Study for Wilsons Promontory Station in Victoria. In Proceedings of the 2015 IEEE 14th International Conference on Machine Learning and Applications, ICMLA 2015, Miami, FL, USA, 9–11 December 2015. [Google Scholar]
  38. Mouatadid, S.; Raj, N.; Deo, R.C.; Adamowski, J.F. Input Selection and Data-Driven Model Performance Optimization to Predict the Standardized Precipitation and Evaporation Index in a Drought-Prone Region. Atmos. Res. 2018, 212, 130–149. [Google Scholar] [CrossRef]
  39. Park, S.; Im, J.; Jang, E.; Rhee, J. Drought Assessment and Monitoring through Blending of Multi-Sensor Indices Using Machine Learning Approaches for Different Climate Regions. Agric. For. Meteorol. 2016, 216, 157–169. [Google Scholar] [CrossRef]
  40. Belayneh, A.; Adamowski, J.; Khalil, B.; Ozga-Zielinski, B. Long-Term SPI Drought Forecasting in the Awash River Basin in Ethiopia Using Wavelet Neural Networks and Wavelet Support Vector Regression Models. J. Hydrol. 2014, 508, 418–429. [Google Scholar] [CrossRef]
  41. Deo, R.C.; Şahin, M. Application of the Artificial Neural Network Model for Prediction of Monthly Standardized Precipitation and Evapotranspiration Index Using Hydrometeorological Parameters and Climate Indices in Eastern Australia. Atmos. Res. 2015, 161–162, 65–81. [Google Scholar] [CrossRef]
  42. Deo, R.C.; Kisi, O.; Singh, V.P. Drought Forecasting in Eastern Australia Using Multivariate Adaptive Regression Spline, Least Square Support Vector Machine and M5Tree Model. Atmos. Res. 2017, 184, 149–175. [Google Scholar] [CrossRef] [Green Version]
  43. Mokhtar, A.; Jalali, M.; He, H.; Al-Ansari, N.; Elbeltagi, A.; Alsafadi, K.; Abdo, H.G.; Sammen, S.S.; Gyasi-Agyei, Y.; Rodrigo-Comino, J. Estimation of SPEI Meteorological Drought Using Machine Learning Algorithms. IEEE Access 2021, 9, 65503–65523. [Google Scholar] [CrossRef]
  44. Yaseen, Z.M.; Ali, M.; Sharafati, A.; Al-Ansari, N.; Shahid, S. Forecasting Standardized Precipitation Index Using Data Intelligence Models: Regional Investigation of Bangladesh. Sci. Rep. 2021, 11, 3435. [Google Scholar] [CrossRef]
  45. Mehdizadeh, S.; Ahmadi, F.; Danandeh Mehr, A.; Safari, M.J.S. Drought Modeling Using Classic Time Series and Hybrid Wavelet-Gene Expression Programming Models. J. Hydrol. 2020, 587, 125017. [Google Scholar] [CrossRef]
  46. Ali, M.; Deo, R.C.; Maraseni, T.; Downs, N.J. Improving SPI-Derived Drought Forecasts Incorporating Synoptic-Scale Climate Indices in Multi-Phase Multivariate Empirical Mode Decomposition Model Hybridized with Simulated Annealing and Kernel Ridge Regression Algorithms. J. Hydrol. 2019, 576, 164–184. [Google Scholar] [CrossRef]
  47. Danandeh Mehr, A.; Kahya, E.; Özger, M. A Gene-Wavelet Model for Long Lead Time Drought Forecasting. J. Hydrol. 2014, 517, 691–699. [Google Scholar] [CrossRef]
  48. Anshuka, A.; van Ogtrop, F.F.; Willem Vervoort, R. Drought Forecasting through Statistical Models Using Standardised Precipitation Index: A Systematic Review and Meta-Regression Analysis. Nat. Hazards 2019, 97, 955–977. [Google Scholar] [CrossRef]
  49. Zhang, R.; Chen, Z.Y.; Xu, L.J.; Ou, C.Q. Meteorological Drought Forecasting Based on a Statistical Model with Machine Learning Techniques in Shaanxi Province, China. Sci. Total Environ. 2019, 665, 338–346. [Google Scholar] [CrossRef]
