Next Article in Journal
Burning Water, Overview of the Contribution of Arjen Hoekstra to the Water Energy Nexus
Next Article in Special Issue
Assessing the Effectiveness of Sustainable Drainage Systems (SuDS): Interventions, Impacts and Challenges
Previous Article in Journal
Reduction of Environmental Impacts Due to Using Permeable Pavements to Harvest Stormwater
Previous Article in Special Issue
Evaluation of Permeable Brick Pavement on the Reduction of Stormwater Runoff Using a Coupled Hydrological Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On the Use of a Real-Time Control Approach for Urban Stormwater Management

1
Department of Environmental Engineering, University of Calabria, 87036 Rende (CS), Italy
2
Department of Civil Engineering, University of Calabria, 87036 Rende (CS), Italy
3
Department of Mechanical, Energy and Management Engineering, University of Calabria, 87036 Rende, Italy
4
CNR-National Research Council of Italy, Institute for High Performance Computing and Networking (ICAR), 87036 Rende (CS), Italy
*
Author to whom correspondence should be addressed.
Water 2020, 12(10), 2842; https://doi.org/10.3390/w12102842
Submission received: 10 September 2020 / Revised: 7 October 2020 / Accepted: 10 October 2020 / Published: 13 October 2020

Abstract

:
The real-time control (RTC) system is a valid and cost-effective solution for urban stormwater management. This paper aims to evaluate the beneficial effect on urban flooding risk mitigation produced by applying RTC techniques to an urban drainage network by considering different control configuration scenarios. To achieve the aim, a distributed real-time system, validated in previous studies, was considered. This approach uses a smart moveable gates system, controlled by software agents, managed by a swarm intelligence algorithm. By running the different scenarios by a customized version of the Storm Water Management Model (SWMM), the findings obtained show a redistribution of conduits filling degrees, exploiting the whole system storage capacity, with a significant reduction of node flooding and total flood volume.

1. Introduction

1.1. Background

As a significant urban infrastructure, combined sewer systems are characterized by complexity and management problems [1]. Nowadays, the relationship between the urban area and Urban Drainage System (UDS) is proving to be potentially damaging for the community and urban structures [2,3,4]. In this regard, the growth of impervious surfaces due to the ongoing urbanization led to a drastic change in the natural hydrological cycle, with a significant surface runoff [5,6,7,8]. Moreover, due to climate change, the intense rainfall phenomena amplify this issue [9,10], pushing the existing drainage networks beyond their capacity [11,12,13,14]. Therefore, the excess flow out of the sewer system intensifies the frequency and amplitude of critical phenomena such as local flooding and combined sewer overflows (CSOs), increasing the vulnerability of urban systems [13,15,16,17]. Urban flooding occurs when the UDS, being overloaded, enters a crisis by threatening property and infrastructure, with adverse consequences for human life, economic activities, and the environment [10,18]. This phenomenon is generally worsened by drainage networks’ design inadequacy and incorrect and low maintenance.
Therefore, a transition from the conventional approaches towards innovative, sustainable, and smart strategies such as Low-Impact Development (LID) techniques and real-time control (RTC) systems is necessary [19,20,21,22,23,24,25,26,27,28,29,30,31,32,33]. In this regard, recent scientific advances are focusing on the use of real-time control (RTC) approaches for the smart optimization of the urban drainage network. This technology’s main advantage is to operate in real-time on the fluctuating conditions of the drainage systems, which are generally designed for a static load but operating under dynamic load. Usually, the drainage network’s global capacity is not totally exploited (some system sections are poorly used or unused while others are overloaded) or is used in an inadequate way, preventing the achievement of the proper management of the system. The potential benefit of the RTC approach grows as well as the difference between the expected “design load” and the actual recorded “service load” [34]. The RTC system is a reliable and cost-effective methodology: it improves the hydraulic performance and helps the drainage system achieve its operational objectives [35,36,37], allowing temporary storage of rainwater volumes directly in existing networks and avoiding huge investments. The use of existing infrastructure for the outflow of wastewater and its control and management makes the economic feature a critical factor in choosing a RTC strategy. RTC cost efficiency is demonstrated by simulation studies [38] and by large-scale applications [39].

1.2. Real-Time Control Approach

In an urban drainage system, the RTC system allows to more easily achieve several operational objectives: minimization of overflow volumes and frequencies, flood prevention, management of discharge peaks to the wastewater treatment plant (WWTP), management of flows during a system disturbance (work deviations, equipment failures, or safety incidents), cost reduction, and other quality aspects [33].
The study carried out in Reference [40] shows that RTC in UDS is an efficient approach to improve performance and reduce the impact on the natural water environment. In recent years, theoretical studies (drainage system + RTC modeling) and technological innovations allowed the spread of RTC systems’ implementation. Methodologies and equipment are available and well adaptable for different UDS uses [41]. The applicability of these devices has widened by recent developments in radar now-casting [42]; moreover, online measurements and available computational capacity have improved these tools’ potential application field.
A RTC system dynamically regulates the drainage system, implementing a feedback loop based on online measurements to achieve specific operational objectives and improve the system’s overall performance. Operational strategies and algorithms regulate the system operations according to the current state and dynamic network conditions detected in “real-time” [43]. A RTC system, equipped with specific devices (sensors, controllers, actuators), can coordinate multiple functionalities, such as (i) monitoring the current state of the network, (ii) comparing the current state against the wanted one, (iii) defining the settings of the control structures to achieve the wanted state, and (iv) determining the physical actions on the final control actuators. Thus, all the equipment allows the system’s dynamic management according to the current conditions and the system’s critical event [44].
The criteria to assess if a RTC strategy is suitable for a UDS are conditioned by several aspects, rather tricky to determine concerning the specific system features [45]. Several studies were carried out to establish the necessary standard and RTC application aspects. An example is the RTC guidelines [46] by the DWA (German Association for Water, Wastewater, and Waste). Moreover, some software tools were developed to support during the decision-making process, e.g., the planning tool called PASST (planning aid for sewer system real-time control) [38,46].
The critical analysis of the above factors allows the definition of a “control strategy”. The RTC possible configurations range from simple and direct controls carried out at the “local” level (punctual or regional) to more complex “global” controls, concerning the whole system, with an “optimal predictive global” configuration [1].
So far, many studies were focused on a centralized RTC’s approach [37,47,48]. However, given the huge amount of data to be read, managed, and processed, this type of control system, although reliable, presents some problems. Generally, in this approach, all data collected by sensors are sent to a central unit that produces the command for the actuators based on a specific strategy. Therefore, based on its functioning, a similar approach requires efficient connections among all elements (sensors and actuators) and a complex mathematical model. Moreover, one node’s failure can compromise the whole system behavior [49].
Recently, Kändler et al. [50] have developed a new concept of a smart in-line storage system that does not need an advanced centralized control system, but it is easy to install and operates by real-time controlled actuators, and it can predict rainfall dynamics.
Moreover, among the literature studies on RTC developed in the last years, some have implemented control algorithms, without network online mathematical models. They use heuristic algorithms, based on externally imposed rules and previous knowledge of the drainage system (derived from experience and/or separate simulation procedures). The two most commonly used algorithms are Rule-Based Control (RBC) [51,52,53] and Fuzzy Logic Control (FLC) [54,55,56,57,58,59,60].
Other examples of optimization control algorithms’ implementation, with mathematical network models, are run online to calculate set-points and determine control actions. These algorithms (reactive and predictive) are quite complex, especially the multi-objective control, which is the most frequent and focuses on the concept of Pareto optimal solution. The most commonly used optimization algorithms are the Linear-Quadratic Regulator (LQR) [61], the Evolutionary Algorithms (EA) [62,63,64,65,66], the Model Predictive Control (MPC) [67,68,69,70,71,72,73,74,75,76,77,78], and the Population Dynamics (PD) [79,80]. Studies focusing on implementing and validating specific software to simulate the RTC control systems’ operation can be found in References [26,81,82,83,84,85].

