Open access
Technical Papers
Nov 9, 2020

Merging Fluid Transient Waves and Artificial Neural Networks for Burst Detection and Identification in Pipelines

Publication: Journal of Water Resources Planning and Management
Volume 147, Issue 1

Abstract

The occurrence of bursts in water pipelines can not only prevent the system from functioning properly, but it can also produce significant water loss that disrupts activities in urban areas. Therefore, the detection and location of bursts in water distribution systems is a vital task for water utilities. Various techniques currently exist to detect the occurrence of these events, but there is a need for a permanent monitoring method that can detect and identify anomalous events quickly and accurately. This paper presents a new technique that uses artificial neural networks (ANNs) to detect and identify bursts in pipelines by interpreting the transient pressure waves that a burst causes along pipelines. The technique is divided into two stages: a model development stage and an application stage. The model development stage includes the generation of transient pressure traces and the training and testing of two different ANNs to (1) detect burst occurrence and (2) identify burst location and size. The application stage includes the processing of a potentially continuous transient pressure trace, analysis by the previously trained ANNs, and then the verification of the results using a transient flow forward numerical model. A numerical application demonstrates the principles of the technique and the potential for merging the use of fluid transient waves and ANNs. The technique has also been validated in the laboratory, indicating that the prediction of the location of the burst is very accurate while the prediction of the burst size requires an additional step to ensure its accuracy.

Introduction

Water transmission and distribution pipelines are critical infrastructure for modern cities. Owing to the sheer size of these pipelines and the fact that most of them are buried underground, the health monitoring and maintenance of this infrastructure is challenging. In addition, some water transmission pipelines cover long distances through remote areas that are not easily inspected on a regular basis. To monitor these systems, different noninvasive techniques have been developed to identify events that may put the functioning of a pipeline at risk. These techniques include visual observations (Thomson and Wang 2009), acoustic monitoring (Shimanskiy et al. 2003; Muggleton et al. 2006; Juliano et al. 2013), thermographic infrared inspection methods (Fahmy and Moselhi 2010), ground-penetrating radar methods (Hunaidi and Giamou 1998), and remote sensing (Agapiou et al. 2016; Martins et al. 2019). However, these techniques are time-consuming, do not provide permanent monitoring of the pipelines, or have a short inspection range along a given pipeline. Therefore, there is a need for a permanent monitoring method capable of identifying anomalous events and, potentially, their associated characteristics in near real time.
Hydraulic-based techniques have been developed based on the understanding of the movement of a fluid along a pipeline and are typically related to the measurement and analysis of two hydraulic variables: flow (or velocity) and pressure in the pipeline. These techniques can include volume-based methods coupled with alert systems (Mounce et al. 2003), pressure and flow analysis using statistical detection (Puust et al. 2008; Li et al. 2014; Lee et al. 2016; Wu et al. 2018b), system state estimation analysis (Andersen and Powell 2000), and transient-based methods (Misiunas et al. 2005). However, while each of these approaches has been moderately successful, they also have associated disadvantages. Some of these methods are only applicable for the detection of leaks, are not able to pinpoint the location of the abnormal event, or require an accurate numerical model of the pipeline, or the resolution of the data limits their ability to detect and locate anomalies in a timely manner.
Among the hydraulic-based techniques, transient-based methods have received attention because they provide for the inspection of a long section of pipe using only pressure measurements (Wang et al. 2002; Brunone et al. 2013; Gong et al. 2014). These methods are based on the interpretation of the effect that the occurrence of an anomalous event has on a measured transient pressure trace. The interpretation of the transient pressure trace can be conducted by visual analysis or by processing the transient pressure trace to identify reflections and detect the occurrence of abnormal events (Misiunas et al. 2007; Srirangarajan et al. 2011; Rashid et al. 2014; Rathnayaka et al. 2016). Nonetheless, a visual analysis cannot be conducted in real time and the processing techniques available require extensive numerical processing that makes a real-time monitoring system difficult.
This paper presents a new transient-based technique that employs artificial neural networks (ANNs) to interpret rapid changes in a transient head trace to identify, locate, and characterize bursts in water pipelines. Unlike current transient-based techniques, the proposed approach is data-driven and relies on no detailed information regarding the analyzed pipeline for the interpretation of the transient pressure trace in near real time. In addition, in contrast to many existing approaches, the proposed technique does not use an ANN as a metamodel of the transient phenomenon but instead uses an ANN as a tool to interpret the measured transient head traces to identify bursts. This technique analyzes short time segments of a potentially continuous transient head trace in previously trained ANNs through two different processes. A first ANN analysis detects whether a particular transient pressure head time window contains abnormal changes that could have been produced by the occurrence of a burst. This initial analysis also determines whether the information contained in the time window is enough to locate and characterize the potential burst. The second ANN analyzes the detected abnormal transient head time windows from the first ANN to predict the location and the characteristics of the occurring burst. Finally, the technique uses a transient flow numerical model to verify that the predictions of location and size of the burst are coherent and are similar to the measured transient head trace.
The approach proposed in this paper works with transient pressure head data to carry out the training of the ANNs. These pressure head data may be obtained from recorded data, from an available numerical model, or from a combination of these two sources. Once the ANNs have been trained, the use of this technique is data-driven in that it only requires measured transient head data to detect and identify a burst (by determining its location and size). Additionally, it can be applied in near real time considering that the computationally expensive processes are concentrated in the training of the ANNs and not in the processing of new transient pressure head traces, which can be tested almost immediately without retraining the ANNs.
One example hydraulic system is considered in this paper to demonstrate the functioning and performance of the proposed technique. A single pipeline with the potential presence of a burst at any point along its length is analyzed. A numerical system with this configuration is used to present the operation of the technique and to train the ANNs. In addition, a laboratory validation is included to demonstrate the promising potential of the technique.
This paper includes a background in hydraulic and transient-based methods for identifying and locating bursts in pipelines. The example hydraulic system is described next, and the methodology to apply the proposed technique is then presented by explaining the two stages involved in the process. The results for the training, testing, and application of the proposed technique are presented for the numerical system. Finally, the experimental validation is described. The proposed technique is shown to be successful in analyzing a potentially continuous transient head trace to identify, locate, and characterize the occurrence of a burst in a pipeline.