  50. Ozger, M.; Mishra, A.K.; Singh, V.P. Estimating Palmer Drought Severity Index Using a Wavelet Fuzzy Logic Model Based on Meteorological Variables. Int. J. Climatol. 2011, 31, 2021–2032. [Google Scholar] [CrossRef]
  51. Masinde, M. Artificial Neural Networks Models for Predicting Effective Drought Index: Factoring Effects of Rainfall Variability. Mitig. Adapt. Strateg. Glob. Chang. 2014, 19, 1139–1162. [Google Scholar] [CrossRef]
  52. Prasad, R.; Deo, R.C.; Li, Y.; Maraseni, T. Input Selection and Performance Optimization of ANN-Based Streamflow Forecasts in the Drought-Prone Murray Darling Basin Region Using IIS and MODWT Algorithm. Atmos. Res. 2017, 197, 42–63. [Google Scholar] [CrossRef]
  53. Liang, Z.; Li, Y.; Hu, Y.; Li, B.; Wang, J. A Data-Driven SVR Model for Long-Term Runoff Prediction and Uncertainty Analysis Based on the Bayesian Framework. Theor. Appl. Climatol. 2018, 133, 137–149. [Google Scholar] [CrossRef]
  54. Bagmar, M.S.H.; Khudri, M.M. Application of Box-Jenkins Models for Forecasting Drought in North-Western Part of Bangladesh. Environ. Eng. Res. 2021, 26, 200294. [Google Scholar] [CrossRef]
  55. Habiba, U.; Shaw, R.; Takeuchi, Y. Farmers’ Adaptive Practices for Drought Risk Reduction in the Northwest Region of Bangladesh. Nat. Hazards 2014, 72, 337–359. [Google Scholar] [CrossRef]
  56. Shahid, S.; Behrawan, H. Drought Risk Assessment in the Western Part of Bangladesh. Nat. Hazards 2008, 46, 391–413. [Google Scholar] [CrossRef]
  57. Paul, B.K. Coping Mechanisms Practised by Drought Victims (1994/5) in North Bengal, Bangladesh. Appl. Geogr. 1998, 18, 355–373. [Google Scholar] [CrossRef]
  58. Kamruzzaman, M.; Cho, J.; Jang, M.-W.; Hwang, S. Comparative Evaluation of Standardized Precipitation Index (SPI) and Effective Drought Index (EDI) for Meteorological Drought Detection over Bangladesh. J. Korean Soc. Agric. Eng. 2019, 61, 145–159. [Google Scholar] [CrossRef]
  59. Mondol, M.A.H.; Ara, I.; Das, S.C. Meteorological Drought Index Mapping in Bangladesh Using Standardized Precipitation Index during 1981–2010. Adv. Meteorol. 2017, 2017, 4642060. [Google Scholar] [CrossRef] [Green Version]
  60. Islam, A.R.M.T.; Karim, M.R.; Mondol, M.A.H. Appraising Trends and Forecasting of Hydroclimatic Variables in the North and Northeast Regions of Bangladesh. Theor. Appl. Climatol. 2021, 143, 33–50. [Google Scholar] [CrossRef]
  61. Selvaraju, R.; Asian Disaster Preparedness Center; Food and Agriculture Organization of the United Nations. Livelihood Adaptation to Climate Variability and Change in Drought-Prone Areas of Bangladesh: Developing Institutions and Options; sian Disaster Preparedness Center: Bangkok, Thailand; Food and Agriculture Organization of the United Nations: Rome, Italy, 2006; ISBN 9789251056028. [Google Scholar]
  62. Akter, K.S.; Rahman, M.M. Spatio-Temporal Quantification and Characterization of Drought Patterns in Bangladesh. J. Water Environ. Technol. 2012, 10, 277–288. [Google Scholar] [CrossRef] [Green Version]
  63. Islam, M.N. Rainfall and Temperature Scenario for Bangladesh. Open Atmos. Sci. J. 2009, 3, 93–103. [Google Scholar] [CrossRef] [Green Version]
  64. Miah, M.G.; Abdullah, H.M.; Jeong, C. Exploring Standardized Precipitation Evapotranspiration Index for Drought Assessment in Bangladesh. Environ. Monit. Assess. 2017, 189, 547. [Google Scholar] [CrossRef]