1.3. Aims of the Study

From this literature analysis, some limitations about the application of a centralized real-time control approach emerge. Therefore, to overcome the previously highlighted limits, this study aims to show the hydrological efficiency of a distributed real-time control (DRTC) approach.
The hydrological behavior of a real drainage system of a highly urbanized area is modeled, and the hydrological response of uncontrolled and RTC-controlled scenarios under different rainfall events is investigated.
This approach aims to exploit the full storage conduits capacity in the drainage network using moveable and smart gates that, managed by the decentralized swarm intelligence algorithm (gossip-based), accumulate the excess stormwater volume in the system sections less overloaded.
The dynamic simulation model (with control) used, which works by customizing SWMM software that communicates in real-time with an external Java multi-agent software, was defined and tested in previous studies [49,86]. However, this study aims to take a step further than the previous ones by focusing on the optimal and most efficient moveable gates location analysis.

2. Materials and Methods

Based on the study’s primary objective, a real-time distributed system control (DRTC) was considered to overcome the problems related to urban stormwater management. The whole methodology used to carry out the analysis, from the data retrieval useful for the hydrodynamic model development and then to the analysis of the system’s criticalities, until the DRTC approach implementation and analysis of its efficiency, is summarized in the flowchart in Figure 1. At the same time, further details can be found in the following methodology’s sub-sections.

2.1. Case Study

In this study, the DRTC approach was applied to a highly urbanized district of an urban catchment located in Paola, Southern Italy (CS), in a Mediterranean climate region (Figure 2), characterized by hot, dry summers, and cool, wet winters.
The general hydrological and hydraulic data and necessary information were taken from a previous study [87]. The specific area considered for the study is a highly urbanized area with a total area of 7.6 ha, out of which 96.0% are impervious, covered by roofing (28.9%), roads, parking lots, and other impermeable surfaces (67.1%), while only about 4.0% of the total surface area is a pervious area [87].
The drainage network, built at different times, is a combined sewer system that is not totally efficient due to frequent rainwater discharge into the sewer, low slopes, undersized conduits, etc. Therefore, the specific examined area reproduces the issues related to a general drainage system in a highly urbanized area, with the advantage that given its dimension, it can be modeled with greater detail. Based on the detailed information retrieved by the previous study [87] for the model development, it was possible to analyze the drainage network’s criticalities, identify the best moveable gates location, and assess the distributed real-time control strategy efficiency.

2.2. Data Analysis

Design storms and measured rainfall events were considered to carry out the analysis.
More in detail, in the first phase, probabilistic analysis was conducted to determine the intensity–duration–frequency (IDF) curves for a Return Period (RP) of 1, 2, 3, 4, 5, 10, 20, 30, 40, and 50 years.
The IDF curves are generally considered a useful tool for standard hydrological risk analysis [88] and for the design, construction, and management process of several hydraulic engineering projects that present natural hazards due to extreme rainfall events [89].
The IDF curve represents a mathematical equation among rainfall intensity, duration, and return period [90]. More in detail, it can be expressed as h d , T = a T d n (where h is the precipitation height, d is the duration, T is the return period, while a(T) and n are two parameters, that have to be estimated by a probabilistic approach) [91] or based on the nomenclature we used as PD D , RP = aD n .
To define the IDF curve, it is necessary historical data of good quality, recorded for a continuous long time [90]. In this regard, for the specific case study, after the analysis of the available historical data for a long continuous period, the IDF relationships were computed by considering the historical records (collected between 1945 and 2005) of the annual maximum series for rainfall durations of 1, 3, 6, 12, and 24 h (obtained from the Regional Environmental Protection Agency of Calabria (ARPACAl) [92], Paola site, cod. 3060). To develop the IDF curve, the Gumbel extreme values distribution was used as cumulative probability function; for each duration and selected return periods, the rainfall depths were calculated and the least-square method was considered to determine the parameters (a and n) of the empirical IDF. The rainfall depth reported in a graph (PD, D) were interpolated by the power function previously described, identifying the IDF curve. More detail on the procedure used to develop the IDF curve can be found in References [90,91].
Based on the defined IDF relationship, the Chicago Hyetographs, proposed by Reference [93] to evaluate the design rainfall for the urban sewer system, was calculated. More in detail, this design storm event, based on two analytical equations, one valid before the peak position and the other after it, was chosen to generate a synthetic rainfall event with the maximum intensity. Therefore, to evaluate the system’s response under different design storms, in this study, starting from the IDF curves of 5, 10, and 20 years return periods, the Chicago Hyetographs of 1 h were developed.
In the second phase, to select recorded rainfall events, 5 min cumulative rainfall series collected from May 2019 to May 2020 on the Paola rain gauge ([92] Paola site, cod. 3060) was considered. More in detail, from all collected data, the individual events were defined as being separated by a continuous dry period of at least six hours, and only the events with a precipitation depth greater than 2 mm were selected [94,95,96].
Therefore, for each selected rainfall event, the hydrological characteristics in terms of precipitation depth (PD), rainfall duration (D), and rainfall intensity (i) were evaluated.
Moreover, since the IDF curves gave information of the frequency of extreme rainfall events for a variety of durations and intensities, the selected measured rainfall events were compared with the historical records obtained from the Regional Environmental Protection Agency of Calabria (ARPACAl) through the intensity–duration–frequency (IDF) previously obtained. At the end of this analysis, only the most significant rainfall events that caused the specific drainage network crises were considered for the RTC simulations.

2.3. The Distributed Real-Time Control Approach

In this study, a real-time distributed system control (DRTC) approach was used. The methodology used involves the use of a specific dynamic simulation model with control. This real-time control strategy refers to the model defined and tested in previous studies [49,86]. The model was created with customization of the Storm Water Management Model (SWMM) (provided by EPA [45] as a hydrodynamic model), to communicate in real-time with a separate Java external software, based on a multi-agent model and a gossip-based algorithm (swarm intelligence).
However, although SWMM allows some real-time control of the parameters of the network (by defining some simple rules, if-then-else), through the “coupling strategy,” the method exploits the potential of the two different software to achieve better system performance and to adjust the hydraulic network parameters dynamically in real-time.
The overall control was distributed over a nodes network, and it was realized using a multi-agent paradigm and a swarm intelligence algorithm. These algorithms define simple entities’ behavior, interacting with each other, produce an emerging property [97]. Thus, simple entities are represented by agents “intelligent” gates and the emerging property is the balancing of conduit capacities to optimize the adequate sewer network storage capacity.
The algorithm structure presents adaptivity and fault-tolerance properties since the procedure’s iterativity ensures that convergence [98] cancels the effect of unforeseen events.
In a real case, to achieve the proposed goals, the network requires the effective distributed implementation of hardware elements, as explained in References [49,86]: sensors, one water level sensor per conduit and a flow sensor on the outfall, computational nodes, which can host and execute the distributed control algorithm, and down-hinged moveable gates, which can be real-time-regulated electronically.
The study carried out by the authors of References [49,86] explains that computational nodes, distributed throughout the network, dynamically regulate the gates according to the information acquired by the sensors in the neighbor areas. Each computational node has a partial view of the network as it can read only from the sensors and actuate only on the gates which can physically reach (peer-to-peer communication). After data reading, the nodes collectively process the acquired information to trigger suitable activation on the gates. The collective computation of the network of nodes supplies the gates with “intelligent” behavior.
The algorithm consists of balancing the water level perceived by the agents. This balance is obtained by the agents that perform two tasks in a repeated sequence: (i) estimate the average water level in the network (gossip-based algorithm), and (ii) properly adjust the specific gate to bring the water level as close to this average as possible.
The gates, mobile and electronically operable, can manage the conduits’ capacity by a Proportional Integrative Derivative (PID) controller [86]. A PID controller consists of a control loop that minimizes an error value consisting of the difference between controlling the process’ output and the desired value (set-point). Through an electromechanical action on the plate, the gates opening degree, as the water flow at the network’s points where the gates are located, is modified based on a “desired” value.
This methodology exploits full storage conduits capacity by accumulating the excess stormwater volume in the system’s less overloaded parts.