Background

The detection and location of bursts in water distribution systems is a complex task for water utilities. Various authors have proposed techniques for locating bursts in water pipelines using hydraulic-based methods. A first group of techniques includes the use of flow and pressure measurements with statistical methods to detect the occurrence of abnormal behavior in a system (Wu and Liu 2017). Ye and Fenner (2011) proposed the use of an adaptive Kalman filtering process to predict flow (or pressure) in a system at a district meter area (DMA) level. This statistical characterization of a dynamic system is able to model the normal hydraulic parameters that are compared to measured data to detect the occurrence of bursts. Similarly, Ahn and Jung (2019) proposed a hybrid statistical model that combines statistical process control with two univariate methods to enhance the performance of the burst detection technique in terms of the detection probability, the rate of false alarms, and the average detection time. Other authors have proposed the use of statistical risk functions (Cheng et al. 2018) and principal component analysis (Palau et al. 2012) to detect bursts in transmission mains and in DMAs. A different approach was described in Wu et al. (2018a), where a clustering-based method was used to identify bursts using only 1 day of historic measured data without using statistical methods to model the expected normal conditions of the system. Although these techniques are successful in detecting the occurrence of bursts, they are unable to accurately pinpoint the location of bursts and their range of effectiveness is often limited to a DMA level.
A second group of techniques uses supervised learning techniques coupled with hydraulic measurements to send an alert regarding the occurrence of a burst. Mounce and Machell (2006) used two ANN architectures (static ANN and time delay ANN) to detect the occurrence of bursts using flow data at a DMA level. The use of ANNs showed potential for identifying changes in the flow that corresponded to unusual fluctuations of this hydraulic variable. Mounce et al. (2010) proposed the use of support vector regression models to predict time series data in a moving time window and compare these series with measured data for the detection of anomalies. The use of this supervised learning technique was applied to historical data proving that 78% of the alerts corresponded to actual abnormal events in the system. Similarly, Romano et al. (2014) proposed a fully automated data-driven methodology at a DMA level using all the pressure and flow measurements available. This approach combined the use of an ANN for the short-term forecasting of hydraulic values and statistical processes to determine whether an abnormal event had occurred. The results obtained showed the potential of data-driven technologies for near-real-time incident management in water distribution systems.
Other authors have proposed the use of artificial immune systems not only for detection but also for an approximate localization of a burst. Tao et al. (2014) proposed the use of pressure data every 10 min and was able to detect and localize a burst in 48% of the cases in a real network. More recently, Huang et al. (2018) proposed the use of a random forest classifier to detect bursts in real time by analyzing successive time windows (every 15 min) of flow data at a DMA level. These contributions demonstrate the use of a supervised learning technique for the detection of bursts in pipelines; however, all the applications to date have been focused at a DMA level and using SCADA flow (or pressure) data, which are often available in intervals between 1 and 15 min.
A different group of techniques involves the use of fluid transients to detect the occurrence of bursts in water systems. These techniques are able to detect and locate anomalies in pipelines such as leaks (Wang et al. 2002; Lee et al. 2007a, b; Capponi et al. 2017), blockages (Rubio Scola et al. 2017), or wall deterioration (Gong et al. 2013) and have obtained accurate results. Liggett and Chen (1994) proposed an inverse transient algorithm to detect leaks using the method of characteristics (MOC) and the Levenberg-Marquardt method to find friction and leak parameters in a system. These authors proposed, for the first time, the use of transient pressure measurements to locate pipe ruptures using an event algorithm with the capacity to approximately pinpoint burst locations to locate additional nodes in the numerical model of systems and conduct inverse transient analysis. Misiunas et al. (2005) used the principles of time domain reflectometry for detecting and locating abrupt pipeline breaks using a single pressure measurement point in a pipeline. A system of continuous monitoring of the pressure at a high sampling frequency, coupled with a cumulative sum test and prefiltering techniques, was used to detect changes in the data. In addition, an offline analysis of a short time window was conducted to interpret the pressure changes to determine the burst location. The results of this research have demonstrated that transient pressure signals can be used to detect bursts in water systems.
More recently, Srirangarajan et al. (2011) described the use of a wavelet-based multiscale analysis combined with a focusing algorithm and a graph-based search algorithm to detect and locate a burst event. This technique showed the potential for application of transient pressure techniques if the characteristics of the system are known and presented the first application in a network layout. Other authors have proposed the use of embedded and distributed event processing algorithms to detect transient events in a system and stroke-based transient recognition algorithms to classify transient events as bursts (Hoskins and Stoianov 2014). Although several authors have proposed techniques that use transient pressure signals to detect and localize bursts in pipelines, the existing techniques require offline processing that can delay the detection of bursts or require specific prior knowledge about the systems.
The analysis of transient pressure signals using supervised learning algorithms has not been widely explored. Bohorquez et al. (2020) presented a technique that uses ANNs to predict the presence of different features (leaks and junctions) in a pipeline after the generation of a controlled transient event. The obtained results demonstrated the potential of combining the principles of transient-based techniques with ANNs to interpret the pressure traces. A similar application was proposed by Perera and Rajapakse (2011) for the identification of transient faults in power transmission networks using hidden Markov models and probabilistic neural networks (PNNs) to classify transients as faults of normal switching events. Considering the current literature, this paper presents a methodology that for the first time merges ANNs and fluid transient waves to detect, locate, and characterize bursts in water pipelines.

Hydraulic System Configuration

The proposed methodology has been applied to a single pipeline under the assumption that a burst can occur at any point along its length. The pipeline, of diameter D, is connected at the upstream end to a reservoir, and at the downstream end there is a closed inline valve (Fig. 1). The length of the pipeline is LT, and the location of the burst is characterized by a distance x measured from the upstream end of the pipeline. The bursts are modeled as circular orifices of diameter DB.
Fig. 1. Single pipeline with a burst system description.
Two pipelines with the configuration presented in Fig. 1 were analyzed. A numerical pipeline was considered with the characteristics presented in Table 1. In this case, the head at the reservoir is defined as H0=55  m at the beginning of the simulation, and a sinusoidal fluctuation of this head is considered to model gradual changes in pressure head that can be observed in pipeline systems. Steady-state friction was considered using a Darcy-Weisbach friction factor f with a pipeline roughness of ε=0.01  mm. The different burst locations and sizes that have been considered are described in what follows, as part of the methodology (Steps A.2 and A.3 in Fig. 2) of the proposed technique.
Table 1. Numerical pipeline characteristics
CharacteristicUnitValue
Length of pipe (LT)(m)1,000
Internal diameter of pipe (D)(mm)587
Wave speed of pipe (a)(m/s)1,111
LT/a time(s)0.9
Fig. 2. Burst detection and identification methodology.
A second pipeline was analyzed as part of the experimental verification of the proposed technique. The characteristics of this pipeline are shown in Table 2 and explained in detail in the experimental results section of this paper.
Table 2. Experimental pipeline characteristics
CharacteristicUnitValue
Length of pipe (L)(m)37.24
Internal diameter of pipe (D)(mm)22.14
Wave speed of pipe (a)(m/s)1,290
Wall thickness (ε)(mm)1.63
LT/a time(s)0.029

Methodology

The effect that a burst has on a transient head trace is characterized by a sharp drop in the head that propagates in both directions away from the burst location (Misiunas et al. 2005; Bohorquez et al. 2018). The proposed methodology for detecting and identifying bursts is presented in Fig. 2. Two stages have been included in this diagram. The first stage comprises the ANN model development (Stage A in Fig. 2), which is carried out first and can be updated regularly depending on the availability of new pressure head data. The application stage (Stage B in Fig. 2) includes the required steps to process and interpret a continuous transient pressure head trace using the ANNs for detecting, identifying, and verifying the occurrence of a burst.

Model Development Stage

To establish a near-real-time monitoring system for a pipeline, the first stage of the proposed methodology includes the development of the burst detection and identification model. This model is composed of two trained ANNs that can predict, first, whether a particular analyzed transient head trace contains abnormal head fluctuations corresponding to the occurrence of a burst in the pipeline and, second, the location and size of this burst. As explained in this section, this model does not represent a metamodel of the transient flow phenomenon that occurs in a pipeline after a burst. The design, training, and testing of these ANNs are focused on interpreting a measured transient head signal to detect the occurrence of a burst.
The steps described in this stage are required only once to set up a monitoring system in a pipeline, which means that the computationally expensive processes are not required for the application stage. However, if new transient pressure head data become available (from numerical modeling or historical measured data), the ANNs can be retrained to include any new information regarding the transient behavior of the pipeline.

