Next Article in Journal
Lag Time and Cumulative Effects of Climate Factors on Drought in North China Plain
Previous Article in Journal
Research Progress and Application Analysis of the Returning Straw Decomposition Process Based on CiteSpace
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Dam System and Reservoir Operational Safety: A Meta-Research

1
Department of Civil Engineering, McMaster University, Hamilton, ON L8S 4L7, Canada
2
Structural Department, Faculty of Engineering, Cairo University, Cairo 12613, Egypt
3
Department of Civil Engineering and School of Computational Science and Engineering, McMaster University, Hamilton, ON L8S 4L7, Canada
*
Author to whom correspondence should be addressed.
Submission received: 28 August 2023 / Revised: 15 September 2023 / Accepted: 21 September 2023 / Published: 29 September 2023
(This article belongs to the Section Water Resources Management, Policy and Governance)

Abstract

:
Dams are critical infrastructure necessary for water security, agriculture, flood risk management, river navigation, and clean energy generation. However, these multiple, and often conflicting, objectives introduce complexity in managing dam operations. In addition, dam infrastructure has been evolving as complex systems-of-systems with multiple interacting components and subsystems, all susceptible to a wide range of uncertainties. Such complexities and uncertainties have triggered extensive research initiatives focused on dam systems and reservoir operational safety. Focusing on the latter, this paper meta-researches (conducts research-on-research) previously published studies to identify the critical research gaps and propose future research directions. In this respect, this paper first performs a quantitative analysis of the pertinent literature, using text mining and subsequent topic modeling, to identify and classify major and uncover latent topics in the field. Subsequently, qualitative analysis is conducted to critically review the identified topics, exploring the concepts, definitions, modeling tools, and major research trends. Specifically, the study identified seven topics: optimization models; climate change; flood risk; inflow forecasting; hydropower generation; water supply management; and risk-based assessment and management. The study also presents three main research gaps associated with the limitations in modeling concepts, modeling tools capabilities, and the lack of resilience-guided management of dam operational safety. Overall, this study presents a road map of the currently available dam and reservoir operational safety research and associated knowledge gaps, as well as potential future research directions to ensure the resilience of such critically important infrastructure, especially in the age of climate change.

1. Introduction

Since the fourth millennium BC, dams have been critical structures constructed to regulate water discharge required for irrigation and to protect people from flood consequences [1]. Attributed to the expansive socio-economic development over the last century, the global demand for water and energy, and, thus, safely operable dams, has been significantly and rapidly increasing [2,3,4,5,6]. In addition, climate change has recently highlighted the role of dams as climate risk mitigation infrastructure [7,8] whereby dams can alleviate the impact of climatological hazards such as droughts and floods. Moreover, several countries increased their reliance on dams for hydropower generation, considered the most efficient clean energy source to reduce greenhouse gas emissions [9]. As a result, hydropower dam construction significantly increased in the past decade to reach nearly 50,000 dams worldwide, with the expectation of continued growth in the future [10].
In parallel to their critical importance, dams are highly complex system-of-systems consisting of several intra-dependencies, including physical (e.g., gates and turbines) and non-physical (e.g., human decisions and maintenance time) components. Dams also have several inter-dependencies with other dams along the same river system (i.e., cascading dams) and other critical infrastructure systems within the communications, energy, food and agriculture, transportation, and water sectors [11]. Such intersectionality of dam operation objectives creates decision-making complexity considering the multi-purpose dam operational rules. For example, flood control operations require a low reservoir water level, while a high reservoir water level is essential for efficient hydropower generation. In addition, on the one hand, dams are considered key for climate risk mitigation [7,12]. On the other hand, climate change is likely to affect dam operational safety, as it introduces uncertainties in the meteorological and hydrological variables that directly impact the temporal and spatial water availability for dam operations [13]. These dynamic, complex, non-linear, and uncertain behaviors introduce serious challenges to dam and reservoir operations, leading to challenges in understanding and predicting dam system operational failures [14]. Given the destructive consequences of such operational failures, extensive research has been established to investigate the safety of dam systems and reservoir operations under multiple stressors.
Several research studies [15,16,17,18,19] reviewed pertinent extensive previous work to identify the critical research gaps in the existing field. However, these traditional reviews were usually subjective in terms of selecting only a subset of documents pertaining to the specific selected sub-topic (e.g., optimization models) within the field. This is prevalent in literature even though the literature showed that topics addressing dam systems and operational safety are very broad due to dam interdependencies and their use’s multi-objectivity. Accordingly, the current study adopts a meta-research (conducting research-on-published research) approach to first identify and classify the main research topics using text analytics (i.e., a machine learning tool that extracts meaningful information from textual data) and subsequently uncover the latent topics and the key research gaps (i.e., ill-covered areas) within the dam system and operational safety. Topic modeling is employed to identify latent topics in different fields such as transportation [20], operational research [21], city resilience [22], and structural engineering [23]. More specifically, this study adopted two stages to represent the meta-research on dam systems and reservoir operational safety. Stage 1 includes quantitative analysis, using topic modeling to identify major topics and uncover the literature’s latent topics (i.e., the topics that do not frequently appear within the literature). Subsequently, Stage 2 includes a qualitative review, examining the definitions, concepts, and recent research trends related to the identified topics. Using quantitative and qualitative analyses, this paper can determine the key research gaps that are not well covered in the literature and can significantly impact dam systems and reservoir operational safety.

2. Quantitative Analysis

2.1. Data Collection

Initially, this study considered more than 1000 journal articles relevant to dam systems and operational safety to be analyzed using text analytics (i.e., topic modeling) to identify the major topics in the field. The data collection process started by searching the titles of the journal articles through the Web of Science (https://www.webofknowledge.com, Last accessed: 1 August 2023) and Engineering Village database (https://www.engineeringvillage.com, Last accessed: 1 August 2023) from 1984 to 2023 using the common keywords related to the field. The considered keywords in the search process are [(Dams OR Reservoirs) AND (Operations OR System) AND (Safety OR Risk OR Resilience OR Uncertainties OR Reliability OR Failure OR Vulnerabilities OR Hazard OR Stochastic)]. The selected articles were subsequently filtered by exploring each article’s abstract and evaluating its relevance to the field. After this abstract-based screening process, the selected articles were further filtered, which yielded 871 journal articles considered for further analysis. The analysis process considers the articles’ abstracts. The abstract is usually adopted by different topic modeling studies [24,25,26] to represent the entire article. Abstracts typically summarize the overall study objective, research questions, methodology, significant findings from the study results, and recommendations through short and precise sentences.

2.2. Pre-Processing

The raw abstracts of the selected articles usually contain various sources of linguistic noise that negatively affect the statistical analysis performed within the topic modeling [27]. This linguistic noise is mainly produced as a result of case type variation (e.g., DAM and dam), special characters (e.g., punctuation), word forms (e.g., operation and operational), and the presence of the English common words (e.g., the, and, of). As such, pre-processing is essential before topic identification to eliminate such linguistic noise. Pre-processing is typically performed through four steps [28]: (1) the transformation step is adopted to change the format of all words to be in lower case; (2) the tokenization step is adopted to convert the unstructured text into words to be analyzed; (3) the treatment step is adopted to act as a filter with ‘stop’ word list to remove the common words out of the text data; (4) the stemming step is adopted to remove all the affixes and return them to the word stem.
Figure 1 shows the word clouds before and after the pre-processing of the raw abstracts data set, showing the words with a minimum of 70 times of occurrence (frequency) within the analyzed text. The size of each word represents its importance, where it reflects the word’s probability of occurrence in the analyzed text. As such, the words with large sizes represent the most frequent (i.e., key) words in the analyzed text. Based on this fact, the comparison between the word clouds in Figure 1 can show that Figure 1a includes highly frequent nontechnical common words (e.g., the, for, and), while Figure 1b eliminates such noise and only represents the technical words related to dam systems and reservoir operational safety (e.g., risk, optim, inflow). This comparison illustrates the critical role of the pre-processing step in text analytics in eliminating the non-essential words that significantly increase the computational cost and affect topic identification.

2.3. Latent Dirichlet Allocation (LDA)

Latent Dirichlet Allocation (LDA) is a generative probabilistic model developed by Blei et al., 2003, [29] based on the probabilistic latent semantic analysis (pLSA) model [30]. The LDA model is used in topic modeling to identify the latent topics of a group of textual documents (more details about how this model works within scientific documents can be found in [31,32]. The main advantage of LDA over other topic models (e.g., Correlated topic models) is the reasonable computational time. LDA significantly reduces the number of dimensions (i.e., words) within the analyzed documents while maintaining the critical relationships between this document’s dimensions and their major topics [29]. Therefore, several studies adopted the LDA model in different fields [33], such as genetic data, music, and structural engineering [23].
The concept of the LDA model is to represent each document (d) (composed of a collection of words (wd)) as a distribution of topics K (θd) (e.g., Two-topic model: document 1 is 70% of topic A and 30% of topic B). Each topic (k) (k K) is characterized by a distribution of words (ψk) (e.g., Topic A consists of 10% of word 1, 20% of word 2, 30% of word 3, etc.). The LDA model aims to estimate the θd and ψk, so the user can infer the major topics represented within the analyzed documents using the frequent words related to each topic. The two distributions, ψk and θd, are evaluated using two Dirichlet distributions Beta (β) (i.e., per-topic-per-word probabilities) and Gamma (α) (i.e., per-document-per-topic probabilities), respectively (for more details about Dirichlet distributions formulation, see [34]). The LDA algorithms procedure can be summarized as shown in Figure 2:
(1)
The model user determines the analyzed documents (D);
(2)
Define the number of topics (K);
(3)
The algorithm randomly assigns a particular topic kdi to each word wdi based on:
  • The word distribution for each topic k (ψk), evaluated by initial Dirichlet (β);
  • The topic distribution for each document d (θd), evaluated by initial Dirichlet (α)
  • where wdi is the word (i) in the word collection (wd) per document (d), d [1, D], i  [1, Nd (number of words in wd)], and kdi is the topic assigned to the word i (wdi) in the document (d), kdi [1, K];
(4)
Using Gibbs sampling [35] with iteration j (j = 1000 in this study), the algorithm improves the topic assignment in terms of enhancing the values of β and α, where the algorithm can calculate the probability that word (wdi) is generated from the topic (kdi);
(5)
The algorithm re-assigns each word (wdi) with the new topic (kdi) based on the previous step calculations.

2.4. Perplexity

One of the most challenging steps in the LDA algorithm is evaluating the optimum number of topics (K) within the analyzed documents (D). This paper adopted perplexity metrics to determine the optimum number of topics. Perplexity is a statistical measurement used as a guide to select the optimum number of topics in topic modeling [29]. Perplexity metrics evaluate the ability of the probability models to predict a sample by calculating the relative degree of uncertainty among such models, where the best model should have the least perplexity value. Figure 3 shows the sensitivity of the perplexity evaluation values to the number of topics. A threshold of 30 topics was chosen as the maximum number of topics to be considered, which is a relatively large number compared to the number of articles under consideration. In this study, we consider 7 as the optimum number of topics, considering the slope of the perplexity curve (dash line) firstly changes at 7, and the differences in the perplexity value between 7 and 30 are relatively small, as shown in Figure 3.

2.5. Topic Identification

Figure 4 and Figure 5 show the identified seven topics, where the word cloud for each topic is represented in Figure 4, while the β value for the top ten words per topic is shown in Figure 5. In Figure 5, β can be interpreted as the probability that each word is generated from one of the identified topics. For example, the β value for “water” in Figure 5e is 0.12, meaning that “water” represents 12% of Topic 5, which consists of a collection of words. The β value can also refer to the level of this word frequency in the topic. The words of high β values in Figure 5, shown large in size for the word clouds in Figure 4, refer to the most frequent words in each topic that can be used to infer the main focus of the topic.
Topic 1 shows high frequency of word stems such as “optim”, “model”, and “algorithm”, which can refer to the (Optimization models) that are used in dam operations. Using the same approach, the high-frequency word for Topic 2, such as “climat”, “change”, “river”, and “flow”, can refer to the variation of the hydrological and meteorological variables caused by (Climate Change) that directly affect dam operational safety. Moreover, the high-frequency word for Topic 3, such as “flood”, “risk”, and “control”, can refer to (Flood Risk) that significantly threaten the safety of dam operations. In addition, the high-frequency word for Topic 4, such as “inflow”, “forecast”, and “uncertainti”, can refer to the effect of (Inflow forecasting) uncertainties for dam operational safety. Furthermore, the high-frequency word stems for Topic 5, such as “hydropower”, “generate”, and “schedule”, can refer to the safety of the (Hydropower generation) operation under various operational conditions. In addition, the high-frequency word stems for Topic 6, such as “water”, “rule”, “hedg”, “drought”, and “suppli”, can refer to the risk of water shortage in (Water Supply Management) due to the increasing water demands and the increasing rate of drought events. Finally, the high-frequency word for Topic 7, such as “risk”, “analysi”, “evalu”, and “assess”, can refer to (Risk-based Assessment and Management) approach that is adopted in dam safety assessment studies.

3. Qualitative Analysis

This section aims to comprehensively review the identified topics in the previous section, citing 376 recent studies related to dam and reservoir operational safety. This section cites such a massive number of the most recent research papers; however, it is worth mentioning that due to space limitations, this section is restricted to concisely stating the definitions, concepts, modeling tools, and major research trends adopted in each of the identified seven topics. For more in-depth details about mathematical formulations and modeling tools application, the paper refers to the recent review papers and textbooks conducted to review each topic specifically. It should also be noted that “reservoir” and “dam” are used interchangeably throughout the literature and this review, referring to the same dam and its reservoir system.

3.1. Topic 1: Optimization Models

Optimization is defined as a mathematical formulation in which algorithms are used to compute the values for a set of decision variables (e.g., outflow, gate positions) that maximize or minimize the objective functions (e.g., flood control, hydropower generation), considering problem constraints (e.g., release constraints, storage constraints). For more details about dam operations optimization model variables, constraints, and objective formulations, see [36]. In the context of dam operations, two approaches are generally adopted to determine the optimal operational rules. One approach is to pre-define the operational rule form (e.g., rule tables, rule curves) and subsequently use the optimization algorithms to optimize the parameters of such defined rules [37]. However, this approach is inflexible as it restrains its results to the pre-defined form. Also, it requires extensive computational time in multi-objective problems. As such, a more flexible approach is adopted, where the optimization models are used first to generate optimal input–output solutions, and then data mining methods are applied to extract the optimal operational rules [38,39]. Accordingly, various complex modeling tools are adopted to conduct such two approaches in optimal dam operations problems. These complex modeling tools, including optimization programming techniques, solution algorithms, and inference models, can be classified into five main categories as shown in Figure 6: (1) implicit stochastic optimization (ISO) models, (2) explicit stochastic optimization (ESO) models, (3) computational intelligence (CI) models, (4) multi-objective optimization (MLO) models, and (5) simulation-optimization (S-O) models.
ISO is an optimization modeling technique that implicitly includes stochastic features of reservoir random variables (e.g., spatial and temporal variations of inflow discharges) using deterministic optimization programming techniques such as linear programming (LP) [40,41,42,43,44] and its extensions (binary LP (BLP), integer LP (ILP), mixed-integer LP (MILP), non-linear programming (NLP) [45,46,47,48,49] including the successive LP (SLP), sequential quadratic programming (SQP), generalized reduced gradient (GRG), deterministic dynamic programming (DDP) [50,51,52,53,54] and its modified models to solve its curse of dimensionality (dynamic programming successive approximation (DPSA), incremental DP (IDP), discrete differential DP (DDDP), and discrete-time optimal control theory (DOCT)) [55,56]. The deterministic optimization models can generate the optimal policies for several historical or synthetically time-series data for reservoir random variables (e.g., inflow). Subsequently, statistical inference methods (e.g., multiple regression analysis) are used to infer the operating rule curves from such generated optimal operating policies, considering current reservoir conditions (e.g., current storage level). However, this approach, ISO modeling, usually has various limitations. For example, such inferred operating rule curves are unique to the input (historical or synthetical) time-series data. Moreover, poor correlations resulting from regression analysis may invalidate the inferred operating rules. Also, using advanced inference methods (e.g., ANNs, FRB) to infer such operating rules (rather than the regression analysis) requires extensive computational cost due to the multiple trials and errors to reach an efficient result.
Unlike ISO, ESO modeling explicitly provides the probabilistic description for the stochastic features of reservoir random variables using stochastic programming techniques such as stochastic linear programming (SLP) [57,58,59], stochastic dynamic programming (SDP) [60,61,62,63], chance-constrained programming (CCP) [64,65], and stochastic optimal control theory (SOCT) [66]. In this respect, the proper probability distribution (parametric or nonparametric) can be fitted for the random reservoir variables based on the statistical frequency analysis. Thus, the inferred advantages of ESO over the ISO are: (1) the optimization process directly considers variables’ uncertainty; (2) The optimal operation rule can be directly determined without using any interference tools. However, ESO programming is more computationally demanding than ISO, particularly in complex multi-objective problems.
Alternatively, CI models solve complex cases that cannot be solved using conventional methods. As such, CI models have a wide range of applications in optimal dam operation complex problems, especially during the development of the current computational capabilities. CI refers to a set of nature-instigated computational models that generally includes three main methods: (1) Evolutionary computation (EC), (2) Artificial neural network (ANN), and (3) Fuzzy logic systems (FLS).
EC models, known as heuristic searching tools, include metaheuristic algorithms such as Genetic algorithms (GA) [67,68,69,70,71], Ant colony optimization (ACO) [72], Particle swarm optimization (PSO) [73,74], Simulated annealing (SA) [75,76], and Honey bees mating optimization (HBMO) [77]. Unlike other algorithms, meta-heuristic algorithms are known for their efficiency in searching for near-optimal (i.e., Pareto optimal) solutions. They can also compute optimal global solutions to problems in which the conventional algorithms fail to converge or become stuck in local optima [51]. As such, EC-based algorithms, using their unique natural structure, can address the complexities of reservoir operation problems such as discontinuity, uncertainty, nonlinearity, multi-objectives, and discreteness. For more information about using metaheuristic algorithms in dam operation, see reviews [78,79,80].
Although ANNs cannot be classified as an optimization model per se, ANN is one of the most powerful data mining methods used to infer the optimal operating rules from optimization model results (e.g., ISO) [51]. ANNs also have a wide range of applications within reservoir operations, such as inflow forecasting (see, Topic 4). ANN is a paralleled-distributed information processing system. It aims to simulate the brain’s information process for thinking and reasoning by resembling the brain’s biological neural networks [81]. In this respect, ANNs typically consist of a large number of nodes (i.e., neurons) arranged in specific patterns (i.e., layers). Within such layers, the information (i.e., data) is typically transmitted between the nodes from the input to the output layers through multiple hidden layers using directional links (i.e., assigned by weights). Each node (other than input layer nodes) is defined by an activation function that evaluates the activation status of each node based on the weighted inputs and added bias. Accordingly, ANNs have multiple architecture types according to the number of layers (e.g., single (Hopfield network), by-linear (adaptive resonance networks), multilayer (backpropagation networks)), and the direction of the information flow (e.g., unidirectional in feed-forward ANN, radial flow in radial basis). More details about ANNs configurations can be found in [82]. It should be noted that most of the recent ANNs studies are cited in Topic 4. More details about ANNs application in dam operation can be found in [19,83].
Similar to ANNs, FLS is considered an advanced inference tool used as an alternative solution to regression analysis to infer operating rules from historical operations or ISO model results [84]. Fuzzy logic, developed by Zadeh in 1965 [85], is a logical-mathematical procedure based on the “If/Then” procedure that simulates the human way of thinking computationally. Unlike complex mathematical models, fuzzy logic models use linguistic descriptions to define the relationships between input data and output decisions. As such, it enables FLS to deal with uncertain decision-making problems. The fuzzy system typically consists of four main components: (a) Fuzzification is a process of transforming the numeric data (i.e., crisp) to fuzzy input; (b) Knowledge base is a collection of fuzzy conditional rules that link the input to the output variables; (c) Inference engine is a process to combine fuzzy rules output; (d) Defuzzification is a process to transform the fuzzy output back into the numeric data. Besides being an inference tool for dam operational rules, Bellman and Zadeh, 1970 [86] introduce the integration of fuzzy sets in the optimization algorithms (i.e., fuzzy optimization models) to represent the imprecision and vagueness in the constraints and objective functions. These fuzzy constraints and objectives are characterized by their membership functions, where the decision is selected based on the intersections of the fuzzy constraints and objectives. This concept is then adopted to develop different fuzzy optimization models (e.g., fuzzy DP [63,87], grey fuzzy SDP [88]) to solve such uncertainty in reservoir operation optimization problems. For more details, see [89].
As most of the recent dam-operation problems are multi-objectives, two approaches: the epsilon-constraint method and the weighting method, are commonly adopted by the optimization models (stated above) to solve the multi-objective operation problem [19]. Regarding the epsilon-constraint method, the multi-objective problem is converted to a single objective by considering the least important operation objectives as the problem constraints. However, for the recent multi-objective dam operations, optimizing all objectives simultaneously using a single solution value is impossible. Instead, a set of efficient non-inferior optimal solutions (i.e., Pareto optimal solutions) is determined to define a boundary beyond which none of the objectives can be improved without sacrificing at least one of the other objectives. Unlike the epsilon-constraint method, the weighting method combines the operation objectives by assigning comparative values (i.e., weights) to each objective, where the Pareto optimal solutions are computed by varying the weights for each objective and solving the problem sequentially. Subsequently, the Pareto front (PF) can be defined as the set of all non-dominated Pareto optimal solutions for the objectives vector. For other details of multi-objective programming, see [90,91,92]. After quantifying the PF, multicriteria decision-making (MCDM) methods are usually employed to rank these Pareto optimal solutions and determine the most preferred solution that needs to be implemented (For more details about MCDM methods and ranking process, see [93,94,95].
Despite the advancement of optimization models, simulation models are still one strategic tool to estimate the dam and reservoir system’s behavior under various conditions. However, simulation models are very time-consuming in complex operation cases with a large number of operational policies. Thus, several studies of large-scale systems use optimization models to preliminary screen such operation policies to significantly improve simulation model efficiency [96,97,98,99]. In this respect, the simulation and the optimization models are preferred to work in a loop sequentially, where the optimization models use the simulation model outputs to find the optimal solution that can then be considered a better-simplified input to the simulation model. Accordingly, several simulation-optimization (S-O) models, such as simulation-GA–GA models [100,101], are adopted in the various dam and reservoir operational problems. More details about S-O modeling techniques application in reservoir operation can be found in [36]. For more details about comparing the recent optimization models adopted in reservoir operations, see the reviews conducted by [102,103,104,105,106].