2.4. Modeling Scenarios

To evaluate the DRTC’s hydraulic performance, the drainage network’s response was modeled by considering different scenarios, depending on moveable gates’ number and positions across the network.
Scenario 0, or reference scenario, corresponding to the uncontrolled drainage system’s current situation, was modeled as reported in Reference [87]. The model area was subdivided into 26 sub-catchments, while the drainage network was modeled by considering 29 nodes and 28 conduits modeled. The network presents variability in conduits’ geometry and materials: concrete conduits with rectangular cross-section of different sizes (diameters 400 × 400 mm and 450 × 450 mm) and stoneware conduits presenting a circular cross-section (diameters of 200 and 300 mm).
To model Scenarios 1 and 2, the software deriving from the SWMM customization with the specific control algorithm (called SWMM-RTC), previously described, was considered. The use of this software has required the network re-modeling, with the introduction of smart moveable gates, modeled in SWMM as a transverse weir with the opening area equal to the conduit section area. More in detail, as shown in Figure 3, Scenario 1 was modeled by applying moveable gates in all conduits, while Scenario 2 presents specific control, with the gates located on the main network, downstream of each sub-network, at the points where the network branches (sub-networks) engage in the main channel.
The Soil Conservation Service (SCS) Curve Number (CN) method was used for the infiltration assessment. More in detail, for the CN estimation, based on the retrieved data for each sub-catchment, the values tabulated in Reference [99] were considered, while the flow routing computations were based on the Dynamic Wave Equations.

2.5. Hydrological Performance Indicators

To evaluate the beneficial effect of the use of the DRTC approach, the results in terms of node flooding (NF), total hours flooded (HF) by nodes, total flood volume (FV), and conduit surcharge (CS) were analyzed, and the performance indicators, reported below, were estimated.
(1)
Flood Volume Reduction (FVRn–m) was expressed as the percentage difference between the total Flood Volume (FV) of n and m scenarios:
FVR n m   % = FV n FV m FV n × 100
(2)
Hours Flooded Reduction (HFRn–m) was determined as the percentage difference between the Hours Flooded (HF) of n and m scenarios:
HFR n m %   = HF n HF m HF n × 100  
(3)
The percentage of nodes flooding (NFn) was calculated concerning the total number of nodes (NNtot):
NF n   % = 100 NN tot NF n NN tot × 100
(4)
The percentage of conduit surcharge (CSn) was evaluated concerning the total number of conduits (NCtot):
CS n %   = 100 NC tot CS n NC tot × 100  
Moreover, to analyze the balancing of the flow within the drainage system for the network branches (sub-networks) and the main channel, a comparative analysis of the conduit’s filling degree was carried out by graphs showing the conduit’s capacity trend over time.

3. Results and Discussion

3.1. Rainfall Events

The design storm events (Chicago Hyetographs), determined as discussed in Section 2.2, present a rainfall duration of 1 h and a return period of 5, 10, and 20 years, with total precipitation of 70.9, 81.6, and 91.9 mm, respectively. The model was first run by considering these three extreme rainfall events characterized by short duration and high intensity to evaluate its efficiency.
While, concerning the evaluation of the recorded rainfall events, as reported in Table 1, the whole studied period was characterized by 54 rainy events, for a total precipitation depth (PD) of 923.4 mm ranging from 2.0 to 105.2 mm with a mean value of 17.1 mm. For the whole dataset: 46.3% of the rainfall events had a precipitation depth less than 10 mm, 25.9% had a precipitation depth between 10 and 20 mm, 7.4% had a precipitation depth between 20 and 30 mm, 11.1% had a precipitation depth between 30 and 40 mm, 5.6% had a precipitation depth between 40 and 50 mm, 1.9% (one event) had a precipitation depth of 74.6 mm, and one event had a precipitation depth of 105.2 mm.
Figure 4 compares all the 54 recorded rainfall events, in terms of total rainfall depth and duration, with the intensity–duration–frequency (IDF) curves. By analyzing this figure, it emerges that most of the rainfall events fall below the one-year return period threshold, eight events present a return period greater than one year, with one greater than two years, one on the ten-year IDF curve, and one relatively near to the 40-year threshold.
The critical real rainfall events, whose hydrological characteristics are summarized in Table 2, were considered to analyze the different scenarios.
First, a comparative evaluation analysis in terms of precipitation depth, mean rainfall intensity, and maximum rainfall intensity was carried for the whole dataset. This analysis was principally based on the different hydrological parameters’ influence on the sewer system’s response. For instance, after a first analysis, the rainfall event with a maximum precipitation depth of 105.2 mm (occurred 25 March 2020) and a mean intensity of 2.2 mm/h did not provide local flooding phenomena. Then, all events with a mean intensity higher than its average value (5.8 mm/h) were considered, obtaining that only the seven rainfall events shown in Table 2 caused network criticalities in terms of node flooding volume for Scenario 0.
Moreover, by analyzing the hydraulic response of the uncontrolled scenario, it was noted that all seven events had a maximum rainfall intensity higher than the average value of the same parameter (21.4 mm/h) for the whole dataset and a precipitation depth higher than the average value (17.1 mm).

3.2. DRTC Approach Efficiency

To evaluate the hydrological performance of the real-time control approach applied in both scenarios (1 and 2), the models’ response, run by considering a total of 10 rainfall events, including the three designs storm events and the seven real ones, was analyzed.
The findings obtained by considering the design storm events, in terms of the number of Node Flooding (NF), Hours Flooded (HF), Conduit Surcharge (CS), and total Flood Volume (FV) for each scenario and each design rainfall event, are shown in Table 3.
By analyzing Table 3 emerges, as expected, growing values of hours flooded and total flood volume as the return period increases.
By comparing the different scenarios for the same return period, it is possible to observe an improvement of the drainage network efficiency with relevant mitigation of flooding produced using a distributed real-time control approach.
Based on the results obtained in Table 3, the performance indexes were evaluated as described in Section 2.5, obtaining the findings described in Table 4, Table 5, Table 6 and Table 7.
Table 4 and Table 5 report the results obtained in terms of Flood Volume Reduction (FVRn–m) and Hour Flooded Reduction (HFRn–m), assessed by comparing the uncontrolled scenario (Scenario 0) with the two real-time controlled scenarios (Scenarios 1 and 2) and then comparing the two controlled scenarios with each other.
While, the results obtained for the third performance indicator (NF), which represents a percentage measure of the number of nodes flooding compared to the total node number, are shown in Table 6.
The results in terms of conduit surcharge for each scenario and design storm event are shown in Table 7.
The findings show that real-time controlled scenarios, 1 and 2, present higher efficiency than the uncontrolled scenario (Scenario 0), achieving the best performance for Scenario 2, where the movable gate’s location was defined after the drainage network hydraulic conditions analysis.
Based on this finding, the analysis was carried out for the seven critical recorded rainfall events by comparing only the uncontrolled Scenario 0 and the real-time controlled Scenario 2, which had presented the better performance indexes for the design storm events. In this regard, Table 8 shows the results obtained in terms of Node flooding (NF), Hours flooded (HF), Total Flood Volume (FV), and Conduits Surcharge (CS) parameters.
By analyzing Table 8, it emerges that Scenario 2 does not only improve the network hydraulic performance but also achieves maximum efficiency in terms of the total reduction of the number of nodes flooding and a considerable reduction of surcharge conduits.
Finally, to evaluate the balancing of the flows within the drainage network, achieved by using the RTC strategy, a comparative analysis was carried out by considering that the filling degree of each conduit belongs to the individual sub-network and the main network. In this regard, the graphs showing the conduit capacity trend over time for RP of 10 years and each scenario are reported in Figure 5 for four conduits (6, 8, 10, and 27) and in the Appendix A (Figure A1, Figure A2, Figure A3, Figure A4, Figure A5, Figure A6, Figure A7 and Figure A8) for all conduits.
These figures confirm that by implementing RTC control strategies (especially in Scenario 2, specific control), achieving a relevant improvement of the drainage network performance is possible. This efficiency is obtained in terms of overflow attenuation and reducing the overflow duration (as shown for conduits 6 and 8 in Figure 5) with a simultaneous capacity of flow stationing inside conduits for a long time (for instance, conduits 27, 6, and 8 in Figure 5; for graphic reasons, the simulation was stopped at one day). While for conduit 10 in Figure 5, it is observed at an almost total reduction of flow.
Observing the graphs for all conduits, a flow balancing with a redistribution of the filling degree within the whole drainage system emerges, which is often below the threshold value of 0.75 in Scenario 2. This significant finding was achieved by exploiting all sewer system reservoir capacity by involving the flows from the overloaded conduits to those less charged. Moreover, in some conduits, a reduction and delay of the peak capacity were achieved.
Table 2 is the best for flow balancing and total flooding volume reduction. Moreover, this scenario also presents considerable economic benefits, with a lower cost of implementation due to the fewer moveable gates installed than in Scenario 1.

