You are on page 1of 17

Journal of Hydrology 564 (2018) 266–282

Contents lists available at ScienceDirect

Journal of Hydrology
journal homepage: www.elsevier.com/locate/jhydrol

Research papers

Simulation and forecasting of streamflows using machine learning models T


coupled with base flow separation

Hakan Tongala, , Martijn J. Booijb
a
Department of Civil Engineering, Engineering Faculty, Süleyman Demirel University, 32260 Isparta, Turkey
b
Department of Water Engineering and Management, Faculty of Engineering Technology, University of Twente, Enschede, the Netherlands

A R T I C LE I N FO A B S T R A C T

This manuscript was handled by G. Syme, Efficient simulation of rainfall-runoff relationships is one of the most complex problems owing to the high
Editor-in-Chief, with the assistance of Bellie number of interrelated hydrological processes. It is well-known that machine learning models could fail in
Sivakumar, Associate Editor simulating streamflows from only meteorological variables in the absence of antecedent streamflow values. The
Keywords: main reason for this could be low and lagged relationships between streamflow and meteorological variables. To
Simulation overcome this inefficiency, for the first time, we developed a simulation framework by coupling a base flow
Forecasting separation method to three machine learning methods. It was demonstrated that separating streamflow into
Base flow separation different components such as base flow and surface flow can be useful for improving simulation and forecasting
Support vector regression
capabilities of machine learning models. We simulated streamflow in four rivers in the United States with
Artificial neural networks
Support Vector Regression (SVR), Artificial Neural Networks (ANNs) and Random Forest (RF) as a function of
Random forest
precipitation (P), temperature (T) and potential evapotranspiration (PET). We concluded that the base flow
separation method improved the simulation performances of the machine learning models to a certain degree.
Apart from the simulation scheme, we also employed a forecasting scheme by using the antecedent observed
discharge values in addition to P, T, and PET. We discussed performances of models in simulation and fore-
casting of streamflow regarding model types, input structures and catchment dynamics in detail.

1. Introduction (i.e. data-driven models) such as Artificial Neural Networks (ANNs) and
Support Vector Machines and Regressions (SVM and SVR) have re-
Streamflow forecasting and simulation is fundamental for wa- ceived increasing attention in hydrology for instance for rainfall-runoff
tershed planning and sustainable water resources management. In the modelling (Hosseini and Mahjouri, 2016; Lin and Chen, 2004; Sudheer
past decades, impacts of climate change and anthropogenic activities et al., 2002; Tiwari and Chatterjee, 2010a; Wu and Chau, 2011; Young
have possibly changed the timing, frequency and intensity of extreme et al., 2017), groundwater level forecasting (Chang et al., 2015;
events, and it is expected that this will continue in the future (Demirel Ebrahimi and Rajaee, 2017; Gholami et al., 2015; Mukherjee and
et al., 2013; Halgamuge and Nirmalathas, 2017; Lansigan, 2009; Naidu Ramachandran, 2018; Suryanarayana et al., 2014) and rainfall fore-
et al., 2012; Sang et al., 2014; Spinoni et al., 2017; Tian et al., 2014; casting (Feng et al., 2015; Hong and Pai, 2007; Lin et al., 2009; Lin and
Vormoor et al., 2016; Yang et al., 2018). Hydrological models can be Jhong, 2015; Nikam and Gupta, 2014; Pai and Hong, 2007; Yu et al.,
used to investigate the impacts of climate change and human activities 2017). We have employed ANNs and SVR in this study since the former
on hydrological processes. Data-driven models can be used for this one is one of the most robust approaches for hydrological forecasting
purpose. These models do not explicitly incorporate hydrological pro- that recognizes the nonlinear relationship between the input and output
cesses (Shortridge et al., 2016) and therefore should be used with care (Adamowski and Sun, 2010; Badrzadeh et al., 2016; Fotovatikhah et al.,
for conditions outside the observed domain. However, some studies 2018). Moreover, several studies indicated that SVR is more powerful
showed that data-driven models can be practical, easy to use, relatively than ANN in capturing nonlinear hydrological relationships due to the
straightforward, can give insight into the governing dynamics and employed structural risk minimization contrary to the empirical risk
eliminate some deficiencies of other types of models (Booker and minimization error (Maity et al., 2010; Wang et al., 2009). There are
Woods, 2014; Han et al., 2007; Humphrey et al., 2016; Masselot et al., several studies indicating that ANN and SVR can be preferred over
2016; Noori and Kalin, 2016). Therefore, machine learning algorithms other machine learning models such as genetic programming and the


Corresponding author.
E-mail addresses: hakantongal@sdu.edu.tr (H. Tongal), m.j.booij@utwente.nl (M.J. Booij).

https://doi.org/10.1016/j.jhydrol.2018.07.004
Received 9 April 2018; Received in revised form 5 June 2018; Accepted 2 July 2018
Available online 05 July 2018
0022-1694/ © 2018 Elsevier B.V. All rights reserved.
H. Tongal, M.J. Booij Journal of Hydrology 564 (2018) 266–282

adaptive neural-based fuzzy inference system (Wang et al., 2009) and several studies focused on simulation models for rainfall-runoff re-
they have been shown to give promising results in many fields of hy- lationships (e.g., Chen et al., 2014; Garbrecht, 2006; Ju et al., 2009; Lee
drology and water resources research (Aqil et al., 2007a). Therefore, we et al., 2008) their models can be largely considered as forecasting
have employed ANN and SVR for the modeling aim in this study. models rather than as simulation models based on the model input
Recently, other data-driven techniques which are variations of de- structure.
cision trees such as Classification And Regression Tree—CART, M5 Separating streamflow into different components such as base flow
model tree and random forest have attracted the attention of re- and surface flow can be useful for improving simulation and forecasting
searchers. Among others, Shortridge et al. (2016) compared several capabilities of machine learning models. In the last decade, to decom-
machine learning algorithms such as generalized additive models, pose the streamflow series into several sub-series the wavelet decom-
multivariate adaptive regression splines, artificial neural networks, position method was combined with several machine learning models
random forest and M5 cubist models in rainfall-runoff simulation for and promising results have been reported in a forecasting scheme (Liu
five rivers in the Lake Tana basin, Ethiopia. They concluded that the et al., 2014; Rezaie-balf et al., 2017). This method is considered to be an
machine learning algorithms reduced the errors in simulation in com- efficient signal processing tool for extracting information of different
parison to the physically-based models developed for the considered time scales, especially in the time-frequency domain with wavelet
basins. Kim and Pachepsky (2010) showed that an ANN model coupled coefficients. However, the optimum level of decomposition of the time
with regression trees could be helpful in the reconstruction of missing series is subjective, and there is no straightforward method for this aim
rainfall data and could enhance streamflow predictions of a physically- (Badrzadeh et al., 2016). Further, application of wavelet analysis re-
based model—the Soil and Water Assessment Tool. Kumar et al. (2016) quires a certain level of expertise, because one should be aware of
forecasted suspended sediment as a function of hydro-meteorological underlying assumptions of different mother wavelet functions such as
variables such as rainfall and streamflow. They compared several ma- Haar, Morlet, and Coiflet1. As indicated by Du et al. (2017), this
chine learning models including least square–SVR, ANN, CART and M5 method may also be widely incorrectly used in literature since the au-
model tree and found that machine learning models successfully fore- thors of these studies may not be aware that they erroneously include
casted suspended sediment. Above studies indicated that machine the information of the future values to be forecasted. However, with the
learning models could efficiently be used for hydrological applications, proposed methodology in this study, the straightforward base flow se-
especially in a forecasting context. It is reported that the random forest paration can be functional for capturing the nonlinear relationship
model produces promising results in comparison to ANN and SVR between streamflow and meteorological variables by introducing a
(Breiman, 2001; Liaw and Wiener, 2002). This is mainly because the temporary memory component to the machine learning models. Be-
random forest model does not have an over-fitting problem with an cause, the base and surface flows provide different useful information
ensemble of permutations (Leasure et al., 2016). The model also cal- about variations, periodicity, and trends. These processes could result
culates importance of variables among predictor variables, and conse- from different hydro-meteorological conditions and using them as input
quently, it is much more interpretable than methods such as ANNs and to the machine learning models providing additional temporal in-
SVR (Baudron et al., 2013). However, to the best of our knowledge, formation about the physical processes can be included in a straight-
there is no study on the comparison of hydrologic forecasting and si- forward manner.
mulation performances of RF, ANN, and SVR. This will eventually Several studies employed the idea of separating streamflow dy-
provide additional insights into the performances of the models in the namics into sub-dynamics. Hu et al. (2001) developed a range-depen-
perspective of water resources planning and management. dent network by considering several ANNs to model the different
To this end, it is essential to discriminate between the terms fore- streamflow regimes (e.g. low, medium, and high) and concluded that a
casting and simulation. As indicated by Klemeš (1986), a hydrological modular approach is better than a single model. Solomatine and Xue
simulation model can be defined as a mathematical function between the (2004) coupled the M5 model tree and ANNs to forecast floods and
output of a hydrological system (e.g. discharge) and other hydro-cli- found that this hybrid model is better than its single counterparts. Wang
matic variables such as precipitation and temperature. However, a hy- et al. (2006) grouped a hydrological time series according to its specific
drological forecasting model connects the output of a hydrological system dynamics with a threshold variable. They reported that different models
not only with hydro-climatic variables but in addition also with ante- for each group could give better results than a single global model fitted
cedent observed values of the considered output of the hydrological to all data. Corzo and Solomatine (2007) employed three different flow
system. It is evident that a simulation model can be operated in a separation techniques to improve the forecasting performance of an
forecasting mode by considering the past observed values of the model ANN model and found that the modular ANN gave a better result than a
output. Simulation of streamflows is especially crucial for specific hy- global ANN model. Rezaie-balf et al. (2017) used a wavelet decom-
drological applications such as obtaining continuous streamflow re- position method to enhance a modular model that is comprised of
cords in ungauged basins where measured streamflow is not available. multivariate adaptive regression splines and M5 model trees to forecast
As indicated by Chen et al. (2014), machine learning models are quite groundwater level. They found that the modular model is better than its
promising in streamflow forecasting by incorporating observed ante- single counterparts based on performance indices. These studies re-
cedent streamflow values but could give a poor performance by only vealed that separating the output of a hydrological system into its dif-
considering other hydro-meteorological variables (i.e. simulation). The ferent components can be used for the improvement of model perfor-
nonlinear relationship with a low correlation between streamflow and mances. However, some drawbacks associated with these studies are:
other hydro-meteorological variables could lead to such poor simula- (1) coupling of different models into a single model could be difficult
tion results. Further, machine learning models do not incorporate for modellers who are unfamiliar with two or more models, (2) in-
physical equations or relationships and do rely on mainly linear re- tegrating several models within a single forecasting scheme could be
lationships between streamflow and other variables. Noise and the re- computationally expensive, and (3) optimization of several models for
latively small length of data sets in addition to different hydro-me- simulation purposes (not for forecasting purposes) necessitates curve
teorological characteristics between training and testing periods known fitting for each model to each component potentially leading to error
as non-stationarity could also deteriorate this relationship. Thus, it is accumulation and considerable uncertainty in the output. Further,
important to analyze, and if necessary and feasible, improve the cap- previous studies have mainly focused on forecasting performances of
abilities of streamflow simulation of machine learning models. To the machine learning models. All these considerations encourage the im-
best of our knowledge the streamflow forecasting and simulation cap- provement of machine learning models by using a simple base flow
abilities of machine learning models like SVR, ANN and random forest separation technique. With the proposed methodology, it is easily
have not been examined rigorously and jointly on a daily scale. While possible to separate streamflow into base and surface flows. Then, these