3.2. Topic 2: Climate Change

Despite the role of dams and reservoirs as a climate-change risk mitigation infrastructure, climate change, in turn, directly affects the safety of dam operational objectives (e.g., flood control, hydropower generation, and water supply management) [7,107,108,109,110,111,112,113]. Climate change intensifies weather extremes that subsequently induce a series of devastating natural hazard events (e.g., floods, droughts). Moreover, the frequent changes in climate conditions increase the uncertainties in the meteorological and hydrological variables, which subsequently introduce risk in dam operational decisions. For example, the rise in water surface temperature reduces snow storage (i.e., snow-pack) and increases the water losses due to evaporation, leading to reduced and earlier spring runoff. Accordingly, dam operators should refill the reservoirs early to ensure adequate summer water supply storage. However, the early refill reduces the flood storage capacity and increases flood risk. In addition to the variation of runoff temporal characteristics, climate change can also affect river runoff discharge due to the rapid variation in precipitation patterns, which introduces uncertainty of floods’ magnitude and frequency. Climate change may also introduce uncertainty in dam components’ operational function (i.e., performance). For instance, intensive rainfalls may excite the soil particles, where the fine sediment particles in water may increase the possibility of gate abrasion and erosion processes, while the oversized particles (e.g., debris, tree branches) can lead to gate blockage. Furthermore, the change in atmospheric temperature can expose gates to additional stresses and deformations, which affect the gates’ maneuvering mechanism. For more information about the impact of climate change on dam operational safety, see [2,107,114,115].
In addition to the significant climate change impact on dam operational safety, most of the existing dams have inefficient construction and operation schemes that neglect the ever-changing climatological conditions [116,117,118]. Accordingly, several models have been developed to provide adaptive operational rules to consider such induced climate change operational risk. In this respect, the integration of climate, hydrology, and optimization models are used to provide adaptive dam operation rules for various climate change scenarios [119,120,121,122,123,124,125].
General circulation models (GCMs) are numerical models that represent atmospheric processes through complex mathematical equations. GCMs are commonly used to generate various climate scenarios [126]. However, GCM outputs have a too-coarse temporal and spatial resolution, which cannot be directly used by the hydrological simulation models. Thus, downscaling techniques are usually adopted to transform the coarse resolution of the GCMs into more acceptable (finer) resolution results. The main idea of downscaling is to convert the large-scale changes in the atmospheric variables (i.e., predictors) generated by GCMs, to local climate variables (i.e., predictands). These downscaling techniques can be classified into dynamic and statistical downscaling models, with some approaches combining the two techniques for higher-resolution results (see [127,128,129]).
Dynamic downscaling usually refers to regional climate models (RCMs) or limited-area models (LAMs), which provide high-resolution regional simulations. RCMs are driven by GCM outputs, as model boundary conditions, to dynamically extrapolate the effect of the large-scale climate pattern to the regional scale. However, the limitation of the dynamic downscaling techniques refers to the computational cost that restricts the downscaling results to a specific resolution level, typically 20–50 km. Subsequently, the model results usually require further downscaling and bias correction for its outputs (see the review in [130]). In contrast, statistical downscaling models require minimal computational time and are easier to implement and interpret. Statistical downscaling models use several statistical-based models (e.g., multiple regression analysis, analog method, LARS-WG) to develop the statistical relationship between the GCM outputs and observed local climate responses, assuming that this relationship does not change in the future under changing climate conditions (i.e., stationarity predictor-predictand relationships) (see [131]). However, statistical downscaling techniques have some limitations due to the lack of observed data with high quality. Therefore, selecting the downscaling technique is a challenging step that depends on the financial and time constraints and the desired temporal and spatial resolutions of the generated climate scenarios (for more details about the comparison between the dynamic and the statical downscaling models, see [132]).
These climatic scenarios, generated by the downscaling techniques, are used as hydrological simulation model inputs. These hydrologic simulation models, such as Variable Infiltration Capacity (VIC) [133,134] and SWAT model [135,136], are employed to simulate the river runoff corresponding to the given climate scenarios. Reservoir response (e.g., storage status) can then be developed for each climate scenario based on the simulated river runoff and the reservoir releases using the operating rule curves generated by the optimization models (stated in Topic 1). The reservoir conditions are then evaluated based on the specified safety and operation criterion, and thus, the best optimal operating rule curve can be selected. This top-down integration approach is one of the widely adopted decision-scaling frameworks in developing dam safety assessment plans, where the reservoir response to the expected future climate scenario drives stakeholder decision-making [137]. Meanwhile, a bottom-up approach [138] is also applied to investigate reservoir system vulnerabilities and the probability of system operational failure. Within the bottom-up approach, the operational failure limits are first defined using the historical records and dam stakeholders’ safety standards. GCMs can then be employed to generate future climate scenarios and subsequently identify the system’s possibility of being pushed into a vulnerable state beyond those limits (see [139]). For more details about the interaction between climate change and dam operations, see the reviews conducted by [2,7,140,141,142].

3.3. Topic 3: Flood Risk

Floods are one of the most destructive natural disasters with severe consequences of human fatalities and economic losses [143]. In the last ten years, floods have caused more than 50,000 deaths and more than USD 400 billion in property damage (Flood Disaster Database, available at https://ourworldindata.org/, Last accessed: 1 December 2022 and https://www.emdat.be/, Last accessed: 1 December 2022). Thus, reservoir flood control operation (RFCO) is a crucial objective in reservoir operations to reduce flood peaks and mitigate flood damage to the downstream and the dam structure itself. RFCO objectives can be specified based on the number of reservoirs, Single or Multi-reservoir systems. For single reservoirs, RFCO aims to minimize the maximum reservoir water level (MRWL) during flood peak discharge, considering the downstream safety against dam releases [144,145,146]. However, a single reservoir may not provide adequate storage capacity for downstream safety. Therefore, multi-reservoir systems are operated together to protect such areas from flood consequences. However, RFCO objectives change for multi-reservoir systems according to the configuration of the reservoirs within the system, cascading (series), parallel, or mixed reservoir system. Cascading reservoirs have hydraulic inter-relationship and storage compensation rules. As such, the downstream reservoir’s safety is affected by the RFCO objectives of the upstream reservoir. In this respect, RFCO aims not only to minimize MRWL to ensure dam safety but also to ensure the safety of different critical flood sections downstream, considering the safety of the downstream area and the following dam structure. On the contrary, parallel reservoirs do not have any hydraulic inter-relationship, where the safety of the downstream area is affected by the coordination in RFCO of the parallel reservoirs. Regarding the mixed reservoir systems, RFCO is more complicated, where the complexity of hydraulic connection for the cascading reservoirs and the coordination of asynchronous peak flows for the parallel reservoirs should be considered simultaneously (see [147,148,149,150,151,152]).
Within each reservoir among these reservoir systems, RFCO objectives can be specified based on the operational stage, which is typically carried out through three main stages [153]: (1) Prior flood event, the operator aims to increase the reservoir storage capacity by pre-releasing reservoir water discharge; (2) During flood event but prior flood peak flow, the operator releases as possible from flood discharge ensuring the safety of reservoir water storage levels; (3) After flood peak flow, the operator aims to balance between maintaining MRWL within the allowed reservoir limits, increasing the reservoir storage to be used in hydropower generation and decreasing dam releases to ensure the safety of the downstream area.
RFCO objectives are typically performed through fixed operating rules that mainly rely on historical or synthetic statistical flood analysis [152,154]. These operating rules can consequently define the RFCO targets (e.g., target reservoir water level) for each operational stage, considering the designed reservoir operational parameters (e.g., designed flood storage). However, these operating rules usually depend on subjective experts’ judgment that cannot provide optimal flood control operations, especially for multi-purpose reservoirs. Therefore, advanced optimization and simulation methods (reviewed in topic 1) were adopted to provide optimal real-time RFCO operating rules under various operational conditions. These studies consider different aspects of the RFCO, including single and multiple reservoirs [147,155,156,157], short-term and long-term scheduling [148,158,159], single and multi-objective approaches [144,160,161,162,163,164], and multi-criteria decision-making models [93,94,165,166]).
However, the obstacles to the safety of RFCO problems are not only related to the complexity of extracting the optimal rules for RFCO operational stages but also the associated uncertainties of RFCO variables (e.g., flood peak flow) and, subsequently, the RFCO targets. Therefore, several studies considered the uncertainties in RFCO and estimated their corresponding risks to support the decision-makers in developing effective dam operational plans under various conditions [161,167,168,169,170,171,172,173,174,175,176,177]. These uncertainties are introduced from four main sources [178,179,180]: (1) hydrological (e.g., inflow) and meteorological variables (e.g., precipitation) uncertainties that introduce randomness in flood frequency and its magnitude; (2) flood forecasting and river flood routing errors, as a result of models structure uncertainty, parameter estimation errors, imperfection of initial and boundary conditions, and simplification of the physical processes; (3) hydraulic uncertainties, such as errors in reservoir outflows (e.g., due to debris accumulations), and errors in reservoir discharge curve (i.e., the difference between actual and designed spillway capacity resulted from flow coefficients and measurement errors of the spillways); and (4) other uncertainties (i.e., deep uncertainties), such as operation time delay, human decision errors, information unavailability, poor communication. These uncertainties were introduced to quantify the corresponding risk in RFCO using risk-based models as reviewed in Topic 7.

3.4. Topic 4: Inflow Forecasting

Inflow is one of the most critical variables for dam and reservoir operations [181]. Accurate inflow forecasting can provide preceding inflow information that can significantly reduce the risk of floods and droughts, improve water allocation schemes for water supply management, and enhance hydropower scheduling plans [182]. Hence, several techniques have been developed to forecast reservoir inflow. Such techniques can be classified as statistical, physical, and intelligent methods. Physical and statistical methods have been widely used in the past decades to predict inflow characteristics. Statistical models, such as Autoregressive, Autoregressive Moving Average (ARMA), and Seasonal ARIMA (see [183] for a comparison between the three models), are used to estimate the temporal inflow variation, assuming that the behavior of the historical data is valid for the future (e.g., seasonal variation) [184]. Although the statistical methods are relatively simple and mature, they suffer from low accuracy in considering the dynamic, non-stationary, and non-linear inflow behavior [183]. Alternatively, the physical models (e.g., soil and water assessment tools), including semi-and fully physically distributed models, use physical-based equations (e.g., mass, momentum, or energy conservation equations) to represent the hydrological characterizations of the inflow data (e.g., [185]). Although physical-based models are more accurate than conceptual hydrological models, they usually require complex mathematical procedures, a large number of parameter calibrations, and complex initial/boundary condition determination. For more details about comparing the physical and conceptual hydrological models, see [186,187]. Alternatively, advanced intelligent methods were adopted to overcome the limitations of the physical and hydrological predictive models, such as artificial neural networks [188], support vector machine [189], decision tree models [190], Bayesian Networks [191], fuzzy inference system [192], grey model [193], wavelet transforms [194], and clustered k-nearest model [195]. The main advantage of such data-driven models is their ability to provide real-time forecasting and include external factors (e.g., weather conditions) that can improve forecasting accuracy (see [196]). Moreover, data-driven models only depend on the historical hydro-meteorological data without directly considering underlying physical processes and thus require much less input and parameter data. For the comparison between the data-driven and physical models, see [197]. However, the forecasting accuracy of the data-driven models can be affected by data representation (i.e., data quality) that affects the extraction and learning features of the historical data [198]. Moreover, using only one data-driven may not be sufficient, as each model has its features that may not be suitable for all rivers’ hydrological characteristics [199]. Therefore, hybrid models have been recently developed [199,200,201,202,203,204,205,206,207,208] to consider the advantages of each forecasting model. Thus, hybrid forecasting models can improve forecasting accuracy and subsequently develop efficient reservoir operation rules. For more details about the application of hybrid models in inflow forecasting, see the comparative study performed by [209] and the review conducted by [210].
Despite the advances in inflow forecasting techniques, their application in reservoir operation is constrained by a limited forecasting horizon (FH) [211,212]. Reservoir operations require a long enough forecasting leading time to provide sufficient information for the decision process horizon. However, a meaningful forecast horizon is usually shorter than the required operation horizon—the shorter the FH, the lower the possibilities of error and uncertainties. Thus, several studies were conducted to determine the effective FH for inflow forecasting models that can provide efficient forecasting leading time for reliable operational decisions (see, [181,213]). Within these studies, posterior evaluation values of forecasts with different lengths of leading time are obtained based on comparing the actual decisions (performed based on the observed inflow) and the numerical decisions generated using the forecast models. These studies concluded that effective FH could be specified based on reservoir operation objectives. The short FH is efficient for reservoirs with small capacities operated for short-term objectives such as flood protection. However, the long FH is essential to large and medium-capacity reservoirs used for long-operating objectives such as water supply and hydropower generation.
The uncertainty of Inflow forecast models not only varies by the forecasting horizon time; but is also affected by the modeling errors, including errors in the initial and boundary conditions, errors in model parameters, poor model resolution, simplifying the physical process, and the imperfect representation of the actual topography [214]. Therefore, nowadays, probabilistic forecasts are preferred over deterministic ones [215,216]. Probabilistic forecasts provide decision-makers with a better trade-off for risks and benefits than the single-value forecast that ignores uncertainties [217,218]. Moreover, probabilistic forecasts can provide reservoir operations with a relatively longer leading time than deterministic forecasts [219,220]. In this respect, the probabilistic distributions are typically generated using ensemble forecasts to randomly sample the uncertainty range in inflow forecasts. However, ensembles are often under-dispersive and, therefore, are typically unreliable in inflow forecast applications [221]. As such, simple uncertainty models, such as the binned probability ensemble and Gaussian uncertainty model, are used to fit the probability distribution functions to the ensembles using the statistical properties of the generated ensemble. Unlike such simple uncertainty models, other complex uncertainty models, such as the generalized likelihood uncertainty estimation and Bayesian recursive estimation, are used to produce probabilistic forecasts by sampling the uncertainty of the hydrology model parameters [222,223,224]. However, due to the assumptions made by the uncertainty models, the resulting probability forecasts should be statistically calibrated for risk-based decision-making. These calibration approaches can be classified into two main groups: (1) ensemble calibration, such as Bayesian model averaging [225], which adjusts the forecasted ensembles to enhance the inflow forecast result; and (2) probability calibration, such as the weighted ranks method [226,227], which directly adjusts the generated probabilities (i.e., uncertainty model results). For more details about calibration techniques and uncertainty models, see [216].

3.5. Topic 5: Hydropower Generation

Power generation is the primary contributor to global warming, producing more than 30% of greenhouse gas emissions (GHG) [228]. Therefore, several countries have increased their reliance on dams as hydropower generation facilities to reach GHG reduction targets. Hydropower is considered one of the largest sources of renewable energy based on the United Nations reports [9,12] and the World Energy Generation database available at https://ourworldindata.org, Last accessed: 1 December 2022. These rates are expected to increase, considering the rapid global energy demand due to the global growth of the population and economy. As such, dam construction has globally boomed in the past decade, with expectations of a continuous increase in construction rates in the future [10]. However, it is more economical to develop advanced dam operational rules to ensure the safety of the existing hydropower power dams and improve their efficiency under various conditions [229,230].
The existing hydropower dams can be classified into Run-of-River (RoR) and reservoir-based hydropower dams. RoR is considered lower in risk than reservoir-based dams. RoR has no or little water storage; instead, it uses the instantaneous flowing discharge of water to generate electricity [231]. However, it is considered a small hydro project that generates limited and unsustainable energy rates. As such, reservoir-based hydropower dams are widely used in hydropower generations that use the reservoir’s storage and the differences in reservoir water elevations to operate the turbines and generate electricity [232]. Although various operation modes are adopted in reservoir-based dams (e.g., peaking, storage–release, pump storage), the main objective of such operations is to maximize hydro generation rates and minimize the production cost under multiple conditions. For more details about the types of hydropower operations, see [233].
Pumped-storage hydropower (PSH) is considered one of the most widely adopted systems for large-scale energy storage. The advantages of such a technique include the high-potential energy stored in the hydro reservoirs, the low production cost per power unit, the efficiency of the energy conversion of the whole cycle, and its flexibility in short-term hydropower operations. The basic idea of the PSH technique is having two water reservoirs at different elevations that can provide a long energy storage period by pumping the water from the lower reservoir to the higher one, converting kinetic energy to stored potential energy. Typically, the water is pumped to the higher reservoir in the low-demand periods (off-peak hours), and it is then released during the high-demand periods (the peak hours). As such, PSH usually takes advantage of electricity price variations (higher prices in peak hours) and ensures the continuity of the energy supply during different seasonal events.
Based on a such basic idea, several studies have been conducted to improve various aspects of PSH operations to maximize the economic benefits and power production efficiency. Table 1 summarizes the recent studies and trends performed on PSH systems between 2020 and 2023. However, for the previous studies, please see the reviews [234,235,236,237,238].
Despite the previous research conducted on pumped-storage hydropower systems, more proper dam safety assessment plans are required to improve the efficiency of this system in energy storage.
Moreover, various studies have been established to employ the advanced optimization models (stated in topic 1) in hydropower generation scheduling through short-term, medium-term, and long-term time intervals. These studies can be summarized in Table 2, stating the main goal and case study for each research work. It should be noted that only the recent studies (from 2023 to 2018) are considered due to the paper space limitations, while for more information about the earlier studies, see reviews conducted by [80,252,253,254,255,256].

3.6. Topic 6: Water Supply Management

Water scarcity is a rapidly increasing problem for many countries that inevitably affects various social and economic sectors [293,294,295]. As a result of environmental degradation, water resources become depleted and subsequently cannot meet the ever-increasing water demands driven by socioeconomic development [296,297]. Moreover, due to global warming, dry and heat weather waves cause severe imbalances in water cycles, causing extreme drought events [298,299,300]. Hence, efficient water supply management in dam operation is critical for regulating the fluctuating water discharge and reducing the imbalances between water supply and water demands. However, the continuous depletion of water resources, dam multi-objective operations, the uncertainty in water demand and hydrological variables, and the possibility of drought events introduce a complexity of water supply operation rules to achieve the optimum water allocation for multiple sectors.
Dams employed several water supply rules such as Pack Rule, Space Rule, New York City Rule, Linear Decision Rule and Standard Operation Policy (SOP), and Hedging Rules (HR) to determine reservoir releases and the starting and ending conditions of supplying water (see, [301]). Among these rules, SOP is widely used in dam operations to satisfy the requirements for water supply management [302]. However, SOP fully releases water demands if the available water is sufficient, saving only the surplus water for future delivery. Consequently, SOP may not be efficient in reserving water discharge for dry seasons and subsequently causes severe water shortages during drought events [303,304,305,306,307]. Accordingly, HR is developed based on optimizing SOP operation rules to reduce the severity of drought events. HRs accept reducing water supply discharge in normal situations and reserve a portion of such water discharge to decrease the expected water shortage during droughts. Thus, effective HRs lead to a balanced water supply operation, which enhances dam performance to mitigate drought risks (see [37,308,309,310,311]).
Two points commonly define HRs: the starting water availability and the ending water availability. Between such two points, the water supply operation follows HRs, and beyond these points, the SOP is followed [312,313]. Thus, several HRs have been developed, such as the Two-point linear hedging rule [314], One-point linear hedging rule [315], Reserve-storage two-point linear hedging rule [316], Three-point linear hedging rule [317], Discrete phased hedging rules [318], to identify the HR parameters within the starting and ending hedging points, aiming to answer two main questions: (1) When to hedge? and (2) How much to hedge? For more information about comparing such types of HRs in dam operation, see [319,320,321,322]. The answer to the two questions is determined according to the parameters of HR trigger and rationing factors, respectively. The trigger can be defined as the initial and terminate threshold at which the HR is to start (i.e., “When to hedge?”), while the rationing factor is the ratio of the reduced water supply from the original target delivery (i.e., How much to hedge). These hedging rule parameters are typically obtained using simulation models, optimization models, or simulation-optimization (S-O) models (for more details, see [323,324,325,326,327,328]) that are summarized in topic 1. For more details about HR parameters quantifications, see [329].