4. Conclusions

Real-time control strategies applied to an existing drainage network represent a solution for optimizing urban stormwater management and minimizing urban flooding risk.
In this study, the behavior of a real drainage network of a highly urbanized area was modeled, and the hydrological response of uncontrolled and RTC controlled scenarios to different extreme rainfall events was analyzed. A distributed real-time control (DRTC) system, defined and tested in previous studies, was used to achieve the aim. This approach was created by customizing SWMM software that communicates in real-time with an external Java multi-agent software. To exploit all conduits’ reservoir capacity in the drainage network, moveable and smart gates, controlled by software agents, managed by a decentralized swarm intelligence algorithm (gossip-based), were implemented in the model.
The results demonstrated the DRTC approach’s validity in system control. A conduit filling degree redistribution, in many cases below 75%, was achieved. Moreover, for some events, a reduction and delay of peak capacity was observed. The CS and NF indicators showed positive findings, with a generalized decrease in conduits’ overload and node flooding. The flow balancing also produced its effects on FVRn-m and HFRn-m indexed reduction, assessed by comparing the scenarios. Scenario 2 achieved the maximum hydrological efficiency for the system in terms of a complete reduction in the number of flooding nodes and an acceptable reduction of surcharge conduits.
Therefore, analyzing the main findings produces a positive qualitative evaluation of the DRTC approach in terms of network hydraulic performance efficiency and environmental vulnerability reduction. This approach allows a significant increase in the system flexibility to react to unexpected scenarios.
Finally, the economic aspect is also crucial for the RTC system choice. This approach allows a general system performance improvement and reduces critical issues, avoiding significant and costly investments through existing infrastructure reorganization.

Author Contributions

Conceptualization, M.M., S.A.P., and P.P.; methodology, M.M., S.A.P., A.V., G.S., and P.P.; software, A.V., G.S., S.A.P., and P.P.; data curation, S.A.P., A.C.B., B.P., and M.T.; formal analysis, S.A.P. and A.C.B.; investigation, M.M., S.A.P., A.C.B., and P.P.; writing—original draft preparation, S.A.P. and A.C.B.; writing—review and editing, M.M., S.A.P., A.C.B., B.P., M.T., and P.P.; supervision: M.M., G.S., and P.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Figure A1. Conduits’ capacity in the main network, highlighted in red in the first picture where the drainage network schematization is shown.
Figure A1. Conduits’ capacity in the main network, highlighted in red in the first picture where the drainage network schematization is shown.
Water 12 02842 g0a1
Figure A2. Conduits’ capacity in the main network.
Figure A2. Conduits’ capacity in the main network.
Water 12 02842 g0a2
Figure A3. Conduits’ capacity in sub-network 1, highlighted in red in the first picture where the drainage network schematization is shown.
Figure A3. Conduits’ capacity in sub-network 1, highlighted in red in the first picture where the drainage network schematization is shown.
Water 12 02842 g0a3
Figure A4. Conduits’ capacity in sub-network 2, highlighted in red in the first picture where the drainage network schematization is shown.
Figure A4. Conduits’ capacity in sub-network 2, highlighted in red in the first picture where the drainage network schematization is shown.
Water 12 02842 g0a4
Figure A5. Conduits’ capacity in sub-network 3, highlighted in red in the first picture where the drainage network schematization is shown.
Figure A5. Conduits’ capacity in sub-network 3, highlighted in red in the first picture where the drainage network schematization is shown.
Water 12 02842 g0a5
Figure A6. Conduits’ capacity in sub-network 4, highlighted in red in the first picture where the drainage network schematization is shown.
Figure A6. Conduits’ capacity in sub-network 4, highlighted in red in the first picture where the drainage network schematization is shown.
Water 12 02842 g0a6
Figure A7. Conduits’ capacity in sub-network 5, highlighted in red in the first picture where the drainage network schematization is shown.
Figure A7. Conduits’ capacity in sub-network 5, highlighted in red in the first picture where the drainage network schematization is shown.
Water 12 02842 g0a7
Figure A8. Conduits’ capacity in sub-network 6, highlighted in red in the first picture where the drainage network schematization is shown.
Figure A8. Conduits’ capacity in sub-network 6, highlighted in red in the first picture where the drainage network schematization is shown.
Water 12 02842 g0a8

