Water pipe networks have a large number of branch joints. Branch joint shunting generates vortices in the fluid, which excite the pipe wall to produce a type of branch noise. The branch noise is coupled with the leak source signal through the pipe. Here, a novel leak location protocol based on the complex-valued FastICA method (C-FastICA) is proposed to address the leak location problem under the branch noise interference. The C-FastICA, a complex-value domain blind deconvolution algorithm, effectively extended the cost function, constraint function, and iteration rules of the instantaneous linear FastICA into the complex-valued domain. The C-FastICA method was used to realize the separation of branch noise and leak source signal. The experimental results showed that the separation efficiency of the C-FastICA was higher than that of time-domain blind convolution separation (T-BCS). Furthermore, the relative location error of the C-FastICA method to the leak point was less than 14.238%, which was significantly lower than in traditional T-BCS and direct cross-correlation (DCC) technology.

  • The paper proposed a procedure based on the complex-valued FastICA method (C-FastICA) to address the leak location problem under the branch noise interference.

  • The accuracy of the proposed C-FastICA leak location method is higher than the traditional cross-correlation location technology and time-domain deconvolution leak location technology.

Leaks in water supply pipelines not only cause economic losses and environmental pollution, but also endanger the health and safety of urban residents. The main causes of pipe leaks are pipeline aging, chemical corrosion, external force damage, and management and maintenance negligence. According to the ‘In-depth Investigation and Investment Prospect Analysis Report of China's Urban Water Supply Industry for 2020–2026’ (Zhiyan Consulting 2019), the annual average rate of pipe leak in mainland China is 15.3% and some cities even exceed 25%, which is higher than the standard in many developed countries.

The acoustic emission (AE) detection method (Kim & Lee 2009; Brennan et al. 2016; Gao et al. 2017; Almeida et al. 2018; Kassab et al. 2019; Xiao et al. 2020) is widely used in detection of pipe leaks due to its low cost, non-intrusiveness, and high efficiency. The classic AE pipe detection is shown in Figure 1; the acoustic or vibration sensors are closely attached to both ends of the leak pipe to capture the leak signals. However, in real-world leak detection, various external and internal noises may interfere with the AE leak signal. External noises include those generated by automobiles, machines, and humans, whereas the internal noises are mainly caused by branch joints, elbow joints, valve joints, and other pipe joints. To effectively remove noise interference, Ahadi & Bakhtiar (2010) proposed a new method that captures the leak sound signal-signature in time domain based on the Short Time Fourier Transforms (STFT). The leak point was then located in the time-frequency domain using the tuned wavelet method. In a different study by Ting et al. (2021), a shift-invariant Dual Tree Complex Wavelet Transform (DTCWT) was used to remove noise interference in water pipe leak locations. Jafari et al. (2020) used an extension version of Kalman filter observer to detect and locate the blockage in the pipeline. Saqib et al. (2017) used a wavelet-based adaptive thresholding scheme to restrain the noise inter-ference within the AE leak signals. Taghvaei et al. (2006) used a cepstrum analysis to detect and locate leaks in pipeline networks.

Figure 1

Classic acoustic emission detection schematic.

Figure 1

Classic acoustic emission detection schematic.

Close modal

Butterfield et al. (2017) investigated the correlation between the leak flow rate and signal processing techniques such as the Vibro-Acoustic Emission (VAE) counts, signal Root Mean Square (RMS), peak in magnitude of the power spectral density, and octave banding. The results showed a strong correlation between the RMS and leak flow rate, prompting the development of a flow prediction model that was based on the RMS parameters. Subsequently, Mostafapour & Davoudi (2013) used the Donnell's non-linear theory, Galerkin method, and Runge–Kutta numerical method to develop an AE model induced by pipe leak vibration. Yu & Li (2017) proposed an experimental investigation into the detection of minor leaks. The time-domain waves, frequency, and energy features of the AE signals were first extracted and compared, before applying the SVM to recognize the minor leaks in the extracted features. In general, these de-noise methods are the procedure of removing or suppressing additive noise. However, these methods have two drawbacks. Firstly, environmental variables in actual leak detection conditions impact the mathematical models, interfering with the task and potentially leading to premature completion. Secondly, these methods generally capture leak signals from a straight leak pipe that is devoid of branch joints, and they consider mixed leak signals to be equivalent to the superposition of noise and leak sources.