3.7. Topic 7: Risk-Based Assessment and Management

Generally, absolute dam and reservoir operational safety is not possible to be attained due to the multiple associated uncertainties, including aleatory uncertainties (e.g., the randomness of nature), epistemic uncertainties (e.g., errors in model parameters), and deep uncertainties (e.g., failure scenario uncertainty) (see [330]). Such uncertainties introduce risk in dam and reservoir system operation decisions. Risk is defined as the potential losses that may occur corresponding to hazardous events [328,331,332,333]. For each hazard event, the risk is responsible for answering three main questions: what can happen (i.e., system failure response), how likely this can happen (i.e., probability of failure), and what are the consequences of this failure (i.e., failure consequence analysis generally). Absolute dam and reservoir operational safety is not possible to be attained due to the multiple associated uncertainties, including aleatory uncertainties (e.g., the randomness of nature), epistemic uncertainties (e.g., errors in model parameters) and deep uncertainties (e.g., failure scenario uncertainty) (see [330]). Such uncertainties introduce risk in dam and reservoir system operation decisions. Risk is defined as the potential losses that may occur corresponding to a hazardous event realization [328,331,332,333]. For each hazard event, the risk is responsible for answering three main questions: what can happen (i.e., system failure response), how likely this can happen (i.e., probability of failure), and what are the consequences of this failure (i.e., failure consequence analysis).
As such, risk has been the main focus of the International Commission on Large Dams (ICOLD) for several years, concluding that risk-based management is key to effective dam safety assessment plans that consider the associated hazard and uncertainty [334]. Risk-based management has recently evolved into “Risk-informed management”, as described by FERC, 2016 [335]. The latter is a process that provides comprehensive safety guidance to dam decision-makers typically established through four main steps [336,337]: (1) Risk analysis aims to estimate and describe the risk posed to the system through a systematic procedure including, the definition of system components and their inter-relations, the identification of potential system hazards and the corresponding system failure mode, the estimation of dam failure risk and the corresponding failure consequences, and the identification of vulnerabilities in system components to be included in risk assessment plans. (2) Risk evaluation provides the understanding and judgment of risk significance to the system. It defines the tolerable level of risk considering the safety standards of dam owners and regulators and the acceptable safety limits to the public, including the economy and human society. This step is considered essential decision guidance for how the estimated risk can be assessed. (3) Risk assessment is the core of the risk management process, which integrates the results for risk analysis (i.e., estimated system risk) and risk evaluation (i.e., tolerable system risk) to assess dam safety level. Accordingly, this step is responsible for assessing the alternatives of risk mitigation plans essential for risk control. (4) Risk control improves the existing risk mitigation and emergency plans according to dam risk assessment and develops the surveillance procedures to monitor the efficiency of the process to avoid or/and control the risk.
Various risk quantification approaches are proposed for dam operation risk with two commonly-adopted approaches [338,339]: (1) Risk is, conceptually, the product of hazard (i.e., probability of hazard occurrence), vulnerabilities (i.e., the possibility of system failure during this hazard event), and exposure (i.e., the possibility of the system to be exposed by this event); or (2) Risk is, conceptually, the product of hazard and its consequences (i.e., severity). Several risk analysis techniques are employed to quantify dam operational failure risk, which can be classified into six main categories. (1) The return period method, introduced by Fuller in 1914 [340], had been widely adopted for decades due to its simplicity in determining dam operation risk in terms of predicting the frequency of hazard events, especially in floods [341,342,343]. In this respect, the return period value is inversely related to the probability of a specific variable (e.g., flood flow) exceeding a particular value (e.g., 100-year flood means the probability that the magnitude of this flood to be exceeded at any year flood is equal 1%). However, it has some limitations in representing the integrated risk of flood control operations (for more details about the return period method, see [344]). (2) The stochastic differential equation method estimates system risk through integral or differential calculations based on the probability density function (PDF) of the uncertain variables [345,346]. (3) Reliability-based methods, such as the first order second moment (FOSM) method and its advanced methods (AFOSM) [347,348,349], calculate system failure risk based on reliability index, β, a relative measure of confidence in the system’s ability to perform its function efficiently. Their advantage over stochastic differential methods is their ability to deal with uncertain variables without a distribution function [350]. However, these methods neglect the higher-order moments so they may underestimate the risk. (4) Stochastic simulation, known as the Monte Carlo (MC) method [351,352], is a widely used statistical sampling technique that generates the PDF of a set of random variables. Using MC simulations, the probability of failure of the system can be calculated based on the joint PDF of its uncertain components (i.e., random variables). MC simulations are considered more accurate than reliability-based methods as they compute the probability of failure of the whole system using more precise PDFs of all uncertain variables. However, the MC results’ accuracy significantly depends on the number of samples generated. The large number of samples is critical for reliable model results, leading to extensive computational costs in complex cases. Thus, several sampling methods, such as latent hypercube sampling and importance sampling, have been developed to decrease the required sampling number and, consequently, decrease the results’ variance (for more details about MC sampling methods, see [353,354]). (5) Tree logical diagrams, such as Event tree analysis (ETA), Fault tree analysis (FTA), and Bow-Tie analysis, are methods of failure risk analysis where the consequences of a single system component failure are identified and tracked through the whole system [355]. Once the dam system is broken down into individual components, the logic diagram can be determined based on asking questions such as “What happens if a component fails?” in ETA (down-top approach), “What are the causes of component failure” in FTA (top-down approach). By answering these questions, tree diagrams can link the failure to its controllers and identify the subsequent probability of system failure using joint probability distributions (For more details about logical tree diagrams, see [336]). (6) The Bayesian belief network (BBN), proposed by Pearl in 1988 [356], is a probabilistic reasoning network that uses the Bayes theorem and graph representation to determine the propagation of system uncertainties through its components [357]. Generally, a BN consists of nodes (vertices) connected by directional links (arcs) that represent the dependencies between the connected nodes. By defining the probability distribution of the basic nodes (i.e., independent variables), the network can represent the propagation of the uncertainties of such basic nodes through the represented system using the Bayes theorem. The advantage of the BBN over the basic tree diagrams is the ability to incorporate updated information and easily be linked to other simulation models to represent dam system behavior (for more details about the advantages of the BBN over the conventional logical tree diagrams, see [358,359,360,361]). However, recent studies adopted hybrid BBN models in dam risk analysis (e.g., Monte Carlo-BBN models [362]; Markov Chain-BBN models [363]) to overcome the drawbacks of the traditional BBN (e.g., static and acyclic representation). For more details about BBN and dynamic BBN, see [360,363].

4. Research Gaps

This section aims to first present the contribution of the identified seven topics to the literature (i.e., the distribution of the seven topics in the analyzed documents). Subsequently, using the qualitative and quantitative analyses stated in the previous sections, this section aims to discover the research gaps that are not well explored (i.e., blue oceans [23]) pertaining to dam and reservoir operational safety. Figure 7 shows the summation of the gamma values for each topic (the probability of each document per topic), representing each topic’s contribution to the analyzed articles. Figure 7 shows that Topic 1 (Optimization Modeling), Topic 3 (Flood Risk), and Topic 7 (Risk-based Assessment and Management) are the highest topics with gamma values, respectively. As such, these topics can be considered the top three topics adopted in the literature related to the remaining five Topics 2, 5, 6, and 4, which have relatively lower gamma values, respectively.
To give a more in-depth understanding, the study identified the following three research gaps as opportunities for future studies. These research gaps are identified by reviewing such massive literature that had been cited in Section 3, including the published research articles, review papers, and textbooks, whereas the results of the quantitative analysis for the most available publications in the field are also used to demonstrate the authors’ perspective about the relatively less-covered areas in the literature. Moreover, within the identified three research gaps, the study presents deeper discussions for such studies that consider the identified relatively less-covered research area, aiming to determine the gaps within such studies by reviewing them qualitatively.
The first research gap is related to the limitation of the modeling concepts adopted in dam operational safety assessment studies. Most adopted models lack multi-hazard interactions, where a primary hazard can trigger one or more secondary hazards within a spatiotemporal scale (for more information about multi-hazard interaction, see [364]). However, they relate dam operational failures to a single external extreme event, where the excepted hazard events are independent. This conclusion can be supported using the quantitative analysis results, where “Multi” or “Multihazard” words do not appear in any of the topic’s word clouds. On the other hand, “Flood” (i.e., single extreme event) is the second higher topic considered by the literature (i.e., Topic 3 (see Figure 7)). Even though, there are several studies (e.g., [365,366,367]) that showed that dam operational failure usually occurs due to a combination of internal (e.g., lack of maintenance, delay in dam access) or external (e.g., windstorm) disturbance events that may also be individually considered within the design envelope.
The second research gap pertains to the capability limitation of the modeling tools adopted in the available dam operational safety assessment studies. The quantitative analysis shows the relatively higher adoption of optimization models (Topic 2 is the highest contributor topic to the literature, see Figure 7) and the risk analysis models (Topic 7 is the third contributor topic to the literature, see Figure 7) compared to employing dynamic dam system simulations in dam operational safety literature. Regarding the risk analysis tools, most of the adopted traditional risk analysis tools (stated in Section 3.7 (e.g., FTA, ETA, B.N.)) are usually static and do not consider the dynamic behavior of dam operation factors (e.g., hydrologic variables, gates operation) [363,368]. This can also be proved from the quantitative analysis results, where the “dynamic” word stem was not one of the most frequent words that appeared in Topic 7-word cloud). Some recent studies tried to represent the dynamic characteristics to predict dam operation failure risk using, for example, the coupling between Monte Carlo simulations and risk analysis techniques (e.g., [362]), dynamic B.N. (e.g., [172]); however, such studies still suffer from significant limitations in terms of the computational cost. In addition, the developed dynamic risk analysis models still need a more general representation approach to fit the different inter-relationships between complex dam systems components (e.g., cyclic or/and acyclic relations), which is a typical feature in dam systems. Moreover, considering the recommendations (e.g., [96,100]) for the essential coupling between the optimization models and dynamic simulation models to capture the dynamic characteristics, in real-time, of dam systems (as discussed in Section 3.1); thus, various researchers (e.g., [365,367,369]) pointed to the importance of system approach in simulating dam system operations.
However, most of the developed dam system dynamic models usually decompose complex dam systems into more manageable systems to be simply analyzed, ignoring system component interactions [14,366]. Thus, the general description of the dam complex system simulation approach has been introduced by King et al., 2017 [365] (using the system dynamics tool) in the context of dam operational safety assessment. The complex system simulation can provide a holistic representation of dam physical and non-physical dam system components/modules and their interrelationships and, thus, can provide a more realistic and more in-depth dynamic system response to various operational scenarios. Complex dam system simulations are then evolved in dam system operation safety (e.g., [370,371]). Nonetheless, the proposed models still suffer from two main shortcomings. (1) The deterministic definition of system components and ignoring the stochastic representation of uncertainties within the system operation modeling. In addition, the limited number of studies that represent the probabilistic definition of dam system components only consider the aleatory uncertainties (e.g., inflow randomness) and the epistemic uncertainty (e.g., model parameters errors) with the lack of representation of deep uncertainties (e.g., human behavior, uncertain failure scenario) that may significantly impact dam operational safety. This can be inferred from the quantitative analysis, where the word “uncertanti” appeared only in Inflow Forecasting and Flood Risk topics, see Figure 4. For more details about aleatory, epistemic, and deep uncertainties, see [330]. (2) The models still need more development to consider the effect of future climate change on the designed dam operational rules. The coupling between the climate models and the complex dam system simulation model can estimate dam system response (in real-time) to multiple climate scenarios. Subsequently, the coupling approach can facilitate the development of more realistic adaptive dam operational rules to mitigate climate change risk.
The third research gap refers to the limitation of adopting the Risk-based Assessment and Management concept in dam operational safety assessment studies. Risk-based management is the third contributor topic to the literature, see Figure 7, whereas the “resilience” word stem does not appear as any of the frequent words within the analyzed documents. The drawbacks of risk-based management can be concluded: (1) Risk definition relies mainly on probabilities estimation of hazard, which is affected by the lack of information and subjective assigning of probabilities; (2) Risk typically combines the probability and severity of hazard events into a single value, which results in similar risk levels for different hazard severity classes (e.g., Low possible event with high severity and high possible event with low severity have the same risk overall value); (3) Risk-based management focused on the immediate response of the system under a specific hazard event; however, the resilience-based management focuses more on the dynamic system functionality (gain or reduction), system recovery and deterioration time (after hazard occurrence), resource allocations pre- and post-hazard realization, and restoration costs [372]; and (4) Bachear et al. (2016) [330] also highlighted the advantages of the resilient-guided dam system in dealing with deep uncertainty that is barely considered in dam safety assessment studies, where the resilience-based assessment ensures that the system can bounce back to its standard level after any hazard event.
As such, the focus of several infrastructure systems (e.g., power, transportation) design standards have been shifted to resilience-based management [373,374]. However, the word “resilience” does not appear in any of the word clouds presented in Figure 1 and Figure 4. Moreover, the limited number of studies introducing resilience-based assessment in dam systems and operational safety (e.g., [371,375]) has some limitations on dam resilience quantification. The developed approaches depend only on simplified dynamic models. However, each system component usually has a different response to the applied hazard, which significantly affects the system’s overall response (including system functionality deterioration and recovery rates) and subsequently affects the quantified system resilience. As a result, the more sophisticated the model is, the more realistic the quantified resilience value for such a system. In addition, the previous studies did not investigate the resilience of dam systems under multiple hazards (consecutive and/or intersected). To overcome such limitations, Badr et al., 2023 [376] introduced the development of a comprehensive resilience-centric system dynamics model for the quantification of dam systems’ dynamic resilience under multi-hazard environments. However, the proposed model was also deterministic and did not explicitly consider the various operational uncertainties and the probabilistic hazard behavior associated with dam system operation under normal and hazardous disruptive conditions. Taking into consideration that the probabilistic resilience quantification is more efficient and realistic for infrastructure systems safety assessment than the deterministic quantification approach [372]. As such, more studies are still required to develop more reliable approaches to provide more efficient resilience-based assessment plans for dam systems’ operational safety.

5. Conclusions

The complexity of dam systems-of-systems, the multi-objective nature of their operations, and their uncertain vulnerability to climatic change pose serious challenges to dam operational management. Notwithstanding the criticality of dam infrastructure to society, such challenges triggered extensive research studies to investigate dam operational safety from various perspectives. Unlike other studies that subjectively considered only specific topics and related documents, this paper meta-researches (quantitively and qualitatively) the entire published literature on dam and reservoir operational safety from 1984 to 2023. This meta-research aims to identify and classify the major topics and subsequently uncover the latent topics in dams and reservoir operational safety. Subsequently, by examining such latent topics qualitatively, the key research gaps are identified as opportunities for future studies. More specifically, using the textual topic modeling, seven key topics are identified from analyzing 871 identified relevant journal articles. Such topics are optimization models; climate change; flood risk; inflow forecasting; hydropower generation; water supply management; and risk-based assessment and management. The contribution of such topics in the pertaining literature is subsequently presented, highlighting that: optimization models, flood risk, and risk-based assessment and management are more relatively considered in the literature than the remaining five topics. The importance of the quantitative analysis (text analytics) is not only limited to identifying the seven topics presented in Section 2, and providing the contribution of each of the seven topics to the literature, as shown in Figure 7; however, the text analytics key value shows up in Section 4, where the authors used the quantitative analysis results to support the authors’ perspectives for the potential research gaps. In addition, the qualitative review of these topics cited 376 research to identify the definitions, concepts, modeling tools, and recent research trends. Using quantitative and qualitative analyses, the study identified three major research gaps in dam safety assessment studies related to the limitations in both modeling concepts and modeling tools, as well as the lack of developing and adopting resilience-guided management strategies. Overall, this study presents a road map of dam and reservoir operational safety literature and associated knowledge gaps. The study also recommended future research directions on dam systems and reservoir operational safety to ensure the resilience of such important infrastructure and water resources in our changing climate.

Author Contributions

Conceptualization and Methodology, A.B., Z.L. and W.E.-D.; Software, Formal analysis, Visualization, Investigation, Data curation, and Writing-original draft preparation, A.B.; Writing—review and editing, and Supervision, Z.L. and W.E.-D.; Funding acquisition, A.B., Z.L. and W.E.-D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Natural Sciences and Engineering Research Council grant number [CREATE/482707-2016], and the Vanier Canada Graduate Scholar (Vanier-CGS) awarded to the corresponding author.

Data Availability Statement

All data, models, and code generated or used during the study appear in the submitted article.