  65. Rahman, A.A.; Alam, M.; Alam, S.S.; Uzzaman, M.R.; Rashid, M.; Rabbani, G. UNDP Human Development Report 2007: Background Paper on Risks, Vulnerability and Adaptation in Bangladesh; United Nations Development Programme: New York, NY, USA, 2008. [Google Scholar]
  66. Habiba, U.; Shaw, R.; Takeuchi, Y. Farmer’s Perception and Adaptation Practices to Cope with Drought: Perspectives from Northwestern Bangladesh. Int. J. Disaster Risk Reduct. 2012, 1, 72–84. [Google Scholar] [CrossRef]
  67. Moritz, S.; Bartz-Beielstein, T. imputeTS: Time Series Missing Value Imputation in R. R J. 2017, 9, 207–218. [Google Scholar] [CrossRef] [Green Version]
  68. 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. Environ. Res. Risk Assess. 2022. [Google Scholar] [CrossRef]
  69. Wolpert, D.H. Stacked Generalization. Neural Netw. 1992, 5, 241–259. [Google Scholar] [CrossRef]
  70. Sikora, R.; Al-Laymoun, O. A Modified Stacking Ensemble Machine Learning Algorithm Using Genetic Algorithms. In Handbook of Research on Organizational Transformations through Big Data Analytics; IGI Global: Hershey, PA, USA, 2014. [Google Scholar]
  71. Heddam, S.; Kisi, O. Modelling Daily Dissolved Oxygen Concentration Using Least Square Support Vector Machine, Multivariate Adaptive Regression Splines and M5 Model Tree. J. Hydrol. 2018, 559, 499–509. [Google Scholar] [CrossRef]
  72. Özger, M.; Başakın, E.E.; Ekmekcioğlu, Ö.; Hacısüleyman, V. Comparison of Wavelet and Empirical Mode Decomposition Hybrid Models in Drought Prediction. Comput. Electron. Agric. 2020, 179, 105851. [Google Scholar] [CrossRef]
  73. Nguyen, L.B.; Li, Q.F.; Ngoc, T.A.; Hiramatsu, K. Adaptive Neuro-Fuzzy Inference System for Drought Forecasting in the Cai River Basin in Vietnam. J. Fac. Agric. Kyushu Univ. 2015, 60, 405–415. [Google Scholar] [CrossRef]
  74. Adarsh, S.; Janga Reddy, M. Evaluation of Trends and Predictability of Short-Term Droughts in Three Meteorological Subdivisions of India Using Multivariate EMD-Based Hybrid Modelling. Hydrol. Process. 2019, 33, 130–143. [Google Scholar] [CrossRef] [Green Version]
  75. 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]
  76. Barzkar, A.; Najafzadeh, M.; Homaei, F. Evaluation of Drought Events in Various Climatic Conditions Using Data-Driven Models and a Reliability-Based Probabilistic Model. Nat. Hazards 2022, 110, 1931–1952. [Google Scholar] [CrossRef]
Figure 1. Study area and weather stations.
Figure 1. Study area and weather stations.
Water 14 01764 g001
Figure 2. Flowchart of SPEI estimation methodology in the study area.
Figure 2. Flowchart of SPEI estimation methodology in the study area.
Water 14 01764 g002
Figure 3. The standardized coefficients of input variables for sensitivity analysis (SPEI-6, 9, 12, and 24).
Figure 3. The standardized coefficients of input variables for sensitivity analysis (SPEI-6, 9, 12, and 24).
Water 14 01764 g003
Figure 4. Taylor diagram representation of model’s performance in predicting SPEIs at Rajshahi station over multiple time scales during the testing period (a) SEEI-6; (b) SEEI-9; (c) SEEI-12; (d) SEEI-24.
Figure 4. Taylor diagram representation of model’s performance in predicting SPEIs at Rajshahi station over multiple time scales during the testing period (a) SEEI-6; (b) SEEI-9; (c) SEEI-12; (d) SEEI-24.