In leak conditions, the mixed sound signal propagates along the pipe wall after the coupling and superimposing of pipe noise and the leak sound signal. Since the propagation mechanism is a convolutional process, the leak signal containing the pipe noise should be separated using a deconvolution method. All the noise generated by the pipe shape, including the noise caused by the branch flow, is a type of posterior noise with unknown time domain, frequency domain, and probability density characteristics. Therefore, the above-mentioned methods are mainly based on power spectrum characteristics, time-frequency characteristics, and other prior noise information that can be established or hypothesized to realize the AE signal processing analysis. Accordingly, the highlighted methods and techniques cannot be directly applied to eliminate branch noise because of the limited amount of prior knowledge of branch type leak detection. In addition, the branch noise and leak source signal have a convolution relationship with the propagation channel (pipe wall). Therefore, isolating the leak source signal from the branch-noise containing leak signal is a typical blind deconvolution separation (BSS) problem (Sun et al. 2015). Without knowing the probability density characteristic, time-domain and frequency-domain characteristics of the leak source signal and the branch noise signal, the blind separation algorithm can still complete the separation of the observation signal (leak mixing signal) obtained by the sound sensor.

In blind separation, FastICA is widely used due to its fast convergence and high separation precision. Therefore, to efficiently extract the leak source signal from the mixed leak signal containing branch noise, this paper proposes a leak detection procedure that is based on the complex-valued FastICA method (C-FastICA). As an improved algorithm of FastICA, C-FastICA effectively extends the cost function, constraint function, and iteration rules of FastICA to the complex-valued domain. In section two of this paper, a computational fluid dynamics (CFD) example is used to describe the current problem in leak location of branch pipe. Section three briefly introduces the C-FastICA theory. Section four covers the leak location experiment, and conclusions are made in section five.

The branch joint that connects a branch pipe to the main pipeline system changes the flow state of the fluid thus generating water eddies near the branch joint. The eddies excite the pipe wall, resulting in a kind of vortex-induced noise referred to in this paper as branch noise. This kind of branch noise has unsteady-state characteristics and can be regarded as an unsteady-state acoustic response caused by branch joint (Wang et al. 2005; Meniconi et al. 2021).

In this section, the computational fluid dynamics (CFD) software from ANSYS Fluent was used to simulate the eddy phenomenon in leak pipe with a branch joint (Figure 2). In preparatory work, the CAD model of leak pipe was built using SolidWorks 2014 software. The three-dimensional mesh was then generated using the ICEM software based on the CAD pipe model. The Fluent software provides a transient large eddy simulation (LES) solver. In calculation, a LES model was selected to simulate the steady flow field. The simulation was performed on a server called Dell Precision T7600 with an Intel Xeon (R) CPU E5-2678 W 0 at 3.10 GHz (16 CPUs) and 128 GB RAM. After the simulated calculation reached a steady state, we used CFD-Post software to post-process the calculation results.

Figure 2

CFD simulation flow chart for leak pipe with a branch joint.

Figure 2

CFD simulation flow chart for leak pipe with a branch joint.

Close modal

Figure 3 shows the distribution of vortex core near the branch joint at four different moments: t = 500 steps, t = 1000 steps, t = 1500 steps, and t = 2000 steps. Figure 3 shows that as the time step increases, the maximum of fluid velocity displays a decreasing trend. According to the definition of vorticity in , where is vorticity of vortex core, is the Hamiltonian, is the rotational velocity of the fluid, the vorticity is positively correlated with the fluid velocity, and the decrease in velocity leads to a decrease in vorticity. Due to the existence of this dynamically vortex core changing (as shown within the small black rectangle), the tube wall is excited to vibrate and to form vortex-induced noise, more details about vortex-induced noise are described in the literature (Chu 2006; Zhang et al. 2015; Zhang et al. 2016). This vortex-induced noise is mixed with the leak signal source and collected by a sensor installed on the pipeline, which becomes a major interference source for leak detection. Thus, we inferred that the leak detection will inevitably be interfered with due to the existence of one or more branch joints.