ANN Architecture Definition

The first step to developing a model capable of detecting and identifying bursts is the definition of the ANN architecture (Step A.1 in Fig. 2). Previous applications of ANNs with transient fluid head traces have shown that a dense network (the most general and widely used ANN structure) is not able to adequately capture the changes in pressure due to the presence of different elements in a pipeline (Bohorquez et al. 2020). Considering this, a one-dimensional (1D) convolutional network architecture was chosen since these networks have fewer weights and are less prone to overfitting. These 1D convolutional networks are designed to perform satisfactorily both in the detection of the occurrence of a burst and in the identification of the burst location and size. This design was developed by modifying different characteristics of the ANN architecture, including the number of convolutional layers, type of activation function, number of filters in each layer, and training batch size.
The final configuration of the designed 1D convolutional networks includes (1) a maximum of seven convolutional layers, (2) the use of leaky rectified linear unit (Leaky ReLU) and Softmax as activation functions, (3) a maximum number of 12 filters that increase in each convolutional layer, (4) a training batch size of 50 samples, and (5) three dense layers of maximum sizes of 21, 9, and 3. With these characteristics, the designed ANNs have between 26,808 and 81,250 weights to be trained.

Transient Head Trace Sample Characteristic Definition

The model development stage includes the training and testing of two different ANNs (Steps A.7 and A.8 in Fig. 2). The data used for these processes are transient head data that can be obtained from recorded data, from an available numerical model, or from a combination of these two sources (Step A.2 in Fig. 2). In this paper, numerically generated data were used for ANN training and testing to demonstrate the application of the proposed technique. In the context of this paper, a sample is the transient head trace that would be observed if a burst (with specific characteristics) occurred at a specific location along a pipeline.
To train and test the ANNs, the spatial distribution and the characteristics of the samples were defined to cover a range of values for the potential burst locations and sizes. For the pipeline described in Fig. 1, the transient head traces were generated by modeling a burst located at random distances on average every 0.2 m along the pipeline. This spatial distribution of samples was selected considering that generating samples with a larger spacing decreases the effectiveness of the ANN training and choosing a smaller spacing does not provide significantly better results, but it does result in a larger computational effort (Bohorquez et al. 2020). In addition, the use of random distances was introduced to avoid the risk that the ANNs would learn only from regularly spaced burst locations. Considering the previously described pipeline, if the burst location is changed every 0.2 m, the total number of samples is 5,000.
Each one of these samples was assigned a random burst size. As was described in Fig. 1, the burst was modeled as a circular orifice with a diameter DB. The range of possible burst sizes was defined considering the head drop that each burst size can cause in the transient head trace. The smallest burst size considered (DB=17  mm) causes a 1-m pressure head drop, and the largest burst (DB=88  mm) causes a 20-m pressure head drop at the burst location.

Transient Head Trace Sample Generation

Various authors have proposed the use of numerically simulated data to train ANNs that may be used to analyze real-time data (Perera and Rajapakse 2011; Zhou et al. 2019). In this paper, the transient pressure head samples for the ANNs training (Steps A.7 and A.8) were generated using the Method of Characteristics (MOC). This transient fluid calculation method transforms the two hyperbolic partial differential equations that govern the behavior of unsteady flow into four ordinary differential equations in order to obtain the variation of flow and head at different points along a pipeline at a given time. Two of these ordinary differential equations describe the relation between the time step (Δt), the spatial resolution of the calculation (Δx), and the wave speed of the transient wave in the pipeline (a). Only steady-state friction was considered in the transient numerical modeling. The effects of unsteady friction were neglected because the transient head trace samples only cover a maximum of 3.5L/a seconds after the occurrence of the burst, and during these first seconds, the effect of unsteady friction is not significant, as has been previously reported (Gong et al. 2014; Zhang et al. 2018).
Considering that the selected spatial separation between bursts is on average only 0.2 m, the selected time step needs to match this spatial resolution. The wave speed in the selected pipeline is 1,111  m/s, therefore the required time step is at least Δt=1.799E4  s. The total simulation time included a period of time before the occurrence of the burst and at least 3.5L/a seconds after the burst occurrence in order to capture the first cycle of reflections of the burst wave in the pipeline and to perform the burst occurrence consistency test (explained below at Step B.5). Thus, the total simulation time was 4.04 s. This means that each of the 5,000 generated transient head traces had more than 20,000 head values.

Downsampling and Processing

The potential size of the input data for the ANN training and testing when the sample spatial distribution and the required time step are considered as described in the previous step is approximately 100 million head values. Considering this, a timewise downsampling (Step A.4 in Fig. 2) was applied to the samples in the input data set since this has proven successful in the training and testing of ANNs with the ability to interpret transient head traces (Bohorquez et al. 2020). The downsampling frequency selected was 256 Hz to match the potential sampling frequency in the field considering existing technology.
Further processing of the downsampled transient head trace is required because the proposed technique is intended to work in near real time. A sliding time window concept (Mounce et al. 2010; Huang et al. 2018) is applied by partitioning each transient head trace into time windows (moving one data point forward at a time) that could contain the first period of transient wave reflections of the occurrence of a burst. To accomplish this, the length of each time window must be at least 2L/a seconds. In this paper, this length was selected as 2.5L/a seconds. Considering the downsampling frequency, the length of each time window, and the total length of the transient head trace (3.5L/a), each trace was transformed into 1,328 time windows. Therefore, the 5,000 transient head traces were transformed into 6.9 million time windows, each one with a length of 577 head values (corresponding to 2.25 s).

Time Window Sample Classification and Selection

Once the partitioning of the transient head traces is complete, the resulting time windows can be classified into three categories depending on the contained head information, as shown in Fig. 3. The first category is defined as Normal Head Condition, or Category N, since these time windows only contain part of the assumed slow sinusoidal head variation before the burst. Fig. 3(a) shows the variation of the head in two different scales to show the part of the sinusoidal head variation for this particular time window. The second category is defined as Abnormal Head Condition with Incomplete Information for Identification, or Category Ab-I. The time windows in this category capture the initial head drop due to the burst, but the burst wave reflection at the upstream reservoir is not included, as shown in Fig. 3(b). The last category is defined as Abnormal Head Condition with Complete Information for Identification, or Category Ab-C. The time windows in this category [as presented in Fig. 3(c)] contain a complete reflection of the transient wave created by the burst at the boundary conditions of the pipeline. Considering this, at Step A.5 of the model development stage, this classification is used to characterize the available time windows. For the pipeline described earlier, there are 3.5 million time windows in Category N, 1.1 million time windows in Category Ab-I, and 2.3 million time windows in Category Ab-C.
Fig. 3. Time window classification.
It is important to recognize that the total number of available time windows in the three categories is very large. With the purpose of reducing the required time for the ANNs training and facilitating the data management, only 20 time windows per category were randomly selected for each of the 5,000 locations of the modeled bursts. By conducting this selection, the total number of time windows was reduced to approximately 300,000. Different numbers of selected time windows were assessed in terms of the final ANN performance, and it was found that 20 time windows per location provides good performance with a reasonable computational time.