Water 14 01764 g004
Figure 5. The scatter plots depict the observed and predicted SPEI values for various models at Rajshahi station over multiple time scales during the testing period, (a) SPEI-6, (b) SPEI-9, (c) SPEI-12, (d) SPEI-24.
Figure 5. The scatter plots depict the observed and predicted SPEI values for various models at Rajshahi station over multiple time scales during the testing period, (a) SPEI-6, (b) SPEI-9, (c) SPEI-12, (d) SPEI-24.
Water 14 01764 g005
Figure 6. Predictive model performance at Rajshahi station over multiple time scales during the testing period, (a) SPEI-6, (b) SPEI-9, (c) SPEI-12, (d) SPEI-24.
Figure 6. Predictive model performance at Rajshahi station over multiple time scales during the testing period, (a) SPEI-6, (b) SPEI-9, (c) SPEI-12, (d) SPEI-24.
Water 14 01764 g006
Table 1. Meteorological stations and descriptive statistics for rainfall and temperature.
Table 1. Meteorological stations and descriptive statistics for rainfall and temperature.
Station NameGeographical
Locations
(Lat × Lon)
Elevation
MSL (m)
Annual Mean Rainfall
(mm)
Tmax
(°C)
Tmin
(°C)
Remarks
Rajshahi24.37 × 88.719.51424.5131.3120.50Model Development station
Bogra24.85 × 89.3717.91738.0530.7521.01Model Validation station
Mymensingh90.43 × 24.72182264.2029.9020.88Model Validation station
Khulna89.53 × 22.783.61834.9231.2821.77Model Validation station
Rangpur89.23 × 25.7332.612248.8229.6320.24Model Validation station
Table 2. The parameters of the machine learning algorithm in modeling SPEI-6, -9, -12 and -24.
Table 2. The parameters of the machine learning algorithm in modeling SPEI-6, -9, -12 and -24.
Model NamesParameter Descriptions
Random Subspace (RSS)Batch size-100, Classifier = REPTree, random seed-1, subspace size = 0.5, numbers of executions slots = 1, number of iterations = 10
Additive Regression (AR)Batch size-100, Classifier = Bagging, shrinkage = 1, number of iterations = 30
M5 Pruned (M5P)Batch size-100, Minimum number of instances = 4
Table 3. The best subset regression analysis for determining the best input combinations to model.
Table 3. The best subset regression analysis for determining the best input combinations to model.
SPEI-6
VariablesMSER2Adjusted R2Mallows’ CpAkaike’s AICAmemiya’s PC
SPEI10.4040.5810.58031.480−417.7260.421
SPEI1/SPEI60.3960.5900.58822.969−425.7560.413
SPEI1/SPEI6/SPEI70.3790.6090.6063.166−445.2820.396
SPEI1/SPEI3/SPEI6/SPEI70.3770.6110.6082.119−446.3710.395
SPEI1/SPEI3/SPEI4/SPEI6/SPEI70.3780.6120.6073.785−444.7110.397