267
H. Tongal, M.J. Booij Journal of Hydrology 564 (2018) 266–282

can be utilized in straightforward simulation and forecasting schemes this seasonal variation of precipitation, it also varies spatially where
for a single machine learning model without adding two or more heavier precipitation is concentrated in the northern part of the state.
models. To this end, the aims of this study are (1) analysing whether a All streamflow time series span from 01.01.1948 to 31.12.2001,
base flow separation technique could improve the simulation cap- since we do not have more recent data after 2001. The data were di-
abilities of standalone machine learning models (i.e. SVR, ANN and vided into three subsets for the aims of training, calibration, and vali-
random forest) and (2) investigating the performances of standalone dation. The first 80% of the record is used for training, about 16% is
machine learning models in simulation and forecasting of rainfall- reserved for the calibration and the remaining two years (about 4% of
runoff relationships on a daily scale. To the best of our knowledge, the total dataset) is used for the validation. We kept the sizes of training
these research aims have not been addressed and tested so far. and calibration periods as large as possible because, generally, machine
learning models require a high amount of data value to capture the
2. Study area and data dynamics of streamflow. To ensure the generalizability, a high amount
of data demand is coupled in the simulation scheme. Because, instead of
The streamflow measurements used in the analyses were acquired using the observed values, simulated values with less correlated to the
from the U.S Geological Survey for four stations in the United States output values were used in the input structure. Further, we also jointly
(US), namely Station #11427000 North Fork Dam—North Fork River, optimize both parameters and structures of the models based on the
Station #12027500 Chehalis River, Station #10309000 East Fork of the training and calibration period.
Carson River, and Station #11342000 Sacramento River. These rivers
have been chosen because their basins have different sizes with distinct 3. Methodology
topographic and hydro-climatic regimes from different regions of the
U.S. However, the developed methodology can be applied to any un- This section includes descriptions of Support Vector Regression
regulated river that has observations of discharge, precipitation and (SVR), Random Forest (RF), and Artificial Neural Networks (ANNs)
temperature. models with information about model parameters and input structures
North Fork American River is the longest branch of the American for streamflow simulation and forecasting. Furthermore, the formulas
River in northern California (Kang and Lee, 2014) and its drainage area of several performance indices and the methods for the estimation of
is about 875 km2. Precipitation is the primary forcing input that varies evapotranspiration and base flow separation are presented.
with elevation where the mean annual precipitation is about 1651 mm
in Blue Canyon at 1676 m above mean sea level (a.m.s.l.), and it is 3.1. Support vector regression
about 813 mm in Auburn at 393 m a.m.s.l. (Jeton et al., 1996; Kang and
Lee, 2014). Dettinger et al. (2004) showed that about two-thirds of the Developed by Vapnik and Chervonenkis (1964) and Vapnik (1995),
streamflow originates from rainfall and snow-melt in winter while less a support vector machine is a kind of machine learning technique that is
than one-third originates from snow-melt in spring. widely applied to classification and regression problems. Support
Located in the south-west region of Washington, the Chehalis River Vector Regression (SVR) stands for utilizing support vector machines
basin is the third largest river basin in the State of Washington with a for the regression. The method embodies structural risk minimization in
drainage area of about 6900 km2. The basin is mainly rain-fed with contrast to the empirical risk minimization approach that is commonly
mean annual precipitation levels of 1520–2030 mm (Henning et al., used by statistical learning methods such as artificial neural networks
2007). Frequent flooding is a significant problem for the basin, there (Lin et al., 2006). The structural risk minimization method minimizes
have been 45 floods since 1945. The Chehalis River basin shows typical empirical risk to achieve a good generalization capacity by reducing the
hydrological characteristics of rivers in western Washington with generalization error as opposed to the training error (Belayneh et al.,
minimal snowmelt input. High flows are directly related to intense 2014; Maity et al., 2010). The basic premise of SVR is the nonlinear
rainfall, and critical low flows can be observed in late summer. Seasonal mapping of the original data into a higher dimensional feature space by
high flows occur from November to March following the typical winter using a kernel function and then performing linear regression in the
rainfall. Flow decreases through the spring and summer months due to feature space (Kalra et al., 2013). Details of this technique are well
relatively dry climate conditions and absence of a snowpack established in the literature and can be found elsewhere (Maity et al.,
(Kimbrough et al., 2006). 2010; Yoon et al., 2011; Yu et al., 2017).
East Fork of the Carson River basin is a snowmelt-dominated There are several kernels that can be used in SVR such as linear,
mountainous basin in the western United States. The western United polynomial, sigmoid and radial basis functions. It is well documented
States has a climate with semi-arid to arid conditions with the highest that the radial basis kernel function performs better than other kernels
precipitation totals occurring in the mountainous region (Hay and (Barzegar et al., 2017; Lin et al., 2006; Liong and Sivapragasam, 2002;
Clark, 2003). Between 50 and 70% of the annual precipitation in these Wu and Wang, 2009; Yu and Liong, 2007). Further as indicated by
areas falls as snow and constitutes the bulk of its annual flow. The Dibike et al. (2001), the radial basis kernel function outperformed other
Carson River has a drainage area of about 920 km2 with elevations kernels in rainfall-runoff modeling. It is more compact than other
between approximately 1600–3000 m (Hay et al., 2000). Frequent rain- kernel functions and reduces the computational time with a higher
on-snow events are a typical feature of the basin. The relation between generalization performance (Choy and Chan, 2003; Maity et al., 2010).
potential evapotranspiration and precipitation is about 12:1 (Grantz Therefore, we used the radial basis kernel function in this study.
et al., 2007).
Sacramento River is the largest river in the State of California by 3.2. Random forest
carrying about 31% of the total surface water of the state, and it is the
second largest U.S. river draining into the Pacific Ocean. In the Random Forests (RFs) are an ensemble of decision trees based on
downstream part, there is the largest storage reservoir of California, bootstrap aggregation with random variable selection (Breiman, 2001).
Shasta Lake. The reservoir supplies water for agricultural, municipal The RF method is a robust nonparametric approach for modeling and
and industrial use, is used for navigation and supports a wide variety of classification of large nonlinear, noisy, and multivariate correlated data
fish habitats (Sapin et al., 2017). The watershed has a Mediterranean (Mohr et al., 2017; Strobl et al., 2008). A single decision tree is well
climate characterized by wet winters and dry summers with a mean known to have a high variance and is prone to noise while showing
monthly temperature varying between 4.5 and 22.4 °C (Chiu et al., statistical instability (Zhu and Pierskalla, 2016). Statistical instability
2017). Little or no precipitation falls between May and October and means that a small perturbation in the training data leads to substantial
most precipitation occurs between November and April. In addition to changes in model outputs. To eliminate over-fitting and statistical

268
H. Tongal, M.J. Booij Journal of Hydrology 564 (2018) 266–282

instability, bootstrap aggregation is used to build many decision trees (e.g., Chen et al., 2016; Das et al., 2008; Goyal et al., 2014; Salerno and
by randomly sampling the observed dataset with replacement (Mohr Tartari, 2009) in addition to the relatively few data requirements. The
et al., 2017). Predictions for unobserved data can be obtained by equation for potential evapotranspiration is expressed below:
averaging predictions from an ensemble of trees, or forest. The RF
ET0 = 0.0023·0.408·Ra (Tavg + 17.8)·TD 0.50 (1)
method advances bootstrap aggregation not only by using a subset of
training data, but also by using randomly sampled input variables or where Ra is the extraterrestrial radiation (MJ m−2 d−1)
and can be ob-
features for splitting tree nodes (Zhu and Pierskalla, 2016). More details tained from equations based on latitude (Allen et al., 1998). The con-
can be found in Xu and Valocchi (2015) and Schnier and Cai (2014), stant 0.408 is used to convert the radiation to evaporation equivalents
among others. in mm. Tavg indicates the average daily temperature (°C), and TD (°C) is
the difference between the maximum and minimum daily temperature.
3.3. Artificial neural networks
3.5. Base flow separation
Artificial Neural Networks (ANNs) are information processing sys-
tems having interconnected neurons to reconstruct complex nonlinear A separation of the hydrograph is useful for determining runoff
input-output relationships. ANNs are composed of simple processing components originating from different dynamics of a catchment such as
elements (nodes) connected by weighted synaptic connections (Tiwari delayed and smoothed responses to rainfall events. Base flow and direct
and Chatterjee, 2010b). ANN models are specified by network structure runoff can be linked with low and high-frequency variability of
or topology, activation functions, and training algorithms. Among dif- streamflow, respectively. As indicated by Eckhardt (2005) it is possible
ferent types of ANNs, multi-layer perceptron feed-forward neural net- to identify base flow with low-pass filtering of streamflow. Eckhardt
works are one of the most frequently used ANNs (e.g., Lohani et al., (2008) compared several base flow separation methods (e.g., HYSEP1,
2011; Singh et al., 2009). In these networks, each neuron is related to HYSEP2, HYSEP3, PART, and UKIH) and concluded that the recursive
all neurons in the preceding layer, and the output of each layer is taken digital filters yield extensively smooth time series of base flow. There-
as an input for the subsequent layer. Several studies indicated that fore, the one-parameter recursive single-pass digital filter method was
Newton or quasi-Newton optimization techniques such as Levenberg- used in this study. Its efficiency was also shown in some hydrological
Marquart (Hagan and Menhaj, 1994) and Broyden–- studies (e.g., Arnold et al., 1995; Sawicz et al., 2011) in addition to
Fletcher–Goldfarb–Shanno algorithms are fast in convergence time, being computationally efficient. A straightforward formulation of the
being less easily trapped in local minima (Aqil et al., 2007b; Badrzadeh one-parameter filter is (Chapman, 1991; Lyne and Hollick, 1979):
et al., 2013), provide enough complexity to capture nonlinear and dy-
α 1−α
namic relationships in hydrological processes and have better gen- bt = bt − 1 + Qt
2−α 2−α (2)
eralization capabilities than the descent gradient back-propagation al-
gorithm (Aqil et al., 2007a). Thus, the three-layer feed-forward neural where bt and bt − 1 are the filtered base flows at time step t and t −1, Qt is
network with the Levenberg-Marquart back-propagation algorithm was the observed streamflow at a time step t , and α is the filter parameter
used in this study. The logistic sigmoid function is more widely used in which was set to 0.925 based on the studies of Eckhardt (2008), Sawicz
hydrological studies than the hyperbolic tangent sigmoid function (e.g., et al. (2011), and Razavi and Coulibaly (2013).
Karunasinghe and Liong, 2006; Melesse et al., 2011; Riad et al., 2004).
However, as indicated by Zadeh et al. (2010) and Maier and Dandy 3.6. Determination of input structure and calibration of model parameters
(1998), the tangent sigmoid function is not only faster in training of the
network, but also provides better predictions than the logistic sigmoid Antecedent values of input variables could be considered to re-
function. Wang et al. (2006) showed that the output of the logistic present the catchment characteristics in a data-driven model
sigmoid function is constrained in a relatively smaller range than the (Solomatine and Xue, 2004). The components of a hydrograph, the base
tangent sigmoid function that leads to the outputs of ANN models with flow and surface flow, are representative of slowly varying groundwater
the logistic sigmoid function being more sensitive to the input interval. and direct response of the basin to the rainfall events, respectively.
Moreover, a linear transfer function in the output neuron in a situation These are related to different storages (i.e. groundwater and surface
with an unknown output range of the time series is advantageous water), and while the surface flow can be considered as a sudden re-
(Haykin, 1999). Therefore, a combination of a hyperbolic tangent sponse, the base flow can be recognized as a slowly varying long-term
function and a linear transfer function in the hidden and output layer response of a basin (Serinaldi et al., 2014). Therefore, it is possible to
was used since this is effective when extrapolating beyond the range of represent different storages and corresponding responses of the catch-
training data (Maier and Dandy, 2000). ment with antecedent values of base flow and surface flow.
The most significant input combination of precipitation (P), poten-
3.4. Estimation of evapotranspiration tial evapotranspiration (PET) and temperature (T) in addition to base
flow and surface flow for the simulation models was identified through
Numerous methods have been developed to estimate potential statistical analyses using cross-correlation, auto-correlation (ACF) and
evapotranspiration that can mainly be divided into two broad groups: partial auto-correlation (PACF) functions in addition to the forward
(1) radiation-based methods (e.g., Priestley-Taylor and Turc methods) stepwise regression method for each catchment. A qualitative analysis
and (2) temperature based methods (e.g., Hargreaves, Blaney-Criddle, of cross-correlation structures revealed which antecedent values of
and Thornthwaite methods). As shown in Tukimat et al. (2012), ra- variables profoundly influence the discharge. Further, ACF and PACF
diation-based methods could be advantageous over the temperature- could suggest the influencing antecedent variables with a possible
based methods. However, these methods are parameter-rich methods periodic structure (Lin et al., 2006). In the forward stepwise regression
and require reliable measurements of meteorological variables. In this method, input variables are subsequently added to the single best input
study, the lack of reliable meteorological data of humidity, wind speed, variable up to the allowed number of antecedent values determined
and solar radiation made it necessary to use a temperature-based from the correlation analyses. Evaluation of changes in adjusted-R2
method. Since the Thornthwaite method (Thornthwaite, 1948) could be indicates whether the added new parameter improves the model more
misleading since it gives zero potential evapotranspiration for tem- than would be expected by chance. While there are several nonlinear
peratures below zero degree Celsius (Rhee and Im, 2017), the widely- methods for determining the input structure especially for a forecasting
used Hargreaves (Hargreaves and Samani, 1985) method can be a re- scheme such as partial mutual information (May et al., 2008) and the
liable alternative since its validity has been shown in numerous studies gamma test (Stefánsson et al., 1997), here, we both qualitatively and