Figure 3

Illustration of vortex core distribution near branch joint.

Figure 3

Illustration of vortex core distribution near branch joint.

Close modal

Building convolution model for leak

In pipe leak accidents, the branch noise is coupled with the leak source signal through the pipe wall, and the leak mixed signal is then superimposed on the background noise to be acquired by a sound sensor. The background noise is mainly composed of environmental noise outside the pipeline, which is additive to the leak mixed signal (Ting et al. 2021; Deep et al. 2022). The mathematical analysis formula of the leak model is as follows:
(1)
where and are observation matrices acquired from two sensors; and are two different transfer matrices; is the unknown mixing source matrix, ; is leak source; is branch noise; and and are the random interference factors. To further simplify Equation (1), we transformed its time domain to frequency domain, resulting in the product form of Equation (1):
(2)
Usually, the z is omitted:
(3)

In the frequency domain, the convolutional mixing model was transformed into a linear instantaneous mixing model, and the C-FastICA was used to separate the mixed signal. Following the completion of the calculations, the frequency domain results were converted into time-domain sources, as shown in Figure 4.

Figure 4

Blind separation sketch of convolutive mixtures in frequency-domain.

Figure 4

Blind separation sketch of convolutive mixtures in frequency-domain.

Close modal

Background of the C-FastICA method

C-FastICA is a complex-valued extension form of FastICA, where the cost function, constraint function, and iterative learning rules of FastICA are effectively extended to the complex-valued domain. In the complex-valued domain, the source signal, mixed signal, and unmixing matrix are all in a complex form, and they are represented as follows:
(4)
where s is the source signal, sr is the real part of s, and si is the imaginary part of s; x is the mixed signal, xr is the real part of x, and xi is the imaginary part of x; W is the unmixing matrix, and Wr is the real part of W, Wi is the imaginary part of W. According to the Kuhn-Tucker conditions (Lv et al. 2007), under the constraint condition , the maximum cost function was obtained by the first order derivative of the following formula:
(5)
where is the nonlinear function, it is defined as . Next, we differentiated the real and imaginary parts of W in Equation (5). The first term on the left side of Equation (5) can be written as:
(6)
The * represents the complex conjugate. Equation (6) can be arranged as a vector form:
(7)
The second term on the left side of Equation (5) can be written as:
(8)
We used the Newton iteration method (Chen 2017) to solve the optimization results of Equation (5). In the first step, we presented the Jacobian matrix approximate form of the matrix in Equation (5):
(9)
Then, like in the first step the Jacobian matrix approximate form of the matrix was:
(10)
Combining Equations (9) and (10) resulted in the Jacobian matrix of the cost function Equation (5) as follows:
(11)
The Newton iteration method was used to iterate on Equation (11):
(12)
We multiplied both sides of Equation (12) by , thus, the iterative learning rule for solving the mixing matrix W was obtained as follows:
(13)
Moreover, in order to avoid the uncertainty of the output vector amplitude, the obtained at each step was normalized:
(14)
Finally, by combining Equations (10) and (11), the time domain algorithm of FastICA was successfully extended to the complex-valued domain, that is, the instantaneous linear blind separation iterative learning rule of C-FastICA was obtained. Then we calculated the frequency-domain decomposition result of the mixed signal x:
(15)
where is the frequency-domain output estimation, the is frequency-domain source signal, Wf is the frequency-domain unmixing matrix. After the calculation of Equation (12), we transformed the complex-valued output into the time domain:
(16)
where is the time-domain output estimation, the proposed C-FastICA is summarized in Algorithm 1.
Algorithm 1: Complex- valued FastICA (C-FastICA) 
(1). The time-domain observation signal is transformed to complex-valued form through FFT:
(2). Initialize
(3). According to the definition , calculate the and
(4). Initialize using Equation (12); 
(5). For iter ≥ 0: 
Calculate 
,  
(6). End for convergence:
(7). Cross-correlation estimation:  
(8). Leak location:  
Algorithm 1: Complex- valued FastICA (C-FastICA) 
(1). The time-domain observation signal is transformed to complex-valued form through FFT:
(2). Initialize
(3). According to the definition , calculate the and
(4). Initialize using Equation (12); 
(5). For iter ≥ 0: 
Calculate 
,  
(6). End for convergence:
(7). Cross-correlation estimation:  
(8). Leak location:  

The main leak location steps are shown in Figure 5, and are listed below as used in the experiment:

Figure 5

Flowchart of the leak location in the branch pipeline system.