Burst Detection ANN Training and Testing

The methodology of the proposed technique includes the use of two different ANNs. First, a burst detection ANN was trained to analyze each time window and allocate it to one of the three categories described previously (Step A.7). The input data set for the training and testing of the burst detection ANN contains the 300,000 time windows selected at Step A.6. This input data set is then randomly divided into two groups, one of which is used for ANN training and the other for ANN testing (50% training and 50% testing). As presented in Fig. 4, the burst detection ANN receives as input one time window at a time with the information of the corresponding category and uses gradient search to find the best combination of weights to describe the training data set. The output of the burst detection ANN is the probability of each time window of belonging to each category and the final category for each time window is assigned as the category with the highest probability. Once the training process is complete, the testing data set is processed on the burst detection ANN to obtain a set of predicted categories, which can be compared to the correct category of each time window.
Fig. 4. Proposed ANNs (N = Normal, Ab-I = Abnormal Head Condition with Incomplete Information for Identification, and Ab-C = Abnormal Head Condition with Complete Information for Identification).

Burst Identification ANN Training and Testing

The second ANN in the proposed methodology is referred to as the burst identification ANN. This ANN was trained to analyze only time windows previously classified as Category Ab-C. The training and testing of this ANN are included in Step A.8. The input data set for this ANN was defined as half of the complete time window data set allocated to the last category at Step A.5 (1.2 million time windows). Similarly to the burst detection ANN, the input data set was randomly divided into two groups for the training and testing processes. The output of the burst identification ANN is the location and the size of a burst, based on the interpretation of a particular time window, as shown in Fig. 4.

Application Stage

The second stage of the technique presented in this paper (Stage B in Fig. 2) comprises the proposed steps that can be carried out to process and interpret a continuous transient head trace in order to detect and identify bursts. The steps presented in this section include the processing of a measured signal to be interpreted by the model created in the model development stage of the technique and could potentially be applied in near real time to a transient head measured signal.

Transient Pressure Head Data Retrieval

Once the burst detection and identification model has been created and validated, it can be applied to a continuous transient head trace obtained from a given measurement source. This transient head trace can be obtained from new and different numerical simulations, measurements from a laboratory setup, or pressure head measurements in a real pipeline. In either of the last two cases, a high-frequency pressure transducer is required for capturing the head variations on a real-time basis. The selected sampling frequency depends on factors such as the pipeline wall properties, the wave speed, and the downsampling frequency selected at Step A.4. In general, the transient head measurements should have at least the same frequency that was selected for the training of the ANNs. The data retrieval system (Step B.1) should also include data acquisition, processing, and communication modules to process the initial data obtained from the pressure transducer.

Sliding Time Window Selection

As presented in the model development stage, the analysis of the transient head trace is conducted by partitioning this trace into short time windows that are shifted one point at a time as the head measurements are obtained. A particular time window must be at least 2L/a seconds long in order to capture the first set of wave reflections caused by the occurrence of the burst. Step B.2 comprises this partitioning and the isolation of one time window at a time to be analyzed following the subsequent steps of the application stage.

Burst Detection Analysis

Once one time window of the transient head trace has been isolated, it is necessary to determine whether this time window contains any abnormal head variation that could have been caused by a burst. This burst detection analysis is composed of two different steps. The first step (Step B.3 in Fig. 2) includes the transformation of the current time window to match the sampling frequency that was selected for the training of the ANNs (Step A.4). As shown earlier, the downsampling frequency is selected based on the sampling frequency available for the head measurements or an estimation of the expected and required sampling frequency in the field.
The downsampled time window obtained from the continuous transient head trace can then be analyzed in the burst detection ANN obtained in Step A.7 (at Step B.4). The result from this ANN is a prediction of a category to characterize the current time window. There are three possible results of this analysis. If the burst detection ANN predicts that the current time window only contains normal head fluctuations, it is not necessary to continue the analysis, and the next time window can be selected (returning to Step B.2). This condition is defined as “Normal Condition” and in an alert system can be represented by the color green.
If the burst detection ANN predicts that the current time window belongs to Category Ab-I, then the condition of the pipeline is now defined as abnormal and can be represented by the color orange. In this case, the analysis continues to the next time window because for the predicted category it is known that this time window does not contain enough information to locate and characterize the burst. Lastly, if the burst detection ANN predicts that the current time window belongs to Category Ab-C, the analysis process continues to the burst identification analysis module. In this case, the condition of the pipeline is changed to “Abnormal Condition–Possible Burst” and may be represented by the color red.

Burst Identification Analysis

Of all the possible time windows analyzed in the application stage, the only ones that are analyzed in the burst identification module are those that are classified as Category Ab-C. The main objective of this module is to determine the location and size of the previously detected burst.
Once one time window has been defined in Category Ab-C, a consistency test is conducted (Step B.5 in Fig. 2). This consistency test determines how many windows have been classified in this same category immediately before the current time window. The main objective of this step is to provide a technique with more robustness to possible misinterpretations from the burst detection ANN. For instance, a particular time window might have been classified as Category Ab-C owing to normal fluctuations of the head in the pipeline, but the following time windows are again classified as Category Ab-I. In this case, the process continues until an invariant classification in Category Ab-C is obtained. The consistency test is considered complete once at least n time windows are continuously classified in this category, where n corresponds to the number of time windows that cover L/a seconds for the analyzed pipeline.
If the consistency test is completed, the analysis process continues to Step B.6, where the previously trained burst identification ANN (at Step A.8) is used to predict a possible location and size for the detected burst. It is important to mention that not just the last time window is used in this step, but the complete batch of time windows that have been included in the consistency test. Given that n time windows are used to obtain a prediction for the location and characteristics of the burst, n different combinations of predicted locations and sizes are obtained. Different statistical measures were considered to achieve a final prediction, and the median was selected because it is less sensitive to possible outliers in the prediction (potentially present due to misinterpretations of the burst identification ANN).

Burst Verification Analysis

It is important to highlight that the proposed technique merges existing knowledge of the impact of bursts in the transient head traces in a pipeline with the use of ANNs to rapidly and more accurately interpret these traces. However, the use of ANNs is also complemented by the use of transient numerical forward models to reinforce the robustness to the predictions. This is the case for the burst verification analysis. This analysis is comprised of two steps: a comparison of the measured transient head trace and a numerical transient head trace generated using the burst identification ANN predictions, and then a potential burst size prediction adjustment.
At Step B.7 (Fig. 2), a numerical simulation of the transient head trace caused by a burst with the predicted characteristics in the previous step is conducted. This transient head trace is then compared with the measured transient head trace using a metric such as the normalized root mean square error (NRMSE) to evaluate their match. If the prediction from the burst identification ANN is accurate enough, the NRMSE between both traces should be under a predefined threshold. In this case, an alert is raised for the pipeline, including the predicted location and size of the burst. For the examples presented in this paper the NRMSE threshold was defined as 2.0%.
In cases where the NRMSE between the two obtained transient head traces is above the threshold, a potential correction in the burst size prediction is conducted. Multiple tests demonstrated that when the predictions of the burst identification ANN are not accurate, most of the time it is due to an error in the prediction of the burst size. To correct for this, at Step B.8 different transient head traces are generated covering the complete range of possible burst sizes to find the burst size that causes a transient pressure trace that is very similar to the analyzed transient head trace (using again NRMSE) and thus adjusts the final prediction. If the value of the comparison metric improves after Step B.8, an alert will be raised with the original burst location prediction and the adjusted burst size prediction. However, if after conducting the burst size adjustment the NRMSE is still large, this means that none of the burst characteristics was predicted accurately, and the analysis will continue with the next time window at Step B.2.
The complete methodology described in this section was assessed using the numerical pipeline described earlier in this paper and was validated with experimental data. The results for these two applications are presented in the following sections.