269
H. Tongal, M.J. Booij Journal of Hydrology 564 (2018) 266–282

quantitatively evaluated the response times (i.e. memory components) were determined by a trial-and-error process by using a grid-search
of catchments as a function of other hydro-climatic variables and cor- technique on the training and calibration datasets. The best parameter
relations among the most significant input variables and the output sets are kept that maximize the objective function value calculated for
variable. After a number of experiments on the training dataset aimed the calibration data set. A range of values for parameter ε between 0.01
at finding the relevant inputs, the following common input structure and 0.5 with an increment of 0.01, for the cost constant (C ) between 1
was selected: and 5 with an increment of 0.1, and for the σ parameter between 0.01
and 1 with an increment 0.01 were evaluated.
Qt = f (Qtg− 1, Qts− 1, PETt , PETt − 1, Tt , Tt − 7, Tt − 14, Tt − 21, Pt , Pt − 1, Pt − 2, ...,Pt − 7)
(3)
3.6.2. Parameters of RF
where Qtg− 1
and are the simulated (not observed) base flow and
Qts− 1 For a RF model, two parameters are needed to be defined, the
surface flow of the antecedent day, respectively. It was also possible to number of variables/features to be used in each tree-building process
obtain different input structures for different catchments and different (mtry ) and the number of trees to be built in the forest to run (ntree ). As
models; however, this would significantly complicate the analysis of the indicated by Yu et al. (2017), sometimes even mtry = 1 can provide a
performance of the proposed methodology since one cannot be sure reasonable performance. For classification and regression tasks, re-
whether the differences in the results came from the input structure, the spectively, one-third and the square root of the total number of input
developed methodology or the employed model. Therefore, we used a variables is suggested (Breiman, 2003; Liaw and Wiener, 2002). In this
common best performing input structure to eliminate additional un- study, for the parameter mtry , several values were considered varying
certainties due to different input variables. from 1 to twice the number of input variables. The generalization error
To analyze whether the simulation with the base flow separation increases beyond an optimal value of ntree and generally, a value of 500
method improved the simulation of streamflow values, we also con- is used (Booker and Woods, 2014; Yu et al., 2017). In this study, dif-
sidered a simulation mode without base flow separation. In this case, ferent values for the parameter ntree varying from 100 to 1000 with an
the model structure is: increment of 100 were tried.
Qt = f (Qtsim
− 1 , PETt , PETt − 1, Tt , Tt − 7, Tt − 14, Tt − 21, Pt , Pt − 1, Pt − 2, ...,Pt − 7 ) (4)
where Qtsimis the simulated discharge at time step t −1. Further, to show 3.6.3. Parameters of ANN
−1
differences in model performances in simulation and forecasting modes, The number of neurons in the input and output layers is dependent
we also analyzed forecasting results by considering the following model on the modeled system/problem, and the number of neurons in the
structure: hidden layer should be determined. The weights between input and
hidden neurons were obtained in the training dataset for various
′ ′
Qt = f (Qtg− 1, Qts− 1 PETt , PETt − 1, Tt , Tt − 7, Tt − 14, Tt − 21, Pt , Pt − 1, Pt − 2, ...,Pt − 7) numbers of hidden neurons up to 10. The performances of trained ANN
(5) models were evaluated for the calibration dataset. Since the random
′ ′ initialization of weights could lead to different performance indices for
where Qtg− 1 and Qts− 1
are observed base flow and surface flow at time step the same number of hidden neurons, we have trained 100 ANNs for
t −1. Prior to the model development, normalization was applied to all each number of hidden neurons in the training stage. Then, the best
input variables as: combination of weights and number of hidden neurons is kept that
X0 −μx maximizes the objective function value calculated for the calibration
Xscaled = data set.
σx (6)

where Xscaled and X0 indicate the scaled and original data, μx and σx
represent the mean and standard deviation of the original data, re- 3.7. Performance indices
spectively. For each machine learning method used in this study, the
optimal model parameters were selected by maximizing the objective To select the appropriate model parameters and to evaluate model
function (Kling-Gupta efficiency, see section 3.7.3) using training and performances, various performance criteria were used. These perfor-
calibration data sets. The Kling-Gupta efficiency objective function mance indices are generally used in hydrological studies because they
eliminates the significant drawbacks of the widely-used Nash-Sutcliffe are reliable and easily interpretable, and are recommended for evalu-
efficiency by considering correlation, deviation and variability, jointly ating the generalizability and reliability of forecasting and simulation
(Muñoz et al., 2014). To this end, in contrast to the time-consuming hydrological models (Doycheva et al., 2017; Legates and McCabe, 1999;
evaluation of different standalone performance indices to select the best Patel and Ramachandran, 2015; Wang et al., 2009). We are aware that
model parameters, the Kling-Gupta efficiency provides a better under- there are also other hydrological metrics, but we intend to give a
standing of model performances by considering different performance straightforward comparison of models in terms of model performances
components obtained from the calibration stage (Gupta et al., 2009). within simulation and forecasting frameworks with the root mean
The models were developed on the training dataset with a range of square error (RMSE), Nash-Sutcliffe coefficient of efficiency (NSE),
parameter values, and the objective function value was calculated for Kling-Gupta efficiency (KGE), volumetric efficiency (VE), index of
the calibration period. The range of parameter values were determined agreement (d), and persistence index (PI) as described below.
by examining several hydrological studies such as Shortridge et al.
(2016) Raghavendra and Deka (2014), Kumar et al. (2016), Naghibi 3.7.1. Root mean square error (RMSE)
et al. (2017), Worland et al. (2018) and from several studies reviewed RMSE is one of the most employed criteria to assess model effi-
by Fotovatikhah et al. (2018). Then, the optimal model parameters ciency:
were kept, and the best models were used to simulate streamflow for
N
the validation period.
RMSE = N−1 ∑ (Oi−Si )2
i=1 (7)
3.6.1. Parameters of SVR
In SVR, three governing parameters, that are the tolerance threshold where N is the length of the data set, Oi and Si are observed and si-
or ε -precision, the regularization parameter (i.e. C, cost constant), and mulated values, respectively. The RMSE gives more importance to
the kernel function parameter of the radial basis kernel function (i.e. σ ), outliers in the data set, and thus it is biased towards errors in the si-
should be optimized. In this study, the values of SVR model parameters mulation of high values (Mekanik et al., 2016).

270
H. Tongal, M.J. Booij Journal of Hydrology 564 (2018) 266–282

3.7.2. Nash-Sutcliffe efficiency (NSE) Table 1


The NSE (Nash and Sutcliffe, 1970) is commonly used in the eva- Parameters of simulation models for four catchments.
luation of performances of hydrological models, and as indicated by Parameters of models SVR ANN RF
Servat and Dezetter (1991) it is one of the best performance metrics for
reflecting the overall fit of a hydrograph. ε C σ Number of hidden mtry ntree
neurons
N
∑ (Oi−Si )2
NSE = 1− iN= 1 North Fork River 0.10 2.0 0.12 3 12 500
∑i = 1 (Oi−O¯ )2 (8) Chehalis River 0.20 2.0 0.09 2 12 500
Carson River 0.20 1.0 0.12 2 1 500
where Ō is the mean of observed values. It varies between −∞ and 1 Sacramento River 0.11 3.4 0.13 5 7 500
(perfect fit). Values between 0 and 1 show an acceptable model per-
formance, while a NSE lower than zero indicates that the mean value of
the series is a better estimation than the constructed model (Moriasi better than the benchmark (McMillan et al., 2013).
et al., 2007).
4. Results and discussion
3.7.3. Kling-Gupta efficiency (KGE)
The Kling-Gupta efficiency is an improvement of the NSE in which In this study, all the models, i.e., SVR, ANN, and RF, were set up for
correlation, bias, and variability are considered jointly to overcome four catchments. To understand whether the base flow separation had
systematic underestimation of peaks and variability in the NSE (Gupta an impact on simulation performances, the simulation models were
et al., 2009). It ranges between −∞ and 1 (perfect fit). The KGE is developed both with and without the base flow separation. Moreover,
defined as: the results for the forecasting mode with base flow separation were
included to show differences between the simulation and forecasting
KGE = 1− (r −1)2 + (α−1)2 + (β−1)2 (9) schemes.
where r is the correlation coefficient, α represents variability and is the
ratio of the standard deviation of simulated and observed time series, 4.1. Simulation results with base flow separation
and β is the bias and obtained by dividing the mean of simulated time
series by the mean of observed time series. The obtained optimal model parameters are shown in Table 1 and
the values of model performance criteria for both calibration and va-
3.7.4. Volumetric efficiency (VE) lidation periods are presented in Tables 2 and 3. It is clear from Table 2
Criss and Winston (2008) proposed the volumetric efficiency (VE) to that ANN models were more efficient in the simulation mode with base
circumvent some problems associated with the Nash-Sutcliffe efficiency flow separation in the calibration stage for the North Fork, Chehalis and
such as overemphasizing high flows relative to other observations due Sacramento rivers while SVR was more efficient for the Carson River.
to the squared deviations. VE varies between −∞ and 1 and is calcu- However, the SVR model outperformed other models in the validation
lated as: period for the North Fork, Carson, and Sacramento rivers; for the
N Chehalis River, the RF model showed a better performance. For the
∑i = 1 |Si−Oi | Chehalis River, the performance statistics of the ANN model are close to
VE = 1− N
∑i = 1 Oi (10) those of the RF model for the validation period.
Interestingly, there are some improvements in the model perfor-
mances in the validation period compared to the calibration period. For
3.7.5. Index of agreement (d) instance, for the North Fork River, while the ANN model has the best
The Index of Agreement (d) (Willmott, 1981) is a standardized performance in the calibration period, the SVR model showed the best
measure of the degree of the agreement between observed and modeled performance in the validation period. Similarly, in the validation per-
values and bounded between 0 (no agreement at all) and 1 (perfect fit). iods of the Chehalis and Sacramento rivers, overall RF and SVR per-
It is given by: formed better than the ANN model which showed the best performance
N
∑i = 1 (Oi−Si )2 in the calibration period for these rivers. In summary, SVR (for the
d = 1− N North Fork, Carson, Sacramento rivers) and RF (for the Sacramento
∑i = 1 (|Si−O¯ | + |Oi−O¯ |)2 (11) River) performed better than the ANN models in the validation period.
Similarly to NSE, the index of agreement is also sensitive to extreme It can be concluded that while better calibrated, the generalization
values due to the squared form. However, it can quantify the additive capability of ANN models is less than that of SVR and RF models for the
and proportional differences in the observed and simulated means and considered rivers. This finding is similar to the finding of Shortridge
variances (Legates and McCabe, 1999). et al. (2016) where the authors simulated monthly discharges of five
rivers with machine learning methods such as ANN and RF. They in-
dicated that RF performed better than the ANN model with the standard
3.7.6. Persistence index (PI)
formulation that includes antecedent values of monthly total pre-
The PI proposed by Kitanidis and Bras (1980) compares the outputs
cipitation, average temperature, and a total percentage of agricultural
of a model with the best estimate for the future as given by the last
land cover.
observation (Randrianasolo et al., 2011). It is assumed that the mod-
It is well known that in some cases the generalization capability for
elled process is a Wiener process in which the variance is a linear
forecasting of ANN models is moderate (Goyal et al., 2014; Yoon et al.,
function of time. In this case, the best estimate of the future is given by
2011). Interestingly, we found a similar result for the simulation cap-
the last observation (Kitanidis and Bras, 1980):
ability of ANN models (Tables 2 and 3). The performance indices of
N
∑i = 1 (Oi−Si )2 ANN models in the calibration period are higher than those in the va-
PI = 1− N lidation period, except RMSE (for all rivers), KGE (for the Chehalis
∑i = 1 (Oi−Oi − L )2 (12)
River) and PI (for the Carson and Sacramento rivers), indicating a poor
where Oi − L is the last observed value at time i minus the lag time L . The generalization capability in streamflow simulation. As indicated by Wu
PI evaluates the model performance against a benchmark (i.e. the last and Chau (2011) a lagged prediction effect could emerge in ANN
observed flow) where a score of 1 indicates a perfect fit and 0 is no modelling studies especially when only antecedent flow values are used