Figure 5

Flowchart of the leak location in the branch pipeline system.

Close modal

Step 1: two mixing leak signal x1(n) and x2(n) were acquired by acceleration sensors.

Step 2: the obtained x1(n) and x2(n) were transformed into frequency domain, then, two frequency-domain signals x1(f) and x2(f) were obtained.

Step 3: the x1(f) and x2(f) were treated as input for C-FastICA separation, and the unmixing matrixes W1(f) and W2(f) were calculated.

Step 4: we further computed the output y1(f) and y2(f) based on the unmixing matrixes W1(f) and W2(f).

Step 5: the output y1(f) and y2(f) were transformed into time domain, then, the time-domain output y1(t) and y2(t) were obtained.

Step 6: the cross-correlation method was used to obtain the time delay estimation (TDE) based on the obtained y1(t) and y2(t) in step 5.

Step 7: we combined the known distance L = 70 m, TDE, and the empirical leak sound velocity c = 1495 m/s to complete the leak location task.

Conditions of the experiment

In this section, we used the acceleration sensors to collect two mixing leak signals containing branch noise, which were then separated using the C-FastICA algorithm. We selected a cast iron water supply pipe with a diameter of 100 mm and a test length of 70 m for the experiment. The branch node has water consumption in the corresponding branch pipeline. The pipe inner pressure was 0.6 MPa, and the inlet flow rate was approximately 2 m/s. The simulated leak point, which was 42 m away from sensor 1, was replaced by a small valve (leakage aperture = 15 mm) and the data sampling rate was set at 10 KHz. More details on the used parameters are shown in Table 1. A schematic diagram of the leak location experiment (Figure 6), and real diagrams of the two sensors and branch joints (Figure 7) are as shown below. Two sensors were placed at the different ends of the main pipe, and the collected leak signals were wirelessly transmitted to the host computer (Figure 6).

Table 1

Detailed parameters of the experiment

ParametersValues
Pipe diameter 100 mm 
Leak valve diameter 15 mm 
Test pipe length: L1 + L2 + L3 70 m 
L1: Sensor 1 distance from the leak point L1 = 42 m 
L2: Branch joint distance from the leak point L2 = 8 m 
L3: Sensor 2 distance from the branch joint L3 = 20 m 
Pipe inner pressure 0.6 MPa 
Flow velocity 2 m/s 
Sampling rate 10 KHz 
ParametersValues
Pipe diameter 100 mm 
Leak valve diameter 15 mm 
Test pipe length: L1 + L2 + L3 70 m 
L1: Sensor 1 distance from the leak point L1 = 42 m 
L2: Branch joint distance from the leak point L2 = 8 m 
L3: Sensor 2 distance from the branch joint L3 = 20 m 
Pipe inner pressure 0.6 MPa 
Flow velocity 2 m/s 
Sampling rate 10 KHz 
Figure 6

Schematic of leak location experiment.

Figure 6

Schematic of leak location experiment.

Close modal
Figure 7

Pictures of branch joint, Sensor 1 and Sensor 2.

Figure 7

Pictures of branch joint, Sensor 1 and Sensor 2.

Close modal

Leak location processing

We chose the leak signal collected by Sensor 1 as an example, and used the C-FastICA algorithm to process it. The time-domain (Figure 8(a)) and frequency-domain (Figure 8(b)) forms of the leak signal containing branch noise are shown below.

Figure 8

(a) Time-domain leak signal from Sensor 1, (b) Leak signal power spectrum from Sensor 1.