Acknowledgments

The authors would like to acknowledge the fruitful discussions with the research teams of the NSERC-CaNRisk-CREATE program, the INTERFACE Institute, and the INViSiONLab, McMaster University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Herzog, M.A.M. Practical Dam Analysis; Thomas Telford Publishing: London, UK, 1998. [Google Scholar] [CrossRef]
  2. Zhang, J.; Shang, Y. Nexus of dams, reservoirs, climate, and the environment: A systematic perspective. Int. J. Environ. Sci. Technol. 2023. [Google Scholar] [CrossRef]
  3. Ma, C.; Xu, X.; Yang, J.; Cheng, L. Safety Monitoring and Management of Reservoir and Dams. Water 2023, 15, 1078. [Google Scholar] [CrossRef]
  4. Tajziehchi, S.; Karbassi, A.; Nabi, G.; Yoo, C.; Ifaei, P. A cost-benefit analysis of Bakhtiari hydropower dam considering the nexus between energy and water. Energies 2022, 15, 871. [Google Scholar] [CrossRef]
  5. Zohuri, B. Nuclear fuel cycle and decommissioning. In Nuclear Reactor Technology Development and Utilization; Woodhead Publishing: Cambridge, UK, 2020; pp. 61–120. [Google Scholar] [CrossRef]
  6. Boretti, A.; Rosa, L. Reassessing the projections of the World Water Development Report. npj Clean Water 2019, 2, 15. [Google Scholar] [CrossRef]
  7. Berga, L. The role of hydropower in climate change mitigation and adaptation: A review. Engineering 2016, 2, 313–318. [Google Scholar] [CrossRef]
  8. Fluixá-Sanmartín, J.; Escuder-Bueno, I.; Morales-Torres, A.; Castillo-Rodríguez Jesica, T. Accounting for climate change uncertainty in Long-Term dam risk management. J. Water Resour. Plan. Manag. 2021, 147, 04021012. [Google Scholar] [CrossRef]
  9. UNEP. The Future We Want. In Proceedings of the Outcome Document of the United Nations Conference on Sustainable Development, Rio, Brazil, 20–22 June 2012; Available online: https://www.un.org/ga/search/view_doc.asp?symbol=A/RES/66/288&Lang=E (accessed on 8 October 2021).
  10. Zarfl, C.; Lumsdon, A.; Tockner, K. A global boom in hydropower dam construction. Aquat. Sci. 2015, 77, 161–170. [Google Scholar] [CrossRef]
  11. Homeland Security, Cybersecurity and Infrastructure Security Agency. National Infrastructure Protection Plan (N.I.P.P.) Dams Sector-Estimating Loss of Life for Dam Failures Scenarios. 2011. Available online: https://damsafety.org/sites/default/files/files/DamsSectorConsequenceEstimation_LossOfLife.pdf (accessed on 7 October 2021).
  12. United Nations World Water Development Report (UN-WWDR). Water and Climate Change. 2020. Available online: https://unesdoc.unesco.org/ark:/48223/pf0000372985.locale=en (accessed on 7 October 2021).
  13. Ehsani, N.; Vörösmarty, C.J.; Fekete, B.M.; Stakhiv, E.Z. Reservoir operations under climate change: Storage capacity options to mitigate risk. J. Hydrol. 2017, 555, 435–446. [Google Scholar] [CrossRef]
  14. Regan, P.J. Dams as systems—A holistic approach to dam safety. In Proceedings of the U.S.S.D. Annual Meeting and Conference 2010, Sacramento, CA, USA, 12–16 April 2010; pp. 1307–1340. [Google Scholar]
  15. Kumar, K.; Saini, R.P. A review of operation and maintenance of hydropower plants. Sustain. Energy Technol. Assess. 2022, 49, 101704. [Google Scholar] [CrossRef]
  16. Hauera, C.; Wagnera, B.; Aignera, J.; Holzapfela, P.; Flödla, P.; Liedermanna, M.; Trittharta, M.; Sindelara, C.; Pulgb, U.; Klöscha, M.; et al. State of the art, shortcomings and future challenges for a sustainable sediment management in hydropower: A review. Renew. Sustain. Energy 2018, 98, 40–55. [Google Scholar] [CrossRef]
  17. Neboh, N.; Adeyemo, J.; Enitan, A.; Olugbara, O. A review of applications of evolutionary algorithms to reservoir operation for hydropower production. Int. J. Geol. Environ. Eng. 2015, 9, 1153–1159. Available online: https://zenodo.org/record/1109762#.YvAJqHbMK38 (accessed on 1 December 2022).
  18. You, L.; Li, C.; Min, X.; Xiaolei, T. Review of dam-break research of earth-rock dam combined with dam safety management. Procedia Eng. 2012, 28, 382–388. [Google Scholar] [CrossRef]
  19. Rani, D.; Moreira, M.M. Simulation–Optimization modeling: A survey and potential application in reservoir systems operation. Water Resour. Manag. 2010, 24, 1107–1138. [Google Scholar] [CrossRef]
  20. Sun, L.; Yin, Y. Discovering themes and trends in transportation research using topic modeling. Transp. Res. Part C Emerg. Technol. 2017, 77, 49–66. [Google Scholar] [CrossRef]
  21. Gatti, C.; Brooks, J.; Nurre, S. A historical analysis of the field of OR/MS using topic models. arXiv 2015, arXiv:1510.05154. [Google Scholar] [CrossRef]
  22. Haggag, M.; Ezzeldin, M.; El-Dakhakhni, W.; Hassini, E. Resilient cities critical infrastructure interdependence: Meta-research. Sustain. Resilient Infrastruct. 2020, 7, 291–312. [Google Scholar] [CrossRef]
  23. Ezzeldin, M.; El-Dakhakhni, W. Meta-researching structural engineering: Trend identification and knowledge gap discovery using text mining. J. Struct. Eng. 2020, 146, 1–22. [Google Scholar] [CrossRef]
  24. Islam, A.; Hassini, S.; El-Dakhakhni, W. A systematic bibliometric review of optimization and resilience within low impact development stormwater management practices. J. Hydrol. 2021, 599, 126457. [Google Scholar] [CrossRef]
  25. Das, S.; Sun, X.; Dutta, A. Text mining and topic modeling of compendiums of papers from transportation research board annual meetings. J. Transp. Res. Board 2016, 2552, 48–56. [Google Scholar] [CrossRef]
  26. Griffiths, T.L.; Steyvers, M. Finding scientific topics. Proc. Natl. Acad. Sci. USA 2004, 101, 5228–5235. [Google Scholar] [CrossRef]
  27. Salloum, S.A.; Al-Emran, M.; Abdel Monem, A.; Shaalan, K. Using text mining techniques for extracting information from research articles. In Intelligent Natural Language Processing: Trends and Applications; Springer: Cham, Switzerland, 2018; pp. 373–397. [Google Scholar] [CrossRef]
  28. Miner, G.; Elder, J.; Hill, T.; Delen, D.; Fast, A.; Nisbet, R.A. Practical Text Mining and Statistical Analysis for Non-Structured Text Data Applications; Academic Press: Cambridge, MA, USA, 2012. [Google Scholar] [CrossRef]
  29. Blei, D.M.; Ng, A.Y.; Jordan, M.I. Latent Dirichlet allocation. J. Mach. Learn. Res. 2003, 3, 993–1022. [Google Scholar]
  30. Hofmann, T. Probabilistic latent semantic indexing. In Proceedings of the 22nd Annual International ACM SIGIR Conference on Research and Development in Information Retrieval (S.I.G.I.R. ‘99), A.C.M., Berkeley, CA, USA, 15–19 August 1999; pp. 50–57. [Google Scholar] [CrossRef]
  31. Amado, A.; Cortez, P.; Rita, P.; Moro, S. Research trends on big data in marketing: A text mining and topic modeling-based literature analysis. Eur. Res. Manag. Bus. Econ. 2018, 24, 1–7. [Google Scholar] [CrossRef]
  32. Steyvers, M.; Griffiths, T. Probabilistic topic models. In Latent Semantic Analysis: A Road to Meaning; Landauer, T., McNamara, D., Dennis, S., Kintsch, W., Eds.; Laurence Erlbaum: Hillsdale, NJ, USA, 2007; Volume 427, pp. 424–440. [Google Scholar]
  33. Blei, D.M. Probabilistic topic models. Commun. ACM 2012, 55, 77–84. [Google Scholar] [CrossRef]
  34. Minka, T. Estimating a Dirichlet distribution. In Technical Rep.; Massachusetts Institute of Technology: Cambridge, MA, USA, 2000. [Google Scholar]
  35. Geman, S.; Geman, D. Stochastic relaxation, Gibbs distributions, and the Bayesian restoration of images. IEEE Trans. Pattern Anal. Mach. Intell. 1984, PAMI-6, 721–741. [Google Scholar] [CrossRef] [PubMed]
  36. Fayaed, S.S.; El-Shafie, A.; Jaafar, O. Reservoir-system simulation and optimization techniques. Stoch. Environ. Res. Risk Assess. 2013, 27, 1751–1772. [Google Scholar] [CrossRef]
  37. Taghian, M.; Rosbjerg, D.; Haghighi, A.; Madsen, H. Optimization of conventional rule curves coupled with hedging rules for reservoir operation. J. Water Resour. Plan. Manag. 2014, 140, 693–698. [Google Scholar] [CrossRef]
  38. Parvez, I.; Shen, J.; Hassan, I.; Zhang, N. Generation of hydro energy by using a data mining algorithm for the cascaded hydropower plant. Energies 2021, 14, 298. [Google Scholar] [CrossRef]
  39. Bozorg-Haddad, O.; Aboutalebi, M.; Ashofteh, P.S. Real-time reservoir operation using data mining techniques. Environ. Monit. Assess. 2018, 190, 594. [Google Scholar] [CrossRef]
  40. Raju, K.; Chandre, G.C.; Karthika, B.S. Optimization of reservoir operation using linear programming. Int. J. Recent Technol. Eng. 2020, 8, 2277–3878. [Google Scholar] [CrossRef]
  41. Ginting, B.M.; Harlan, D.; Taufik, A. Optimization of reservoir operation using the linear program, a case study of Riam Jerawi Reservoir, Indonesia. Int. J. River Basin Manag. 2017, 15, 187–198. [Google Scholar] [CrossRef]
  42. Heydari, M.; Othman, F.; Qaderi, K. Developing optimal reservoir operation for multiple and multipurpose reservoirs using mathematical programming. Math. Probl. Eng. 2015, 2015, 435752. [Google Scholar] [CrossRef]
  43. Needham, J.T.; David, W.; Watkins, D.W.; Lund, J.R.; Nanda, S.K. Linear programming for flood control in the Iowa and Des Moines rivers. Water Resour. Plan. Manag. 2000, 126, 118–127. [Google Scholar] [CrossRef]
  44. Houck, M.H.; Cohon, J.L. Sequential explicitly stochastic linear programming models: A proposed method for design and management of multi-purpose reservoir system. Water Resour. Res. 1978, 14, 161–168. [Google Scholar] [CrossRef]
  45. Wu, X.; Wu, Y.; Cheng, X.; Chuntian Cheng Li, Z.; Wu, Y. A mixed-integer linear programming model for hydro unit commitment considering operation constraint priorities. Renew. Energy 2023, 204, 507–520. [Google Scholar] [CrossRef]
  46. Dogan, M.S.; Lund, J.R.; Josue Medellin-Azuara, J.M. Hybrid linear and nonlinear programming model for hydropower reservoir optimization. J. Water Resour. Plan. Manag. 2021, 147, 06021001. [Google Scholar] [CrossRef]
  47. Arunkumar, R.; Jothiprakash, V. Optimal reservoir operation for hydropower generation using a non-linear programming model. J. Inst. Eng. Ser. A 2012, 93, 111–120. [Google Scholar] [CrossRef]
  48. Gu, J.J.; Guo, P.; Huang, G.H.; Shen, N. Optimization of the industrial structure facing sustainable development in resource-based cities subjected to water resources under uncertainty. Stoch. Environ. Res. Risk Assess. 2012, 27, 659–673. [Google Scholar] [CrossRef]
  49. Simonovic, S.P.; Marino, M.A. Reliability programming in reservoir management, single multipurpose reservoir. Water Resour. Res. 1980, 16, 844–848. [Google Scholar] [CrossRef]
  50. Nandalal, K.D.W.; Bogardi, J.J. Dynamic Programming-Based Operation of Reservoirs: Applicability and Limits; Cambridge University Press: Cambridge, UK, 2007. [Google Scholar] [CrossRef]
  51. Labadie, J.W. Optimal operation of multi-reservoir systems: A state-of-the-art review. J. Water Resour. Plan. Manag. 2004, 130, 93–111. [Google Scholar] [CrossRef]
  52. Shim, K.-C.; Fontane, D.; Labadie, J. Spatial decision support system for integrated river basin flood control. J. Water Resour. Plan. Manag. 2002, 128, 190–201. [Google Scholar] [CrossRef]
  53. Yakowitz, S. Dynamic programming applications in water resources. Water Resour. Res. 1982, 18, 673–696. [Google Scholar] [CrossRef]
  54. Bellman, R. Dynamic Programming; Princeton University Press: Princeton, NJ, USA, 1957; ISBN 0-691-07951-X. [Google Scholar]
  55. Mizyed, N.; Loftis, J.; Fontane, D. Operation of large multi-reservoir systems using optimal-control theory. J. Water Resour. Plan. Manag. 1992, 118, 371–387. [Google Scholar] [CrossRef]
  56. Pontryagin, L.; Boltyanskii, V.; Gamdrelidze, R.; Mishchenko, E. The Mathematical Theory of Optimal Processes; Interscience: New York, NY, USA, 1962. [Google Scholar]
  57. Chen, C.; Kang, C.; Wang, J. Stochastic linear programming for reservoir operation with constraints on reliability and vulnerability. Water 2018, 10, 175. [Google Scholar] [CrossRef]
  58. Lee, Y.; Kim, S.K.K.; Ko, I.H. Two-stage stochastic linear programming model for coordinated multi-reservoir operation. In Proceedings of the Operations Management 2006: Operating Reservoirs in Changing Conditions, Sacramento, CA, USA, 14–16 August 2006; Zimbelman, D., Loehlein, W.C., Eds.; American Society of Civil Engineers: Reston, VA, USA, 2006; pp. 400–410, ISBN 978-0-7844-0875-9. [Google Scholar] [CrossRef]
  59. Seifi, A.; Hipel, K. Interior-point method for reservoir operation with stochastic inflows. J. Water Resour. Plan. Manag. 2001, 127, 48–57. [Google Scholar] [CrossRef]
  60. Juárez ML, A.; Ramírez, R.M.; Mora, R.D.; Reséndiz, A.M.; Esteban, R.V.; De la Cruz Courtois, O.A.; Elizondo, E.C. Guide curves in stochastic dynamic programming to optimize operation policies in sequential hydropower dams. Stoch. Environ. Res. Risk Assess. 2023. [Google Scholar] [CrossRef]
  61. Fayaed, S.S.; Fiyadh, S.S.; Khai, W.J.; Ahmed, A.N.; Afan, H.A.; Ibrahim, R.K.; Fai, C.M.; Koting, S.; Mohd, N.S.; Jaafar, W.Z.B.; et al. Improving dam and reservoir operation rules using stochastic dynamic programming and artificial neural network integration model. Sustainability 2019, 11, 5367. [Google Scholar] [CrossRef]
  62. Liu, P.; Zhao, J.; Li, L.; Shen, Y. Optimal reservoir operation using stochastic dynamic programming. J. Water Resour. Prot. 2012, 4, 342–345. [Google Scholar] [CrossRef]
  63. Tilmant, A.; Faouzi, E.H.; Vanclooster, M. Optimal operation of multi-purpose reservoirs using flexible stochastic dynamic programming. Appl. Soft Comput. 2002, 2, 61–74. [Google Scholar] [CrossRef]
  64. Azaiez, M.N.; Hariga, M.; Al-Harkan, I. A chance-constrained multi-period model for a special multi-reservoir system. Comput. Oper. Res. 2005, 32, 1337–1351. [Google Scholar] [CrossRef]
  65. Ouarda, T.; Labadie, J. Chance-constrained optimal control for multi-reservoir system optimization and risk analysis. Stoch. Environ. Res. Risk Assess. 2001, 15, 185–204. [Google Scholar] [CrossRef]
  66. Kitanidis, P.K. A first-order approximation to stochastic optimal control of reservoirs. Stoch. Hydrol. Hydraul. 1987, 1, 169–184. [Google Scholar] [CrossRef]
  67. Mezenner, N.; Dechemi, N.; Bermad, A.; Benkaci, T. Optimized reservoir operation using genetic algorithm and simulated inflows to reservoir based principal components analysis: The case of Cheffia reservoir—Algeria. Model. Earth Syst. Environ. 2023. [Google Scholar] [CrossRef]
  68. Chandra, V.S.S.; Hareendran, S.A.; Sankar, S.S. Optimal reservoir optimization using multiobjective genetic algorithm. In Advances in Swarm Intelligence. I.C.S.I. 2020; Tan, Y., Shi, Y., Tuba, M., Eds.; Lecture Notes in Computer Science 12145; Springer: Cham, Switzerland, 2020. [Google Scholar] [CrossRef]
  69. Olukanni, D.O.; Adejumo, T.A.; Salami, A.W.; Adedeji, A.A.; Alavi, A.H. Optimization-based reliability of a multipurpose reservoir by genetic algorithms: Jebba hydropower dam, Nigeria. Cogent Eng. 2018, 5, 143874. [Google Scholar] [CrossRef]
  70. Al-Aqeeli, Y.H.; Lee, T.S.; Abd Aziz, S. Enhanced genetic algorithm optimization model for a single reservoir operation based on hydropower generation: A case study of Mosul reservoir, northern Iraq. SpringerPlus 2016, 5, 797. [Google Scholar] [CrossRef]
  71. Sharif, M.; Wardlaw, R. Multireservoir systems optimization using genetic algorithms: Case study. J. Comput. Civ. Eng. 2000, 14, 255–263. [Google Scholar] [CrossRef]
  72. Kumar, D.N.; Reddy, M.J. Ant colony optimization for multi-purpose reservoir operation. Water Resour. Manag. 2006, 20, 879–898. [Google Scholar] [CrossRef]
  73. Al-Aqeeli, Y.H.; Mahmood Agha, O.M. Optimal operation of a multi-reservoir system for hydropower production using particle swarm optimization algorithm. Water Resour. Manag. 2020, 34, 3099–3112. [Google Scholar] [CrossRef]
  74. Moradi, A.M.; Dariane, A.B. Particle swarm optimization: Application to reservoir operation problems. In Proceedings of the IEEE International Advance Computing Conference, Patiala, India, 6–7 March 2009; pp. 1048–1051. [Google Scholar] [CrossRef]
  75. Georgiou, P.E.; Papamichail, D.M.; Vougioukas, S.G. Optimal irrigation reservoir operation and simultaneous multi-crop cultivation area selection using simulated annealing. Irrig. Drain. 2006, 55, 129–144. [Google Scholar] [CrossRef]
  76. Teegavarapu, R.S.; Simonovic, S.P. Optimal operation of reservoir systems using simulated annealing. Water Resour. Manag. 2002, 16, 401–428. [Google Scholar] [CrossRef]
  77. Haddad, O.B.; Afshar, A.; Mariño, M.A. Honeybees mating optimization (H.B.M.O.) algorithm: A new heuristic to roach for water resources optimization. J. Water Resour. Manag. 2006, 20, 661–680. [Google Scholar] [CrossRef]
  78. Beiranvand, B.; Ashofteh, P.S. A Systematic Review of Optimization of Dams Reservoir Operation Using the Meta-heuristic Algorithms. Water Resour. Manag. 2023, 37, 3457–3526. [Google Scholar] [CrossRef]
  79. Sharifi, M.R.; Akbarifard, S.; Qaderi, K. Comparative analysis of some evolutionary-based models in optimization of dam reservoir operation. Sci. Rep. 2021, 11, 15611. [Google Scholar] [CrossRef] [PubMed]
  80. Azad, A.S.; Rahaman Md, S.A.; Watada, J.; Vasant, P.; Vintaned, J.A.G. Optimization of the hydropower energy generation using meta-heuristic approaches: A review. Energy Rep. 2020, 6, 2230–2248. [Google Scholar] [CrossRef]
  81. Ahmad, A.; El-Shafie, A.; Razali, S.F.M.; Mohamad, Z.S. Reservoir optimization in water resources: A review. Water Resour. Manag. 2014, 28, 3391–3405. [Google Scholar] [CrossRef]
  82. Aggarwal, C.C. Neural Networks and Deep Learning: A Textbook; Springer International Publishing: Berlin/Heidelberg, Germany, 2018. [Google Scholar] [CrossRef]
  83. Zhanga, D.; Lina, J.; Penga, Q.; Wangb, D.; Yang, T.; Sorooshian, S.; Liua, X.; Zhuang, J. Modeling and simulating reservoir operation using the artificial neural network, support vector regression, and deep learning algorithm. J. Hydrol. 2018, 565, 720–736. [Google Scholar] [CrossRef]
  84. Moorthi, P.V.P.; Singh, A.P.; Agnivesh, P. Regulation of water resources systems using fuzzy logic: A case study of Amaravathi dam. Appl. Water Sci. 2018, 8, 132. [Google Scholar] [CrossRef]
  85. Zadeh, L.A. Fuzzy sets. Inf. Control 1965, 8, 338–353. [Google Scholar] [CrossRef]
  86. Bellman, R.E.; Zadeh, L.A. Decision-making in a fuzzy environment. Manag. Sci. 1970, 17, B-141–B-273. [Google Scholar] [CrossRef]
  87. Mousavi, S.J.; Karamouz, M.; Menhadj, M.B. Fuzzy-State Stochastic Dynamic Programming for Reservoir Operation. J. Water Resour. Plan. Manag. 2004, 130, 460–470. [Google Scholar] [CrossRef]
  88. Dehghani, M.; Riahi-Madvar, H.; Hooshyaripor, F.; Mosavi, A.; Shamshirband, S.; Zavadskas, E.K.; Chau, K.-W. Prediction of Hydropower Generation Using Grey Wolf Optimization AdaptiveNeuro-Fuzzy Inference System. Energies 2019, 12, 289. [Google Scholar] [CrossRef]
  89. Trillas, E.; Eciolaza, L. Fuzzy Logic, An Introductory Course for Engineering Students; Springer: Cham, Switzerland, 2015. [Google Scholar] [CrossRef]
  90. Ferdowsi, A.; Singh, V.P.; Ehteram, M.; Mirjalili, S. Multi-objective optimization approaches for design, planning, and management of water resource systems. In Essential Tools for Water Resources Analysis, Planning, and Management; Bozorg-Haddad, O., Ed.; Springer Water: Singapore, 2021. [Google Scholar] [CrossRef]
  91. Goicoechea, A.; Hansen, D.; Duckstein, L. Multiobjective Decision Analysis with Engineering and Business Applications; Wiley: New York, NY, USA, 1982. [Google Scholar]
  92. Cohon, J.L. Multiobjective Programming and Planning; Academic Press: New York, NY, USA, 1978. [Google Scholar]