271
H. Tongal, M.J. Booij Journal of Hydrology 564 (2018) 266–282

Table 2
Performance indices obtained from the simulation models for all rivers for the calibration period with base flow separation.
North Fork River Chehalis River Carson River Sacramento River

SVR ANN RF SVR ANN RF SVR ANN RF SVR ANN RF

RMSE 1.46 1.38 1.44 1.65 1.32 1.94 0.68 0.87 1.21 2.84 2.09 2.38
NSE 0.83 0.85 0.84 0.91 0.94 0.88 0.80 0.67 0.37 0.70 0.83 0.79
KGE 0.91 0.91 0.86 0.94 0.91 0.94 0.81 0.76 0.35 0.74 0.89 0.75
VE 0.72 0.74 0.72 0.77 0.81 0.78 0.67 0.57 0.27 0.70 0.72 0.62
d 0.96 0.96 0.95 0.98 0.99 0.97 0.96 0.90 0.69 0.90 0.95 0.93
PI 0.39 0.46 0.41 0.27 0.53 −0.02 −5.59 −9.75 −19.86 −0.48 0.19 −0.04

Bold numbers indicate the best value of the considered performance indices for the rivers.

as input. In this study, we found no lag effect in the simulated discharge RF model gave somewhat unacceptably biased simulation values be-
values as can be seen in Figs. 1–4. tween 0 and 0.5 m3 s−1. Possible reasons for this result could be the
Observed and simulated streamflows using the optimal models are better generalization capability of SVR due to the structural risk mini-
shown in Figs. 1–4. A good overall match between the daily observed mization approach that leads to an optimal global solution (see Gizaw
and simulated values can be seen except for the Carson River. All and Gan, 2016; Kumar et al., 2016; Shrestha and Shukla, 2015) and
models show their worst performance for this river, especially the RF complex rainfall-runoff relationships in the Carson River, which will be
model. The poor simulation results of the RF model could be due to the discussed in Section 4.4. To understand whether the base flow separa-
model parameters and catchment dynamics. It is clear that the RF tion method increased efficiencies of the models, we also developed
model, of which the mtry parameter is equal to 1, is unable to simulate simulation models without the base flow separation method in the
the catchment dynamics of the Carson River. Some studies indicated following section.
that a RF model with one variable/feature (i.e., mtry = 1) resulted in a
good accuracy (e.g., Liaw and Wiener, 2002), whereas one variable
4.2. Simulation results without base flow separation
seems to be insufficient to obtain good simulation results in this study.
It should be noted that the random forest model is a rule-based non-
Simulation results without base flow separation were obtained by
parametric approach where the model output is obtained with a mean
using Eq. (4) to analyse the contribution of the proposed simulation
value of an ensemble of trees. In some cases, the majority of classes
method. From Tables 1–5, it is clear that models developed using the
could lead to a bias in the output due to the averaging process.
base flow separation method are superior to the models without base
Therefore, rather than an averaging method, linear models can be
flow separation in both calibration and validation periods with the
employed in each leaf of ensemble trees as in the M5 model tree.
exceptions of the RF model for the Chehalis and Carson rivers. The RF
Several studies indicated that the M5 model tree could be successfully
model of the Chehalis River gave better results in the calibration period
applied to hydrological problems (e.g., Bhattacharya and Solomatine,
without base flow separation (Table 4). In the validation period, the
2005; Esmaeilzadeh et al., 2017; Rezaie-balf et al., 2017; Solomatine
KGE value (i.e., 0.92 in Table 5) is better than its counterpart (i.e., 0.89
and Xue, 2004), thus, can be a reliable alternative to a random forest
in Table 3) for this model. For the Carson River, all performance indices
model in simulation of streamflow which deserves future investigation.
of the RF model in the calibration (Table 4) and validation (Table 5)
Additionally, the catchment dynamics of the Carson River could have
without base flow separation are better than those in the calibration
significant impacts on the model performances. As we discuss in detail
(Table 2) and validation (Table 3) obtained by using base flow se-
in Section 4.4, the Carson River is a snow-melt dominated basin, and
paration. Especially, the validation results of this river indicated that
the input structure seems to be insufficient to simulate the lagged re-
the base flow separation considerably deteriorated the simulation per-
lationship between snowmelt and runoff. It is clear that there is a need
formance of the RF model. Possible reasons for obtaining a poor si-
for future work to improve the simulation capabilities of machine
mulation result with the RF model for the Carson River with base flow
learning models for snow-dominated basins. The worst performances of
separation could be the value of the model parameter (i.e., mtry = 1),
all models were obtained for this river in both calibration and valida-
input redundancy, and collinearity. When mtry = 1, the RF model evi-
tion periods.
dently could not simulate basin dynamics with a high number of input
By comparing the performances of all models for the Carson River, it
features with complex interrelations. Increases in the majority of classes
is evident that SVR showed the best performance. Notably, the scatter
with a high number of features could lead to bias in the random forest
diagram of observed and simulated values is acceptable only for the
due to the averaging process (Zhu and Pierskalla, 2016). Therefore, a
SVR model. The scatter diagrams of the other models indicate con-
high number of input features could lead to bias in the outputs. Input
sistent biases. For instance, the ANN model consistently underestimated
redundancy could also lead to such a result. Input redundancy could
and overestimated observed values between 0.5 and 3 m3 s−1. Also, the
result from a situation where specific input features provide redundant

Table 3
Performance indices obtained from the simulation models for all rivers for the validation period with base flow separation.
North Fork River Chehalis River Carson River Sacramento River

SVR ANN RF SVR ANN RF SVR ANN RF SVR ANN RF

RMSE 1.04 1.32 1.24 1.39 0.99 0.94 0.34 0.53 1.09 1.50 1.59 1.70
NSE 0.83 0.73 0.76 0.85 0.93 0.93 0.84 0.61 −0.67 0.84 0.83 0.80
KGE 0.88 0.81 0.73 0.89 0.94 0.89 0.77 0.72 −0.43 0.90 0.85 0.83
VE 0.67 0.57 0.64 0.65 0.76 0.80 0.64 0.46 −0.46 0.73 0.70 0.64
d 0.95 0.93 0.94 0.96 0.98 0.98 0.96 0.87 0.68 0.96 0.95 0.94
PI 0.60 0.35 0.43 −0.15 0.42 0.48 −2.81 −8.14 −38.25 0.51 0.45 0.37

Bold numbers indicate the best value of the considered performance indices for the rivers.

272
H. Tongal, M.J. Booij Journal of Hydrology 564 (2018) 266–282

50
North Fork River
Observation
40 Support vector regression
Artificial neural networks
Random forest

Discharge (m3s )
-1
30

20

10

0
0 100 200 300 400 500 600 700
Days

Fig. 1. Simulated and observed hydrographs and scatter diagrams of the North Fork River for the validation period.

information (Wei, 2016). This phenomenon could lead to more complex 4.3. Comparison of simulation and forecasting results
rules in a random forest and consequently over-fitting. An over-fitted
random forest might be susceptible to error propagation in streamflow In this section, we compare the simulation and forecasting perfor-
simulation since each leaf of a decision tree in the random forest is mances of the models. To get a valid evaluation of the difference be-
dependent on simple decision rules and seemingly small perturbations tween the model structures for simulation and forecasting, the same
in decisions lead to high errors in the final value (i.e. the simulated model parameters for the simulation are used in the forecasting stage.
value). However, it is clear that developing the RF model with less input To this end, the used input structure of Eq. (3) was changed to the
features (i.e. without base flow separation) with mtry = 1 improved the structure of Eq. (5). The forecasting performance of the models can be
model simulation performance significantly. Except for the issues found in Tables 6 and 7 for the calibration and validation period, re-
mentioned above, we can infer that base flow separation improved the spectively. In the calibration period, the RF (North Fork and Sacra-
simulation performances of all models. This can be better seen in the mento rivers) and ANN (Chehalis and Carson rivers) models performed
performances of the selected best models according to the validation better than the SVR models. Similarly, in the validation period, the RF
results with and without base flow separation (Fig. 5). (Chehalis and Sacramento rivers) and ANN (North Fork and Carson
It is evident from Fig. 5 that the simulation with base flow separa- rivers) models outperformed the SVR model. However, it should be
tion improved the performances of the selected best models except for noted that forecasting performances of the SVR model are still rea-
the RF model for the Chehalis River with only a small decrease in the sonable and are close to those of RF and ANN.
KGE index from 0.92 to 0.89. Overall, the highest increase was obtained Using one-day antecedent observed base flow and surface flow in
for the SVR model for the Carson River and the improvement for all the forecasting, rather than the simulated values as in the simulation
rivers can be seen especially from the normalized RMSE and PI values. mode, significantly increased the performances of the models in the
Except for the Chehalis River, we obtained large decreases in the nor- calibration and validation periods. The highest increase in performance
malized RMSEs and significant increases in PI values. Fig. 5 shows that from simulation to forecasting was observed for the RF model. For in-
the smallest performance gain was obtained for the Chehalis River stance, in the validation period, it was the best forecasting model for
where performances of all models are quite good for the simulation, the rivers of Chehalis and Sacramento (Table 7). It is evident that in-
possibly due to the less complex catchment dynamics (see Section 4.4). troducing observed correlated variables that are observed base flow and
surface flow at time step t −1 into the model structure increases the RF

273
H. Tongal, M.J. Booij Journal of Hydrology 564 (2018) 266–282

Fig. 2. Simulated and observed hydrographs and scatter diagrams of the Chehalis River for the validation period.

model performance. The main reason for this could be that correlated irrelevant input variables, because a variable selection is an integrated
variables increased the performance of capturing dynamics by splitting part of RF in the tree growing process. However, since the highest
the feature space of each tree in RF. Further, in the simulation phase, performance gain is obtained for the RF model by only changing the
the model output is based on the previous output whereas the observed input structure (i.e. from Eqs. (3)–(5)) it can be concluded that the RF
values are integrated into the forecasting phase. The simulation results model is the most sensitive model to the input structure. Our simulation
showed that the RF model without base flow separation is better than and forecasting results point out that the RF model is sensitive to the
the simulation with base flow separation. It seems that error accumu- input structure of the model, and attention should be given to the se-
lation is higher for RF than the other models when the previous output lection of input features.
is utilized in the simulation phase. It yielded better results when highly The highest increases in model performance from simulation to
correlated observed values were included in the model structure. As forecasting schemes were obtained for the Carson River. In the fol-
indicated by Granata and de Marinis (2017), the error rate decreases if lowing section, we provide a detailed discussion about the catchment
the strength of the individual trees in the RF increases. The majority of dynamics of the Carson River and the obtained model performances for
the classes increases the bias of the regular random forests method this river.
because the method is constructed to obtain an overall minimum error
rate with equal considerations for all classes (Zhu and Pierskalla, 2016).
Including auto-correlated one-day antecedent observed base flow and 4.4. Catchment dynamics of the Carson River
surface flow in the forecasting, rather than the simulated values as in
the simulation mode, could have possibly increased the strengths of From Fig. 6, it can be seen that the Carson River is a snowmelt
individual trees. Weighted random forests that give different weights to dominated basin and most of the annual runoff comes from snowmelt
different classes (i.e. different dynamics in hydro-climatic variables) runoff. The basin is influenced by a significant snowmelt season in
can be a potential alternative to the regular random forest in the si- May–June and streamflow is less dependent on rainfall and firmly
mulation of streamflows which is a beyond the scope of this study. Xu connected with air temperature variations (Hay et al., 2000). The
and Valocchi (2015) indicated that all potential inputs could be in- maximum Pearson correlation calculated for 365 days between tem-
tegrated to a random forest since it is insensitive to the presence of perature and streamflow (r = 0.50) is higher than the one between
rainfall and streamflow (r = 0.16) and potential evapotranspiration and