Numerical Results

The methodology presented in Fig. 2 was applied to the system described in Fig. 1 and Table 1. A burst detection ANN and a burst identification ANN were designed, trained, and tested to detect and identify bursts. This section presents the results for both the model development stage and the application stage.

Model Development Stage

Considering the multiple steps involved in the model development stage (Step A.8 in Fig. 2), samples of numerical transient head traces were generated at random distances on average every 0.2 m along the pipeline to complete 5,000 transient head traces in total. Once all the transient head traces were generated, these were downsampled to a sampling frequency of 256 Hz and divided into time windows of length 2.5L/a seconds. A total of 6.9 million time windows were obtained (from Step A.5) and 20 time windows per burst location were selected for the training and testing of the burst detection ANN.
At Step A.7, a burst detection ANN was trained and tested using this input data set. Fig. 5 shows the classification performance of the burst detection ANN into two of the three possible categories. This accuracy is presented in terms of the percentage of time windows that are misclassified at each considered burst location. In each figure, the results are presented for the training and testing processes separately. It is important to note that the total accuracy of classification in Category Ab-I is 96.07% for the testing data set and 99.17% for the classification in Category Ab-C.
Fig. 5. Percentage of misclassified time windows for (a) Category Ab-I (incomplete); and (b) Category Ab-C (complete).
Fig. 5(a) presents the percentage of time windows misclassified in Category Ab-I. As can be seen in this figure, the behavior of the results is similar in the training and testing of the burst detection ANN, showing that there is no overfitting of the ANN to the training data. In addition, it is possible to observe that when the burst is located very close to the upstream end of the pipeline, the burst detection ANN has difficulty classifying a time window as Category Ab-I. This is explained by the fact that the transient head response of a burst located at this end only contains a quick head drop that, after which it almost immediately recovers. Fig. 5(b) shows the percentage of misclassified time windows for Category Ab-C. This figure demonstrates that the burst detection ANN does not present any significant difficulties in correctly classifying a time window in this category in either the training or testing steps of the development of the ANN.
Plots of the results of the accuracy in classifying a time window of Category N are not presented because, for the training and testing of the burst detection ANN, the accuracy in this category was 100%. At Step A.8, the burst identification ANN has now been trained and tested. Results for this step are presented in Fig. 6.
Fig. 6. Median error in estimation of (a) burst location; and (b) burst size along pipeline.
Fig. 6 shows the median error in the estimation of the burst location [Fig. 6(a)] and size [Fig. 6(b)], depending on the actual location of the burst (burst position). The median error was selected to present these results because several time windows corresponding to the same burst location were used for the training and testing purposes, and the use of this metric makes it possible to analyze an estimation of the distribution of the predictions. These two figures demonstrate that the burst identification ANN does not overfit to the training data set given that the testing data set predictions follow the same trend. The maximum median errors in the prediction of the burst location were found when the burst was located at the end of the pipeline. However, 99% of the median location prediction errors are smaller than 6.33 m (0.63% of the pipeline length). On the other hand, the maximum median error in the prediction of the burst size was found at the beginning of the pipeline, when the burst was located at only 0.36 m from the upstream end of the pipeline and most of the median size predictions were between 0.5 and +0.5  mm. Once the model development stage was completed, meaning that the two ANNs had been trained and tested, new numerical traces were generated to be tested at the application stage of the proposed technique.

Application Stage

To assess the performance of the burst detection and location ANNs and to demonstrate the application stage of the proposed technique, numerical traces corresponding to different bursts were generated to replicate Step B.1. This section presents the results for one of these traces as an example. The obtained transient head trace is presented in Fig. 7. This transient head trace corresponds to a burst located at 125.75 m along the pipeline with an orifice burst size of 56 mm. Fig. 7 also shows the length of a typical time window that would be selected at Step B.2. For the dimensions of the pipeline described in Fig. 1, the length of the time windows is 2.25 s.
Fig. 7. Numerical transient head trace for burst located at 125.75 m along pipeline and burst size of 56 mm.
Once the first time window has been selected, the burst detection analysis is conducted by downsampling the time window to match the sampling frequency of 256 Hz (Step B.3) and by processing this time window through the burst detection ANN (obtained at Step A.7). The results from this analysis are presented in Fig. 8 for the first 1,210 time windows of the full transient head trace of Fig. 7.
Fig. 8. Prediction of head condition categories in burst detection ANN for a numerical transient head trace.
Fig. 8 presents the category predicted by the burst detection ANN for each time window from Fig. 7. In addition, this plot includes the correct category in black for comparison. This figure demonstrates that the burst detection ANN classifies correctly all the time windows that correspond to Category N, and for the windows that correspond to Categories Ab-I and Ab-C, there is a short lag (of four time windows maximum) before the burst detection ANN correctly recognizes a time window. This classification lag is not significant and would represent a lag of 0.016 s in the detection of a burst.
Once a time window has been identified as Category Ab-C for the first time, the consistency test (over L/a seconds) is triggered (Step B.4 in Fig. 2). As an illustration, Fig. 8 includes a rectangle that encloses the 230 time windows that belong to the consistency test and have been processed in the burst identification ANN at Step B.6. The results for the prediction of the location and size of the burst are presented in Fig. 9.
Fig. 9. Distribution of prediction in burst identification ANN for burst (a) location; and (b) size for numerical transient head trace.
Considering that 230 time windows are included in the consistency test (covering L/a seconds of the transient head trace), the results from the burst identification ANN include 230 location and size predictions for the burst registered in the transient head trace. Fig. 9(a) shows that the obtained distribution of the burst location predictions is between 111.4 and 135.5 m when the real location is 125.75 m. This plot also shows that the obtained distribution includes predictions before and after the real burst location and that the median prediction is only 0.85 m away from the real burst location (which represents 0.085% of the total length of the 1,000-m-long pipeline). Fig. 9(b) presents the distribution of the predictions for the burst size, from which it is possible to determine that the median size of the burst is predicted to be 0.17 mm (an error of 0.30% from the actual burst size of 56 mm).
To complete the application stage, a burst verification analysis was conducted. Using the median predictions for the burst location and size, a transient head trace is generated and compared to the transient head trace presented in Fig. 7. This comparison is shown in Fig. 10.
Fig. 10. Burst verification analysis for numerical transient head trace.
From this figure it is possible to observe that the two transient head traces are quite similar, even considering that the initial head variation is different. In addition, the NRMSE was calculated, obtaining a difference of 1.1% between both traces. This proves that the proposed methodology results in accurate predictions of the burst location and size. Figs. 5 and 6 also demonstrate that the methodology yields good results when applied to transient head traces not used for the training of the ANNs.