SPEI1/SPEI3/SPEI4/SPEI5/SPEI6/SPEI70.3780.6120.6075.141−443.3670.398
SPEI1/SPEI3/SPEI4/SPEI5/SPEI6/SPEI7/SPEI80.3790.6120.6067.072−441.4370.400
SPEI1/SPEI2/SPEI3/SPEI4/SPEI5/SPEI6/SPEI7/SPEI80.3800.6120.6059.000−439.5110.401
SPEI-9
SPEI10.2460.7450.74411.701−642.4580.256
SPEI1/SPEI80.2430.7490.7486.194−647.9010.253
SPEI1/SPEI3/SPEI80.2410.7520.7502.947−651.1760.251
SPEI1/SPEI3/SPEI4/SPEI80.2400.7530.7503.533−650.6080.252
SPEI1/SPEI3/SPEI4/SPEI7/SPEI80.2400.7530.7514.020−650.1450.252
SPEI1/SPEI2/SPEI3/SPEI4/SPEI7/SPEI80.2400.7540.7505.394−648.7820.253
SPEI1/SPEI2/SPEI3/SPEI4/SPEI6/SPEI7/SPEI80.2410.7540.7507.016−647.1680.254
SPEI1/SPEI2/SPEI3/SPEI4/SPEI5/SPEI6/SPEI7/SPEI80.2410.7540.7509.000−645.1840.255
SPEI-12
SPEI10.1510.8440.8446.523−861.5810.156
SPEI1/SPEI80.1480.8480.847−0.607−868.7520.154
SPEI1/SPEI5/SPEI80.1490.8480.8470.857−867.2970.154
SPEI1/SPEI6/SPEI7/SPEI80.1490.8480.8472.160−866.0050.155
SPEI1/SPEI3/SPEI5/SPEI7/SPEI80.1490.8480.8463.684−864.4890.155
SPEI1/SPEI3/SPEI4/SPEI6/SPEI7/SPEI80.1490.8480.8465.127−863.0570.156
SPEI1/SPEI3/SPEI4/SPEI5/SPEI6/SPEI7/SPEI80.1490.8480.8467.023−861.1640.156
SPEI1/SPEI2/SPEI3/SPEI4/SPEI5/SPEI6/SPEI7/SPEI80.1500.8480.8469.000−859.1870.157
SPEI-24
SPEI10.0810.9160.9161.071−1116.4400.084
SPEI1/SPEI50.0810.9170.9161.240−1116.2870.084
SPEI1/SPEI5/SPEI70.0810.9170.9170.835−1116.7240.084
SPEI1/SPEI5/SPEI7/SPEI80.0810.9180.9171.403−1116.1820.084
SPEI1/SPEI5/SPEI6/SPEI7/SPEI80.0810.9180.9173.057−1114.5350.084
SPEI1/SPEI2/SPEI5/SPEI6/SPEI7/SPEI80.0810.9180.9175.033−1112.5590.085
SPEI1/SPEI2/SPEI4/SPEI5/SPEI6/SPEI7/SPEI80.0810.9180.9167.025−1110.5680.085
SPEI1/SPEI2/SPEI3/SPEI4/SPEI5/SPEI6/SPEI7/SPEI80.0810.9180.9169.000−1108.5930.085
SPI1 = SPI (t-1), SPI8 = SPI (t-8); Bold: indicates the selected input combinations.
Table 4. Input variables selected for multiscale SPEI prediction.
Table 4. Input variables selected for multiscale SPEI prediction.
OutputInput Variables
SPEI6SPEI (t-1), SPEI (t-6), SPEI (t-7)
SPEI9SPEI (t-1), SPEI (t-2), SPEI (t-3), SPEI (t-4), SPEI (t-6), SPEI (t-7), SPEI (t-8)
SPEI12SPEI (t-1), SPEI (t-3), SPEI (t-4), SPEI (t-5), SPEI (t-6), SPEI (t-7), SPEI (t-8)
SPEI24SPEI (t-1)
Table 5. Computed statistical index values for the eight machine learning models (individual and hybrid) during training and testing stages.
Table 5. Computed statistical index values for the eight machine learning models (individual and hybrid) during training and testing stages.