  93. Zhu, F.; Zhong, P.-A.; Xu, B.; Chen, J.; Sun, Y.; Liu, W.; Li, T. Stochastic multi-criteria decision-making based on stepwise weight information for real-time reservoir operation. J. Clean. Prod. 2020, 257, 120554. [Google Scholar] [CrossRef]
  94. Zhu, F.; Zhong, P.-A.; Sun, Y.; Xu, B. Selection of criteria for multi-criteria decision-making of reservoir flood control operation. J. Hydro Inform. 2017, 19, 558–571. [Google Scholar] [CrossRef]
  95. Malekmohammadi, B.; Zahraie, B.; Kerachian, R. Ranking solutions of multi-objective reservoir operation optimization models using multi-criteria decision analysis. Expert Syst. Appl. 2011, 38, 7851–7863. [Google Scholar] [CrossRef]
  96. Wu, X.; Li, S.; Cheng, C.; Miao, S. Simulation-Optimization model to derive operation rules of multiple cascaded reservoirs for Nash equilibrium. J. Water Resour. Plan. Manag. 2019, 145, 04019013. [Google Scholar] [CrossRef]
  97. Jamali, S.; Jamali, B. Cascade hydropower systems optimal operation: Implications for Iran’s Great Karun hydropower systems. Appl. Water Sci. 2019, 9, 66. [Google Scholar] [CrossRef]
  98. Niewiadomska-Szynkiewicz, E.; Blaszczyk, J. Simulation-based optimization methods applied to large-scale water systems control. In Proceedings of the Intl IEEE Conferences on Ubiquitous Intelligence and Computing, Advanced and Trusted Computing, Scalable Computing and Communications, Cloud and Big Data Computing, Internet of People, and Smart World Congress (UIC/ATC/ScalCom/CBDCom/IoP/SmartWorld), Toulouse, France, 18–21 July 2016; pp. 649–656. [Google Scholar] [CrossRef]
  99. Chaturvedi, M.C.; Srivastava, D.K. Study of complex water-resources system with screening and simulation models. Water Resour. Res. 1981, 17, 783–794. [Google Scholar] [CrossRef]
  100. Anand, J.; Gosain, A.K.; Khosa, R. Optimization of multipurpose reservoir operation by coupling soil and water assessment tool (SWAT) and genetic algorithm for optimal operating policy (Case study: Ganga River Basin). Sustainability 2018, 10, 1660. [Google Scholar] [CrossRef]
  101. Suiadee, W.; Tingsanchali, T. A combined simulation-genetic algorithm optimization model for optimal rule curves of a reservoir: A case study of the Nam on Irrigation Project, Thailand. Hydrol. Process 2007, 21, 3211–3225. [Google Scholar] [CrossRef]
  102. Lai, V.; Huang, Y.F.; Koo, C.H.; Ahmed, A.N.; El-Shafie, A. A Review of Reservoir Operation Optimisations: From Traditional Models to Metaheuristic Algorithms. Arch Comput. Methods Eng 2022, 29, 3435–3457. [Google Scholar] [CrossRef]
  103. Saab, S.M.; Othman, F.B.; Tan, C.G.; Allawi, M.F.; El-Shafie, A. Review of generating optimal operation for dam and reservoir water system: Simulation models and optimization algorithms. Appl. Water Sci. 2022, 12, 73. [Google Scholar] [CrossRef]
  104. Dobson, B.; Wagener, T.; Pianosi, F. An argument-driven classification and comparison of reservoir operation optimization methods. Adv. Water Resour. 2019, 128, 74–86. [Google Scholar] [CrossRef]
  105. Macian-Sorribes, H.; Pulido-Velazquez, M. Inferring efficient operating rules in multi-reservoir water resource systems: A review. WIREs Water 2020, 7, e1400. [Google Scholar] [CrossRef]
  106. Giuliani, M.; Lamontagne, J.R.; Reed, P.M.; Castelletti, A. A state-of-the-art review of optimal reservoir control for managing conflicting demands in a changing world. Water Resour. Res. 2021, 57, e2021WR029927. [Google Scholar] [CrossRef]
  107. Dalcin, A.P.; Brêda, J.; Marques, G.; Tilmant, A.; de Paiva, R.; Kubota, P. The Role of Reservoir Reoperation to Mitigate Climate Change Impacts on Hydropower and Environmental Water Demands. J. Water Resour. Plan. Manag. 2023, 149, 04023005. [Google Scholar] [CrossRef]
  108. Aljoda, A.; Jain, S. Uncertainties and risks in reservoir operations under changing hydroclimatic conditions. J. Water Clim. Change 2021, 12, 1708–1723. [Google Scholar] [CrossRef]
  109. Zhong, W.; Guo, J.; Chen, L.; Zhou, J.; Zhang, J.; Wang, D. Future hydropower generation prediction of large-scale reservoirs in the upper Yangtze River basin under climate change. J. Hydrol. 2020, 588, 125013. [Google Scholar] [CrossRef]
  110. Liu, B.; Lund, J.R.; Liu, L.; Liao, S.; Li, G.; Cheng, C. Climate change impacts on hydropower in Yunnan, China. Water 2020, 12, 197. [Google Scholar] [CrossRef]
  111. Fluixá-Sanmartín, J.; Morales-Torres, A.; Escuder-Bueno, I. Quantification of climate change impact on dam failure risk under hydrological scenarios: A case study from a Spanish dam. Nat. Hazards Earth Syst. Sci. 2019, 19, 2117–2139. [Google Scholar] [CrossRef]
  112. Zolghadr-Asli, B.; Bozorg-Haddad, O.; Chu, X. Effects of the uncertainties of climate change on the performance of hydropower systems. J. Water Clim. Change 2019, 10, 591–609. [Google Scholar] [CrossRef]
  113. Boehlert, B.; Strzepek, K.M.; Gebretsadik, Y.; Swanson, R.; McCluskey, A.; Neumann, J.E.; McFarland, J.; Martinich, J. Climate change impacts and greenhouse gas mitigation effects on U.S. hydropower generation. Appl. Energy 2016, 183, 1511–1519. [Google Scholar] [CrossRef]
  114. Fluixá-Sanmartín, J.; Altarejos-García, L.; Morales-Torres, A.; Escuder-Bueno, I. Review article: Climate change impacts on dam safety. Nat. Hazards Earth Syst. Sci. 2018, 18, 2471–2488. [Google Scholar] [CrossRef]
  115. Watts, R.J.; Richter, B.D.; Opperman, J.J.; Bowmer, K.H. Dam reoperation in an era of climate change. Mar. Freshw. Res. 2011, 62, 321–327. [Google Scholar] [CrossRef]
  116. Li, W.; Li, Z.; Ge, W.; Wu, S. Risk evaluation model of life loss caused by dam-break flood and its application. Water 2019, 11, 1359. [Google Scholar] [CrossRef]
  117. Chernet, H.; Knut Alfredsen, M.; Midttømme, G. Safety of hydropower dams in a changing climate. J. Hydraul. Eng. 2014, 19, 569–582. [Google Scholar] [CrossRef]
  118. Ahmadi, M.; Bozorg Haddad, O.; Loáiciga, H.A. Adaptive reservoir operation rules under climatic change. Water Resour. Manag. 2014, 29, 1247–1266. [Google Scholar] [CrossRef]
  119. Lompi, M.; Mediero, L.; Soriano, E.; Caporali, E. Climate change and hydrological dam safety: A stochastic methodology based on climate projections. Hydrol. Sci. J. 2023, 68, 745–763. [Google Scholar] [CrossRef]
  120. Nourani, V.; Rouzegaria, N.; Molajouc, A.; Baghanama, A.H. An integrated simulation-optimization framework to optimize the reservoir operation adapted to climate change scenarios. J. Hydrol. 2020, 587, 125018. [Google Scholar] [CrossRef]
  121. Tukimat, N.N.A.; Harun, S. Comparative study on the reservoir operation planning with the climate change adaptation. SN Appl. Sci. 2019, 1, 1449. [Google Scholar] [CrossRef]
  122. Chang, J.; Wang, X.; Li, Y.Y.; Wang, Y.; Zhang, H. Hydropower plant operation rules optimization response to climate change. Energy 2018, 160, 886–897. [Google Scholar] [CrossRef]
  123. Ehteram, M.; Mousavi, S.F.; Karami, H.; Farzin, S.; Singh, V.P.; Chau, K.-W.; El-Shafie, A. Reservoir operation based on evolutionary algorithms and multi-criteria decision-making under climate change and uncertainty. J. Hydro Inform. 2018, 20, 332–355. [Google Scholar] [CrossRef]
  124. Yang, G.; Guo, S.; Li, L.; Hong, X.; Wang, L. Multi-objective operating rules for Danjiangkou reservoir under climate change. Water Resour. Manag. 2016, 30, 1183–1202. [Google Scholar] [CrossRef]
  125. Zhou, Y.; Guo, S. Incorporating ecological requirements into multipurpose reservoir operating rule curves for adaptation to climate change. J. Hydrol. 2013, 498, 153–164. [Google Scholar] [CrossRef]
  126. Forrest, K.; Tarroja, B.; Chiang, F.; AghaKouchak, A.; Samuelsen, S. Assessing climate change impacts on California hydropower generation and ancillary services provision. Clim. Change 2018, 151, 395–412. [Google Scholar] [CrossRef]
  127. Anaraki, M.V.; Farzin, S.; Mousavi, S.F.; Karami, H. Uncertainty analysis of climate change impacts on flood frequency by using hybrid machine learning methods. Water Resour. Manag. 2021, 35, 199–223. [Google Scholar] [CrossRef]
  128. Wang, J.; Fonseca, R.M.; Rutledge, K. A hybrid statistical-dynamical downscaling of air temperature over Scandinavia using the W.R.F. Model. Adv. Atmos. Sci. 2020, 37, 57–74. [Google Scholar] [CrossRef]
  129. Tang, J.; Niu, X.; Wang, S.; Gao, H.; Wang, X.; Wu, J. Statistical downscaling and dynamical downscaling of regional climate in China: Present climate evaluations and future climate projections. J. Geophys. Res. Atmos. 2016, 121, 2110–2129. [Google Scholar] [CrossRef]
  130. Xu, Z.; Han, Y.; Yang, Z. Dynamical downscaling of regional climate: A review of methods and limitations. Sci. China Earth Sci. 2019, 62, 365–375. [Google Scholar] [CrossRef]
  131. Tabari, H.; Paz, S.M.; Buekenhout, D.; Willems, P. Comparison of statistical downscaling methods for climate change impact analysis on precipitation-driven drought. Hydrol. Earth Syst. Sci. 2021, 25, 3493–3517. [Google Scholar] [CrossRef]
  132. Trzaska, S.; Schnarr, E. A Review of Downscaling Methods for Climate Change Projections; Center for International Earth Science, Information Network (C.I.E.S.I.N.): Burlington, VT, USA, 2014. [Google Scholar]
  133. Dang, T.D.; Vu, D.T.; Chowdhury, A.F.M.K.; Galelli, S. A software package for the representation and optimization of water reservoir operations in the V.I.C. hydrologic model. Environ. Model. Softw. 2020, 126, 104673. [Google Scholar] [CrossRef]
  134. Dang, T.D.; Vu, D.T.; Chowdhury, A.F.M.K.; Galelli, S. On the representation of water reservoir storage and operations in large-scale hydrological models: Implications on model parameterization and climate change impact assessments. Hydrol. Earth Syst. Sci. 2020, 24, 397–416. [Google Scholar] [CrossRef]
  135. Beharry, S.L.; Gabriels, D.; Lobo, D.; Ramsewak, D.; Clarke, R.M. Use of the SWAT model for estimating reservoir volume in the Upper Navet watershed in Trinidad. Appl. Sci. 2021, 3, 163. [Google Scholar] [CrossRef]
  136. Shrestha, J.P.; Pahlow, M.; Cochrane, T.A. Development of a SWAT hydropower operation routine and its application to assessing hydrological alterations in the Mekong. Water 2020, 12, 2193. [Google Scholar] [CrossRef]
  137. Wilby, R.L.; Dessai, S. Robust adaptation to climate change. Weather 2010, 65, 180–185. [Google Scholar] [CrossRef]
  138. Brown, C.; Ghile, Y.; Laverty, M.; Li, K. Decision scaling: Linking bottom-up vulnerability analysis with climate projections in the water sector. Water Resour. Res. 2012, 48, W09537. [Google Scholar] [CrossRef]
  139. Mateus, M.C.; Tullos, D. Reliability, sensitivity, and vulnerability of reservoir operations under climate change. J. Water Resour. Plan. Manag. 2017, 143, 04016085. [Google Scholar] [CrossRef]
  140. Becker, B.; Ochterbeck, D.; Piovesan, T. A comparison of the homotopy method with linearization approaches for a non-linear optimization problem of operations in a reservoir cascade. Energy Syst. 2023. [Google Scholar] [CrossRef]
  141. Şen, Z. Reservoirs for Water Supply Under Climate Change Impact—A Review. Water Resour. Manag. 2021, 35, 3827–3843. [Google Scholar] [CrossRef]
  142. Lu, S.; Dai, W.; Tang, Y.; Guo, M. A review of the impact of hydropower reservoirs on global climate change. Sci. Total Environ. 2020, 711, 134996. [Google Scholar] [CrossRef]
  143. United Nations Office for Disaster Risk Reduction (UNDRR). Flood Hazard and Risk Assessment, in Words into Action Guidelines: National Disaster Risk Assessment, Hazard Specific Risk Assessment. 2017. Available online: www.preventionweb.net/files/52828_04floodhazardandriskassessment.pdf (accessed on 8 November 2021).
  144. Zhang, X.; Luo, J.; Sun, X.; Xie, J. Optimal reservoir flood operation using a decomposition-based multi-objective evolutionary algorithm. Eng. Optim. 2019, 51, 42–62. [Google Scholar] [CrossRef]
  145. Qi, Y.; Yu, J.; Li, X.; Wei, Y.; Miao, Q. Reservoir flood control operation using a multi-objective evolutionary algorithm with decomposition and preferences. Appl. Soft Comput. 2016, 50, 21–33. [Google Scholar] [CrossRef]
  146. Hsu, N.S.; Wei, C.C. A multipurpose reservoir real-time operation model for flood control during typhoon invasion. J. Hydrol. 2007, 336, 282–293. [Google Scholar] [CrossRef]
  147. Jain, S.K.; Shilpa, L.S.; Rani, D.; Sudheer, K.P. State-of-the-art review: Operation of multi-purpose reservoirs during flood season. J. Hydrol. 2023, 618, 129165. [Google Scholar] [CrossRef]
  148. Güvengir, U.; Savasaneril, S.; Altan-Sakarya, A.B.; Buhan, S. Short-term flood control and Long-Term energy maximization in multi-reservoir systems using improved particle swarm optimization. Water Resour. Manag. 2021, 35, 4293–4307. [Google Scholar] [CrossRef]
  149. Xu, B.; Zhong, P.-A.; Lu, Q.; Zhu, F.; Huang, X.; Ma, Y.; Fu, J. Multiobjective stochastic programming with recourses for real-time flood water conservation of a multi-reservoir system under uncertain forecasts. J. Hydrol. 2020, 590, 125513. [Google Scholar] [CrossRef]
  150. Maliwal, S.; Murmu, M.; Yadu, L.K.; Verma, M.K. Multi-reservoir flood control operation by optimization technique: A Review. Int. J. Eng. Res. Technol. 2019, 8, 681–685. [Google Scholar]
  151. Lei, X.; Zhang, J.; Wang, H.; Wang, M.; Khu, S.-T.; Li, Z.; Tan, Q. Deriving mixed reservoir operating rules for flood control based on weighted non-dominated sorting genetic algorithm II. J. Hydrol. 2018, 564, 967–983. [Google Scholar] [CrossRef]
  152. Zhou, C.; Sun, N.; Chen, L.; Ding, Y.; Zhou, J.; Zha, G.; Luo, G.; Dai, L.; Yang, X. Optimal operation of cascade reservoirs for flood control of multiple areas downstream: A case study in the Upper Yangtze River basin. Water 2018, 10, 1250. [Google Scholar] [CrossRef]
  153. Chou, F.N.; Wu, C.-W. Stage-wise optimizing operating rules for flood control in a multi-purpose reservoir. J. Hydrol. 2014, 521, 245–260. [Google Scholar] [CrossRef]
  154. Liu, Y.; Qin, H.; Mo, L.; Wang, Y.; Chen, D.; Pang, S.; Yin, X. Hierarchical flood operation rules optimization using a multi-objective cultured evolutionary algorithm based on decomposition. Water Resour. Manag. 2019, 33, 337–354. [Google Scholar] [CrossRef]
  155. Zhang, J.; Cai, X.; Lei, X.; Liu, P.; Wang, H. Real-time reservoir flood control operation enhanced by data assimilation. J. Hydrol. 2021, 598, 126426. [Google Scholar] [CrossRef]
  156. Cuevas-Velásquez, V.; Sordo-Ward, A.; García-Palacios, J.H.; Bianucci, P.; Garrote, L. A probabilistic model for real-time flood operation of a dam based on a deterministic optimization model. Water 2020, 12, 3206. [Google Scholar] [CrossRef]
  157. Rahimi, H.; Ardakani, M.K.; Ahmadian, M.; Tang, X. Multi-reservoir utilization planning to optimize hydropower energy and flood control simultaneously. Environ. Process. 2020, 7, 41–52. [Google Scholar] [CrossRef]
  158. Porse, E.; Sandoval-Solis, S.; Lane, B. Integrating environmental flows into multi-objective reservoir management for a transboundary, water-scarce river basin: Rio Grande/bravo. Water Resour. Manag. 2015, 29, 2471–2484. [Google Scholar] [CrossRef]
  159. Li, Q.; Ouyang, S. Research on multi-objective joint optimal flood control model for cascade reservoirs in river basin system. Nat. Hazards 2015, 77, 2097–2115. [Google Scholar] [CrossRef]
  160. Wan, X.; Xue, Y.; Hua, L.; Wu, Q. Multi-objective collaborative decision-making for flood resource utilization in a reservoir. Stoch. Environ. Res. Risk Assess. 2023. [Google Scholar] [CrossRef]
  161. Wan, W.; Liu, Y.; Zheng, H.; Zhao, J.; Zhao, F.; Lu, Y. Optimization of Multi-Reservoir Flood Control Operating Rules: A Case Study for the Chaobai River Basin in China. Water 2023, 15, 2817. [Google Scholar] [CrossRef]
  162. Hu, H.-X.; Shao, L.-H.; Hu, Q.; Zhang, Y.; Hu, Z.-Y. Multi-objective reservoir optimal operation based on GCN and NSGA-II algorithm. In Proceedings of the 4th International Conference on Advanced Electronic Materials, Computers and Software Engineering (A.E.M.C.S.E.), Changsha, China, 26–28 March 2021; pp. 546–551. [Google Scholar] [CrossRef]
  163. Liu, X.; Luo, J. A dynamic multi-objective optimization model with interactivity and uncertainty for real-time reservoir flood control operation. Appl. Math. Model. 2019, 74, 606–620. [Google Scholar] [CrossRef]
  164. Moridi, A.; Yazdi, J. Optimal allocation of flood control capacity for multi-reservoir systems using a multi-objective optimization approach. Water Resour. Manag. 2017, 31, 4521–4538. [Google Scholar] [CrossRef]
  165. Zhu, F.; Zhong, P.-A.; Sun, Y. Multi-criteria group decision making under uncertainty: Application in reservoir flood control operation. Environ. Model. Softw. 2018, 100, 236–251. [Google Scholar] [CrossRef]
  166. Zhu, F.; Zhong, P.-A.; Wu, Y.-N.; Sun, Y.; Chen, J.; Jia, B. SMAA-based stochastic multi-criteria decision-making for reservoir flood control operation. Stoch. Environ. Res. Risk Assess. 2017, 31, 1485–1497. [Google Scholar] [CrossRef]
  167. Yu, X.; Xu, Y.-P.; Gu, H.; Guo, Y. Multi-objective robust optimization of reservoir operation for real-time flood control under forecasting uncertaint. J. Hydrol. Part A 2023, 620, 129421. [Google Scholar] [CrossRef]
  168. Huang, X.; Xu, B.; Zhong, P.-A.; Yao, H.; Yue, H.; Zhu, F.; Lu, Q.; Sun, Y.; Mo, R.; Li, Z.; et al. Robust multiobjective reservoir operation and risk decision-making model for real-time flood control coping with forecast uncertainty. J. Hydrol. 2022, 605, 127334. [Google Scholar] [CrossRef]
  169. Wang, Q.; Zhou, J.; Dai, L.; Huang, K.; Zha, G. Risk assessment of multi-reservoir joint flood control system under multiple uncertainties. J. Flood Risk Manag. 2021, 14, e12740. [Google Scholar] [CrossRef]
  170. Lu, Q.; Zhong, P.-A.; Xu, B.; Zhu, F.; Ma, Y.; Wang, H.; Xu, S. Risk analysis for reservoir flood control operation considering two-dimensional uncertainties based on Bayesian network. J. Hydrol. 2020, 589, 125353. [Google Scholar] [CrossRef]
  171. Wang, Q.; Zhou, J.; Huang, K.; Dai, L.; Zha, G.; Chen, L.; Qin, H. Risk assessment and decision-making based on Mean-CVaR-Entropy for flood control operation of large-scale reservoirs. Water 2019, 11, 649. [Google Scholar] [CrossRef]
  172. Chen, J.; Zhong, P.-A.; An, R.; Zhu, F.-L.; Xu, B. Risk analysis for real-time flood control operation of a multi-reservoir system using a dynamic Bayesian network. Environ. Model. Softw. 2019, 111, 409–420. [Google Scholar] [CrossRef]
  173. Huang, K.; Ye, L.; Chen, L.; Wang, Q.; Dai, L.; Zhou, J.; Singh, V.P.; Huang, M.; Zhang, J. Risk analysis of flood control reservoir operation considering multiple uncertainties. J. Hydrol. 2018, 565, 672–684. [Google Scholar] [CrossRef]
  174. Chen, J.; Zhong, P.-A.; Wang, M.-L.; Zhu, F.-L.; Wan, X.-Y.; Zhang, Y. A risk-based model for real-time flood control operation of a cascade reservoir system under emergency conditions. Water 2018, 10, 167. [Google Scholar] [CrossRef]
  175. Zhu, F.; Zhong, P.-A.; Sun, Y.; Yeh, W.W.-G. Real-time optimal flood control decision-making and risk propagation under multiple uncertainties. Water Resour. Res. 2017, 53, 10635–10654. [Google Scholar] [CrossRef]
  176. Zhou, Y.; Guo, S. Risk analysis for flood control operation of seasonal flood-limited water level incorporating inflow forecasting error. Hydrol. Sci. J. 2014, 59, 1006–1019. [Google Scholar] [CrossRef]
  177. Yan, B.; Guo, S.; Chen, L. Estimation of reservoir flood control operation risks while considering inflow forecasting errors. Stoch. Environ. Res. Risk Assess. 2014, 28, 359–368. [Google Scholar] [CrossRef]