Experimental Results

A series of experiments in the Robin Hydraulics Laboratory of the University of Adelaide were conducted to validate the proposed methodology for detecting and identifying bursts in pipelines using ANNs. The layout of the experimental pipeline system replicates the system configuration presented in Fig. 1. The upstream end of the pipeline is connected to a pressurized tank, and the downstream end has a closed inline valve. The characteristics of the pipeline are shown in Table 2.
To validate the methodology described in Fig. 2, all the steps of the model development stage were replicated for the pipeline in the laboratory. The architecture of the ANNs and the number of samples to create the input data set for the ANNs were not modified; however, new transient head traces were numerically generated taking into consideration the pipeline characteristics presented in Table 2.
The fast opening of a side-discharge solenoid valve was used to model the occurrence of a burst in the pipeline. Therefore, the numerical modeling of the transient head traces included a typical opening curve for a solenoid valve with an opening time of 10 ms. The downsampled frequency selected was 5 kHz considering that the pressure head in the experiment was measured at the downstream end of the pipeline using a PDCR 810 (Druck, Groby, Leicester) pressure transducer with a 10-kHz sampling rate.
Once the numerical transient head traces were generated, the time windows were obtained, classified, and used to train and test the burst detection and burst identification ANNs. Because the same procedure was explained for the numerical results, the results of the training and testing of these ANNs are not included here. However, considering that stochastic gradient descent algorithms were used for the ANN training, each time that this process is conducted, the resulting weights are different. Therefore, both ANNs were trained multiple times (without modifying its architecture) to assess the robustness of the obtained results. In the case of the burst detection ANN, the results were not sensitive to changes in the ANN weights, so only the results from one training attempt are presented. For the burst identification ANN, results will be presented in what follows in this section for three of these training processes. Each training process for the burst detection ANN took approximately 3 h, while the training of the burst identification ANN took 15 h on average.
The transient head trace obtained from the experimental test is shown in Fig. 11. This trace corresponds to a burst (modeled as the opening of a side-discharge solenoid valve) located 11.06 m downstream of the pressurized tank and has a size of approximately 2.0 mm. The initial pressure head before the occurrence of the burst was 29.48 m. Fig. 11 also presents the length of the time windows for this pipeline. The length selected was 2.5L/a seconds, which corresponds to 0.072 s.
Fig. 11. Experimental transient head trace.
The application stage was validated on the transient head presented in Fig. 11, following an individual analysis of each time window in sequence. Results from the burst detection analysis (Step B.4 in Fig. 2) are presented in Fig. 12. This figure presents the category predicted by the burst detection ANN for the first 577 time windows of the complete transient head trace. In addition, the correct category for each of these time windows is included with the objective of illustrating where the burst detection ANN misclassified a time window.
Fig. 12. Prediction of head condition categories in burst detection ANN for an experimental transient head trace.
As can be seen in the figure, most of the time windows are correctly classified, except for one time window that belongs to Category N and was classified as Category Ab-I. In addition, some other time windows were classified in Category N, but belong to Category Ab-I (showing a time lag in the ability of the burst detection ANN to identify the occurrence of a burst), and two time windows that belong to Category Ab-I were classified into Category Ab-C (not visible in the plot).
This figure also shows the time windows that were considered in the consistency test. A total of 144 time windows were classified as Category Ab-C and so are enclosed in the rectangle in the figure. These 144 time windows were processed by the three different burst identification ANNs, and the distribution of the predictions for the burst location and size are presented in Fig. 13.
Fig. 13. Distribution of predictions in burst identification ANN for (a) burst location (along 37.24-m pipeline); and (b) burst for experimental transient head trace.
Fig. 13(a) shows that the three burst identification ANNs provide different distributions of results, but for the three cases, the median predicted location is within 1.50% of error in terms of the total length of the pipeline (37.24 m), as shown in Table 3. The closest prediction to the real location of the burst was obtained for the third burst identification ANN in which the burst location was predicted to be only 0.06 m away from the real location. In a similar way, Fig. 13(b) presents the distribution in the predictions of the burst size. The three burst identification ANNs predicted different burst sizes, resulting in a median error between 6% and 16% of the true burst size. Although these results are less accurate than the burst location results, it is important to note that the largest median error corresponds to an absolute error of only 0.32 mm. In addition, the computational time required to obtain these results was only a few seconds.
Table 3. Results predicted by burst-identification ANNs. LT=37.4  m
Burst-identification ANNTrue location (m)True size (mm)Median location (m)Median size (mm)Median location error (%)Median size error (%)NRMSE (%)
111.062.0010.632.191.159.502.58
210.552.321.3716.005.61
311.002.120.166.002.83
Lastly, a burst verification analysis was conducted. Three different transient head traces were generated taking into consideration the prediction of the three burst identification ANNs and they were compared to the experimental head trace presented in Fig. 11. The comparison is shown in Fig. 14(a), and the obtained values for the NRMSE are presented in Table 3. There is significant agreement between the transient head trace obtained in the laboratory and the numerical transient head traces generated from the predictions of the burst identification ANNs in terms of the occurrence of the head drop and recovery due to the arrival of the burst wave to the measurement point. This is due to the accurate prediction of the burst location. However, differences are visible in the magnitude of the head drop for the three predictions. These differences are due to the error in the prediction of the burst size given that the magnitude of this head drop is directly related to the burst size. Following the proposed methodology, a burst prediction adjustment was applied taking into consideration the discrepancy between the transient head traces in Fig. 14(a) and the fact that the values for NRMSE are larger than the selected threshold.
Fig. 14. Burst verification analysis: (a) numerical transient head comparison; and (b) results after burst prediction adjustment.
Taking into consideration the prediction location obtained from the burst identification ANNs, a series of transient head traces with different burst sizes were generated covering the complete range of possible burst sizes (ranging from 0.4 to 3.5 mm in increments of 0.1 mm). For all of these transient head traces, the NRMSE was calculated when compared to the experimental transient head trace. Fig. 14(b) shows the transient head trace with the lowest NRMSE, which corresponds to a burst size of 2.0 mm (obtaining an error of 1.47%). It is important to highlight that only 31 transient head traces were generated for this adjustment, which does not represent an impractical computational time. Thus, the proposed technique was able to locate and characterize the occurrence of a burst in the laboratory with enough accuracy based only on the interpretation of the measured transient head trace.

Conclusions

This paper presents a novel technique to detect and characterize the occurrence of bursts in pipelines by merging the use of fluid transient waves and artificial neural networks. A complete methodology was described and divided into two main stages: model development and model application. The model development stage includes the generation of a transient head sample data set and the design, training, and testing processes to obtain two different ANNs for (1) the detection and (2) identification of a burst. The application stage includes the processing of a potentially continuous transient head trace into individual time windows to be processed through the two ANNs and a final stage of verification of the obtained results.
The technique was applied to a numerical example and validated with data obtained from a laboratory test at the University of Adelaide. In both cases, the technique was successful at detecting and identifying the occurrence of a burst in a pipeline. In the numerical application, a sharp burst was considered. The burst detection ANN accurately classified the analyzed time windows in one of three possible categories with only a short time lag of 0.016 s to identify transient head traces in the third category. The burst identification ANN predicted the occurrence of a burst 0.85 m from the real burst location (an error of 0.085% on the 1,000-m pipeline) with an error in the prediction of the burst size of 0.17 mm.
For the experimental laboratory application, the burst was modeled as a circular orifice that did not open instantaneously in order to test the technique with transient head traces that do not contain perfectly sharp transient waves. In this case, the burst detection ANN accurately classified most of the time windows. However, there was a slightly longer time lag in the identification of the first head deviation from normal conditions. For the burst identification ANN, three different trainings were conducted to test the robustness of the predictions. Each burst identification ANN predicted different burst locations and sizes; however, all the errors in the location prediction were within 1.5% of the total length of the pipeline. The burst size predictions were less accurate, oscillating between 6% and 16% of the real burst size.
Because this is the first application using fluid transient waves and ANNs, the training samples were obtained from a numerical model. However, future applications could use historical data from real systems and consider more general situations in terms of the burst characteristics and hydraulic systems analyzed. An application of this technique in a more complex hydraulic system might involve the development of more robust ANNs, the generation of more training samples, and the placement of more sensors in the system to collect enough information.