ModelsTraining Period (1980–2009)Testing Period (2010–2018)
RMAERMSERAE
(%)
RRSE
(%)
RMAERMSERAE
(%)
RRSE
(%)
SPEI6
AR0.9290.2680.36334.2037.700.6750.5940.77066.9071.90
M5P0.7850.4430.59656.5861.970.7570.4910.67455.2662.92
RSS0.8090.4310.56754.9958.950.6870.5710.74064.3269.12
AR-M5P0.7860.4530.59757.8362.060.7070.5420.71560.9466.81
AR-RSS0.7880.4580.59358.4661.660.6420.5970.77067.2071.97
RSS-M5P0.79470.4440.58656.7261.020.7130.5360.70860.2966.16
AR-M5P-RSS0.76970.4730.61860.3764.270.6970.5470.72261.5467.43
SPEI9
AR0.9680.1680.24621.3525.250.7470.4440.60053.9757.32
M5P0.8790.3230.46541.0347.610.7630.3970.57148.3154.53
RSS0.8930.3160.44440.1445.480.7620.4260.57351.7754.76
AR-M5P0.8580.3460.50344.0151.550.7360.4470.615 54.3458.75
AR-RSS0.9220.2840.38736.1439.650.7180.4600.61955.9559.08
RSS-M5P0.8820.3200.46140.7047.230.7420.4380.61253.2358.44
AR-M5P-RSS0.8590.3450.50243.9151.420.7400.4600.61955.9559.08
SPEI12
AR0.9770.1330.21016.9521.450.8330.3390.48340.6245.21
M5P0.9250.2540.37332.3338.060.8690.2810.42433.5639.70
RSS0.9090.3210.43440.7444.260.7110.4720.61756.4357.82
AR-M5P0.9260.2490.37031.6737.800.8510.3060.4436.6241.62
AR-RSS0.9540.2090.29826.5530.410.8020.4070.5348.6449.42
RSS-M5P0.9340.2430.35030.8635.710.8490.3130.4537.4442.20
AR-M5P-RSS0.9280.2460.36531.2137.220.8470.3240.45638.7442.75
SPEI24
AR0.9750.1400.19719.2822.190.9080.2270.40318.1828.92
M5P0.9600.1680.24723.0927.900.9280.2080.38916.6027.89
RSS0.9690.1560.22121.4724.910.9150.2320.40118.5428.75
AR-M5P0.9580.1820.25425.0928.710.8840.2690.43621.4931.30
AR-RSS0.9590.1830.25125.1328.380.8780.2800.44122.3731.62
RSS-M5P0.9630.1720.24123.7027.160.8740.2830.44022.6831.58
AR-M5P-RSS0.9620.1730.24223.8427.320.8860.2700.43321.5931.09
Numbers in boldface indicate the ideal values.
Table 6. Statistical performance of M5P model in predicting SPEIs at the test stations.
Table 6. Statistical performance of M5P model in predicting SPEIs at the test stations.
ModelsStatistical Indices
RMAERMSERAERRSE
SEPI6-Bogra0.8020.440.5854.5059.47
SEPI6-Khulna0.7930.450.6054.1460.65
SEPI6-Mymensingh0.7870.460.6155.3761.37
SEPI6-Rangpur0.7910.460.6156.0960.85
SEPI9-Bogra0.8820.310.4638.0346.65
SEPI9-Khulna0.8500.370.5243.5852.32
SEPI9-Mymensingh0.8800.330.4740.8146.99
SEPI9-Rangpur0.8610.350.5142.7950.38
SEPI12-Bogra0.9380.240.3429.0434.34
SEPI12-Khulna0.8990.300.4436.1543.37
SEPI12-Mymensingh0.9250.260.3731.5737.43
SEPI12-Rangpur0.9290.240.3730.1336.65
SEPI24-Bogra0.9620.170.2720.6026.45
SEPI24-Khulna0.9270.240.3728.2536.17
SEPI24-Mymensingh0.9610.180.2721.9426.80
SEPI24-Rangpur0.9660.160.2619.2825.17
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Elbeltagi, A.; AlThobiani, F.; Kamruzzaman, M.; Shaid, S.; Roy, D.K.; Deb, L.; Islam, M.M.; Kundu, P.K.; Rahman, M.M. Estimating the Standardized Precipitation Evapotranspiration Index Using Data-Driven Techniques: A Regional Study of Bangladesh. Water 2022, 14, 1764. https://doi.org/10.3390/w14111764

AMA Style

Elbeltagi A, AlThobiani F, Kamruzzaman M, Shaid S, Roy DK, Deb L, Islam MM, Kundu PK, Rahman MM. Estimating the Standardized Precipitation Evapotranspiration Index Using Data-Driven Techniques: A Regional Study of Bangladesh. Water. 2022; 14(11):1764. https://doi.org/10.3390/w14111764

Chicago/Turabian Style

Elbeltagi, Ahmed, Faisal AlThobiani, Mohammad Kamruzzaman, Shamsuddin Shaid, Dilip Kumar Roy, Limon Deb, Md Mazadul Islam, Palash Kumar Kundu, and Md. Mizanur Rahman. 2022. "Estimating the Standardized Precipitation Evapotranspiration Index Using Data-Driven Techniques: A Regional Study of Bangladesh" Water 14, no. 11: 1764. https://doi.org/10.3390/w14111764

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