References

  1. United States Environmental Protection Agengy (US EPA). Real Time Control of Urban Drainage Networks; US EPA: Washington, DC, USA, 2006. [Google Scholar]
  2. Kaźmierczak, A.; Cavan, G. Surface water flooding risk to urban communities: Analysis of vulnerability, hazard and exposure. Landsc. Urban Plan. 2011, 103, 185–197. [Google Scholar] [CrossRef]
  3. Aronica, G.T.; Palla, A.; Lanza, G.L. Drenaggio delle acque meteoriche e rischio di allagamento in area urbana. In Focus su AcQuE E AmbIEntE uRbAno. Qualità Dell’ambiente Urbano IX Rapporto; ISPRA—Istituto Superiore per la Protezione e la Ricerca Ambientale: Roma, Italy, 2013; pp. 189–195. ISBN 9788844806224. [Google Scholar]
  4. Piro, P.; Carbone, M. A modelling approach to assessing variations of total suspended solids (tss) mass fluxes during storm events. Hydrol. Process. 2014, 28, 2419–2426. [Google Scholar] [CrossRef]
  5. Pirouz, B.; Palermo, S.A.; Turco, M.; Piro, P. New Mathematical Optimization Approaches for LID Systems. In NUMTA 2019: Numerical Computations: Theory and Algorithm; Lecture Notes in Computer Science, Volume 11973; Sergeyev, Y., Kvasov, D., Eds.; Spinger: Cham, Switzerland, 2020; pp. 583–595. [Google Scholar] [CrossRef]
  6. Tebaldi, C.; Hayhoe, K.; Arblaster, J.M.; Meehl, G.A. Going to the Extremes. Clim. Chang. 2006, 79, 185–211. [Google Scholar] [CrossRef]
  7. Jacobson, C.R. Identification and quantification of the hydrological impacts of imperviousness in urban catchments: A review. J. Environ. Manag. 2011, 92, 1438–1448. [Google Scholar] [CrossRef]
  8. Palermo, S.A.; Talarico, V.C.; Pirouz, B. Optimizing Rainwater Harvesting Systems for Non-potable Water Uses and Surface Runoff Mitigation. In NUMTA 2019: Numerical Computations: Theory and Algorithm; Lecture Notes in Computer Science, Volume 11973; Sergeyev, Y., Kvasov, D., Eds.; Springer: Cham, Switzerland, 2020; pp. 570–582. [Google Scholar] [CrossRef]
  9. Pichs Madruga, R.; Seyboth, K.; Eickemeier, P.; Matschoss, P.; Hansen, G.; Kadner, S.; Schlömer, S.; Zwickel, T.; Von Stechow, C. Renewable Energy Sources and Climate Change Mitigation Special Report of the Intergovernmental Panel on Climate Change; Sokona, O.E.Y., Ed.; Cambridge University Press: Cambridge, UK, 2012; ISBN 9781107607101. [Google Scholar]
  10. Füssel, H.-M.; Jol, A.; Kurnik, B.; Hemming, D. Climate Change, Impacts and Vulnerability in Europe 2012: An Indicator-Based Report; European Environment Agency: Copenhagen, Denmark, 2012; ISBN 9789292133467. [Google Scholar] [CrossRef]
  11. Starzec, M.; Dziopak, J.; Słyś, D.; Pochwat, K.; Kordana, S. Dimensioning of Required Volumes of Interconnected Detention Tanks Taking into Account the Direction and Speed of Rain Movement. Water 2018, 10, 1826. [Google Scholar] [CrossRef] [Green Version]
  12. Kordana, S. The identification of key factors determining the sustainability of stormwater systems. E3S Web Conf. 2018, 45, 00033. [Google Scholar] [CrossRef] [Green Version]
  13. Piro, P.; Turco, M.; Palermo, S.A.; Principato, F.; Brunetti, G. A Comprehensive Approach to Stormwater Management Problems in the Next Generation Drainage Networks. In Internet of Things for Smart Urban Ecosystems; Springer: Cham, Switzerland, 2019; pp. 275–304. [Google Scholar] [CrossRef]
  14. Pumo, D.; Arnone, E.; Francipane, A.; Caracciolo, D.; Noto, L.V. Potential implications of climate change and urbanization on watershed hydrology. J. Hydrol. 2017, 554, 80–99. [Google Scholar] [CrossRef]
  15. Berz, G.; Kron, W.; Loster, T.; Rauch, E.; Schimetschek, J.; Schmieder, J.; Siebert, A.; Smolka, A.; Wirtz, A. World map of natural hazards—A global view of the distribution and intensity of significant exposures. Nat. Hazards 2001, 23, 443–465. [Google Scholar] [CrossRef]
  16. Milly, P.C.D.; Wetherald, R.T.; Dunne, K.A.; Delworth, T.L. Increasing risk of great floods in a changing climate. Nature 2002, 415, 514–517. [Google Scholar] [CrossRef]
  17. Semadeni-Davies, A.; Hernebring, C.; Svensson, G.; Gustafsson, L.G. The impacts of climate change and urbanisation on drainage in Helsingborg, Sweden: Combined sewer system. J. Hydrol. 2008, 350, 100–113. [Google Scholar] [CrossRef]
  18. United Nations, Department of Economic and Social Affairs. World Urbanization Prospects The 2018 Revision; United Nations: New York, NY, USA, 2018. [Google Scholar]
  19. Turco, M.; Brunetti, G.; Palermo, S.A.; Capano, G.; Grossi, G.; Maiolo, M.; Piro, P. On the environmental benefits of a permeable pavement: metals potential removal efficiency and Life Cycle Assessment. Urban Water J. 2020, 1–9. [Google Scholar] [CrossRef]
  20. Oberascher, M.; Zischg, J.; Palermo, S.A.; Kinzel, C.; Rauch, W.; Sitzenfrei, R. Smart Rain Barrels: Advanced LID Management Through Measurement and Control. In UDM 2018: New Trends in Urban Drainage Modelling; Green Energy and Technology; Mannina, G., Ed.; Springer: Cham, Switzerland, 2019; pp. 777–782. [Google Scholar] [CrossRef]
  21. Oberascher, M.; Zischg, J.; Kastlunger, U.; Schöpf, M.; Kinzel, C.; Zingerle, C.; Rauch, W.; Sitzenfrei, R. Advanced Rainwater Harvesting through Smart Rain Barrels. In Proceedings of the World Environmental and Water Resources Congress 2019, Pittsburgh, PA, USA, 19–23 May 2019; American Society of Civil Engineers: Reston, VA, USA, 2019; pp. 75–82. [Google Scholar]
  22. Sadler, J.M.; Goodall, J.L.; Behl, M.; Bowes, B.D.; Morsy, M.M. Exploring real-time control of stormwater systems for mitigating flood risk due to sea level rise. J. Hydrol. 2020, 583, 124571. [Google Scholar] [CrossRef]
  23. Marchioni, M.; Becciu, G. Experimental results on permeable pavements in Urban areas: A synthetic review. Int. J. Sustain. Dev. Plan. 2015, 10, 806–817. [Google Scholar] [CrossRef]
  24. Piro, P.; Carbone, M.; Morimanno, F.; Palermo, S.A. Simple flowmeter device for LID systems: From laboratory procedure to full-scale implementation. Flow Meas. Instrum. 2019. [Google Scholar] [CrossRef]
  25. Weyand, M. Real-time control in combined sewer systems in Germany—Some case studies. Urban Water 2002, 4, 347–354. [Google Scholar] [CrossRef]
  26. Butler, D.; Schütze, M. Integrating simulation models with a view to optimal control of urban wastewater systems. Environ. Model. Softw. 2005, 20, 415–426. [Google Scholar] [CrossRef]
  27. Martin, C.; Ruperd, Y.; Legret, M. Urban stormwater drainage management: The development of a multicriteria decision aid approach for best management practices. Eur. J. Oper. Res. 2007, 181, 338–349. [Google Scholar] [CrossRef]
  28. Turco, M. Modelling the Hydraulic Behaviour of Permeable Pavements through a Reservoir Element Model. In Proceedings of the International Multidisciplinary Scientific GeoConference Surveying Geology and Mining Ecology Management; SGEM: Sofia, Bulgaria, 2018. [Google Scholar] [CrossRef]
  29. Carbone, M.; Turco, M.; Nigro, G.; Piro, P. Modeling of hydraulic behaviour of green roof in catchment scale. In Proceedings of the International Multidisciplinary Scientific GeoConference Surveying Geology and Mining Ecology Management; SGEM: Albena, Bulgaria, 2014. [Google Scholar] [CrossRef]
  30. Maiolo, M.; Pirouz, B.; Bruno, R.; Palermo, S.A.; Arcuri, N.; Piro, P. The Role of the Extensive Green Roofs on Decreasing Building Energy Consumption in the Mediterranean Climate. Sustainability 2020, 12, 359. [Google Scholar] [CrossRef] [Green Version]
  31. Brunetti, G.; Papagrigoriou, I.-A.; Stumpp, C. Disentangling model complexity in green roof hydrological analysis: A Bayesian perspective. Water Res. 2020, 182, 115973. [Google Scholar] [CrossRef]