  178. Chen, J.; Zhong, P.-A.; Zhang, W.; Zhu, F.; Zhang, Y. Improved risk-assessment model for real-time reservoir flood-control operation. J. Water Resour. Plan. Manag. 2020, 146, 04020007. [Google Scholar] [CrossRef]
  179. Sun, Y.; Zhu, F.; Chen, J.; Li, J. Risk analysis for reservoir real-time optimal operation using the scenario tree-based stochastic optimization method. Water 2018, 10, 606. [Google Scholar] [CrossRef]
  180. Chen, J.; Zhong, P.-A.; Zhang, Y.; Navar, D.; Yeh, W.W.-G. A decomposition-integration risk analysis method for real-time operation of a complex flood control system. Water Resour. Res. 2017, 53, 2490–2506. [Google Scholar] [CrossRef]
  181. Zhao, Q.; Cai, X.; Li, Y. Determining inflow forecast horizon for reservoir operation. Water Resour. Res. 2019, 55, 4066–4081. [Google Scholar] [CrossRef]
  182. Kim, T.; Shin, J.-Y.; Kim, H.; Kim, S.; Heo, J.-H. The use of large-scale climate indices in monthly reservoir inflow forecasting and its application on time series and artificial intelligence models. Water 2019, 11, 374. [Google Scholar] [CrossRef]
  183. Valipour, M.; Banihabib, M.E.; Behbahani, S.M.R. Comparison of the A.R.M.A., A.R.I.M.A., and the autoregressive artificial neural network models in forecasting the monthly inflow of the Dez dam reservoir. J. Hydrol. 2013, 476, 433–441. [Google Scholar] [CrossRef]
  184. Karamouz, M.; Araghinejad, S. Advanced Hydrology; Amirkabir University Press: Tehran, Iran, 2005. [Google Scholar]
  185. Liang, Z.; Tang, T.; Li, B.; Liu, T.; Wang, J.; Hu, Y. Long-term streamflow forecasting using SWAT through the integration of the random forest’s precipitation generator: A case study of Danjiangkou. Reserv. Hydrol. Res. 2018, 49, 1513–1527. [Google Scholar] [CrossRef]
  186. Jaiswal, R.K.; Ali, S.; Bharti, B. Comparative evaluation of conceptual and physical rainfall–runoff models. Appl. Water Sci. 2020, 10, 48. [Google Scholar] [CrossRef]
  187. Tegegne, G.; Park, D.K.; Kim, Y.-O. Comparison of hydrological models for the assessment of water resources in a data-scarce region, the Upper Blue Nile River Basin. J. Hydrol. 2017, 14, 49–66. [Google Scholar] [CrossRef]
  188. Apaydin, H.; Feizi, H.; Sattari, M.T.; Colak, M.S.; Shamshirband, S.; Chau, K.-W. Comparative analysis of recurrent neural network architectures for reservoir inflow forecasting. Water 2020, 12, 1500. [Google Scholar] [CrossRef]
  189. Babaei, M.; Moeini, R.; Ehsanzadeh, E. Artificial neural network and support vector machine models for inflow prediction of dam reservoir (Case study: Zayandehroud dam reservoir). Water Resour. Manag. 2019, 33, 2203–2218. [Google Scholar] [CrossRef]
  190. Charoenporn, P. Reservoir inflow forecasting using ID3 and C4.5 decision tree model. In Proceedings of the 2017 3rd IEEE International Conference on Control Science and Systems Engineering (ICCSSE), Beijing, China, 17–19 August 2017; pp. 698–701. [Google Scholar] [CrossRef]
  191. Noorbeh, P.; Roozbahani, A.; Kardan Moghaddam, H. Annual and Monthly Dam Inflow Prediction Using Bayesian Networks. Water Resour. Manag. 2020, 34, 2933–2951. [Google Scholar] [CrossRef]
  192. Lin, Y.H.; Chiu, C.C.; Lee, P.C.; Lin, Y.J. Applying fuzzy grey modification model on inflow forecasting. Eng. Appl. Artif. Intell. 2012, 25, 734–743. [Google Scholar] [CrossRef]
  193. Tikhamarine, Y.; Souag-Gamane, D.; Ahmed, A.N.; Kisi, O.; El-Shafie, A. Improving artificial intelligence models accuracy for monthly streamflow forecasting using grey Wolf optimization (G.W.O.) algorithm. J. Hydrol. 2020, 582, 124435. [Google Scholar] [CrossRef]
  194. Ashrafi, S.M.; Mostaghimzadeh, E.; Adib, A. Applying wavelet transformation and artificial neural networks to develop forecasting-based reservoir operating rule curves. Hydrol. Sci. J. 2020, 65, 2007–2021. [Google Scholar] [CrossRef]
  195. Ebrahimi, E.; Shourian, M. River flow prediction using dynamic method for selecting and prioritizing K-Nearest Neighbors based on data features. J. Hydrol. Eng. 2020, 25, 04020010. [Google Scholar] [CrossRef]
  196. Latif, S.D.; Ahmed, A.N.; Sathiamurthy, E.; Huang, Y.F.; El-Shafie, A. Evaluation of deep learning algorithm for inflow forecasting: A case study of Durian Tunggal Reservoir, Peninsular Malaysia. Nat. Hazards 2021, 109, 351–369. [Google Scholar] [CrossRef]
  197. Chua, L.H.C. Considerations for data-driven and physically based hydrological models in flow forecasting. IFAC Proc. Vol. 2012, 45, 1025–1030. [Google Scholar] [CrossRef]
  198. Längkvist, M.; Karlsson, L.; Loutfi, A. A review of unsupervised feature learning and deep learning for time-series modeling. Pattern Recognit. Lett. 2014, 42, 11–24. [Google Scholar] [CrossRef]
  199. Bai, Y.; Chen, Z.; Xie, J.; Li, C. Daily reservoir inflow forecasting using multiscale deep feature learning with hybrid models. J. Hydrol. 2016, 532, 193–206. [Google Scholar] [CrossRef]
  200. Dong, N.; Guan, W.; Cao, J.; Zou, Y.; Yang, M.; Wei, J.; Chen, L.; Wang, H. A hybrid hydrologic modeling framework with data-driven and conceptual reservoir operation schemes for reservoir impact assessment and predictions. J. Hydrol. 2023, 619, 129246. [Google Scholar] [CrossRef]
  201. Khorram, S.; Jehbez, N. A Hybrid CNN-LSTM Approach for Monthly Reservoir Inflow Forecasting. Water Resour. Manag. 2023, 37, 4097–4121. [Google Scholar] [CrossRef]
  202. Tian, D.; He, X.; Srivastava, P.K.; Kalin, L. A hybrid framework for forecasting monthly reservoir inflow based on machine learning techniques with dynamic climate forecasts, satellite-based data, and climate phenomenon information. Stoch. Environ. Res. Risk Assess. 2022, 36, 2353–2375. [Google Scholar] [CrossRef]
  203. Attar, N.F.; Pham, Q.B.; Nowbandegani, S.F.; Rezaie-Balf, M.; Fai, C.M.; Ahmed, A.N.; Pipelzadeh, S.; Dung, T.D.; Nhi, P.T.T.; Khoi, D.N.; et al. Enhancing the prediction accuracy of data-driven models for monthly streamflow in the Urmia Lake basin based upon the autoregressive conditionally heteroskedastic time-series model. Appl. Sci. 2020, 10, 571. [Google Scholar] [CrossRef]
  204. Banihabib, M.E.; Bandari, R.; Valipour, M. Improving daily peak flow forecasts using hybrid Fourier-series autoregressive integrated moving average and recurrent artificial neural network models. AI 2020, 1, 263–275. [Google Scholar] [CrossRef]
  205. Luo, B.; Fang, Y.; Wang, H.; Zang, D. Reservoir inflow prediction using a hybrid model based on deep learning. IOP Conf. Ser. Mater. Sci. Eng. 2020, 715, 012044. [Google Scholar] [CrossRef]
  206. Chen, N.; Xiong, C.; Du, W.; Wang, C.; Lin, X.; Chen, Z. An improved genetic algorithm coupling a Back-Propagation neural network model (IGA-BPNN) for Water-Level predictions. Water 2019, 11, 1795. [Google Scholar] [CrossRef]
  207. Banihabib, M.E.; Ahmadian, A.; Valipour, M. Hybrid MARMA-NARX model for flow forecasting based on large-scale climate signals, sea-surface temperatures, and rainfall. Hydrol. Res. 2018, 49, 1788–1803. [Google Scholar] [CrossRef]
  208. Awan, J.A.; Bae, D. Application of adaptive neuro-fuzzy inference system for dam inflow prediction using long-range weather forecast. In Proceedings of the Eighth International Conference on Digital Information Management (ICDIM 2013), Islamabad, Pakistan, 10–12 September 2013; pp. 247–251. [Google Scholar] [CrossRef]
  209. Alquraish, M.M.; Abuhasel, K.A.; Alqahtani, A.S.; Khadr, M. A comparative analysis of hidden Markov model, hybrid support vector machines, and hybrid artificial neural fuzzy inference system in reservoir inflow forecasting (Case study: King Fahd Dam, Saudi Arabia). Water 2021, 13, 1236. [Google Scholar] [CrossRef]
  210. Ibrahim, K.S.M.H.; Huang, Y.F.; Ahmed, A.N.; Koo, C.H.; El-Shafie, A. A review of the hybrid artificial intelligence and optimization modeling of hydrological streamflow forecasting. Alex. Eng. J. 2022, 61, 279–303. [Google Scholar] [CrossRef]
  211. You, J.Y.; Cai, X. Determining forecast and decision horizons for reservoir operations under hedging policies. Water Resour. Res. 2008, 44, W11430. [Google Scholar] [CrossRef]
  212. Zhao, T.; Yang, D.; Cai, X.; Zhao, J.; Wang, H. Identifying effective forecast horizon for real-time reservoir operation under a limited inflow forecast. Water Resour. Res. 2012, 48, W01540. [Google Scholar] [CrossRef]
  213. Turner, S.W.D.; Xu, W.; Voisin, N. Inferred inflow forecast horizons guiding reservoir release decisions across the United States. Hydrol. Earth Syst. Sci. 2020, 24, 1275–1291. [Google Scholar] [CrossRef]
  214. Zhao, T.; Cai, X.; Yang, D. Effect of streamflow forecast uncertainty on real-time reservoir operation. Adv. Water Resour. 2011, 34, 495–504. [Google Scholar] [CrossRef]
  215. Zhong, Y.; Guo, Y.; Xiong, F.; Liu, D.; Ba, H.; Wu, X. Probabilistic forecasting based on ensemble forecasts and EMOS method for T.G.R. inflow. Front. Earth Sci. 2020, 14, 188–200. [Google Scholar] [CrossRef]
  216. Bourdin, D.R.; Nipen, T.N.; Stull, R.B. Reliable probabilistic forecasts from an ensemble reservoir inflow forecasting system. Water Resour. Res. 2014, 50, 3108–3130. [Google Scholar] [CrossRef]
  217. Barbetta, B.; Sahoo, B.; Bonaccorsi, B.; Nanda, T.; Chatterjee, C.; Moramarco, T.; Todini, E. Addressing effective real-time forecasting inflows to dams through predictive uncertainty estimate. J. Hydrol. 2023, 620, 129512. [Google Scholar] [CrossRef]
  218. Zhou, Y.; Guo, S.; Xu, C.-Y.; Chang, F.-J.; Yin, J. Improving the reliability of probabilistic multi-step-ahead flood forecasting by fusing an unscented Kalman filter with a recurrent neural network. Water 2020, 12, 578. [Google Scholar] [CrossRef]
  219. Li, Y.; Wu, Z.; He, H.; Lu, G. Deterministic and probabilistic evaluation of sub-seasonal precipitation forecasts at various spatiotemporal scales over China during the boreal summer monsoon. Atmosphere 2021, 12, 1049. [Google Scholar] [CrossRef]
  220. Zhong, Y.; Guo, S.; Ba, H.; Xiong, F.; Chang, F.-J.; Lin, K. Evaluation of the B.M.A. probabilistic inflow forecasts using T.I.G.G.E. numeric precipitation predictions based on artificial neural network. Hydrol. Res. 2018, 49, 1417–1433. [Google Scholar] [CrossRef]
  221. Bourdin, D.R.; Stull, R.B. Bias-corrected short-range Member-to-Member ensemble forecasts of reservoir inflow. J. Hydrol. 2013, 502, 77–88. [Google Scholar] [CrossRef]
  222. Yang, X.; Zhou, J.; Fang, W.; Wang, Y. An ensemble flow forecast method based on an autoregressive model and hydrological uncertainty processer. Water 2020, 12, 3138. [Google Scholar] [CrossRef]
  223. Thi, P.C.; Ball, J.E.; Dao, N.H. Uncertainty estimation using the Glue and Bayesian approaches in flood estimation: A case study—Ba River, Vietnam. Water 2018, 10, 1641. [Google Scholar] [CrossRef]
  224. Thiemann, M.; Trosset, M.; Gupta, H.; Sorooshian, S. Bayesian recursive parameter estimation for hydrologic models. Water Resour. Res. 2001, 37, 2521–2535. [Google Scholar] [CrossRef]
  225. Parrish, M.A.; Moradkhani, H.; DeChant, C.M. Toward reduction of model uncertainty: Integration of Bayesian model averaging and data assimilation. Water Resour. Res. 2012, 48, W03519. [Google Scholar] [CrossRef]
  226. Hamill, T.M. Interpretation of rank histograms for verifying ensemble forecasts. Mon. Weather Rev. 2001, 129, 550–560. Available online: http://n2t.net/ark:/85065/d71z450p (accessed on 1 January 2022). [CrossRef]
  227. Hamill, T.M.; Colucci, S.J. Verification of Eta-RSM short-range ensemble forecasts. Mon. Weather Rev. 1997, 125, 1312–1327. [Google Scholar] [CrossRef]
  228. CAIT (Climate Analysis Indicators Tool). Climate Data Explorer; Data for the Year 2016; World Resources Institute: Washington, DC, USA, 2016; Available online: http://cait2.wri.org/ (accessed on 1 October 2020).
  229. Kumar, P.; Dhanya, T.; Chahar, B. A PSO approach for optimum design of dynamic inversion controller in water distribution systems. J. Water Supply Res. Technol. AQUA 2016, 65, 580–591. [Google Scholar] [CrossRef]
  230. Fallah-Mehdipour, E.; Bozorg Haddad, O.; Mariño, M.A. Developing reservoir operational decision rule by genetic programming. J. Hydro Inform. 2013, 15, 103–119. [Google Scholar] [CrossRef]
  231. Dias, V.; da Luz, M.P.; Medero, G.M.; Nascimento, D.T.F. An overview of hydropower reservoirs in Brazil: Current situation, Future perspectives and impacts of climate change. Water 2018, 10, 592. [Google Scholar] [CrossRef]
  232. Tkáč, S. Hydropower plants, an overview of the current types and technology. J. Civ. Eng. 2018, 13, 115–126. [Google Scholar] [CrossRef]
  233. McManamay, R.A.; Oigbokie, C.O.; Kao, S.-C.; Bevelhimer, M.S. Classification of US Hydropower Dams by their Modes of Operation. River Res. Appl. 2016, 32, 1450–1468. [Google Scholar] [CrossRef]
  234. Das, P.; Das, B.K.; Mustafi, N.N.; Sakir, T. A review on pump-hydro storage for renewable and hybrid energy systems applications. Energy 2020, 3, e223. [Google Scholar] [CrossRef]
  235. Pitorac, L.; Vereide, K.; Lia, L. Technical Review of Existing Norwegian Pumped Storage Plants. Energies 2020, 13, 4918. [Google Scholar] [CrossRef]
  236. Hoffstaedt, J.P.; Truijen, D.P.K.; Fahlbeck, J.; Gans, L.H.A.; Qudaih, M.; Laguna, A.J.; De Kooning, J.D.M.; Stockman, K.; Nilsson, H.; Storli, P.-T.; et al. Low head pumped hydro storage: A review of applicable technologies for design, grid integration, control, and modeling. Renew. Sustain. Energy 2022, 158, 112119. [Google Scholar] [CrossRef]
  237. Blakers, A.; Stocks, M.; Lu, B.; Cheng, C. A review of pumped hydro energy storage. Prog. Energy 2021, 3, 022003. [Google Scholar] [CrossRef]
  238. Pérez-Díaz, J.I.; Chazarra, M.; García-González, J.; Cavazzini, G.; Stoppato, A. Trends and challenges in the operation of pumped-storage hydropower plants. Renew. Sustain. Energy 2015, 44, 767–784. [Google Scholar] [CrossRef]
  239. Wang, Z.; Fang, G.; Wen, X.; Tan, Q.; Zhang, P.; Liu, Z. Coordinated operation of conventional hydropower plants as hybrid pumped storage hydropower with wind and photovoltaic plants. Energy Convers. Manag. 2023, 277, 116654. [Google Scholar] [CrossRef]
  240. Ge, X.; Ma, Y.; Li, Y.; Jiao, Y.; Wang, Z.; Wu, F.; Zhao, B.; Xue, X. Daily peak shaving operation of mixed pumped-storage hydro plants considering cascade hydraulic coupling. In Proceedings of the 3rd International Conference on Power and Electrical Engineering (ICPEE 2022), Singapore 29–31 December 2022; pp. 971–978. [Google Scholar] [CrossRef]
  241. Bredeson, L.; Cicilio, P. Hydropower and Pumped Storage Hydropower Resource Review and Assessment for Alaska’s Railbelt Transmission System. Energies 2023, 16, 5494. [Google Scholar] [CrossRef]
  242. Nasab, M.A.; Zand, M.; Padmanabhan, S.; Bhaskar, M.S.; Guerrero, J.M. An efficient, robust optimization model for the unit commitment considering renewable uncertainty and pumped-storage hydropower. Comput. Electr. Eng. 2022, 100, 107846. [Google Scholar] [CrossRef]
  243. Nasir, J.; Javed, A.; Ali, M.; Ullah, K.; Kazmi, S.A.A. Capacity optimization of pumped storage hydropower and its impact on an integrated conventional hydropower plant operation. Appl. Energy 2022, 323, 119561. [Google Scholar] [CrossRef]
  244. Zhao, J.-F.; Oh, U.-J.; Park, J.-C.; Park, E.S.; Im, H.-B.; Lee, K.Y.; Choi, J.-S. A Review of World-wide Advanced Pumped Storage Hydropower Technologies. IFAC-Pap. 2022, 55, 170–174. [Google Scholar] [CrossRef]
  245. Yuan, W.; Xin, W.; Su, C.; Cheng, C.; Yan, D.; Wu, Z. Cross-regional integrated transmission of wind power and pumped-storage hydropower considering the peak shaving demands of multiple power grids. Renew. Energy 2022, 190, 1112–1126. [Google Scholar] [CrossRef]
  246. Li, Y.; Wu, F.; Li, J.; Yin, Y.; Li, Z.; Ai, L. Chance-constrained energy management for pumped storage hydropower plant to compensate for wind power uncertainties. In Proceedings of the 2021 IEEE Power & Energy Society General Meeting (PESGM), Washington, DC, USA, 26–29 July 2021; pp. 1–5. [Google Scholar] [CrossRef]
  247. Dong, Z.; Tan, J.; Muljadi, E.; Nelms, R.M.; St-Hilaire, A.; Pevarnik, M.; Jacobson, M.D. Developing of Quaternary Pumped Storage Hydropower for Dynamic Studies. In IEEE Transactions on Sustainable Energy; IEEE: New York, NY, USA, 2020; Volume 11, pp. 2870–2878. [Google Scholar] [CrossRef]
  248. Bayazıt, Y.; Bakış, R.; Koç, C. A study on the transformation of multi-purpose dams into pumped storage hydroelectric power plants by using GIS model. Int. J. Green Energy 2021, 18, 308–318. [Google Scholar] [CrossRef]
  249. Alvarez, G.E. Operation of pumped storage hydropower plants through optimization for power systems. Energy 2020, 202, 17797. [Google Scholar] [CrossRef]
  250. Ma, X.; Wu, D.; Wang, D.; Huang, B.; Desomber, K.; Fu, T.; Weimar, M. Optimizing pumped storage hydropower for multiple grid services. J. Energy Storage 2022, 51, 104440. [Google Scholar] [CrossRef]
  251. Hunt, J.D.; Zakeri, B.; Lopes, R.; Barbosa, P.S.F.; Nascimento, A.; de Castro, N.J.; Brandão, R.; Schneider, P.S.; Wada, Y. Reviews Existing and new arrangements of pumped-hydro storage plants. Renew. Sustain. Energy 2020, 129, 109914. [Google Scholar] [CrossRef]
  252. Villeneuve, Y.; Séguin, S.; Chehri, A. AI-Based Scheduling Models, Optimization, and Prediction for Hydropower Generation: Opportunities, Issues, and Future Directions. Energies 2023, 16, 3335. [Google Scholar] [CrossRef]
  253. Bernardes, J., Jr.; Santos, M.; Abreu, T.; Prado, L., Jr.; Miranda, D.; Julio, R.; Viana, P.; Fonseca, M.; Bortoni, E.; Bastos, G.S. Hydropower Operation Optimization Using Machine Learning: A Systematic Review. AI 2022, 3, 78–99. [Google Scholar] [CrossRef]
  254. Bordin, C.; Skjelbred, H.I.; Kong, J.; Yang, Z. Machine learning for hydropower scheduling: State of the art and future research directions. Procedia Comput. Sci. 2020, 176, 1659–1668. [Google Scholar] [CrossRef]
  255. Hammid, A.T.; Awad, O.I.; Sulaiman, M.H.; Gunasekaran, S.S.; Mostafa, S.A.; Kumar, N.M.; Khalaf, B.A.; Al-Jawhar, Y.A.; Abdulhasan, R.A. A review of optimization algorithms in solving hydro generation scheduling problems. Energies 2020, 13, 2787. [Google Scholar] [CrossRef]
  256. Singh, V.K.; Singal, S.K. Operation of hydropower plants review. Renew. Sustain. Energy Rev. 2017, 69, 610–619. [Google Scholar] [CrossRef]
  257. Zhang, R.; Zhang, S.; Wen, X.; Jing, Z. Refined Scheduling Based on Dynamic Capacity Model for Short-term Hydropower Generation. Water Resour Manag. Manag. 2023, 37, 21–35. [Google Scholar] [CrossRef]
  258. Brito, B.H.; Finardi, E.C.; Takigawa, F.Y.K.; Pereira, A.I. Exploring symmetry in a short-term hydro scheduling problem: The case of the Santo Antônio hydro plant. J. Water Resour. Plan. Manag. 2022, 148, 05021026. [Google Scholar] [CrossRef]
  259. Daadaa, M.; Séguin, S.; Demeester, K.; Anjos, M.F. An optimization model to maximize energy generation in short-term hydropower unit commitment using efficiency points. Int. J. Electr. Power Energy Syst. 2021, 125, 106419. [Google Scholar] [CrossRef]
  260. Mathur, P.; Swartz, C.L.E.; Zyngier, D.; Welt, F. Robust online scheduling for optimal short-term operation of cascaded hydropower systems under uncertainty. J. Process Control 2021, 98, 52–65. [Google Scholar] [CrossRef]
  261. Guisández, I.; Pérez-Díaz, J.I. Mixed integer linear programming formulations for the hydro production function in a unit-based short-term scheduling problem. Int. J. Electr. Power Energy Syst. 2021, 128, 106747. [Google Scholar] [CrossRef]