274
H. Tongal, M.J. Booij Journal of Hydrology 564 (2018) 266–282

Fig. 3. Simulated and observed hydrographs and scatter diagrams of the Carson River for the validation period.

streamflow (r = 0.37) (not shown). The linear relationship between forecasting scheme. The observed flow at time step t −1 can be con-
rainfall and streamflow is insignificant for the Carson River (r = 0.16) sidered as a representative of the memory or water storage in the
while there is a linear relationship for other rivers with a correlation of system and is the integral response of the system for a specific combi-
0.47, 0.57, and 0.61 for the North Fork, Chehalis and Sacramento nation of storages (Serinaldi et al., 2014). Since the employed machine
rivers, respectively. This weak relationship could lead to the poor si- learning models do not have an individual memory component related
mulation results for this river. to dynamics, including the one-antecedent discharge value can act as a
As indicated by Siou et al. (2011), the correlation between rainfall memory of the system that is a representative of the specific response of
and discharge for different time lags provides information about the the catchment.
basin response to rainfall. A correlation drop below 0.2 can be con-
sidered as a “memory effect” in hydrology. The memory effect indicates
the possibility of predicting conditions in the future based on the his- 5. Future directions of research
torical records. The only adverse memory effect (i.e. correlation below
zero) is consistently observed for the Carson River after lag one (not To improve the simulation capabilities of the machine learning
shown). In other words, there is a negative relationship between rainfall models, in future work, two techniques can be integrated; (1) pre-
and discharge one or more days ahead. This finding indicates that there processing methods such as singular spectrum analysis (Lisi et al., 1995;
is indeed a delay between temperatures above zero and snowmelt Marques et al., 2006), and (2) integrating a multi-model ensemble
runoff and, future discharges are negatively related with the previous technique (Araghinejad et al., 2011; Duan et al., 2007). In the former
rainfall. Thus, it can be concluded that coupled in this mountainous approach, streamflow time series can be decomposed into low and high-
region, due to snowmelt runoff and a negative memory effect, the frequency components that represent a specific level of the temporal
runoff dynamics of the Carson River are difficult to simulate by the characteristics of streamflow. Also, as reported by Danandeh Mehr et al.
employed machine learning models that lack a memory (or storage) (2013), the wavelet decomposition method as a pre-processing tech-
component. However, by including the one-antecedent observed base nique can successfully overcome non-stationarity in streamflow, and
′ ′ the obtained decomposed parts of the original series represent a specific
and surface flows (i.e., Qtg− 1 and Qts− 1) to the input structure, we obtained
time feature of the original series which could be helpful for simulation.
significant increases in the performances of the models in the
A multi-model ensemble technique can also be useful for improving

275
H. Tongal, M.J. Booij Journal of Hydrology 564 (2018) 266–282

Fig. 4. Simulated and observed hydrographs and scatter diagrams of the Sacramento River for the validation period.

performances of the machine learning models. In this approach, addi- analysis is beyond the scope of this study and since the number of
tional simulation models that belong to the same class of models or parameters and complexity of model structures do not have a direct
other models that are successful in capturing different aspects of relationship with the goodness of fit and the uncertainty range we
streamflow dynamics can be used jointly. Before applying a multi- cannot reach a conclusion about the predictive uncertainty of the
model ensemble technique, some robust feature extraction methods in models. However, the number of parameters could have a more sig-
addition to clustering analysis can be applied to identify different hy- nificant contribution than the model structure uncertainty (Tian et al.,
drological processes (Nourani, 2017). However, it should be noted that 2014). In this study, ANN is the most complex model structure and has
additional uncertainty sources could be introduced by using different the highest number of parameters, i.e. weights and biases. Further,
models in the multi-model ensemble technique in addition to un- Tongal and Booij (2017) showed that parametric uncertainty could be
certainty sources of input and model parameters. The uncertainty closely related to the dynamics of streamflow. A possible higher

Table 4
Performance indices obtained from the simulation models for all rivers for the calibration period without base flow separation.
North Fork River Chehalis River Carson River Sacramento River

SVR ANN RF SVR ANN RF SVR ANN RF SVR ANN RF

RMSE 2.21 2.05 1.98 1.71 1.50 1.74 1.01 1.18 1.01 3.29 2.60 2.59
NSE 0.62 0.67 0.70 0.91 0.93 0.90 0.56 0.40 0.56 0.59 0.75 0.75
KGE 0.80 0.84 0.84 0.90 0.95 0.92 0.71 0.33 0.41 0.67 0.75 0.78
VE 0.56 0.54 0.62 0.75 0.81 0.81 0.55 0.39 0.50 0.56 0.58 0.64
d 0.90 0.91 0.92 0.98 0.98 0.98 0.91 0.71 0.79 0.86 0.92 0.92
PI −0.38 −0.20 −0.11 0.22 0.39 0.19 −13.46 −18.78 −13.59 −0.99 −0.24 −0.23

Bold numbers indicate the best value of the considered performance indices for the rivers.

276
H. Tongal, M.J. Booij Journal of Hydrology 564 (2018) 266–282

Table 5
Performance indices obtained from the simulation models for all rivers for the validation period without base flow separation.
North Fork River Chehalis River Carson River Sacramento River

SVR ANN RF SVR ANN RF SVR ANN RF SVR ANN RF

RMSE 1.71 2.06 1.97 1.47 1.08 1.08 0.78 0.58 0.50 2.27 2.12 1.85
NSE 0.55 0.34 0.40 0.83 0.91 0.91 0.14 0.54 0.65 0.64 0.69 0.76
KGE 0.71 0.51 0.41 0.89 0.93 0.92 0.21 0.61 0.55 0.78 0.64 0.81
VE 0.34 0.24 0.28 0.63 0.76 0.79 0.25 0.34 0.40 0.55 0.45 0.65
d 0.87 0.84 0.86 0.96 0.98 0.98 0.85 0.84 0.87 0.89 0.90 0.93
PI −0.08 −0.58 −0.44 −0.30 0.30 0.30 −19.21 −9.91 −7.34 −0.13 0.01 0.25

Bold numbers indicate the best value of the considered performance indices for the rivers.

(a) North Fork River with SVR (b) Chehalis River with RF

(c) Carson River with SVR* (d) Sacramento River with SVR

Fig. 5. Comparison of performance indices with and without base flow separation of the best simulation models during the validation period with base flow
separation. *The PI index was scaled by dividing by the absolute smallest PI value (i.e., 19.21). **NRMSE indicates the normalized RMSE with the observed discharge
of the considered river.

Table 6
Performance indices obtained from the forecast models for all rivers for the calibration period with base flow separation.
North Fork River Chehalis River Carson River Sacramento River

SVR ANN RF SVR ANN RF SVR ANN RF SVR ANN RF

RMSE 1.19 1.19 1.19 1.19 0.86 0.91 0.26 0.23 0.38 2.65 1.66 1.57
NSE 0.89 0.89 0.89 0.95 0.98 0.97 0.97 0.98 0.94 0.74 0.90 0.91
KGE 0.93 0.93 0.94 0.96 0.97 0.97 0.98 0.98 0.81 0.73 0.90 0.85
VE 0.85 0.86 0.88 0.86 0.89 0.91 0.89 0.91 0.83 0.80 0.85 0.87
d 0.97 0.97 0.97 0.99 0.99 0.99 0.99 0.99 0.98 0.91 0.97 0.97
PI 0.60 0.60 0.60 0.62 0.80 0.78 0.04 0.23 −1.10 −0.29 0.50 0.55

Bold numbers indicate the best value of the considered performance indices for the rivers.

277
H. Tongal, M.J. Booij Journal of Hydrology 564 (2018) 266–282

Table 7
Performance indices obtained from the forecast models for all rivers for the validation period with base flow separation.
North Fork River Chehalis River Carson River Sacramento River

SVR ANN RF SVR ANN RF SVR ANN RF SVR ANN RF

RMSE 0.86 0.79 0.89 0.91 0.65 0.51 0.18 0.16 0.23 1.33 1.25 1.18
NSE 0.89 0.90 0.88 0.94 0.97 0.98 0.95 0.96 0.93 0.88 0.89 0.90
KGE 0.88 0.94 0.92 0.93 0.97 0.98 0.94 0.98 0.79 0.90 0.88 0.90
VE 0.79 0.82 0.85 0.82 0.86 0.92 0.82 0.88 0.73 0.83 0.86 0.88
d 0.97 0.97 0.97 0.98 0.99 0.99 0.99 0.99 0.98 0.97 0.97 0.97
PI 0.72 0.77 0.70 0.51 0.75 0.84 −0.07 0.14 −0.70 0.61 0.66 0.69

Bold numbers indicate the best value of the considered performance indices for the rivers.

prediction uncertainty due to parametric uncertainty can be obtained Also, they used different model structures for the simulation and fore-
for streamflow that has a high coefficient of variation showing complex casting which makes the comparison questionable. However, it should
hydrograph features with fluctuations in rising and falling limbs (i.e. be noted that there is still some room for improvement of our fore-
North Fork River with ANN model), and vice versa, which needs a casting models, since these have the same model structure (i.e. the same
thorough analysis. The outcomes of this study are case-sensitive and number of hidden neurons) as the simulation models. Thus, there is a
including predictive uncertainty analysis may change or support the need for further investigation of the forecasting capabilities of machine
reported results herein. In a future study, we also aim to investigate the learning models for various catchments that have different hydro-cli-
prediction uncertainty resulting from the input, model structure and matic features.
parameters for different lead times by employing the generalized like- The results obtained by Taormina et al. (2015) indicated that the
lihood uncertainty estimation method (see Tian et al. (2014); Tongal base flow separation method did not improve the forecasting cap-
and Booij (2017)). abilities of artificial neural networks and extreme learning machines for
In conclusion, the forecasting capabilities of the developed models nine different gauging stations in the United States. They employed two
are more accurate and reliable than the simulation capabilities. different approaches for forecasting purposes that use base flow se-
However, it should be noted that Bozorg-Haddad et al. (2016) found an paration where (1) the base flow separation and model parameters are
opposite result where simulation and forecasting ANN models of optimized through global optimization and (2) the recession parameter
monthly runoff were developed by considering the combination of is not estimated during the optimization process similarly to our study.
temperature, evapotranspiration, precipitation and discharge and their They reported that the second approach yield similar and in some cases
antecedent values on a monthly scale. They found better simulation better results than the former one. Further, optimization of the reces-
results than forecasting results. The model structure, utilization of dif- sion parameter gave substantially smaller base flow index parameters
ferent optimization algorithms, different time scales and consideration than the recommended values for most of the examined watersheds. As
of different catchments could lead to such a difference. For instance, in the authors mentioned, this result indicated that setting the filter
their study, the best model structures consist of two hidden layers, and parameters according to expert knowledge could result in accurate base
the authors indicated that one hidden layer is insufficient for simulation flow separation. In this study, we employed an exhaustive-grid search
and forecasting in their case study in contrast to our finding. Their best optimization that is quite time-consuming to tune the structures and
simulation and forecasting models with two hidden layers include re- parameters of the machine learning models. However, recently, popu-
spectively 18–10 and 10–24 neurons which is a rather high number. lation-based optimization algorithms have been successfully applied to

Fig. 6. Annual hydrographs of the rivers for the period 1948–2003.