  32. Pirouz, B.; Arcuri, N.; Pirouz, B.; Palermo, S.A.; Turco, M.; Maiolo, M. Development of an Assessment Method for Evaluation of Sustainable Factories. Sustainability 2020, 12, 1841. [Google Scholar] [CrossRef] [Green Version]
  33. Bevilacqua, P.; Mazzeo, D.; Arcuri, N. Thermal inertia assessment of an experimental extensive green roof in summer conditions. Build. Environ. 2018, 131, 264–276. [Google Scholar] [CrossRef]
  34. Schütze, M.; Campisano, A.; Colas, H.; Vanrolleghem, P.; Schilling, W. Real-Time Control of Urban Water Systems. In Proceedings of the International Conference on Pumps, Electromechanical Devices and Systems Applied to Urban Water Management PEDS, Lencia, Spain, 22–25 April 2003. [Google Scholar]
  35. Schütze, M.; Campisano, A.; Colas, H.; Schilling, W.; Vanrolleghem, P.A. Real time control of urban wastewater systems—Where do we stand today? J. Hydrol. 2004, 299, 335–348. [Google Scholar] [CrossRef]
  36. Bach, P.M.; Rauch, W.; Mikkelsen, P.S.; McCarthy, D.T.; Deletic, A. A critical review of integrated urban water modelling—Urban drainage and beyond. Environ. Model. Softw. 2014, 54, 88–107. [Google Scholar] [CrossRef]
  37. Beeneken, T.; Erbe, V.; Messmer, A.; Reder, C.; Rohlfing, R.; Scheer, M.; Schuetze, M.; Schumacher, B.; Weilandt, M.; Weyand, M. Real time control (RTC) of urban drainage systems—A discussion of the additional efforts compared to conventionally operated systems. Urban Water J. 2013, 10, 293–299. [Google Scholar] [CrossRef]
  38. Dirckx, G.; Schütze, M.; Kroll, S.; Thoeye, C.; De Gueldre, G.; Van De Steene, B. RTC versus static solutions to mitigate CSO’s impact. In Proceedings of the 12nd International Conference on Urban Drainage, Porto Alegre, Brazil, 10–15 September 2011. [Google Scholar]
  39. Fradet, O.; Pleau, M.; Marcoux, C. Reducing CSOs and giving the river back to the public: Innovative combined sewer overflow control and riverbanks restoration of the St Charles River in Quebec City. Water Sci. Technol. 2011, 63, 331–338. [Google Scholar] [CrossRef]
  40. Rauch, W.; Bertrand-Krajewski, J.-L.; Krebs, P.; Mark, O.; Schilling, W.; Schütze, M.; Vanrolleghem, P.A. Deterministic modelling of integrated urban drainage systems. Water Sci. Technol. 2002, 45, 81–94. [Google Scholar] [CrossRef]
  41. Vezzaro, L.; Grum, M. A generalised Dynamic Overflow Risk Assessment (DORA) for Real Time Control of urban drainage systems. J. Hydrol. 2014, 515, 292–303. [Google Scholar] [CrossRef] [Green Version]
  42. Thorndahl, S. Radar Based Flow and Water Level Forecasting in Sewer Systems: A danisk case study. In Urbanrain09: 8th International Workshop on Precipitation in Urban Areas: St. Moritz, Switzerland; Eidgenössische Technische Hochschule Zürich, Laboratorium für Lebensmittel-Verfahrenstechnik: Zürich, Switzerland, 2009; ISBN 9783909386277. [Google Scholar]
  43. Campisano, A.; Cabot Ple, J.; Muschalla, D.; Pleau, M.; Vanrolleghem, P.A. Potential and limitations of modern equipment for real time control of urban wastewater systems. Urban Water J. 2013, 10, 300–311. [Google Scholar] [CrossRef]
  44. Schilling, W.; Andersson, B.; Nyberg, U.; Aspegren, H.; Rauch, W.; Harremoës, P. Real time control of wastewater systems. J. Hydraul. Res. 1996, 34, 785–797. [Google Scholar] [CrossRef]
  45. Rossman, L.A. Storm Water Management Model, Quality Assurance Report: Dynamic Wave Flow Routing; US Environmental Protection Agency, Office of Research and Development, National Risk Management Research Laboratory: Cincinnati, OH, USA, 2006. [Google Scholar]
  46. Erbe, V.; Schütze, M.; Haas, U. Application of A Guideline Document for Sewer System Real Time Control; Novatech: Lyon, France, 2007. [Google Scholar]
  47. Pleau, M.; Colas, H.; Lavallee, P.; Pelletier, G.; Bonin, R. Global optimal real-time control of the Quebec urban drainage system. Environ. Model. Softw. 2005, 20, 401–413. [Google Scholar] [CrossRef]
  48. Darsono, S.; Labadie, J.W. Neural-optimal control algorithm for real-time regulation of in-line storage in combined sewer systems. Environ. Model. Softw. 2007, 22, 1349–1361. [Google Scholar] [CrossRef]
  49. Garofalo, G.; Giordano, A.; Piro, P.; Spezzano, G.; Vinci, A. A distributed real-time approach for mitigating CSO and flooding in urban drainage systems. J. Netw. Comput. Appl. 2017, 78, 30–42. [Google Scholar] [CrossRef]
  50. Kändler, N.; Annus, I.; Vassiljev, A.; Puust, R. Real time controlled sustainable urban drainage systems in dense urban areas. J. Water Supply Res. Technol. 2020, 69, 238–247. [Google Scholar] [CrossRef]
  51. Aulinas, M.; Nieves, J.C.; Cortés, U.; Poch, M. Supporting decision making in urban wastewater systems using a knowledge-based approach. Environ. Model. Softw. 2011, 26, 562–572. [Google Scholar] [CrossRef]
  52. Klepiszewski, K.; Schmitt, T.G. Comparison of conventional rule based flow control with control processes based on fuzzy logic in a combined sewer system. Water Sci. Technol. 2002, 46, 77–84. [Google Scholar] [CrossRef] [PubMed]
  53. Shishegar, S.; Duchesne, S.; Pelletier, G. An integrated optimization and rule-based approach for predictive real time control of urban stormwater management systems. J. Hydrol. 2019, 577, 124000. [Google Scholar] [CrossRef]
  54. Regneri, M.; Klepiszewski, K.; Seiffert, S.; Vanrolleghem, P.A.; Ostrowski, M. Transport Sewer Model Calibration by Experimental Generation of Discrete Discharges from Individual CSO Structures. In Proceedings of the 6th International Congress on Environmental Modelling and Software, Leipzig, Germany, 1–5 July 2012. [Google Scholar]
  55. Tang, W.; Wang, Z.; Feng, Q.; Wang, M. Application of fuzzy expert control to APMP Pulping Wastewater treatment process of aerobic. In Proceedings of the 2010 IEEE International Conference on Mechatronics and Automation, Xi’an, China, 4–7 August 2010; pp. 339–344. [Google Scholar]
  56. Seggelke, K.; Löwe, R.; Beeneken, T.; Fuchs, L. Implementation of an integrated real-time control system of sewer system and waste water treatment plant in the city of Wilhelmshaven. Urban Water J. 2013, 10, 330–341. [Google Scholar] [CrossRef]
  57. Chiang, Y.M.; Chang, L.C.; Tsai, M.J.; Wang, Y.F.; Chang, F.J. Auto-control of pumping operations in sewerage systems by rule-based fuzzy neural networks. Hydrol. Earth Syst. Sci. 2011, 15, 185–196. [Google Scholar] [CrossRef] [Green Version]
  58. Tagherouit, W.B.; Bennis, S.; Bengassem, J. A Fuzzy Expert System for Prioritizing Rehabilitation of Sewer Networks. Comput. Civ. Infrastruct. Eng. 2011, 26, 146–152. [Google Scholar] [CrossRef]
  59. Fuchs, L.; Beeneken, T.; Spönemann, P.; Scheffer, C. Model based real-time control of sewer system using fuzzy-logic. Water Sci. Technol. 1997, 36, 343–347. [Google Scholar] [CrossRef]
  60. Mounce, S.R.; Shepherd, W.; Ostojin, S.; Abdel-Aal, M.; Schellart, A.N.A.; Shucksmith, J.D.; Tait, S.J. Optimisation of a fuzzy logic-based local real-time control system for mitigation of sewer flooding using genetic algorithms. J. Hydroinf. 2020. [Google Scholar] [CrossRef]
  61. Lemos, J.M.; Pinto, L.F. Distributed Linear-Quadratic Control of Serially Chained Systems: Application to a Water Delivery Canal [Applications of Control]. IEEE Control Syst. Mag. 2012, 32, 26–38. [Google Scholar] [CrossRef]
  62. Muschalla, D. Optimization of integrated urban wastewater systems using multi-objective evolution strategies. Urban Water J. 2008, 5, 59–67. [Google Scholar] [CrossRef]
  63. Cho, J.H.; Seok Sung, K.; Ryong Ha, S. A river water quality management model for optimising regional wastewater treatment using a genetic algorithm. J. Environ. Manag. 2004, 73, 229–242. [Google Scholar] [CrossRef]