  262. Skjelbred, H.I.; Kong, J.; Fosso, O.B. Dynamic incorporation of nonlinearity into M.I.L.P. formulation for short-term hydro scheduling. Int. J. Electr. Power Energy Syst. 2020, 116, 105530. [Google Scholar] [CrossRef]
  263. Su, C.; Yuan, W.; Cheng, C.; Wang, P.; Sun, L.; Zhang, T. Short-term generation scheduling of cascade hydropower plants with strong hydraulic coupling and head-dependent prohibited operating zones. J. Hydrol. 2020, 591, 125556. [Google Scholar] [CrossRef]
  264. Zhou, L.; Cheng, C.; Liao, S.; Wang, J. Multiobjective scheduling method for short-term peak shaving operation of cascade hydro plants. J. Water Resour. Plan. Manag. 2020, 146, 04020073. [Google Scholar] [CrossRef]
  265. Yang, Z.; Yang, K.; Wang, Y.; Su, L.; Hu, H. multi-objective short-term hydropower generation operation for cascade reservoirs and stochastic decision-making under multiple uncertainties. J. Clean. Prod. 2020, 276, 122995. [Google Scholar] [CrossRef]
  266. Shang, L.; Shang, Y.; Hu, L.; Li, J. Performance of genetic algorithms with different selection operators for solving short-term optimized reservoir scheduling problems. Soft Comput. 2020, 24, 6771–6785. [Google Scholar] [CrossRef]
  267. Cheng, X.; Tang, Y.-L.; Liu, J.; Zhong, H.; Li, W.-W.; Wu, Y.-L. Ultrashort-Term scheduling of inter-basin cascaded hydropower plants to rapidly balance the load demand. IEEE Access 2020, 8, 32737–32756. [Google Scholar] [CrossRef]
  268. Liao, S.; Zhang, Y.; Liu, B.; Liu, Z.; Fang, Z.; Li, S. Short-term peak-shaving operation of head-sensitive cascaded hydropower plants based on spillage adjustment. Water 2020, 12, 3438. [Google Scholar] [CrossRef]
  269. Hu, H.; Yang, K.; Liu, L.; Su, L.; Yang, Z. Short-term hydropower generation scheduling using an improved cloud adaptive quantum-inspired binary social spider optimization algorithm. Water Resour. Manag. 2019, 33, 2357–2379. [Google Scholar] [CrossRef]
  270. Hermida, G.; Castronuovo, E.D. On the hydropower short-term scheduling of large basins, considering nonlinear programming, stochastic inflows, and heavy ecological restrictions. Int. J. Electr. Power Energy Syst. 2018, 97, 408–417. [Google Scholar] [CrossRef]
  271. Ge, X.; Xia, S.; Lee, W.-J.; Chung, C.Y. A successive approximation approach for short-term cascaded hydro scheduling with variable water flow delay. Electr. Power Syst. Res. 2018, 154, 213–222. [Google Scholar] [CrossRef]
  272. Wang, J.; Hu, W.; Liu, S. Short-term hydropower scheduling model with two coupled temporal scales. J. Water Resour. Plan. Manag. 2018, 144, 04017095. [Google Scholar] [CrossRef]
  273. Liu, S.; Luo, J.; Chen, H.; Wang, Y.; Li, X.; Zhang, J.; Wang, J. Third-Monthly Hydropower Scheduling of Cascaded Reservoirs Using Successive Quadratic Programming in Trust Corridor. Water 2023, 15, 716. [Google Scholar] [CrossRef]
  274. Li, G.; Lu, J.; Yang, R.; Cheng, C. Igdt-based medium-term optimal cascade hydropower operation in multimarket with hydrologic and economic uncertainties. J. Water Resour. Plan. Manag. 2021, 147, 05021015. [Google Scholar] [CrossRef]
  275. Beltrán, F.; Finardi, E.C.; de Oliveira, W. Two-stage and multi-stage decompositions for the medium-term hydrothermal scheduling problem: A computational comparison of solution techniques. Int. J. Electr. Power Energy Syst. 2021, 127, 106659. [Google Scholar] [CrossRef]
  276. Pérez-Díaz, J.I.; Guisández, I.; Chazarra, M.; Helseth, A. Medium-term scheduling of a hydropower plant participating as a price-maker in the automatic frequency restoration reserve market. Electr. Power Syst. Res. 2020, 185, 106399. [Google Scholar] [CrossRef]
  277. Ge, X.; Xia, S.; Lee, W. An efficient stochastic algorithm for mid-term scheduling of cascaded hydro systems. J. Mod. Power Syst. Clean Energy 2019, 7, 163–173. [Google Scholar] [CrossRef]
  278. Hjelmeland, M.N.; Zou, J.; Helseth, A.; Ahmed, S. Nonconvex medium-term hydropower scheduling by stochastic dual dynamic integer programming. IEEE Trans. Sustain. Energy 2019, 10, 481–490. [Google Scholar] [CrossRef]
  279. Feng, S.; Zheng, H.; Qiao, Y.; Yang, Z.; Wang, J.; Liu, S. Weekly hydropower scheduling of cascaded reservoirs with hourly power and capacity balances. Appl. Energy 2022, 311, 118620. [Google Scholar] [CrossRef]
  280. He, Z.; Wang, C.; Wang, Y.; Wei, B.; Zhou, J.; Zhang, H.; Qin, H. Dynamic programming with successive approximation and relaxation strategy for long-term joint power generation scheduling of large-scale hydropower station group. Energy 2021, 222, 119960. [Google Scholar] [CrossRef]
  281. Mbeutcha, Y.; Gendreau, M.; Emiel, G. The benefit of PARMA modeling for long-term hydroelectric scheduling using stochastic dual dynamic programming. J. Water Resour. Plan. Manag. 2021, 147, 05021002. [Google Scholar] [CrossRef]
  282. Ding, X.; Mo, X.; Zhou, J.; Bi, S.; Jia, B.; Liao, X. Long-term scheduling of cascade reservoirs considering inflow forecasting uncertainty based on a disaggregation model. Water Resour. Manag. 2021, 35, 645–660. [Google Scholar] [CrossRef]
  283. He, Z.; Zhou, J.; Qin, H.; Jia, B.; He, F.; Liu, G.; Feng, K. A fast water level optimal control method based on two-stage analysis for long-term power generation scheduling of hydropower station. Energy 2020, 210, 118531. [Google Scholar] [CrossRef]
  284. Yan, Z.; Liao, S.; Cheng, C.; Zhou, L. Long-term optimization of large-scale hydropower system operations based on decomposition coordination. J. Water Resour. Plan. Manag. 2020, 146, 04020078. [Google Scholar] [CrossRef]
  285. Hu, H.; Yang, K. Multiobjective long-term generation scheduling of cascade hydroelectricity system using a quantum-behaved particle swarm optimization based on decomposition. IEEE Access 2020, 8, 100837–100856. [Google Scholar] [CrossRef]
  286. Cao, R.; Shen, J.; Cheng, C.; Wang, J. Optimization model for the Long-Term operation of an interprovincial hydropower plant incorporating peak shaving demands. Energies 2020, 13, 4804. [Google Scholar] [CrossRef]
  287. Su, L.; Yang, K.; Hu, H.; Yang, Z. Long-term hydropower generation scheduling of large-scale cascade reservoirs using chaotic adaptive multi-objective bat algorithm. Water 2019, 11, 2373. [Google Scholar] [CrossRef]
  288. Hu, H.; Yang, K.; Su, L.; Yang, Z. A novel adaptive multi-objective particle swarm optimization based on decomposition and dominance for long-term generation scheduling of cascade hydropower system. Water Resour. Manag. 2019, 33, 4007–4026. [Google Scholar] [CrossRef]
  289. Wen, X.; Zhou, J.; He, Z.; Wang, C. Long-term scheduling of large-scale cascade hydropower stations using improved differential evolution algorithm. Water 2018, 10, 383. [Google Scholar] [CrossRef]
  290. Hammid, A.T.; Bin Sulaiman, M.H. Series division method based on PSO and F.A. to optimize Long-Term Hydro Generation Scheduling. Sustain. Energy Technol. Assess. 2018, 29, 106–118. [Google Scholar] [CrossRef]
  291. Luo, B.; Miao, S.; Cheng, C.; Lei, Y.; Chen, G.; Gao, L. Long-term generation scheduling for cascade hydropower plants considering price correlation between multiple markets. Energies 2019, 12, 2239. [Google Scholar] [CrossRef]
  292. Kang, C.; Chen, C.; Wang, J. An efficient linearization method for long-term operation of cascaded hydropower reservoirs. Water Resour. Manag. 2018, 32, 3391–3404. [Google Scholar] [CrossRef]
  293. Ehteram, M.; Karami, H.; Mousavi, S.F.; Farzin, S.; Kisi, O. Evaluation of contemporary evolutionary algorithms for optimization in reservoir operation and water supply. J. Water Supply Res. Technol. 2018, 67, 54–67. [Google Scholar] [CrossRef]
  294. UN-Water. Coping with Water Scarcity—The Challenge of the Twenty-First Century. 2007. Available online: www.fao.org/3/aq444e/aq444e.pdf (accessed on 8 November 2021).
  295. Syme, G.J. Acceptable risk and social values: Struggling with uncertainty in Australian water allocation. Stoch. Environ. Res. Risk Assess. 2013, 28, 113–121. [Google Scholar] [CrossRef]
  296. Chen, J.; Shi, H.Y.; Sivakumar, B.; Peart, M.R. Population, water, food, energy, and dams. Renew. Sustain. Energy Rev. 2016, 56, 18–28. [Google Scholar] [CrossRef]
  297. Chang, F.; Wang, K.W. A systematical water allocation scheme for drought mitigation. J. Hydrol. 2013, 507, 124–133. [Google Scholar] [CrossRef]
  298. Payus, C.; Huey, L.A.; Adnan, F.; Rimba, A.B.; Mohan, G.; Chapagain, S.K.; Roder, G.; Gasparatos, A.; Fukushi, K. Impact of extreme drought climate on water security in north Borneo: Case study of Sabah. Water 2020, 12, 1135. [Google Scholar] [CrossRef]
  299. Wu, J.; Chen, X.; Yao, H.; Gao, L.; Chen, Y.; Liu, M. Non-linear relationship of hydrological drought responding to meteorological drought and impact of a large reservoir. J. Hydrol. 2017, 551, 495–507. [Google Scholar] [CrossRef]
  300. World Meteorological Organization. Drought Monitoring and Warning: Concepts, Progress and Future Challenges; No. 1006; WMO Publication: Geneva, Switzerland, 2006; pp. 6–9. Available online: http://www.wamis.org/agm/pubs/brochures/WMO1006e.pdf (accessed on 1 August 2022).
  301. Huang, C.; Zhao, J.; Wang, Z.; Shang, W. Optimal hedging rules for two-objective reservoir operation: Balancing water supply and environmental flow. J. Water Resour. Plan. Manag. 2016, 142, 04016053. [Google Scholar] [CrossRef]
  302. Ji, Y.; Lei, X.; Cai, S.; Wang, X. Hedging rules for water supply reservoirs based on the model of simulation and optimization. Water 2016, 8, 249. [Google Scholar] [CrossRef]
  303. Ngamsert, R.; Techarungruengsakul, R.; Kaewplang, S.; Hormwichian, R.; Prasanchum, H.; Sivanpheng, O.; Kangrang, A. Optimizing Solution in Decision Supporting System for River Basin Management Consisting of a Reservoir System. Water 2023, 15, 2510. [Google Scholar] [CrossRef]
  304. Choi, Y.; Ji, J.; Lee, E.; Lee, S.; Yi, S.; Yi, J. Developing Optimal Reservoir Rule Curve for Hydropower Reservoir with an add-on Water Supply Function Using Improved Grey Wolf Optimizer. Water Resour. Manag. 2023, 37, 2063–2082. [Google Scholar] [CrossRef]
  305. Wang, J.; Cheng, C.; Wu, X.; Shen, J. Optimal Hedging for Hydropower Operation and End-of-Year Carryover Storage Values. J. Water Resour. Plan. Manag. 2019, 145, 04019003. [Google Scholar] [CrossRef]
  306. Xu, B.; Zhong, P.-A.; Huang, Q.; Wang, J.; Yu, Z.; Zhang, J. Optimal Hedging Rules for Water Supply Reservoir Operations under Forecast Uncertainty and Conditional Value-at-Risk Criterion. Water 2017, 9, 568. [Google Scholar] [CrossRef]
  307. Felfelani, F.; Movahed, A.J.; Zarghami, M. Simulating hedging rules for effective reservoir operation by using system dynamics: A case study of Dez Reservoir, Iran. Lake Reserv. Manag. 2013, 29, 126–140. [Google Scholar] [CrossRef]
  308. Thiha, S.; Shamseldin, A.Y.; Melville, B.W. Optimal reservoir operation using the improved multi-step-ahead time-varying hedging rule under climate and land-use changes. Hydrol. Sci. J. 2023, 68, 1139–1161. [Google Scholar] [CrossRef]
  309. Chang, J.; Guo, A.; Wang, Y.; Ha, Y.; Zhang, R.; Xue, L.; Tu, Z. Reservoir operations to mitigate drought effects with a hedging policy triggered by the drought prevention limiting water level. Water Resour. Res. 2019, 55, 904–922. [Google Scholar] [CrossRef]
  310. Hu, T.; Zhang, X.-Z.; Zeng, X.; Wang, J. A two-step approach for analytical optimal hedging with two triggers. Water 2016, 8, 52. [Google Scholar] [CrossRef]
  311. Neelakantan, T.R.; Sasireka, K. Hydropower reservoir operation using standard operating and standard hedging policies. Int. J. Eng. Technol. 2013, 5, 1191–1196. [Google Scholar]
  312. You, J.Y.; Cai, X. Hedging rules for reservoir operations: 1. A theoretical analysis. Water Resour. Res. 2008, 44, W01415. [Google Scholar] [CrossRef]
  313. You, J.Y.; Cai, X. Hedging rules for reservoir operations: 2. A numerical model. Water Resour. Res. 2008, 44, W01416. [Google Scholar] [CrossRef]
  314. Kumar, K.; Kasthurirengan, S. Generalized linear two-point hedging rule for water supply reservoir operation. J. Water Resour. Plan. Manag. 2018, 144, 04018051. [Google Scholar] [CrossRef]
  315. Shih, J.S.; ReVelle, C. Water-supply operations during drought: Continuous hedging rule. J. Water Resour. Plan. Manag. 1994, 120, 613–629. [Google Scholar] [CrossRef]
  316. Shiau, J.T. Analytical optimal hedging with explicit incorporation of reservoir release and carryover storage targets. Water Resour. Res. 2011, 47, W01515. [Google Scholar] [CrossRef]
  317. Rittima, A. Hedging policy for reservoir system operation: A case study of Mun Bon and Lam Chae reservoirs. J. Nat. Sci. 2009, 43, 833–842. [Google Scholar]
  318. Seo, S.B.; Kim, Y.O.; Kang, S.U. Time-varying discrete hedging rules for drought contingency plan considering long-range dependency in streamflow. Water Resour. Manag. 2019, 33, 2791–2807. [Google Scholar] [CrossRef]
  319. Jin, Y.; Lee, S. Comparative effectiveness of reservoir operation applying hedging rules based on available water and beginning storage to cope with droughts. Water Resour. Manag. 2019, 33, 1897–1911. [Google Scholar] [CrossRef]
  320. Men, B.; Wu, Z.; Liu, H.; Li, Y.; Zhao, Y. Research on hedging rules based on water supply priority and benefit loss of water shortage—A case study of Tianjin, China. Water 2019, 11, 778. [Google Scholar] [CrossRef]
  321. Tayebiyan, A.; Mohammad, T.A.; Al-Ansari, N.; Malakootian, M. Comparison of optimal hedging policies for hydropower reservoir system operation. Water 2019, 11, 121. [Google Scholar] [CrossRef]
  322. Lee, S.; Jin, Y.; Park, J. Comparison of hedging rule curves depending on available water and current storage in reservoir operation under droughts. Eur. Water 2017, 60, 227–232. [Google Scholar]
  323. Mansouri, S.; Fathian, H.; Nikbakht Shahbazi, A.; Asadi Lour, M.; Asareh, A. Multi-objective simulation–optimization operation of dam reservoir in low water regions based on hedging principles. Environ. Sci. Pollut. Res. 2023, 30, 41581–41590. [Google Scholar] [CrossRef]
  324. Jalilian, A.; Heydari, M.; Azari, A.; Shabanlou, S. Extracting optimal rule curve of dam reservoir based on stochastic inflow. Water Resour. Manag. 2022, 36, 1763–1782. [Google Scholar] [CrossRef]
  325. Bayesteh, M.; Azari, A. Stochastic optimization of reservoir operation by applying hedging rules. J. Water Resour. Plan. Manag. 2021, 147, 04020099. [Google Scholar] [CrossRef]
  326. Jamshidi, J.; Shourian, M. Hedging rules-based optimal reservoir operation using bat algorithm. Water Resour. Manag. 2019, 33, 4525–4538. [Google Scholar] [CrossRef]
  327. Srinivasan, K.; Kumar, K. Multi-Objective Simulation-Optimization Model for Long-term Reservoir Operation using Piecewise Linear Hedging Rule. Water Resour Manag. 2018, 32, 1901–1911. [Google Scholar] [CrossRef]
  328. Srivastava, A. A computational framework for dam safety risk Assessment with uncertainty analysis. In All Graduate Theses and Dissertations; Utah State University: Logan, UT, USA, 2013; p. 1480. [Google Scholar] [CrossRef]
  329. Neelakantan, T.R.; Sasireka, K. Review of hedging rules applied to reservoir operation. Int. J. Eng. Technol. 2015, 7, 1571–1580. [Google Scholar]
  330. Baecher, G. Uncertainty in dam safety risk analysis. Georisk Assess. Manag. Risk Eng. Syst. Geohazards 2016, 10, 92–108. [Google Scholar] [CrossRef]
  331. ICOLD. Bulletin 130: Risk Assessment in Dam Safety Management; International Commission on Large Dams: Paris, France, 2005. [Google Scholar]
  332. Ljungqvist, K. A Probabilistic Approach to Risk Analysis. A Comparison between Undesirable Indoor Events and Human Sensitivity. Ph.D. Thesis, Luleå University of Technology, Luleå, Sweden, 2005. [Google Scholar]
  333. Stedinger, J.R.; Heath, D.C.; Thompson, K. Risk Analysis for Dam Safety Evaluation: Hydrologic Risk; IWR Report 96-R-13.1996; Cornell University: Ithaca, NY, USA, 1996. [Google Scholar]
  334. Babu, G.L.S.; Srivastava, A. Reliability Analysis of Earth Dams. J. Geotech. Geoenviron. Eng. 2010, 136, 995–998. [Google Scholar] [CrossRef]
  335. Federal Energy Regulatory Commission (FERC). Risk-Informed Decision-Making Guidelines. Version 4.1. 2016. Available online: https://www.ferc.gov/sites/default/files/2020-04/Introduction-to-Risk-Informed-Decision-Making.pdf (accessed on 1 August 2022).
  336. Hartford, D.N.D.; Baecher, G.B. Risk and Uncertainty in Dam Safety; Thomas Telford Publishing: London, UK, 2004. [Google Scholar] [CrossRef]
  337. Federal Emergency Management Agency (FEMA). Federal Guidelines for Dam Safety Risk Management. FEMA P-1025. 2015. Available online: https://www.fema.gov/sites/default/files/2020-08/fema_dam-safety_risk-management_P-1025.pdf (accessed on 27 August 2022).
  338. Simonovic, S.P. From risk management to quantitative disaster resilience- A paradigm shift. Int. J. Saf. Secur. Eng. 2016, 6, 85–95. [Google Scholar] [CrossRef]
  339. Tong, H. A Network Approach to Interdependent Infrastructure Resilience Assessment for Natural Hazards. Ph.D. Thesis, The University of Western Ontario, London, ON, Canada, 2019; p. 6047. Available online: https://ir.lib.uwo.ca/etd/6047 (accessed on 27 August 2023).
  340. Fuller, W. Flood flows. Trans. Am. Soc. Civ. Eng. 1914, 77, 564–617. [Google Scholar] [CrossRef]
  341. Du, T.; Xiong, L.; Xu, C.; Gippel, C.; Guo, S.; Liu, P. Return period and risk analysis of nonstationary low-flow series under climate change. J. Hydrol. 2015, 527, 234–250. [Google Scholar] [CrossRef]
  342. Read, L.K.; Vogel, R.M. Reliability, return periods, and risk under nonstationary. Water Resour. Res. 2015, 51, 6381–6398. [Google Scholar] [CrossRef]
  343. Poduje, A.C.C.; Belli, A.; Haberlandt, U. Dam risk assessment based on univariate versus bivariate statistical approaches: A case study for Argentina. Hydrol. Sci. J. 2014, 59, 2216–2232. [Google Scholar] [CrossRef]
  344. Volpi, E.; Fiori, A.; Grimaldi, S.; Lombardo, F.; Koutsoyiannis, D. One hundred years of return period: Strengths and limitations. Water Resour. Res. 2015, 51, 8570–8585. [Google Scholar] [CrossRef]
  345. Dai, L.; Zhou, J.; Chen, L.; Huang, K.; Wang, Q.; Zha, G. Flood-risk analysis based on a stochastic differential equation method. J. Flood Risk Manag. 2019, 12, e12515. [Google Scholar] [CrossRef]
  346. Jiang, S. Application of stochastic differential equations in risk assessment for flood releases. Hydrol. Sci. J. 1998, 43, 349–360. [Google Scholar] [CrossRef]
  347. Adarsh, S.; Janga Reddy, M. Reliability analysis of composite channels using first-order approximation and Monte-Carlo simulations. Stoch. Environ. Res. Risk Assess. 2013, 27, 477–487. [Google Scholar] [CrossRef]
  348. Marengo, H.; Arreguin, F.; Aldama, A.; Morales, V. Case study: Risk analysis by overtopping of diversion works during dam construction: The La Yesca hydroelectric project, Mexico. Struct. Saf. 2013, 42, 26–34. [Google Scholar] [CrossRef]
  349. Park, Y.; Yeghiazarian, L.; Stedinger, J.; Montemagno, C.D. Numerical approach to cryptosporidium risk assessment using reliability method. Stoch. Environ. Res. Risk Assess. 2008, 22, 169–183. [Google Scholar] [CrossRef]
  350. Ganji, A.; Jowkarshorijeh, L. Advance first order second moment (A.F.O.S.M.) method for single reservoir operation reliability analysis: A case study. Stoch. Environ. Res. Risk Assess. 2012, 26, 33–42. [Google Scholar] [CrossRef]
  351. Xin, C.; Chongshim, G. Risk analysis of gravity dam instability using credibility theory Monte Carlo simulation model. SpringerPlus 2016, 5, 778. [Google Scholar] [CrossRef]