Figure 8

(a) Time-domain leak signal from Sensor 1, (b) Leak signal power spectrum from Sensor 1.

Close modal

To prove that the C-FastICA algorithm outperforms the traditional algorithm, we separated the mixing leak signal using the well-known time-domain blind convolution separation (T-BCS) algorithm. The signal had to be preprocessed before separating the mixing leak signal to suppress the influence of the additive noise n1 and n2 in Equation (1). After removing the additive noise in the mixing leak signal, the leak acoustic source and branch noise are the main components in the mixing leak signal. Then, the C-FastICA was used to perform blind separation for the mixing leak signal in the frequency domain. For more information on suppressing the additive noise, see Bell & Sejnowski (1995). The C-FastICA algorithm revealed that the spectral components of y1 and y2 were mainly occupied in the range of 1700–2600 and 1100–4000 Hz, respectively (Figure 9(a)). The T-BCS algorithm results indicated that the spectral characteristics of y1 and y2 were between 1600–2200 and 1000–4000 Hz, respectively (Figure 9(b)). This denotes that y2 has a wider frequency bandwidth than y1. Therefore, we can infer from the previous leak theory(Diao et al. 2020; Scussel et al. 2021) that the randomness of the leak signal is often greater than that of other stable noise signals, except white noise, implying that the bandwidth of the leak source signal is wider than other noise signals. Consequently, we concluded that y2 was the leak source signal obtained by the separation algorithm, and y1 was the branch noise. When the number of branch pipe nodes is more than one, there will be multiple narrow frequency components in the separation result, and these narrow frequency components are considered branch noise.

Figure 9

Separated results of C-FastICA and T-BCS: (a) separated components y1 and y2 based on C-FastICA; (b) separated components y1 and y2 based on T-BCS.

Figure 9

Separated results of C-FastICA and T-BCS: (a) separated components y1 and y2 based on C-FastICA; (b) separated components y1 and y2 based on T-BCS.

Close modal
To evaluate the convergence speed of the algorithms, the inter-symbol interference (ISI) concept was introduced and defined as follows (Jenq 1979):
(17)
where is the (i, j)th element in the global system matrix C. The algorithm is more convergent when the ISI output curve of each channel is closer to zero. The ISI output curves indicated that each of the two channels converged at about 300 steps in C-FastICA (Figure 10(a)) and at about 400 steps in T-BCS (Figure 10(b)). Thus, convergence is faster in the C-FastICA algorithm than in T-BCS arigorithm.
Figure 10

ISI changes between C-FastICA channels and T-BCS channels.

Figure 10

ISI changes between C-FastICA channels and T-BCS channels.

Close modal
The crosstalk error PI was introduced to compare the separated degree of the C-FastICA and the T-BCS. The closer the PI value is to zero, the higher the separation degree of the mixed signal. The PI was defined as follows (Leung & Siu 2007):
(18)
where M is the number of source components, cik is the element of C = W·A, A is the mixing matrix, and W is the separation matrix obtained by separation algorithms. The C-FastICA algorithm and the T-BCS algorithm were used to separate the leak signal containing the branch noise for another 50 times. It should be emphasized that the leak mixing signal and other specific experimental parameters used in these 50 times experiments remain the same. The PI results are shown in Figure 11.
Figure 11

PI values of C-FastICA and T-BCS were blind separated for 50 times.

Figure 11

PI values of C-FastICA and T-BCS were blind separated for 50 times.

Close modal
The PI coefficients of C-FastICA algorithm were smaller, with a median value of about 0.192, than those of the T-BCS algorithm which had a median value of about 0.237 (Figure 11). Therefore, C-FastICA had a higher separation degree for the mixing leak signals, and the obtained leak source contained more leak characteristics. Following the separation calculation for the mixing leak signal collected by Sensor 1, we processed the mixing leak signal collected by Sensor 2 in the same way. The cross-correlation method was then used to obtain the TDE τ based on the two leak source signals:
(19)
where and are two leak sources obtained by blind separation, is from Sensor 1, is from Sensor 2, is the delay parameter. The leak hole location was computed using the distance L1 = 42 m, TDE τ, and the known leak sound velocity c = 1495 m/s as follows:
(20)
where L is equivalent to . The peak value obtained using the C-FastICA algorithm (Figure 12(a)) was the closest to the true leak location compared to peak values shown by the T-BCS algorithm (Figure 12(b)) and the direct cross-correlation method (DCC; Figure 12(c)).
Figure 12

(a) Leak location using the C-FastICA method, is equivalent to 36.82 m; (b) Leak location using the T-BCS method, is equivalent to 35.71 m; (c) Leak location using the DCC method, is equivalent to 61.24 m.