  64. Fu, G.; Butler, D.; Khu, S.-T. Multiple objective optimal control of integrated urban wastewater systems. Environ. Model. Softw. 2008, 23, 225–234. [Google Scholar] [CrossRef]
  65. Barreto, W.; Vojinovic, Z.; Price, R.; Solomatine, D. Multiobjective Evolutionary Approach to Rehabilitation of Urban Drainage Systems. J. Water Resour. Plan. Manag. 2010, 136, 547–554. [Google Scholar] [CrossRef]
  66. Ngamalieu-Nengoue, U.A.; Iglesias-Rey, P.L.; Martínez-Solano, F.J. Urban Drainage Networks Rehabilitation Using Multi-Objective Model and Search Space Reduction Methodology. Infrastructures 2019, 4, 35. [Google Scholar] [CrossRef] [Green Version]
  67. Ocampo-Martinez, C. Model Predictive Control of Wastewater Systems; Springer Science & Business Media: Berlin, Germany, 2010. [Google Scholar]
  68. Toro, R.; Ocampo-Martínez, C.; Logist, F.; Van Impe, J.; Puig, V. Tuning of predictive controllers for drinking water networked systems. IFAC Proc. Vol. 2011, 44, 14507–14512. [Google Scholar] [CrossRef]
  69. Xu, M.; van Overloop, P.J.; van de Giesen, N.C. On the study of control effectiveness and computational efficiency of reduced Saint-Venant model in model predictive control of open channel flow. Adv. Water Resour. 2011, 34, 282–290. [Google Scholar] [CrossRef]
  70. Zamora, C.; Giraldo, J.M.; Leirens, S. Model predictive control of water transportation networks. In Proceedings of the 2010 IEEE ANDESCON, Bogota, Colombia, 15–17 September 2010. [Google Scholar]
  71. Ocampo-Martinez, C.; Ingimundarson, A.; Bemporad, A.; Puig, V. Suboptimal Hybrid Model Predictive Control: Application to Sewer Networks. IFAC Proc. Vol. 2008, 41, 10021–10026. [Google Scholar] [CrossRef] [Green Version]
  72. Van Heeringen, K.-J.; Gooijer, J.; Schwanenberg, D. Practical application of drainage system control by using MPC in Noorderzijlvest. In Proceedings of the EGU General Assembly, Vienna, Austria, 7–12 April 2013; Volume 15. [Google Scholar]
  73. Ocampo-Martinez, C.; Puig, V.; Grosso, J.M.; Montes-de-Oca, S. Multi-layer Decentralized MPC of Large-scale Networked Systems. Intell. Syst. Control Autom. Sci. Eng. 2014, 69, 495–515. [Google Scholar] [CrossRef] [Green Version]
  74. Leirens, S.; Zamora, C.; Negenborn, R.R.; De Schutter, B. Coordination in urban water supply networks using distributed model predictive control. In Proceedings of the 2010 American Control Conference, Baltimore, MD, USA, 30 June–2 July 2010; pp. 3957–3962. [Google Scholar]
  75. Negenborn, R.R.; van Overloop, P.J.; Keviczky, T.; De Schutter, B. Distributed model predictive control of irrigation canals. Netw. Heterog. Media 2009, 4, 359–380. [Google Scholar] [CrossRef]
  76. Lihui, C.; Yugeng, X.; Dewei, L. Aggregation-based model predictive control of open channel networks. In Proceedings of the 29th Chinese Control Conference, Beijing, China, 29–31 July 2010. [Google Scholar]
  77. Malaterre, P.O.; Rodellar, J. Multivariable predictive control of irrigation canals. Design and evaluation on a 2-pool model. In Proceedings of the International Workshop on Regulation of Irrigation Canals: State of the Art of Research and Applications, Marrakesh, Morroco, 22–24 April 1997. [Google Scholar]
  78. Puig, V.; Cembrano, G.; Romera, J.; Quevedo, J.; Aznar, B.; Ramón, G.; Cabot, J. Predictive optimal control of sewer networks using CORAL tool: application to Riera Blanca catchment in Barcelona. Water Sci. Technol. 2009, 60, 869–878. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Ramirez-Llanos, E.; Quijano, N. A population dynamics approach for the water distribution problem. Int. J. Control 2010, 83, 1947–1964. [Google Scholar] [CrossRef]
  80. Barreiro-Gomez, J.; Obando, G.; Riano-Briceno, G.; Quijano, N.; Ocampo-Martinez, C. Decentralized Control for Urban Drainage Systems via population dynamics: Bogotá case study. In Proceedings of the 2015 European Control Conference, Linz, Austria, 15–17 July 2015; pp. 2426–2431. [Google Scholar]
  81. Messmer, A.; Schütze, M.; Ogurek, M. A Demonstration Software Tool For Real Time Control Of Urban Drainage Systems. In Proceedings of the 11th International Conference on Urban Drainage, Edinburgh, UK, 31 August–5 September 2008. [Google Scholar]
  82. Riaño-Briceño, G.; Barreiro-Gomez, J.; Ramirez-Jaime, A.; Quijano, N.; Ocampo-Martinez, C. MatSWMM—An open-source toolbox for designing real-time control of urban drainage systems. Environ. Model. Softw. 2016, 83, 143–154. [Google Scholar] [CrossRef] [Green Version]
  83. Li, J.; Burian, S.; Oroza, C. Exploring the potential for simulating system-level controlled smart stormwater system. In Proceedings of the World Environmental and Water Resources Congress 2019: Water Wastewater, and Stormwater; Urban Water Resources; and Municipal Water Infrastructure, Pittsburgh, PA, USA, 19–23 May 2019; pp. 46–56. [Google Scholar]
  84. Yuan, J.M.; Tilford, K.A.; Jiang, H.Y.; Cluckie, I.D. Real-time urban drainage system modelling using weather radar rainfall data. Phys. Chem. Earth Part B 1999, 24, 915–919. [Google Scholar] [CrossRef]
  85. Duchesne, S.; Mailhot, A.; Dequidt, E.; Villeneuve, J.-P. Mathematical modeling of sewers under surcharge for real time control of combined sewer overflows. Urban Water 2001, 3, 241–252. [Google Scholar] [CrossRef]
  86. Giordano, A.; Spezzano, G.; Vinci, A.; Garofalo, G.; Piro, P. A cyber-physical system for distributed real-time control of urban drainage networks in smart cities. In IDCS 2014: Internet and Distributed Computing Systems; Springer: Cham, Switzerland, 2014. [Google Scholar] [CrossRef]
  87. Palermo, S.A.; Talarico, V.C.; Turco, M. On the LID systems effectiveness for urban stormwater management: case study in Southern Italy. IOP Conf. Ser. 2020, 410, 012012. [Google Scholar] [CrossRef]
  88. García-Marín, A.P.; Ayuso-Muñoz, J.L.; Jiménez-Hornero, F.J.; Estévez, J. Selecting the best IDF model by using the multifractal approach. Hydrol. Process. 2013, 27, 433–443. [Google Scholar] [CrossRef]
  89. Afrin, S.; Islam, M.M.; Rahman, M.M. Development of IDF Curve for Dhaka City Based on Scaling Theory under Future Precipitation Variability Due to Climate Change. Int. J. Environ. Sci. Dev. 2015. [Google Scholar] [CrossRef] [Green Version]
  90. Ewea, H.A.; Elfeki, A.M.; Al-Amri, N.S. Development of intensity–duration–frequency curves for the Kingdom of Saudi Arabia. Geomatics, Nat. Hazards Risk 2017, 8, 570–584. [Google Scholar] [CrossRef] [Green Version]
  91. De Paola, F.; Giugni, M.; Topa, M.E.; Bucchignani, E. Intensity-Duration-Frequency (IDF) rainfall curves, for data series and climate projection in African cities. SpringerPlus 2014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Arpacal. 2020. Available online: http://www.cfd.calabria.it/index.php/dati-stazioni/dati-storici (accessed on 12 June 2020).
  93. Keifer, C.; Chu, H.H. Synthetic Storm Pattern for Drainage Design. J. Hydraul. Div. 1957, 83, 1–25. [Google Scholar]
  94. Voyde, E.; Fassman, E.; Simcock, R. Hydrology of an extensive living roof under sub-tropical climate conditions in Auckland, New Zealand. J. Hydrol. 2010, 394, 384–395. [Google Scholar] [CrossRef]
  95. Stovin, V.; Vesuviano, G.; Kasmin, H. The hydrological performance of a green roof test bed under UK climatic conditions. J. Hydrol. 2012, 414, 148–161. [Google Scholar] [CrossRef]
  96. Palermo, S.A.; Turco, M.; Principato, F.; Piro, P. Hydrological effectiveness of an extensive green roof in Mediterranean climate. Water 2019, 11, 1378. [Google Scholar] [CrossRef] [Green Version]