278
H. Tongal, M.J. Booij Journal of Hydrology 564 (2018) 266–282

increase the forecasting capabilities of data-driven models (Chau, 2007; machine learning methods. As recommended by Benninga et al.
Chen et al., 2015; Wu and Chau, 2006). Further, as indicated by Chau (2017), the reliability of streamflow forecasts can be enlarged by
(2007), combining the advantages of a global optimization method including uncertainties in the hydrological model parameters and
such as particle swarm optimization and local fast convergence of Le- the initial conditions, and by enlarging the dispersion of the me-
venberg-Marquardt algorithm could increase the forecasting perfor- teorological input forecasts. The reported results herein are case-
mance of a data-driven model in a much shorter time. Investigating this sensitive and can change by including the uncertainty analysis.
paradigm could lessen the enormous computational costs and time for 2. We found some improvements in model performances in the vali-
the simulation scheme with the base flow separation method in this dation period compared to the calibration period. In addition to the
study. To the best of our knowledge, there is not any study that applied other discussed issues, this could also result from the relatively small
population-based optimization algorithms in the simulation of stream- length of the validation period in comparison to the calibration
flows with data-driven models. Therefore, it would be valuable to in- period. Therefore, additional study is required to estimate the effects
vestigate the applicability of population-based optimization methods in of data length of calibration and validation periods on the results.
simulation of streamflows coupled with base flow separation in a future 3. The base flow recession parameter was assumed to be constant to
study. avoid additional computational costs. In addition to employing
other types of base flow separation methods, the cost-effective po-
6. Conclusion pulation-based optimization algorithms could also be integrated into
the proposed methodology.
In the context of water resources planning and management, effi- 4. Due to the limited number of hydro-meteorological variables, we
cient utilization of water resources requires accurate and successful employed the Hargreaves method to estimate potential evapo-
simulation and forecasting of streamflow. In this study, a simulation transpiration. More physically-based radiation-based methods can
framework has been developed to explore efficiencies of different ma- be employed to check additional performance improvement.
chine learning methods (i.e. support vector machines, artificial neural 5. In this study, we only simulated and forecasted one-day ahead
networks and random forest) in streamflow simulation for four rivers in streamflow values. Investigation of simulation and forecasting cap-
the United States. In this simulation framework, we analyzed the per- abilities of machine learning models for additional lead times and
formances of SVR, ANN, and RF coupled with a one-parameter base comparison with conceptual models would be interesting.
flow separation method proposed by Eckhardt (2005). To the authors’ 6. To improve the simulation capabilities of machine learning
knowledge, so far these models have not been tested and compared for methods, pre-processing methods such as singular spectrum and
simulation of streamflow by coupling with a base flow separation wavelet analyses could be integrated into the proposed metho-
method. The simulation with base flow separation mainly indicated dology.
that: 7. Finally, for better simulation results especially in snow-dominated
basins, one can integrate a multi-model ensemble modelling tech-
1. The RF model performed worse than SVR and ANN in the calibration nique into the proposed approach by using additional simulation
period. The ANN model performed best in three basins (i.e. North models that are successful in capturing different aspects of stream-
Fork, Chehalis, Sacramento rivers) while the SVR model was best in flow dynamics.
the Carson River.
2. We found some improvements in model performances in the vali- References
dation period compared to the calibration period. In the validation
period, SVR and RF performed better than the ANN model due to Adamowski, J., Sun, K., 2010. Development of a coupled wavelet transform and neural
poor generalization capabilities of the ANN model. network method for flow forecasting of non-perennial rivers in semi-arid watersheds.
J. Hydrol. 390 (1–2), 85–91.
3. All models yield the worst simulation results for the Carson River, Allen, R.G., Pereira, L.S., Raes, D., Smith, M., 1998. Crop evapotranspiration-Guidelines
particularly the RF model. This could be associated with the com- for computing crop water requirements-FAO Irrigation and drainage paper 56 FAO,
plex streamflow dynamics resulting from the lag between tem- Rome, vol. 300, 9, D05109.
Aqil, M., Kita, I., Yano, A., Nishiyama, S., 2007a. A comparative study of artificial neural
perature and snowmelt runoff. The SVR gave acceptable results for networks and neuro-fuzzy in continuous modeling of the daily and hourly behaviour
this river. of runoff. J. Hydrol. 337 (1–2), 22–34.
Aqil, M., Kita, I., Yano, A., Nishiyama, S., 2007b. Neural networks for real time catchment
flow modeling and prediction. Water Resour. Manage. 21 (10), 1781–1796.
We also analyzed whether the base flow separation method im- Araghinejad, S., Azmi, M., Kholghi, M., 2011. Application of artificial neural network
proved the simulation performances of the machine learning models. ensembles in probabilistic hydrological forecasting. J. Hydrol. 407 (1–4), 94–104.
This was achieved by employing the simulation framework by only Arnold, J., Allen, P., Muttiah, R., Bernhardt, G., 1995. Automated base flow separation
and recession analysis techniques. Groundwater 33 (6), 1010–1018.
using the simulated discharge — it was not separated into base flow and
Badrzadeh, H., Sarukkalige, R., Jayawardena, A., 2016. Improving ann-based short-term
direct flow. We concluded that the base flow separation improved the and long-term seasonal river flow forecasting with signal processing techniques.
simulation capabilities of the machine learning methods except for the River Res. Appl. 32 (3), 245–256.
RF model for the Carson River. Hence, the use of hydrological knowl- Badrzadeh, H., Sarukkalige, R., Jayawardena, A.W., 2013. Impact of multi-resolution
analysis of artificial intelligence models inputs on multi-step ahead river flow fore-
edge such as base flow separation improved the simulation results. casting. J. Hydrol. 507, 75–85.
Finally, to differentiate the simulation and forecasting capabilities, we Barzegar, R., Moghaddam, A.A., Adamowski, J., Fijani, E., 2017. Comparison of machine
employed a forecasting scheme. In this scheme, we only changed the learning models for predicting fluoride contamination in groundwater. Stoch.
Environ. Res. Risk Assess. 31 (10), 2705–2718.
input structure of the models by keeping the model parameters the Baudron, P., Alonso-Sarría, F., García-Aróstegui, J.L., Cánovas-García, F., Martínez-
same as in the simulation scheme. It was found that the highest increase Vicente, D., Moreno-Brotóns, J., 2013. Identifying the origin of groundwater samples
in performance was obtained for the RF model. Also, the highest in- in a multi-layer aquifer system with Random Forest classification. J. Hydrol. 499,
303–315.
creases in the model performances were observed for the Carson River. Belayneh, A., Adamowski, J., Khalil, B., Ozga-Zielinski, B., 2014. Long-term SPI drought
Some limitations and future research directions related to the pro- forecasting in the Awash River Basin in Ethiopia using wavelet neural network and
posed methodology are: wavelet support vector regression models. J. Hydrol. 508, 418–429.
Benninga, H.-J.F., Booij, M.J., Romanowicz, R.J., Rientjes, T.H., 2017. Performance of
ensemble streamflow forecasts under varied hydrometeorological conditions. Hydrol.
1. Inevitably, there are several uncertainty sources such as input, Earth Syst. Sci. 21 (10), 5273–5291.
model and parameter uncertainty that contribute to uncertainty in Bhattacharya, B., Solomatine, D.P., 2005. Neural networks and M5 model trees in mod-
elling water level–discharge relationship. Neurocomputing 63, 381–396.
simulated values. Thus, in future work, there is a need for quanti-
Booker, D., Woods, R., 2014. Comparing and combining physically-based and
fying the uncertainty in simulated values obtained from the different