Figure 12

(a) Leak location using the C-FastICA method, is equivalent to 36.82 m; (b) Leak location using the T-BCS method, is equivalent to 35.71 m; (c) Leak location using the DCC method, is equivalent to 61.24 m.

Close modal

Since the branch noise was not removed before the analysis using the DCC method, it had a significant impact on the cross-correlation estimation thus yielding a leak location value with the largest error (Figure 12(c)) compared to the other two methods. To further verify that the separation performance of C-FastICA for mixed leak signals was better than T-BCS, this experiment was replicated eight times. The leak location results for each method are shown in Table 2. In the eight times leak location experiments in Table 2, eight acquisitions of leak mixing signal were carried out, the leak mixing signal used in each experiment is different.

Table 2

Leak location results based on C-FastICA, T-BCS and DCC

NumberC-FastICA
T-BCS
DCC
Distance (m)Distance (m)Distance (m)
36.82 12.333 35.71 14.976 61.24 45.810 
36.95 12.024 34.07 18.881 NaN NaN 
37.05 11.786 34.59 17.643 NaN NaN 
37.54 10.619 32.82 21.857 58.96 40.381 
37.12 11.619 33.29 20.738 66.53 58.405 
36.76 12.476 35.17 16.262 NaN NaN 
36.02 14.238 34.69 17.405 NaN NaN 
37.33 11.119 34.18 18.619 56.82 35.286 
37.98 9.571 35.02 16.619 NaN NaN 
NumberC-FastICA
T-BCS
DCC
Distance (m)Distance (m)Distance (m)
36.82 12.333 35.71 14.976 61.24 45.810 
36.95 12.024 34.07 18.881 NaN NaN 
37.05 11.786 34.59 17.643 NaN NaN 
37.54 10.619 32.82 21.857 58.96 40.381 
37.12 11.619 33.29 20.738 66.53 58.405 
36.76 12.476 35.17 16.262 NaN NaN 
36.02 14.238 34.69 17.405 NaN NaN 
37.33 11.119 34.18 18.619 56.82 35.286 
37.98 9.571 35.02 16.619 NaN NaN 

The relative location error (δ) of C-FastICA did not exceed 14.238%, while in the T-BCS algorithm it reached 21.857%, indicating that C-FastICA algorithm had a higher position accuracy than T-BCS algorithm. However, noise had a significant impact on the DCC method whose relative error value was greater than 58%, and the leak location could not even be completed in numbers 2, 3, 6, 7, and 9.

This paper proposed the use of the C-FastICA method to separate the mixing leak signal and to solve the branch noise interference problem in the leak location of branch water pipes. Following the separation calculation, the leak sources were identified, and the leak location was completed by combining the prior leak sound velocity and the known pipe length. The feasibility of the proposed leak location method was proved through experiments, and the following conclusions were reached:

  • (1)

    When there was a branch joint in a water pipeline, the leak signal was mixed with the branch noise, and the T-BCS and DCC methods for locating the leak point had a significant error or were unable to complete the location. It shows that the influence of the branch noise cannot be ignored on the branch pipe leak location.

  • (2)

    The experiments proved that the C-FastICA algorithm has a faster convergence speed and a higher separation level for the mixed leak signals than the T-BCS algorithm.

This project was supported by the National Natural Science Foundation of China (No. 51675069), the National Natural Science Foundation of China (No. 51775070), and the Fundamental Research Funds for the Central Universities (Nos. 2018CDQYGD0020, cqu2018CDHB1A05).