  352. Goodarzi, E.; Teang, M.; Ziaei, L. Risk and uncertainty analysis for dam overtopping Case Study: The Doroudzan dam, Iran. J. Hydro-Environ. Res. 2013, 8, 50–61. [Google Scholar] [CrossRef]
  353. Hariri-Ardebili, M.A. Risk, Reliability, Resilience (R3) and beyond in dam engineering: A state-of-the-art review. Int. J. Disaster Risk Reduct. 2018, 31, 806–831. [Google Scholar] [CrossRef]
  354. Baecher, G.B.; Christian, J.T. Reliability and Statistics in Geotechnical Engineering; Wiley: New York, NY, USA, 2003; ISBN 978-0-471-49833-9. [Google Scholar]
  355. Chen, S.T. Statistics on dam failure. In Reliability and Uncertainty Analyses in Hydraulic Design; Yen, B., Tung, Y., Eds.; ASCE: New York, NY, USA, 1993; pp. 97–105. [Google Scholar]
  356. Pearl, J. Probabilistic Reasoning in Intelligent Systems; Morgan-Kaufmann: San Mateo, CA, USA, 1988. [Google Scholar] [CrossRef]
  357. Hanea, M.; Remus, G. Non-parametric Bayesian networks for parameter estimation in reservoir simulation: A graphical take on the ensemble Kalman filter (Part I). Comput. Geosci. 2013, 17, 929–949. [Google Scholar] [CrossRef]
  358. Tang, X.; Chen, A.; He, J. Optimized variable selection of Bayesian network for dam risk analysis: A case study of earth dams in the United States. J. Hydrol. 2023, 61, 129091. [Google Scholar] [CrossRef]
  359. Wang, T.; Li, Z.; Ge, W.; Zhang, Y.; Jiao, Y.; Sun, H.; Zhang, H. Calculation of dam risk probability of cascade reservoirs considering risk transmission and superposition. J. Hydrol. 2022, 609, 127768. [Google Scholar] [CrossRef]
  360. Chen, Y.; Lin, P. Bayesian network of risk assessment for a super-large dam exposed to multiple natural risk sources. Stoch. Environ. Res. Risk Assess. 2018, 33, 581–592. [Google Scholar] [CrossRef]
  361. Zhang, L.M.; Xu, Y.; Jia, J.; Zhao, C. Diagnosis of embankment dam distresses using Bayesian networks Part, I. Global-level characteristics based on a dam distress database. Can. Geotech. J. 2011, 48, 1630–1644. [Google Scholar] [CrossRef]
  362. El-Awady, A.; Ponnambalam, K. Integration of simulation and Markov Chains to support Bayesian Networks for probabilistic failure analysis of complex systems. Reliab. Eng. Syst. Saf. 2021, 211, 107511. [Google Scholar] [CrossRef]
  363. Badr, A.; Yosri, A.; Hassini, S.; El-Dakhakhni, W. Coupled Continuous-Time Markov Chain–Bayesian Network Model for Dam Failure Risk Prediction. J. Infrastruct. Syst. 2021, 27, 04021041. [Google Scholar] [CrossRef]
  364. Gill, J.C.; Malamud, B.D. Reviewing and visualizing the interactions of natural hazards. Rev. Geophys. 2014, 52, 680–722. [Google Scholar] [CrossRef]
  365. King, L.M.; Simonovic, S.P.; Hartford, D.N.D. Using system dynamics simulation for assessment of hydropower system safety. Water Resour. Res. 2017, 53, 7148–7174. [Google Scholar] [CrossRef]
  366. Hartford, D.N.D.; Baecher, G.B.; Zielinski, P.A.; Patev, R.C.; Ascila, R.; Rytters, K. Operational Safety of Dams Reservoirs; ICE Publishing: London, UK, 2016. [Google Scholar] [CrossRef]
  367. Komey, A.; Deng, Q.; Baecher, G.B.; Zielinski, P.A.; Atkinson, T. Systems Reliability of Flow Control in Dam Safety. In Proceedings of the 12th International Conference on Application of Statistics and Probability in Civil Engineering, ICASP12, Vancouver, BC, Canada, 12–15 July 2015; pp. 1–8. [Google Scholar]
  368. Kwag, S.; Gupta, A. Probabilistic risk assessment framework for structural systems under multiple hazards using Bayesian statistics. Nucl. Eng. Des. 2017, 315, 20–34. [Google Scholar] [CrossRef]
  369. Zomorodian, M.; Lai, S.H.; Homayounfar, M.; Ibrahim, S.; Fatemi, S.E.; El-Shafie, A. The state-of-the-art system dynamics application in integrated water resources modeling. J. Environ. Manag. 2018, 227, 294–304. [Google Scholar] [CrossRef] [PubMed]
  370. Lee, S.; Kang, D. Analyzing the Effectiveness of a Multi-Purpose Dam Using a System Dynamics Model. Water 2020, 12, 1062. [Google Scholar] [CrossRef]
  371. Ignjatovic, L.; Stojkovic, M.; Ivetic, D.; Milašinovic, M.; Milivojevic, N. Quantifying multi-parameter dynamic resilience for complex reservoir systems using failure simulations: A case study of the Pirot reservoir system. Water 2021, 13, 3157. [Google Scholar] [CrossRef]
  372. Salem, S.; Siam, A.; El-Dakhakhni, W.; Tait, M. Probabilistic resilience-guided infrastructure risk management. J. Manag. Eng. 2020, 36, 04020073. [Google Scholar] [CrossRef]
  373. Linkov, I.; Bridges, T.; Creutzig, F.; Decker, J.; Fox-Lent, C.; Kröger, W.; Lambert, J.H.; Levermann, A.; Montreuil, B.; Nathwani, J.; et al. Changing the resilience paradigm. Nat. Clim. Change 2014, 4, 407–409. [Google Scholar] [CrossRef]
  374. Lewis, T. Critical Infrastructure Protection in Homeland Security: Defending a Networked Nation; Wiley: Hoboken, NJ, USA, 2020. [Google Scholar]
  375. Simonovic, S.P.; Arunkumar, R. Comparison of static and dynamic resilience for a multipurpose reservoir operation. Water Resour. Res. 2016, 52, 8630–8649. [Google Scholar] [CrossRef]
  376. Badr, A.; Zoe, L.; El-Dakhakhni, W. Dynamic Resilience Quantification of Hydropower Infrastructure in Multihazard Environments. J. Infrastruct. Syst. 2023, 29, 04023012. [Google Scholar] [CrossRef]
Figure 1. A comparison between the word cloud of the most frequent words in the analyzed documents before (a) and after (b) pre-processing step. (a) Word cloud before pre-processing; (b) Word cloud after pre-processing.
Figure 1. A comparison between the word cloud of the most frequent words in the analyzed documents before (a) and after (b) pre-processing step. (a) Word cloud before pre-processing; (b) Word cloud after pre-processing.
Water 15 03427 g001
Figure 2. Graphical representation of the LDA algorithm’s procedure.
Figure 2. Graphical representation of the LDA algorithm’s procedure.
Water 15 03427 g002
Figure 3. Perplexity plot (Perplexity–Number of topics relationship).
Figure 3. Perplexity plot (Perplexity–Number of topics relationship).
Water 15 03427 g003
Figure 4. The word cloud of the most frequent words per topic. (a) Topic 1: Optimization Models; (b) Topic 2: Climate Change; (c) Topic 3: Flood Risk; (d) Topic 4: Inflow Forecasting; (e) Topic 5: Hydropower Generation; (f) Topic 6: Water Supply Management; (g) Topic 7: Risk-based Assessment and Management.
Figure 4. The word cloud of the most frequent words per topic. (a) Topic 1: Optimization Models; (b) Topic 2: Climate Change; (c) Topic 3: Flood Risk; (d) Topic 4: Inflow Forecasting; (e) Topic 5: Hydropower Generation; (f) Topic 6: Water Supply Management; (g) Topic 7: Risk-based Assessment and Management.
Water 15 03427 g004aWater 15 03427 g004b
Figure 5. Beta Values (β) (the probability of each word occurrence) per topic. (a) Topic 1: Optimization Models; (b) Topic 2: Climate Change; (c) Topic 3: Flood Risk; (d) Topic 4: Inflow Forecasting; (e) Topic 5: Hydropower Generation; (f) Topic 6: Water Supply Management; (g) Topic 7: Risk-based Assessment and Management.
Figure 5. Beta Values (β) (the probability of each word occurrence) per topic. (a) Topic 1: Optimization Models; (b) Topic 2: Climate Change; (c) Topic 3: Flood Risk; (d) Topic 4: Inflow Forecasting; (e) Topic 5: Hydropower Generation; (f) Topic 6: Water Supply Management; (g) Topic 7: Risk-based Assessment and Management.
Water 15 03427 g005aWater 15 03427 g005b
Figure 6. Classification of the optimization modeling tools used in dam operations.
Figure 6. Classification of the optimization modeling tools used in dam operations.
Water 15 03427 g006
Figure 7. Summation of the Gamma values (the probability of each topic occurrence in each analyzed document) of the identified seven topics.
Figure 7. Summation of the Gamma values (the probability of each topic occurrence in each analyzed document) of the identified seven topics.
Water 15 03427 g007
Table 1. Summary of the recent studies conducted on PSH systems (2023–2020).
Table 1. Summary of the recent studies conducted on PSH systems (2023–2020).
Ref.Main GoalCase Study
[239]Investigate the economic and energy efficiency of the coordinated operations between a hybrid PSH, integrating a pumping station between cascading conventional hydropower stations, with wind and photovoltaic systems at different time scales.Wujiang River, China
[240]Investigate the peak shaving of cascade hydropower with mixed PSH to decrease the variation of the residual load of the external power grid, considering the hydraulic coupling of multiple reservoirs and water delay time in the hydropower model.Southwest China
[241]Review and assess the research conducted on the operation, cost, and desired locations of PSH resources in the Railbelt transmission system.Alaska, USA
[242]Develop a new mixed-integer linear optimization model for allocating the PSH unit on the next day of the unit commitment program, using the Autoregressive integral Moving Average (ARIMA) and Stationary Autoregressive integral Moving Average (SARIMA) to predict the market prices of the next day, considering the electricity market uncertainties.Ontario, Canada
[243]Develop a comprehensive power and energy model for PSH integrated with conventional hydropower (CH), aiming to maximize the energy production of the CH, improve PSH operations to consider the discharge variation, achieve a monthly balance between PSH and CH operations, determine the optimum value for PSH and CH based on the range of the charge and discharge ratio, compare capacity optimization of PSH and CH with and without water stream.Balakot hydropower plant, Pakistan
[244]Review the recent advanced pumped storage hydropower technologies, including conventional pumped storage hydropower, adjustable speed pumped storage hydropower, and ternary pumped storage hydropower. N.A.
[245]Investigate the short-term operation of a hybrid wind farm-PSH system transmitting power to multiple cross-regional power grids through ultra-high-voltage transmission lines.Eastern China
[246]Propose an optimal unit model for a coordinated wind-PSH generation system, considering wind uncertainties using a chance-constrained optimization model.Northwest, China
[247]Develop a Quaternary PSH combining the Conventional-PSH and Adjustable speed-PSH technologies, aiming to provide fast power support and efficiently consider the uncertainties of power generation.N.A.
[248]Develop a GIS-based model to transfer a conventional hydropower dam to PSH, determining the most suitable second reservoir location, and calculating its hydroelectricity energy potential. Porsuk Dam, Turkey
[249]Develop a new mixed integer linear programming model for PSH, allowing a greater number of breakpoints, leading to more realistic solutions, and reducing computational efforts.Argentine Republic
[250]Proposed a generic method to model and optimize small-scale PSH that can be applied to different PSH technologies, including adjustable-speed and hydraulic short circuits and PSH configurations such as ternary, reversible, and separate turbines and pumps. Shell Energy North America (SENA) modular PSH concept
[251]Review the existing types of PSH, highlighting the advantages and disadvantages of each PSH system. N.A.
Table 2. Summary of the recent hydropower scheduling studies (2023–2018).
Table 2. Summary of the recent hydropower scheduling studies (2023–2018).
Ref.Main GoalCase Study
Short-Term Hydropower Scheduling (STHS)
[257]Propose a dynamic capacity short-term scheduling model based on one-dimensional unsteady flow for Three Gorges in China.Three Gorges, China
[258]Propose a two-step technique based on mixed-integer linear programming (MILP) to solve the STHS by exploring the symmetry related to the identical generating units in San Antonio and obtaining the optimal load distribution.San Antonio, Brazil
[259]Develop a linear mixed-integer optimization model to maximize total hydropower generation among all periods of the operation year, considering the production rate at the maximum storage for all possible operational scenarios of the turbine and the most efficient point of water discharge.Saguenay-Lac-St-Jean, Quebec, Canada.
[260]Develop a rolling horizon robust online scheduling scheme that employs stochastic optimization within a feedback model to consider the uncertainties in hydropower operation, including electricity prices, power demand, inflow, and model parameters.Ontario, Canada
[261]Compare five mixed-integer linear programming formulations for efficient STHS.Hypothetical three hydropower plant system
[38]Identify a quick decision approach for STHS for cascaded hydropower plants using data mining algorithms.Tianshengqiao, Guangdong, China.
[262]Propose a dynamic determination process for the breakpoints of the piecewise linear approximation method for solving STHS, considering the intake loss, efficiency of the head-dependent turbine, penstock loss, the varying time head effect, and tailrace loss.Northern Norway
[263]Develop mixed-linear programming for hourly hydropower scheduling applied to a cascading hydropower system with hydraulic coupling and multiple prohibited zones for head-dependent operation.Tianshengqiao, Hongshui River, southwest China
[264]Develop a multi-objective mixed-integer non-linear programming for the peak shaving operations to enhance power efficiency and control the peak loads.Beipanjiang, China
[265]Propose a stochastic multicriteria decision-making framework for STHS under multiple uncertainties.Qingjiang, Hubei, China
[266]Determine the efficiency of genetic algorithms with different selection operators for optimizing the STHS problem.Gezhouba, China
[267]Develop a rolling optimal operation model for hourly (ultra) STHS in real-time for cascading hydropower dam system.Yunnan, China
[268]Develop a short-term peak-shaving method using fuzzy clustering analysis and linear mixed-integer programming for cascading hydropower systems with sensitive heads, considering water spillage adjustments.Hongshui, China
[269]Develop an improved cloud adaptive quantum-inspired binary social spider optimization model to optimize short-term scheduling sub-problems, economic load dispatch, and unit commitment.Three Gorges hydropower station, Yangtze, China.
[270]Develop an advanced optimization model to improve the hydropower economic profit of a large chain of reservoirs, considering ecological restrictions, European regulations, and operation uncertainties.Guadalquivir, southern Spain.
[271]Develop a successive approximation approach to consider the STHS non-linear characteristics.Ten Reservoir System, China
[272]Used linear and dynamic programming to determine the number of operating units and the power in each quarter-hour and for each hydro plant. Yunnan, China
Mid-Term Hydropower Scheduling (MTHS)
[273]Develop a successive quadratic programming model with linearization updated by the non-linear constraints of the hydropower facility composed of cascading reservoirs.Jinsha River, China
[274]Develop a bi-level stochastic model based on information gap decision-making theory for MTHS of cascading hydropower systems, considering the economic and hydrologic uncertainties in hydropower market participation.Southwest China
[275]Develop a two-stage decomposition approach for the MTHS problem, where every state is represented as a multi-period stochastic model.Brazil
[276]Develop an MTHS approach acting as a price-maker in the automatic frequency restoration reserve market.Northwest Spain
[277]Develop a stochastic optimization model for mid-term scheduling using Latin hypercube sampling and Cholesky decomposition coupled with sensitivity analysis, considering the uncertainty of natural inflows.China
[278]Apply stochastic dual dynamic integer programming (SDDIP) to a non-convex MTHS problem.Norwegian hydropower
Long-Term Hydropower Scheduling (LTHS)
[279]Improves the traditional weekly hydropower scheduling by integrating the hourly power and capacity balances (HPCB), which are formulated in the form of mixed-integer linear constraints involving the spare, reserve, maintenance, disabled, and working capacities, besides optimizing the levels and orders of hydro plants in peak-shaving the hourly power load curve.Lancang River, China
[280]Develop improved dynamic programming with successive approximation at each stage and relaxation strategy to solve the joint optimal operation problem of the large-scale hydropower plan groups.Yangtze River, China
[281]Illustrate the advantages of coupling the periodic autoregressive and the moving average (PARMA) over using only the periodic autoregressive (PAR) model to represent inflow uncertainties within the SDDP optimization model of LTHS.Quebec, Canada
[46]Develop a hybrid linear and nonlinear hydropower reservoir optimization model to provide an efficient and faster solution for the LTHS problem.California, USA
[282]Develop the LHTS approach to consider inflow forecasting uncertainty based on the adaptive nearest neighbor Gaussian temporal disaggregation method.Xiluodu-Xiangjiaba, China
[283]Develop an improved optimal method to control water levels, considering the two-stage analysis of LTHS and adjustable policy for the target outflow.Xiluodu and Three Gorges, China
[284]Develop a hybrid decomposition-coordination and discrete differential dynamic programming model (IDC–DDDP) for solving the LTHS for large-scale hydropower systems.Southwestern China
[285]Develop a multi-objective quantum-behaved particle swarm optimization model based on improved Tchebycheff decomposition with a modified generator of direction vectors to maximize the total production rate and firm the hydropower output.Three Gorges and Gezhouba, China
[286]Develop a linear mixed-integer optimization model to consider peak shaving demands, aiming to maximize the power generation profit within the LTHS of an interprovincial hydropower plant.Xiluodu, China
[287]Develop a chaotic adaptive multi-objective bat algorithm for cascade hydropower dams.Qingjiang, southern China
[288]Develop an adaptive multi-objective particle swarm optimization model based on the dominance and decomposition of a multi-objective long-term generation.Three Gorges, China
[289]Develop an improved differential evolution algorithm based on the LSHADE evolutionary algorithm, using new mutation strategies to provide a wider search range and accelerate convergence.Jinsha River, western China.
[290]Develop a series division method based on particle swarm optimization and firefly algorithm for the LTHS problem.Himreen lake, Diyala/Iraq
[291]Develop a novel optimization model based on copula theory to consider the uncertainty of electricity prices and the correlation between multiple markets.Wu river, southwestern
[292]Develop a novel piecewise linearization method for LTHS that converts the non-linear problem to a linear programming problem without using integer variables.Lancang River, southwest of China.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Badr, A.; Li, Z.; El-Dakhakhni, W. Dam System and Reservoir Operational Safety: A Meta-Research. Water 2023, 15, 3427. https://0-doi-org.brum.beds.ac.uk/10.3390/w15193427

AMA Style

Badr A, Li Z, El-Dakhakhni W. Dam System and Reservoir Operational Safety: A Meta-Research. Water. 2023; 15(19):3427. https://0-doi-org.brum.beds.ac.uk/10.3390/w15193427

Chicago/Turabian Style

Badr, Ahmed, Zoe Li, and Wael El-Dakhakhni. 2023. "Dam System and Reservoir Operational Safety: A Meta-Research" Water 15, no. 19: 3427. https://0-doi-org.brum.beds.ac.uk/10.3390/w15193427

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

Article Metrics

Back to TopTop