Data Availability Statement

The numerical samples used for the ANN training and the laboratory data that support the findings of this study are available from the corresponding author upon reasonable request.

References

Agapiou, A., D. D. Alexakis, K. Themistocleous, and D. G. Hadjimitsis. 2016. “Water leakage detection using remote sensing, field spectroscopy and GIS in semiarid areas of Cyprus.” Urban Water J. 13 (3): 221–231. https://doi.org/10.1080/1573062X.2014.975726.
Ahn, J., and D. Jung. 2019. “Hybrid statistical process control method for water distribution pipe burst detection.” J. Water Resour. Plann. Manage. 145 (9): 06019008. https://doi.org/10.1061/(ASCE)WR.1943-5452.0001104.
Andersen, J. H., and R. S. Powell. 2000. “Implicit state-estimation technique for water network monitoring.” Urban Water 2 (2): 123–130. https://doi.org/10.1016/S1462-0758(00)00050-9.
Bohorquez, J., B. Alexander, A. R. Simpson, and M. F. Lambert. 2020. “Leak detection and topology identification in pipelines using both fluid transients and artificial neural networks.” J. Water Resour. Plann. Manage. 146 (6): 04020040. https://doi.org/10.1061/(ASCE)WR.1943-5452.0001187.
Bohorquez, J., A. R. Simpson, and M. F. Lambert. 2018. “Characterization of transient pressure traces due to the effects of different anomalies and features in water pipelines.” In Proc., 13th Pressure Surges Conf., 151–169. Canfield, UK: BHR Group.
Brunone, B., M. Ferrante, S. Meniconi, and C. Massari. 2013. “Effectiveness assessment of pipe systems by means of transient test-based techniques.” Procedia Environ. Sci. 19 (Jan): 814–822. https://doi.org/10.1016/j.proenv.2013.06.090.
Capponi, C., M. Ferrante, A. C. Zecchin, and J. Gong. 2017. “Leak detection in a branched system by inverse transient analysis with the admittance matrix method.” Water Resour. Manage. 31 (13): 4075–4089. https://doi.org/10.1007/s11269-017-1730-6.
Cheng, W., H. Fang, G. Xu, and M. Chen. 2018. “Using SCADA to detect and locate bursts in a long-distance water pipeline.” Water 10 (12): 1727. https://doi.org/10.3390/w10121727.
Fahmy, M., and O. Moselhi. 2010. “Automated detection and location of leaks in water mains using infrared photography.” J. Perform. Constr. Facil. 24 (3): 242–248. https://doi.org/10.1061/(ASCE)CF.1943-5509.0000094.
Gong, J., M. F. Lambert, A. R. Simpson, and A. C. Zecchin. 2014. “Detection of localized deterioration distributed along single pipelines by reconstructive MOC analysis.” J. Hydraul. Eng. 140 (2): 190–198. https://doi.org/10.1061/(ASCE)HY.1943-7900.0000806.
Gong, J., A. R. Simpson, M. F. Lambert, A. C. Zecchin, Y. I. Kim, and A. Tijsseling. 2013. “Detection of distributed deterioration in single pipes using transient reflections.” J. Pipeline Syst. Eng. Pract. 4 (1): 32–40. https://doi.org/10.1061/(ASCE)PS.1949-1204.0000111.
Hoskins, A., and I. Stoianov. 2014. “InfraSense: A distributed system for the continuous analysis of hydraulic transients.” Procedia Eng. 70 (Jan): 823–832. https://doi.org/10.1016/j.proeng.2014.02.090.
Huang, P., N. Zhu, D. Hou, J. Chen, Y. Xiao, J. Yu, G. Zhang, and H. Zhang. 2018. “Real-time burst detection in district metering areas in water distribution system based on patterns of water demand with supervised learning.” Water 10 (12): 1765. https://doi.org/10.3390/w10121765.
Hunaidi, O., and P. Giamou. 1998. “Ground-penetrating radar for detection of leaks in curried plastic water distribution pipes.” In Proc., 7th Int. Conf. on Ground-Penetrating Radar Lawrence. Kansas: Radar Systems and Remote Sensing Laboratory.
Juliano, T., J. Meegoda, and D. Watts. 2013. “Acoustic emission leak detection on a metal pipeline buried in sandy soil.” J. Pipeline Syst. Eng. Pract. 4 (3): 149–155. https://doi.org/10.1061/(ASCE)PS.1949-1204.0000134.
Lee, P., M. Lambert, A. Simpson, J. Vitkovsky, and D. Misiunas. 2007a. “Leak location in single pipelines using transient reflections.” Australas. J. Water Resour. 11 (1): 53–65. https://doi.org/10.1080/13241583.2007.11465311.
Lee, P., J. P. Vitkovsky, M. F. Lambert, A. R. Simpson, and J. A. Liggett. 2007b. “Leak location in pipelines using the impulse response function.” J. Hydraul. Res. 45 (5): 643–652. https://doi.org/10.1080/00221686.2007.9521800.
Lee, S. J., G. Lee, J. C. Suh, and J. M. Lee. 2016. “Online burst detection and location of water distribution systems and its practical applications.” J. Water Resour. Plann. Manage. 142 (1): 04015033. https://doi.org/10.1061/(ASCE)WR.1943-5452.0000545.
Li, R., H. Huang, K. Xin, and T. Tao. 2014. “A review of methods for burst/leakage detection and location in water distribution systems.” Water Supply 15 (3): 429–441. https://doi.org/10.2166/ws.2014.131.
Liggett, J. A., and L. Chen. 1994. “Inverse transient analysis in pipe networks.” J. Hydraul. Eng. 120 (8): 934–955. https://doi.org/10.1061/(ASCE)0733-9429(1994)120:8(934).
Martins, R., A. Azevedo, A. B. Fortunato, E. Alves, A. Oliveira, and A. Carvalho. 2019. An innovative and reliable water leak detection service supported by data-intensive remote sensing processing, 96–108. Berlin: Springer.
Misiunas, D., M. F. Lambert, A. R. Simpson, and G. Olsson. 2007. “Assessing water mains condition using hydraulic transients.” Proc. Inst. Civ. Eng. Water Manage. 160 (2): 89–94. https://doi.org/10.1680/wama.2007.160.2.89.
Misiunas, D., J. P. Vítkovský, G. Olsson, A. R. Simpson, and M. F. Lambert. 2005. “Pipeline breaks detection using pressure transient monitoring.” J. Water Resour. Plann. Manage. 131 (4): 316–325. https://doi.org/10.1061/(ASCE)0733-9496(2005)131:4(316).
Mounce, S. R., A. Khan, A. S. Wood, A. J. Day, P. D. Widdop, and J. Machell. 2003. “Sensor-fusion of hydraulic data for burst detection and location in a treated water distribution system.” Inf. Fusion 4 (3): 217–229. https://doi.org/10.1016/S1566-2535(03)00034-4.