Data cannot be made publicly available; readers should contact the corresponding author for details.

Ahadi
M.
&
Bakhtiar
M. S.
2010
Leak detection in water-filled plastic pipes through the application of tuned wavelet transforms to Acoustic Emission signals
.
Applied Acoustics
71
,
634639
.
https://doi.org/10.1016/j.apacoust.2010.02.006
.
Almeida
F. C. L.
,
Brennan
M. J.
,
Joseph
P. F.
,
Gao
Y.
&
Paschoalini
A. T.
2018
The effects of resonances on time delay estimation for water leak detection in plastic pipes
.
Journal of Sound and Vibration
420
,
315
329
.
https://doi.org/10.1016/j.jsv.2017.06.025
.
Brennan
M. J.
,
Kroll De Lima
F.
,
De Almeida
F. C. L.
,
Joseph
P. F.
&
Paschoalini
A. T.
2016
A virtual pipe rig for testing acoustic leak detection correlators: proof of concept
.
Applied Acoustics
102
,
137
145
.
https://doi.org/10.1016/j.apacoust.2015.09.015
.
Butterfield
J. D.
,
Krynkin
A.
,
Collins
R. P.
&
Beck
S. B. M.
2017
Experimental investigation into vibro-acoustic emission signal processing techniques to quantify leak flow rate in plastic water distribution pipes
.
Applied Acoustics
119
,
146
155
.
http://dx.doi.org/10.1016/j.apacoust.2017.01.002.
Chen
Y. Z.
2017
Evaluation of the degenerate scale in Laplace equation by using Newton iteration method
.
Engineering Analysis with Boundary Elements
80
(
December 2016
),
105
107
.
http://dx.doi.org/10.1016/j.enganabound.2017.02.013
.
Chu
A. K. H.
2006
Passive transport of atomic gases within a vortex tube induced by a noise wave
.
Physics Letters, Section A: General, Atomic and Solid State Physics
349
(
1–4
),
135
145
.
Deep
K.
,
Fong
S.
&
Narasimhan
S.
2022
Active acoustic leak detection and localization in a plastic pipe using time delay estimation
.
Applied Acoustics
187
,
108482
.
https://doi.org/10.1016/j.apacoust.2021.108482
.
Diao
X.
,
Jiang
J.
,
Shen
G.
,
Chi
Z.
,
Wang
Z.
,
Ni
L.
,
Mebarki
A.
,
Bian
H.
&
Hao
Y.
2020
An improved variational mode decomposition method based on particle swarm optimization for leak detection of liquid pipelines
.
Mechanical Systems and Signal Processing
143
,
106787
.
https://doi.org/10.1016/j.ymssp.2020.106787
.
Gao
Y.
,
Brennan
M. J.
,
Liu
Y.
,
Almeida
F. C. L.
&
Joseph
P. F.
2017
Improving the shape of the cross-correlation function for leak detection in a plastic water distribution pipe using acoustic signals
.
Applied Acoustics
127
,
24
33
.
http://dx.doi.org/10.1016/j.apacoust.2017.05.033
.
Jafari
R.
,
Razvarz
S.
&
Gegov
A.
2020
Blockage detection in pipeline based on the extended kalman filter observer
.
Electronics
9
(
91
),
161
174
.
http://dx.doi.org/10.3390/electronics9010091
.
Jenq
Y. C.
1979
Probability of error in digital fiber optic systems with inter-symbol interference and signal dependent additive noise
.
Journal of the Franklin Institute
307
(
5
),
291
303
.
https://www.sciencedirect.com/science/article/pii/0016003279900334
.
Kassab
S.
,
Michel
F.
&
Maxit
L.
2019
Water experiment for assessing vibroacoustic beamforming gain for acoustic leak detection in a sodium-heated steam generator
.
Mechanical Systems and Signal Processing
134
,
106332
.
https://doi.org/10.1016/j.ymssp.2019.106332
.
Kim
M. S.
&
Lee
S. K.
2009
Detection of leak acoustic signal in buried gas pipe based on the time-frequency analysis
.
Journal of Loss Prevention in the Process Industries