279
H. Tongal, M.J. Booij Journal of Hydrology 564 (2018) 266–282

empirically-based approaches for estimating the hydrology of ungauged catchments. Exp. Syst. Appl. 41 (11), 5267–5276.
J. Hydrol. 508, 227–239. Granata, F., de Marinis, G., 2017. Machine learning methods for wastewater hydraulics.
Bozorg-Haddad, O., Zarezadeh-Mehrizi, M., Abdi-Dehkordi, M., Loáiciga, H.A., Mariño, Flow Meas. Instrum. 57, 1–9.
M.A., 2016. A self-tuning ANN model for simulation and forecasting of surface flows. Grantz, K., Rajagopalan, B., Zagona, E., Clark, M., 2007. Water management applications
Water Resour. Manage. 30 (9), 2907–2929. of climate-based hydrologic forecasts: case study of the Truckee-Carson River Basin.
Breiman, L., 2003. Manual–setting up, using, and understanding random forests V4. 0. J. Water Resour. Plann. Manage. 133 (4), 339–350.
< http://oz.berkeley.edu/users/breiman.Using_random_forests_v4.0.pdf > . Gupta, H.V., Kling, H., Yilmaz, K.K., Martinez, G.F., 2009. Decomposition of the mean
Breiman, L., 2001. Random forests. Mach. Learn. 45 (1), 5–32. squared error and NSE performance criteria: Implications for improving hydrological
Chang, J., Wang, G., Mao, T., 2015. Simulation and prediction of suprapermafrost modelling. J. Hydrol. 377 (1), 80–91.
groundwater level variation in response to climate change using a neural network Hagan, M.T., Menhaj, M.B., 1994. Training feedforward networks with the Marquardt
model. J. Hydrol. 529, 1211–1220. algorithm. IEEE Trans. Neural Networks 5 (6), 989–993.
Chapman, T.G., 1991. Comment on “Evaluation of automated techniques for base flow Halgamuge, M.N., Nirmalathas, A., 2017. Analysis of large flood events: based on flood
and recession analyses” by RJ Nathan and TA McMahon. Water Resour. Res. 27 (7), data during 1985–2016 in Australia and India. Int. J. Disaster Risk Reduct. 24, 1–11.
1783–1784. Han, D., Kwong, T., Li, S., 2007. Uncertainties in real-time flood forecasting with neural
Chau, K., 2007. A split-step particle swarm optimization algorithm in river stage fore- networks. Hydrol. Process. 21 (2), 223–228.
casting. J. Hydrol. 346 (3–4), 131–135. Hargreaves, G.H., Samani, Z.A., 1985. Reference crop evapotranspiration from tem-
Chen, L., Singh, V.P., Guo, S., Zhou, J., Ye, L., 2014. Copula entropy coupled with arti- perature. Appl. Eng. Agric. 1 (2), 96–99.
ficial neural network for rainfall-runoff simulation. Stoch. Env. Res. Risk Assess. 28 Hay, L.E., Clark, M., 2003. Use of statistically and dynamically downscaled atmospheric
(7), 1755–1767. model output for hydrologic simulations in three mountainous basins in the western
Chen, X., Chau, K., Busari, A., 2015. A comparative study of population-based optimi- United States. J. Hydrol. 282 (1–4), 56–75.
zation algorithms for downstream river flow forecasting by a hybrid neural network Hay, L.E., Wilby, R.L., Leavesley, G.H., 2000. A comparison of delta change and down-
model. Eng. Appl. Artif. Intell. 46, 258–268. scaled GCM scenarios for three mountainous basins in the United States. JAWRA J.
Chen, Y.D., Zhang, Q., Xiao, M., Singh, V.P., Zhang, S., 2016. Probabilistic forecasting of Am. Water Resour. Assoc. 36 (2), 387–397.
seasonal droughts in the Pearl River basin, China. Stoch. Environ. Res. Risk Assess. 30 Haykin, S., 1999. Neural Networks-A Comprehensive Foundation. Prentice Hall Inc, New
(7), 2031–2040. Jersey.
Chiu, M.-C., Hunt, L., Resh, V.H., 2017. Climate-change influences on the response of Henning, J.A., Gresswell, R.E., Fleming, I.A., 2007. Use of seasonal freshwater wetlands
macroinvertebrate communities to pesticide contamination in the Sacramento River, by fishes in a temperate river floodplain. J. Fish Biol. 71 (2), 476–492.
California watershed. Sci. Total Environ. 581, 741–749. Hong, W.C., Pai, P.F., 2007. Potential assessment of the support vector regression tech-
Choy, K., Chan, C.W., 2003. Modelling of river discharges and rainfall using radial basis nique in rainfall forecasting. Water Resour. Manage. 21 (2), 495–513.
function networks based on support vector regression. Int. J. Syst. Sci. 34 (14–15), Hosseini, S.M., Mahjouri, N., 2016. Integrating Support Vector Regression and a geo-
763–773. morphologic Artificial Neural Network for daily rainfall-runoff modeling. Appl. Soft
Corzo, G., Solomatine, D., 2007. Baseflow separation techniques for modular artificial Comput. 38, 329–345.
neural network modelling in flow forecasting. Hydrol. Sci. J. 52 (3), 491–507. Hu, T.S., Lam, K.C., Ng, S.T., 2001. River flow time series prediction with a range-de-
Criss, R.E., Winston, W.E., 2008. Do Nash values have value? discussion and alternate pendent neural network. Hydrol. Sci. J. 46 (5), 729–745.
proposals. Hydrol. Process. 22 (14), 2723–2725. Humphrey, G.B., Gibbs, M.S., Dandy, G.C., Maier, H.R., 2016. A hybrid approach to
Danandeh Mehr, A., Kahya, E., Olyaie, E., 2013. Streamflow prediction using linear ge- monthly streamflow forecasting: Integrating hydrological model outputs into a
netic programming in comparison with a neuro-wavelet technique. J. Hydrol. 505, Bayesian artificial neural network. J. Hydrol. 540, 623–640.
240–249. Jeton, A.E., Dettinger, M.D., Smith, J.L., 1996. Potential effects of climate change on
Das, T., Bárdossy, A., Zehe, E., He, Y., 2008. Comparison of conceptual model perfor- streamflow, eastern and western slopes of the Sierra Nevada, California and Nevada.
mance using different representations of spatial variability. J. Hydrol. 356 (1–2), US Department of the Interior, US Geological Survey.
106–118. Ju, Q., Yu, Z., Hao, Z., Ou, G., Zhao, J., Liu, D., 2009. Division-based rainfall-runoff
Demirel, M.C., Booij, M.J., Hoekstra, A.Y., 2013. Impacts of climate change on the sea- simulations with BP neural networks and Xinanjiang model. Neurocomputing 72
sonality of low flows in 134 catchments in the River Rhine basin using an ensemble of (13–15), 2873–2883.
bias-corrected regional climate simulations. Hydrol. Earth Syst. Sci. 17 (10), Kalra, A., Ahmad, S., Nayak, A., 2013. Increasing streamflow forecast lead time for
4241–4257. snowmelt-driven catchment based on large-scale climate patterns. Adv. Water
Dettinger, M.D., Cayan, D.R., Meyer, M.K., Jeton, A.E., 2004. Simulated hydrologic re- Resour. 53, 150–162.
sponses to climate variations and change in the Merced, Carson, and American River Kang, K., Lee, J.H., 2014. Hydrologic modelling of the effect of snowmelt and tempera-
basins, Sierra Nevada, California, 1900–2099. Clim. Change 62 (1–3), 283–317. ture on a mountainous watershed. J. Earth Syst. Sci. 123 (4), 705–713.
Dibike, Y.B., Velickov, S., Solomatine, D., Abbott, M.B., 2001. Model induction with Karunasinghe, D.S., Liong, S.-Y., 2006. Chaotic time series prediction with a global model:
support vector machines: introduction and applications. J. Comput. Civil Eng. 15 (3), artificial neural network. J. Hydrol. 323 (1), 92–105.
208–216. Kim, J.-W., Pachepsky, Y.A., 2010. Reconstructing missing daily precipitation data using
Doycheva, K., Horn, G., Koch, C., Schumann, A., König, M., 2017. Assessment and regression trees and artificial neural networks for SWAT streamflow simulation. J.
weighting of meteorological ensemble forecast members based on supervised ma- Hydrol. 394 (3), 305–314.
chine learning with application to runoff simulations and flood warning. Adv. Eng. Kimbrough, R.A., Ruppert, G., Wiggins, W., Smith, R., Kresch, D., 2006. Water Resources
Inf. 33, 427–439. Data-Washington Water Year 2005. U.S, Geological Survey.
Du, K., Zhao, Y., Lei, J., 2017. The incorrect usage of singular spectral analysis and dis- Kitanidis, P.K., Bras, R.L., 1980. Real-time forecasting with a conceptual hydrologic
crete wavelet transform in hybrid models to predict hydrological time series. J. model 2. Applications and results. Water Resour. Res. 16, 1034–1044.
Hydrol. 552, 44–51. Klemeš, V., 1986. Operational testing of hydrological simulation models. Hydrol. Sci. J.
Duan, Q., Ajami, N.K., Gao, X., Sorooshian, S., 2007. Multi-model ensemble hydrologic 31 (1), 13–24.
prediction using Bayesian model averaging. Adv. Water Resour. 30 (5), 1371–1386. Kumar, D., Pandey, A., Sharma, N., Flügel, W.-A., 2016. Daily suspended sediment si-
Ebrahimi, H., Rajaee, T., 2017. Simulation of groundwater level variations using wavelet mulation using machine learning approach. Catena 138, 77–90.
combined with neural network, linear regression and support vector machine. Global Lansigan, F., 2009. Frequency analysis of extreme hydrologic events and assessment of
Planet. Change 148, 181–191. water stress in a changing climate in the Philippines. In: Taniguchi, M., Burnett, W.C.,
Eckhardt, K., 2005. How to construct recursive digital filters for baseflow separation. Fukushima, Y., Haigh, M., Umezawa, Y., (Eds.), pp. 497–501.
Hydrol. Process. 19 (2), 507–515. Leasure, D., Magoulick, D.D., Longing, S., 2016. Natural flow regimes of the Ozark-
Eckhardt, K., 2008. A comparison of baseflow indices, which were calculated with seven Ouachita interior highlands region. River Res. Appl. 32 (1), 18–35.
different baseflow separation methods. J. Hydrol. 352 (1), 168–173. Lee, K.T., Hung, W.-C., Meng, C.-C., 2008. Deterministic insight into ANN model per-
Esmaeilzadeh, B., Sattari, M.T., Samadianfard, S., 2017. Performance evaluation of ANNs formance for storm runoff simulation. Water Resour. Manage. 22 (1), 67–82.
and an M5 model tree in Sattarkhan Reservoir inflow prediction. ISH J. Hydraul. Eng. Legates, D.R., McCabe, G.J., 1999. Evaluating the use of “goodness-of-fit” Measures in
23 (3), 283–292. hydrologic and hydroclimatic model validation. Water Resour. Res. 35 (1), 233–241.
Feng, Q., Wen, X., Li, J., 2015. Wavelet analysis-support vector machine coupled models Liaw, A., Wiener, M., 2002. Classification and regression by random. Forest R news 2 (3),
for monthly rainfall forecasting in arid regions. Water Resour. Manage. 29 (4), 18–22.
1049–1065. Lin, G.-F., Chen, G.-R., Wu, M.-C., Chou, Y.-C., 2009. Effective forecasting of hourly ty-
Fotovatikhah, F., Herrera, M., Shamshirband, S., Chau, K.-W., Faizollahzadeh Ardabili, S., phoon rainfall using support vector machines. Water Resour. Res. 45 (8).
Piran, M.J., 2018. Survey of computational intelligence as basis to big flood man- Lin, G.-F., Chen, L.-H., 2004. A non-linear rainfall-runoff model using radial basis function
agement: challenges, research directions and future work. Eng. Appl. Comput. Fluid network. J. Hydrol. 289 (1–4), 1–8.
Mech. 12 (1), 411–437. Lin, G.-F., Jhong, B.-C., 2015. A real-time forecasting model for the spatial distribution of
Garbrecht, J.D., 2006. Comparison of three alternative ANN designs for monthly rainfall- typhoon rainfall. J. Hydrol. 521, 302–313.
runoff simulation. J. Hydrol. Eng. 11 (5), 502–505. Lin, J.-Y., Cheng, C.-T., Chau, K.-W., 2006. Using support vector machines for long-term
Gholami, V., Chau, K.W., Fadaee, F., Torkaman, J., Ghaffari, A., 2015. Modeling of discharge prediction. Hydrol. Sci. J. 51 (4), 599–612.
groundwater level fluctuations using dendrochronology in alluvial aquifers. J. Liong, S.Y., Sivapragasam, C., 2002. Flood stage forecasting with support vector ma-
Hydrol. 529, 1060–1069. chines. JAWRA J. Am. Water Resour. Assoc. 38 (1), 173–186.
Gizaw, M.S., Gan, T.Y., 2016. Regional flood frequency analysis using support vector Lisi, F., Nicolis, O., Sandri, M., 1995. Combining singular-spectrum analysis and neural
regression under historical and future climate. J. Hydrol. 538, 387–398. networks for time series forecasting. Neural Process. Lett. 2 (4), 6–10.
Goyal, M.K., Bharti, B., Quilty, J., Adamowski, J., Pandey, A., 2014. Modeling of daily Liu, Z., Zhou, P., Chen, G., Guo, L., 2014. Evaluating a coupled discrete wavelet transform
pan evaporation in sub tropical climates using ANN, LS-SVR, Fuzzy Logic, and ANFIS. and support vector regression for daily and monthly streamflow forecasting. J.