Mounce, S. R., and J. Machell. 2006. “Burst detection using hydraulic data from water distribution systems with artificial neural networks.” Urban Water J. 3 (1): 21–31. https://doi.org/10.1080/15730620600578538.
Mounce, S. R., R. B. Mounce, and J. B. Boxall. 2010. “Novelty detection for time series data analysis in water distribution systems using support vector machines.” J. Hydroinf. 13 (4): 672–686. https://doi.org/10.2166/hydro.2010.144.
Muggleton, J. M., M. J. Brennan, R. J. Pinnington, and Y. Gao. 2006. “A novel sensor for measuring the acoustic pressure in buried plastic water pipes.” J. Sound Vib. 295 (3): 1085–1098. https://doi.org/10.1016/j.jsv.2006.01.032.
Palau, C. V., F. J. Arregui, and M. Carlos. 2012. “Burst detection in water networks using principal component analysis.” J. Water Resour. Plann. Manage. 138 (1): 47–54. https://doi.org/10.1061/(ASCE)WR.1943-5452.0000147.
Perera, N., and A. Rajapakse. 2011. “Recognition of fault transients using a probabilistic neural-network classifier.” In Proc., 2011 IEEE Power and Energy Society General Meeting, 1–11. New York: IEEE.
Puust, R., Z. Kapelan, D. Savic, and T. Koppel. 2008. “Probabilistic leak detection in pipe networks using the SCEM-UA algorithm.” In Proc., Water Distribution Systems Analysis Symp. 2006, 1–12. Reston, VA: ASCE.
Rashid, S., S. Qaisar, H. Saeed, and E. Felemban. 2014. “A method for distributed pipeline burst and leakage detection in wireless sensor networks using transform analysis.” Int. J. Distrib. Sens. Netw. 10 (7): 939657. https://doi.org/10.1155/2014/939657.
Rathnayaka, S., B. Shannon, P. Rajeev, and J. Kodikara. 2016. “Monitoring of pressure transients in water supply networks.” Water Resour. Manage. 30 (2): 471–485. https://doi.org/10.1007/s11269-015-1172-y.
Romano, M., Z. Kapelan, and D. A. Savić. 2014. “Automated detection of pipe bursts and other events in water distribution systems.” J. Water Resour. Plann. Manage. 140 (4): 457–467. https://doi.org/10.1061/(ASCE)WR.1943-5452.0000339.
Rubio Scola, I., G. Besançon, and D. Georges. 2017. “Blockage and leak detection and location in pipelines using frequency response optimization.” J. Hydraul. Eng. 143 (1): 04016074. https://doi.org/10.1061/(ASCE)HY.1943-7900.0001222.
Shimanskiy, S., T. Iijima, and Y. Naoi. 2003. “Development of microphone leak detection technology on fugen NPP.” Prog. Nucl. Energy 43 (1): 357–364. https://doi.org/10.1016/S0149-1970(03)00043-X.
Srirangarajan, S., M. Iqbal, H. B. Lim, M. Allen, A. Preis, and A. J. Whittle. 2011. “Water main burst event detection and localization.” In Proc., Water Distribution Systems Analysis 2010, 1324–1335. Reston, VA: ASCE.
Tao, T., H. Huang, F. Li, and K. Xin. 2014. “Burst detection using an artificial immune network in water-distribution systems.” J. Water Resour. Plann. Manage. 140 (10): 04014027. https://doi.org/10.1061/(ASCE)WR.1943-5452.0000405.
Thomson, J., and L. Wang. 2009. State of technology review report of condition assessment of ferrous water transmission and distribution systems. Cincinnati, OH: USEPA.
Wang, X. J., M. F. Lambert, A. R. Simpson, J. Liggett, and J. Vítkovský. 2002. “Leak detection in pipelines using the damping of fluid transients.” J. Hydraul. Eng. 128 (7): 697–711. https://doi.org/10.1061/(ASCE)0733-9429(2002)128:7(697).
Wu, Y., and S. Liu. 2017. “A review of data-driven approaches for burst detection in water distribution systems.” Urban Water J. 14 (9): 972–983. https://doi.org/10.1080/1573062X.2017.1279191.
Wu, Y., S. Liu, K. Smith, and S. Wang. 2018b. “Using correlation between data from multiple monitoring sensors to detect bursts in water distribution systems.” J. Water Resour. Plann. Manage. 144 (2): 04017084. https://doi.org/10.1061/(ASCE)WR.1943-5452.0000870.
Wu, Y., S. Liu, and X. Wang. 2018a. “Distance-based burst detection using multiple pressure sensors in district metering areas.” J. Water Resour. Plann. Manage. 144 (11): 06018009. https://doi.org/10.1061/(ASCE)WR.1943-5452.0001001.
Ye, G., and R. A. Fenner. 2011. “Kalman filtering of hydraulic measurements for burst detection in water distribution systems.” J. Pipeline Syst. Eng. Pract. 2 (1): 14–22. https://doi.org/10.1061/(ASCE)PS.1949-1204.0000070.
Zhang, C., J. Gong, A. Zecchin, M. F. Lambert, and A. R. Simpson. 2018. “Faster inverse transient analysis with a head-based method of characteristics and a flexible computational grid for pipeline condition assessment.” J. Hydraul. Eng. 144 (4): 04018007. https://doi.org/10.1061/(ASCE)HY.1943-7900.0001438.
Zhou, X., Z. Tang, W. Xu, F. Meng, X. Chu, K. Xin, and G. Fu. 2019. “Deep learning identifies accurate burst locations in water distribution networks.” Water Res. 166 (Dec): 115058. https://doi.org/10.1016/j.watres.2019.115058.

Information & Authors

Information

Published In

Go to Journal of Water Resources Planning and Management
Journal of Water Resources Planning and Management
Volume 147Issue 1January 2021

History

Received: Mar 5, 2020
Accepted: Jun 15, 2020
Published online: Nov 9, 2020
Published in print: Jan 1, 2021
Discussion open until: Apr 9, 2021

Authors

Affiliations

Ph.D. Candidate, School of Civil, Environmental and Mining Engineering, Univ. of Adelaide, Adelaide 5005, Australia. ORCID: https://orcid.org/0000-0001-5071-8676. Email: [email protected]
Angus R. Simpson, M.ASCE [email protected]
Professor, School of Civil, Environmental and Mining Engineering, Univ. of Adelaide, Adelaide 5005, Australia (corresponding author). Email: [email protected]
Professor, School of Civil, Environmental and Mining Engineering, Univ. of Adelaide, Adelaide 5005, Australia. ORCID: https://orcid.org/0000-0001-8272-6697. Email: [email protected]
Senior Lecturer, School of Computer Science, Univ. of Adelaide, Adelaide 5005, Australia. ORCID: https://orcid.org/0000-0003-4118-2798. Email: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share