22
(
6
),
990
994
.
http://dx.doi.org/10.1016/j.jlp.2008.08.009
.
Lv
Y.
,
Hu
T.
,
Wang
G.
&
Wan
Z.
2007
A penalty function method based on Kuhn-Tucker condition for solving linear bilevel programming
.
Applied Mathematics and Computation
188
(
1
),
808
813
.
Meniconi
S.
,
Cifrodelli
M.
,
Capponi
C.
,
Duan
H.-F.
&
Brunone
B.
2021
Transient response analysis of branched pipe systems toward a reliable skeletonization
.
Journal of Water Resources Planning and Management
147
(
2
),
04020109
.
Mostafapour
A.
&
Davoudi
S.
2013
Analysis of leakage in high pressure pipe using acoustic emission method
.
Applied Acoustics
74
(
3
),
335
342
.
http://dx.doi.org/10.1016/j.apacoust.2012.07.012
.
Saqib
N. U.
,
Mysorewala
M. F.
&
Cheded
L.
2017
A novel multi-scale adaptive sampling-based approach for energy saving in leak detection for WSN-based water pipelines
.
Measurement Science and Technology
28
(
12
),
125102
.
Scussel
O.
,
Brennan
M. J.
,
Almeida
F. C. L.
,
Muggleton
J. M.
,
Rustighi
E.
&
Joseph
P. F.
2021
Estimating the spectrum of leak noise in buried plastic water distribution pipes using acoustic or vibration measurements remote from the leak
.
Mechanical Systems and Signal Processing
147
,
107059
.
https://doi.org/10.1016/j.ymssp.2020.107059
.
Sun
J.
,
Xiao
Q.
,
Wen
J.
&
Yang
Y.
2015
Underdetermined blind source separation of pipeline leak vibration signals based on empirical mode decomposition and joint approximate diagonalization of eigenmatrices
.
Journal of Vibroengineering
17
(
3
),
1326
1340
.
Taghvaei
M.
,
Beck
S. B. M.
&
Staszewski
W. J.
2006
Leak detection in pipelines using cepstrum analysis
.
Measurement Science and Technology
17
(
2
),
367
372
.
Ting
L. L.
,
Tey
J. Y.
,
Tan
A. C.
,
King
Y. J.
&
Abd Rahman
F.
2021
Water leak location based on improved dual-tree complex wavelet transform with soft thresholding de-noising
.
Applied Acoustics
174
,
107751
.
https://doi.org/10.1016/j.apacoust.2020.107751
.
Wang
X. J.
,
Lambert
M. F.
&
Simpson
A. R.
2005
Behavior of short lateral dead ends on pipeline transients: a lumped parameter model and an analytical solution
.
Journal of Fluids Engineering, Transactions of the ASME
127
(
3
),
529
535
.
Xiao
R.
,
Joseph
P.
&
Li
J.
2020
The leak noise spectrum in gas pipeline systems: theoretical and experimental investigation
.
Journal of Sound and Vibration
488
,
115646
.
https://doi.org/10.1016/j.jsv.2020.115646
.
Zhang
T.
,
Zhang
Y. O.
&
Ouyang
H.
2015
Structural vibration and fluid-borne noise induced by turbulent flow through a 90° piping elbow with/without a guide vane
.
International Journal of Pressure Vessels and Piping
125
,
66
77
.
http://dx.doi.org/10.1016/j.ijpvp.2014.09.004
.
Zhang
N.
,
Xie
H.
,
Wang
X.
&
Wu
B. S.
2016
Computation of vortical flow and flow induced noise by large eddy simulation with FW-H acoustic analogy and Powell vortex sound theory
.
Journal of Hydrodynamics
28
(
2
),
255
266
.
http://dx.doi.org/10.1016/S1001-6058(16)60627-3
.
Zhiyan Consulting
2019
In-depth Investigation and Investment Prospect Analysis Report of China's Urban Water Supply Industry for 2020–2026
.
Beijing
.

Author notes

These authors contributed equally to this work.

This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the original work is properly cited (http://creativecommons.org/licenses/by-nc-nd/4.0/).