280
H. Tongal, M.J. Booij Journal of Hydrology 564 (2018) 266–282

Hydrol. 519, 2822–2831. classification: empirical analysis of hydrologic similarity based on catchment func-
Lohani, A.K., Goel, N., Bhatia, K., 2011. Comparative study of neural network, fuzzy logic tion in the eastern USA. Hydrol. Earth Syst. Sci. 15 (9), 2895–2911.
and linear transfer function techniques in daily rainfall-runoff modelling under dif- Schnier, S., Cai, X., 2014. Prediction of regional streamflow frequency using model tree
ferent input domains. Hydrol. Process. 25 (2), 175–193. ensembles. J. Hydrol. 517, 298–309.
Lyne, V., Hollick, M., 1979. Stochastic time-variable rainfall-runoff modelling. Inst. Eng. Serinaldi, F., Zunino, L., Rosso, O.A., 2014. Complexity–entropy analysis of daily stream
Australia Nat. Conf. 89–93. flow time series in the continental United States. Stoch. Environ. Res. Risk Assess. 28
Maier, H.R., Dandy, G.C., 1998. The effect of internal parameters and geometry on the (7), 1685–1708.
performance of back-propagation neural networks: an empirical study. Environ. Servat, E., Dezetter, A., 1991. Selection of calibration objective fonctions in the context of
Model. Software 13 (2), 193–209. rainfall-ronoff modelling in a Sudanese savannah area. Hydrol. Sci. J. 36 (4),
Maier, H.R., Dandy, G.C., 2000. Neural networks for the prediction and forecasting of 307–330.
water resources variables: a review of modelling issues and applications. Environ. Shortridge, J.E., Guikema, S.D., Zaitchik, B.F., 2016. Machine learning methods for em-
Model. Software 15 (1), 101–124. pirical streamflow simulation: a comparison of model accuracy, interpretability, and
Maity, R., Bhagwat, P.P., Bhatnagar, A., 2010. Potential of support vector regression for uncertainty in seasonal watersheds. Hydrol. Earth Syst. Sci. 20 (7), 2611.
prediction of monthly streamflow using endogenous property. Hydrol. Process. 24 Shrestha, N., Shukla, S., 2015. Support vector machine based modeling of evapo-
(7), 917–923. transpiration using hydro-climatic variables in a sub-tropical environment. Agric.
Marques, C.A.F., Ferreira, J.A., Rocha, A., Castanheira, J.M., Melo-Gonçalves, P., Vaz, N., For. Meteorol. 200, 172–184.
Dias, J.M., 2006. Singular spectrum analysis and forecasting of hydrological time Singh, K.P., Basant, A., Malik, A., Jain, G., 2009. Artificial neural network modeling of the
series. Phys. Chem. Earth, Parts A/B/C 31 (18), 1172–1179. river water quality—a case study. Ecol. Model. 220 (6), 888–895.
Masselot, P., Dabo-Niang, S., Chebana, F., Ouarda, T.B.M.J., 2016. Streamflow fore- Siou, L.K.A., Johannet, A., Borrell, V., Pistre, S., 2011. Complexity selection of a neural
casting using functional regression. J. Hydrol. 538, 754–766. network model for karst flood forecasting: the case of the Lez Basin (southern
May, R.J., Maier, H.R., Dandy, G.C., Fernando, T.G., 2008. Non-linear variable selection France). J. Hydrol. 403 (3–4), 367–380.
for artificial neural networks using partial mutual information. Environ. Modell. Solomatine, D.P., Xue, Y., 2004. M5 model trees and neural networks: application to flood
Software 23 (10), 1312–1326. forecasting in the upper reach of the Huai River in China. J. Hydrol. Eng. 9 (6),
McMillan, H.K., Hreinsson, E.Ö., Clark, M.P., Singh, S.K., Zammit, C., Uddstrom, M.J., 491–501.
2013. Operational hydrological data assimilation with the recursive ensemble Spinoni, J., Naumann, G., Vogt, J.V., 2017. Pan-European seasonal trends and recent
Kalman filter. Hydrol. Earth Syst. Sci. 17 (1), 21–38. changes of drought frequency and severity. Glob. Planet. Change 148, 113–130.
Mekanik, F., Imteaz, M., Talei, A., 2016. Seasonal rainfall forecasting by adaptive net- Stefánsson, A., Končar, N., Jones, A.J., 1997. A note on the gamma test. Neural Comput.
work-based fuzzy inference system (ANFIS) using large scale climate signals. Clim. Appl. 5 (3), 131–133.
Dynam. 46 (9–10), 3097–3111. Strobl, C., Boulesteix, A.-L., Kneib, T., Augustin, T., Zeileis, A., 2008. Conditional variable
Melesse, A.M., Ahmad, S., McClain, M.E., Wang, X., Lim, Y.H., 2011. Suspended sediment importance for random forests. BMC Bioinform. 9 (1), 307.
load prediction of river systems: an artificial neural network approach. Agric. Water Sudheer, K.P., Gosain, A.K., Ramasastri, K.S., 2002. A data-driven algorithm for con-
Manag. 98 (5), 855–866. structing artificial neural network rainfall-runoff models. Hydrol. Process. 16 (6),
Mohr, C.H., Manga, M., Wang, C.-Y., Korup, O., 2017. Regional changes in streamflow 1325–1330.
after a megathrust earthquake. Earth Planet. Sci. Lett. 458, 418–428. Suryanarayana, C., Sudheer, C., Mahammood, V., Panigrahi, B.K., 2014. An integrated
Moriasi, D.N., Arnold, J.G., Van Liew, M.W., Bingner, R.L., Harmel, R.D., Veith, T.L., wavelet-support vector machine for groundwater level prediction in Visakhapatnam,
2007. Model evaluation guidelines for systematic quantification of accuracy in wa- India. Neurocomputing 145, 324–335.
tershed simulations. Trans. ASABE 50 (3), 885–900. Taormina, R., Chau, K.-W., Sivakumar, B., 2015. Neural network river forecasting
Mukherjee, A., Ramachandran, P., 2018. Prediction of GWL with the help of GRACE TWS through baseflow separation and binary-coded swarm optimization. J. Hydrol. 529,
for unevenly spaced time series data in India: analysis of comparative performances 1788–1797.
of SVR, ANN and LRM. J. Hydrol. 558, 647–658. Thornthwaite, C.W., 1948. An approach toward a rational classification of climate. Geog.
Muñoz, E., Tume, P., Ortíz, G., 2014. Uncertainty in rainfall input data in a conceptual Rev. 38 (1), 55–94.
water balance model: effects on outputs and implications for predictability. Earth Sci. Tian, Y., Booij, M., Xu, Y.-P., 2014. Uncertainty in high and low flows due to model
Res. J. 18 (1), 69–75. structure and parameter errors. Stoch. Environ. Res. Risk Assess. 28 (2), 319–332.
Naghibi, S.A., Ahmadi, K., Daneshi, A., 2017. Application of support vector machine, Tiwari, M.K., Chatterjee, C., 2010a. Development of an accurate and reliable hourly flood
random forest, and genetic algorithm optimized random forest models in ground- forecasting model using wavelet–bootstrap–ANN (WBANN) hybrid approach. J.
water potential mapping. Water Resour. Manage. 31 (9), 2761–2775. Hydrol. 394 (3), 458–470.
Naidu, C., Satyanarayana, G.C., Durgalakshmi, K., Rao, L.M., Mounika, G.J., Raju, A.D., Tiwari, M.K., Chatterjee, C., 2010b. Uncertainty assessment and ensemble flood fore-
2012. Changes in the frequencies of northeast monsoon rainy days in the global casting using bootstrap based artificial neural networks (BANNs). J. Hydrol. 382 (1),
warming. Global Planet. Change 92, 40–47. 20–33.
Nash, J.E., Sutcliffe, J.V., 1970. River flow forecasting through conceptual models, Part I Tongal, H., Booij, M.J., 2017. Quantification of parametric uncertainty of ANN models
– a discussion of principles. J. Hydrol. 10, 282–290. with GLUE method for different streamflow dynamics. Stoch. Env. Res. Risk Assess.
Nikam, V., Gupta, K., 2014. SVM-based model for short-term rainfall forecasts at a local 31 (4), 993–1010.
scale in the Mumbai Urban Area, India. J. Hydrol. Eng. 19 (5), 1048–1052. Tukimat, N.N.A., Harun, S., Shahid, S., 2012. Comparison of different methods in esti-
Noori, N., Kalin, L., 2016. Coupling SWAT and ANN models for enhanced daily stream- mating potential evapotranspiration at Muda Irrigation Scheme of Malaysia. J. Agric.
flow prediction. J. Hydrol. 533, 141–151. Rural Develop. Trop. Subtrop. (JARTS) 113 (1), 77–85.
Nourani, V., 2017. An Emotional ANN (EANN) approach to modeling rainfall-runoff Vapnik, V., Chervonenkis, A., 1964. A note on one class of perceptrons. Autom. Remote
process. J. Hydrol. 544, 267–277. Control 25 (1), 103.
Pai, P.F., Hong, W.C., 2007. A recurrent support vector regression model in rainfall Vapnik, V.N., 1995. The Nature of Statistical Learning Theory. Springer-Verlag, New York
forecasting. Hydrol. Process. 21 (6), 819–827. Inc, pp. 188.
Patel, S., Ramachandran, P., 2015. A comparison of machine learning techniques for Vormoor, K., Lawrence, D., Schlichting, L., Wilson, D., Wong, W.K., 2016. Evidence for
modeling river flow time series: the case of Upper Cauvery River Basin. Water Resour. changes in the magnitude and frequency of observed rainfall vs. snowmelt driven
Manage. 29 (2), 589–602. floods in Norway. J. Hydrol. 538, 33–48.
Raghavendra, N.S., Deka, P.C., 2014. Support vector machine applications in the field of Wang, W.-C., Chau, K.-W., Cheng, C.-T., Qiu, L., 2009. A comparison of performance of
hydrology: a review. Appl. Soft Comput., 19(Supplement C),pp. 372–386. several artificial intelligence methods for forecasting monthly discharge time series.
Randrianasolo, A., Ramos, M.H., Andrêassian, V., 2011. Hydrological ensemble fore- J. Hydrol. 374 (3–4), 294–306.
casting at ungauged basins: using neighbour catchments for model setup and up- Wang, W., Gelder, P.H.A.J.M.V., Vrijling, J.K., Ma, J., 2006. Forecasting daily streamflow
dating. Adv. Geosci. 1 (11). using hybrid ANN models. J. Hydrol. 324 (1–4), 383–399.
Razavi, T., Coulibaly, P., 2013. Classification of Ontario watersheds based on physical Wei, C.-C., 2016. Comparing single-and two-segment statistical models with a conceptual
attributes and streamflow series. J. Hydrol. 493, 81–94. rainfall-runoff model for river streamflow prediction during typhoons. Environ.
Rezaie-balf, M., Naganna, S.R., Ghaemi, A., Deka, P.C., 2017. Wavelet coupled MARS and Modell. Software 85, 112–128.
M5 Model Tree approaches for groundwater level forecasting. J. Hydrol. 553, Willmott, C.J., 1981. On the validation of models. Phys. Geogr. 2, 184–194.
356–373. Worland, S.C., Farmer, W.H., Kiang, J.E., 2018. Improving predictions of hydrological
Rhee, J., Im, J., 2017. Meteorological drought forecasting for ungauged areas based on low-flow indices in ungaged basins using machine learning. Environ. Modell.
machine learning: using long-range climate forecast and remote sensing data. Agric. Software 101, 169–182.
For. Meteorol. 237, 105–122. Wu, C., Chau, K., 2006. A flood forecasting neural network model with genetic algorithm.
Riad, S., Mania, J., Bouchaou, L., Najjar, Y., 2004. Rainfall-runoff model usingan artificial Int. J. Environ. Pollut. 28 (3–4), 261–273.
neural network approach. Math. Comput. Modell. 40 (7–8), 839–846. Wu, C., Chau, K., 2011. Rainfall-runoff modeling using artificial neural network coupled
Salerno, F., Tartari, G., 2009. A coupled approach of surface hydrological modelling and with singular spectrum analysis. J. Hydrol. 399 (3), 394–409.
Wavelet Analysis for understanding the baseflow components of river discharge in Wu, K.-P., Wang, S.-D., 2009. Choosing the kernel parameters for support vector ma-
karst environments. J. Hydrol. 376 (1–2), 295–306. chines by the inter-cluster distance in the feature space. Pattern Recogn. 42 (5),
Sang, Y.F., Wang, Z., Liu, C., 2014. Spatial and temporal variability of precipitation ex- 710–717.
trema in the Haihe River Basin, China. Hydrol. Process. 28 (3), 926–932. Xu, T., Valocchi, A.J., 2015. Data-driven methods to improve baseflow prediction of a
Sapin, J., Rajagopalan, B., Saito, L., Caldwell, R.J., 2017. A K-Nearest neighbor based regional groundwater model. Comput. Geosci. 85, 124–136.
stochastic multisite flow and stream temperature generation technique. Environ. Yang, J., Chang, J., Wang, Y., Li, Y., Hu, H., Chen, Y., Huang, Q., Yao, J., 2018.
Modell. Software 91, 87–94. Comprehensive drought characteristics analysis based on a nonlinear multivariate
Sawicz, K., Wagener, T., Sivapalan, M., Troch, P.A., Carrillo, G., 2011. Catchment drought index. J. Hydrol. 557, 651–667.

281
H. Tongal, M.J. Booij Journal of Hydrology 564 (2018) 266–282

Yoon, H., Jun, S.-C., Hyun, Y., Bae, G.-O., Lee, K.-K., 2011. A comparative study of ar- Hydrol. 552, 92–104.
tificial neural networks and support vector machines for predicting groundwater Yu, X., Liong, S.-Y., 2007. Forecasting of hydrologic time series with ridge regression in
levels in a coastal aquifer. J. Hydrol. 396 (1), 128–138. feature space. J. Hydrol. 332 (3–4), 290–302.
Young, C.-C., Liu, W.-C., Wu, M.-C., 2017. A physically based and machine learning hy- Zadeh, M.R., Amin, S., Khalili, D., Singh, V.P., 2010. Daily outflow prediction by multi
brid approach for accurate rainfall-runoff modeling during extreme typhoon events. layer perceptron with logistic sigmoid and tangent sigmoid activation functions.
Appl. Soft Comput. 53, 205–216. Water Resour. Manage. 24 (11), 2673–2688.
Yu, P.-S., Yang, T.-C., Chen, S.-Y., Kuo, C.-M., Tseng, H.-W., 2017. Comparison of random Zhu, J., Pierskalla, W.P., 2016. Applying a weighted random forests method to extract
forests and support vector machine for real-time radar-derived rainfall forecasting. J. karst sinkholes from LiDAR data. J. Hydrol. 533, 343–352.

282

You might also like