  97. Bonabeau, E.; Dorigo, M.; Theraulaz, G. Swarm Intelligence: From Natural to Artificial Systems; Oxford University Press: Oxford, UK, 1999; Volume 4, ISBN 0195131592. [Google Scholar]
  98. Jelasity, M.; Montresor, A.; Babaoglu, O. Gossip-based aggregation in large dynamic networks. ACM Trans. Comput. Syst. 2005, 23, 219–252. [Google Scholar] [CrossRef] [Green Version]
  99. Cronshey, R. Urban Hydrology for Small Watersheds; US Department of Agriculture: Washington, DC, USA, 1986. [Google Scholar]
Figure 1. The flowchart of the methodology.
Figure 1. The flowchart of the methodology.
Water 12 02842 g001
Figure 2. Case study location. A map of Italy with Calabria Region (left), the location of Paola village in Calabria Region (middle), and a zoom-in of Paola (in yellow box) and of the specific highly urbanized area considered as a case study (in green box) (right).
Figure 2. Case study location. A map of Italy with Calabria Region (left), the location of Paola village in Calabria Region (middle), and a zoom-in of Paola (in yellow box) and of the specific highly urbanized area considered as a case study (in green box) (right).
Water 12 02842 g002
Figure 3. Drainage network schematization. From left to right: Scenario 0, Scenario 1, and Scenario 2.
Figure 3. Drainage network schematization. From left to right: Scenario 0, Scenario 1, and Scenario 2.
Water 12 02842 g003
Figure 4. Rainfall characteristics for the 12-month data series (54 measured rainfall data between May 2019 and May 2020) compared with the intensity–duration–frequency (IDF) return period curves estimated for Paola (CS).
Figure 4. Rainfall characteristics for the 12-month data series (54 measured rainfall data between May 2019 and May 2020) compared with the intensity–duration–frequency (IDF) return period curves estimated for Paola (CS).
Water 12 02842 g004
Figure 5. The drainage network response, in terms of conduits capacity over the time, to the design storm event (RP = 10 years) for each modeled scenario.
Figure 5. The drainage network response, in terms of conduits capacity over the time, to the design storm event (RP = 10 years) for each modeled scenario.
Water 12 02842 g005
Table 1. Hydrological characteristics of each rainfall event recorded from May 2019 to May 2020. PD: precipitation depth, D: rainfall duration, Mean i: mean rainfall intensity, and Max i: maximum rainfall intensity.
Table 1. Hydrological characteristics of each rainfall event recorded from May 2019 to May 2020. PD: precipitation depth, D: rainfall duration, Mean i: mean rainfall intensity, and Max i: maximum rainfall intensity.
No.DatePDDMax iMean i
(day/month/year)(hour:minute)(mm)(h)(mm/h)(mm/h)
104/05/201908:0031.46.58244.8
206/05/201906:004.81.589.63.0
312/05/201905:002.80.924.83.0
413/05/201900:0016.21.543.210.8
515/05/201923:008.82.021.64.4
628/05/201902:009.42.521.63.8
701/06/201901:007.22.07.23.6
816/07/201901:0031.02.0857.614.9
902/09/201918:0013.20.8350.415.9
1008/09/201911:0021.61.2545.617.3
1119/09/201913:007.81.0833.67.2
1224/09/201902:004.00.54.88.0
1326/09/201914:0014.61.6652.88.8
1403/10/201904:00101.9216.85.2
1505/10/201913:0017.43.4216.85.1
1607/10/201906:0040.63.7576.810.8
1716/10/201908:004.20.3328.812.7
1831/10/201919:004.21.254.83.4
1904/11/201901:0011.42.4224.04.7
2006/11/201913:0033.26.5821.65.0
2109/11/201907:0032.25.1731.26.2
2211/11/201919:0017.004.0826.44.2
2313/11/201900:0021.03.4231.26.1
2417/11/201907:0034.04.1743.28.2
2519/11/201918:0024.64.7519.25.2
2624/11/201907:0074.612.2533.66.1
2728/11/201905:0016.22.1738.47.5
2806/12/201904:003.41.254.82.7
2907/12/201901:007.40.7521.69.9
3009/12/201912:0018.43.679.65.0
3112/12/201902:0042.28.8336.04.8
3220/12/201903:003.20.927.23.5
3321/12/201908:0012.23.337.23.7
3422/12/201900:0012.43.08124.0
3525/12/201902:005.21.427.23.7
3618/01/202010:0011.024.016.80.5
3729/01/202020:003.41.016.83.4
3802/02/202003:002.81.084.82.6
3905/02/202002:002.40.5812.04.1
4010/02/202007:004.21.674.82.5
4114/02/202012:005.81.0812.05.4
4220/02/202000:004.80.7519.26.4
4328/02/202002:007.01.8314.43.8
4403/03/202002:0021.05.6721.63.7
4504/03/202020:003.60.59.67.2
4607/03/202011:0016.64.339.63.8
4722/03/202010:004.01.54.82.7
4825/03/202004:00105.224.519.24.3
4901/04/202004:008.01.3312.06.0
5003/04/202006:006.22.254.82.8
5120/04/202002:0015.65.337.22.9
5221/04/202022:0047.69.0836.05.2
5319/05/202015:004.21.584.82.7
5420/05/202011:0032.23.5831.29.0
Maximum105.224.576.817.3
Minimum2.40.334.80.5
Mean17.13.621.45.8
Sum923.4195.05
Table 2. Rainfall events for simulations.
Table 2. Rainfall events for simulations.
No.DatePDDMax iMean i
(day/month/year)(mm)(h)(mm/h)(mm/h)
113/05/201916.21.543.210.8
216/07/2019312.0857.614.9
308/09/201921.61.2545.617.3
407/10/201940.63.7576.810.8
517/11/2019344.1743.28.2
624/11/201974.612.2533.66.1
720/05/202032.23.5831.29
Table 3. Node Flooding (NF), Hours Flooded (HF), Conduit Surcharge (CS), and total Flood Volume (FV) for each scenario (S0, S1, S2) and each design rainfall event of 5, 10, and 20-year return periods (RP).
Table 3. Node Flooding (NF), Hours Flooded (HF), Conduit Surcharge (CS), and total Flood Volume (FV) for each scenario (S0, S1, S2) and each design rainfall event of 5, 10, and 20-year return periods (RP).
RP = 5 YearsRP = 10 YearsRP = 20 Years
S0S1S2S0S1S2S0S1S2
NF (-) 126512871287
HF (h) 3.871.070.494.431.630.964.892.091.13
CS (-)19871910919129
FV (106 L) 0.880.190.041.080.330.091.280.490.14
Table 4. Flood Volume Reduction (FVR) index for design storm events.
Table 4. Flood Volume Reduction (FVR) index for design storm events.
RPFVR (%)
FVR0–1FVR0–2FVR1–2
5 years78.4195.4578.95
10 years69.4491.6772.73
20 years61.7289.0671.43
Table 5. Hour Flooded Reduction (HFR) index for design storm events.
Table 5. Hour Flooded Reduction (HFR) index for design storm events.
RPHFR (%)
HFR0–1HFR0–2HFR1–2
5 years72.3587.3454.21
10 years63.2178.3341.1
20 years57.2676.8945.93
Table 6. Nodes flooding (NF) index for design storm events.
Table 6. Nodes flooding (NF) index for design storm events.
RPNF (%)
NF 0NF 1NF 2
5 years41.3820.6917.24
10 years41.3827.5924.14
20 years41.3827.5924.14
Table 7. Conduit surcharge (CS) index for design storm events.
Table 7. Conduit surcharge (CS) index for design storm events.
RPCS (%)
CS 0CS 1CS 2
5 years67.8628.5725.00
10 years67.8635.7132.14
20 years67.8642.8632.14
Table 8. Node Flooding (NF), Hours Flooded (HF), total Flood Volume (FV), and Conduits Surcharge (CS) for Scenario 0 (S0) and Scenario 2 (S2).
Table 8. Node Flooding (NF), Hours Flooded (HF), total Flood Volume (FV), and Conduits Surcharge (CS) for Scenario 0 (S0) and Scenario 2 (S2).
No.NF (-)HF (-)FV (106 L)CS (-)
S0S2S0S2S0S2S0S2
1501.550.000.140.00141
2815.620.010.960.00155
3603.850.000.350.00141
4401.550.000.100.00141
5604.220.000.400.00141
6401.560.000.20.00141
7602.940.000.260.00141

Share and Cite

MDPI and ACS Style

Maiolo, M.; Palermo, S.A.; Brusco, A.C.; Pirouz, B.; Turco, M.; Vinci, A.; Spezzano, G.; Piro, P. On the Use of a Real-Time Control Approach for Urban Stormwater Management. Water 2020, 12, 2842. https://doi.org/10.3390/w12102842

AMA Style

Maiolo M, Palermo SA, Brusco AC, Pirouz B, Turco M, Vinci A, Spezzano G, Piro P. On the Use of a Real-Time Control Approach for Urban Stormwater Management. Water. 2020; 12(10):2842. https://doi.org/10.3390/w12102842

Chicago/Turabian Style

Maiolo, Mario, Stefania Anna Palermo, Anna Chiara Brusco, Behrouz Pirouz, Michele Turco, Andrea Vinci, Giandomenico Spezzano, and Patrizia Piro. 2020. "On the Use of a Real-Time Control Approach for Urban Stormwater Management" Water 12, no. 10: 2842. https://doi.org/10.3390/w12102842

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