Next Article in Journal
Cloud Modelling of Property-Level Flood Exposure in Megacities
Previous Article in Journal
Magnetic Metal–Organic Framework Enhanced Inorganic Coagulation for Water Purification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study of Navigable Flow Conditions in the Intermediate Channel of Decentralized Cascade Locks

1
Key Laboratory of Transport Technology in Navigation Building Construction, Nanjing Hydraulic Research Institute, Nanjing 210029, China
2
The National Key Laboratory of Water Disaster Prevention, Nanjing Hydraulic Research Institute, Nanjing 210029, China
*
Author to whom correspondence should be addressed.
Water 2023, 15(19), 3394; https://doi.org/10.3390/w15193394
Submission received: 31 August 2023 / Revised: 20 September 2023 / Accepted: 21 September 2023 / Published: 27 September 2023
(This article belongs to the Section Hydraulics and Hydrodynamics)

Abstract

:
In this study, the effects of the different conveyance modes of the intermediate channel in decentralized cascade locks on navigation flow conditions were investigated. A new hybrid numerical simulation method was established to evaluate navigable flow conditions in intermediate channels at different water conveyance modes. This hybrid numerical simulation method was reliably compared by physical modeling tests. We used the 33.73 m class high-head intermediate channel filled with water as a study case. The study used the maximum water surface slope and maximum flow velocity as evaluation indexes for navigable flow conditions. The results showed that the navigable flow conditions of the centralized water conveyance mode were worse compared to the decentralized water conveyance mode in the intermediate channel. Especially in the upstream region of the intermediate channel with a centralized outflow, the navigable flow conditions were exceptionally harsh. We recommend the decentralized outflow mode in the high-head intermediate channel. This study provides an effective numerical simulation method for optimizing the water conveyance mode of the high-head intermediate channel of decentralized cascade locks and saving project investment.

1. Introduction

Inland waterway freight systems play a key role in the trade and commerce of many countries. It has low cost and low carbon emission compared to road and rail transportation [1,2,3]. However, the first problem of inland waterway freight systems is how to overcome the huge dams with complex terrains that have large differences between upstream and downstream heights. Some dams are over 100 m tall, and others are over 200 m [4]. For example, four world-class high dams, namely Wudongde, Baihetang, Xiluodu, and Xiangjiaba, have been built in the lower reaches of the Jinsha River in southwest China [5]. Ship locks can help ships successfully overcome high dams (see Figure 1). When the ship sailed downstream in decentralized cascade ship locks, it entered the upstream chamber from the upstream approach channel and closed the upstream gate. The chamber emptied with water into the intermediate channel to lower the water level of the chamber. Once the water level in the upstream chamber matched that of the intermediate channel, the downstream gate was opened, allowing the ship to enter the intermediate channel. The process was repeated in a similar manner when passing through the downstream lock. Sailing upstream was analogous to sailing downstream for ships.
The decentralized cascade lock was a navigation building connecting locks at both ends through the restricted intermediate channel. It effectively reduced the working head of single-stage locks and made full use of the complex terrain [6]. When the upstream and downstream locks filled and emptied water into the intermediate channel, unsteady flow fluctuations occurred in the intermediate channel. Specifically, during water discharge from the upstream lock into the intermediate channel, a positive surge was generated in the intermediate channel [7]. During water discharge from the intermediate channel to the downstream lock, a negative surge formed within the intermediate channel [8]. Unsteady flow fluctuations were reflected and superimposed in propagation to form a complex flow state, thus causing the vessel’s transverse rocking, longitudinal tilt, and droop effect, affecting vessel navigation safety [9]. It was crucial to limit the fluctuations caused by the filling and emptying of locks to ensure the vessel’s safety, where the water surface slope and flow velocity are the important parameters to measure whether the vessel can navigate safely [10].
Traditional intermediate channels used a centralized outflow mode where the filling and emptying of water did not share a common water transmission culvert [11]. It exhibited excellent adaptability to intermediate channels with a small head. However, the concentrated outflow mode seriously deteriorated the intermediate channel’s flow state with a high head. It posed a considerable risk to the safety of vessel mooring and navigation. To optimize the flow condition, we proposed a novel water conveyance mode for water transmission in the intermediate channel—the dispersed outflow. Intermediate channel filling and emptying water shared a common culvert. Water was no longer concentrated in an outflow from one end, but distributed from different locations.
In previous studies, the three-dimensional model was established to calculate the hydraulic characteristic value for the filling and emptying of navigation structures with water [12,13,14,15]. However, the scale of the intermediate channel was significantly larger than that of the ship lock, making it difficult to create a three-dimensional mathematical model with both high quality and efficiency.
A hybrid numerical simulation method of the intermediate channel of dispersed outflow was established in this paper, which was applied to evaluate the effect of different water conveyance modes on the navigable flow condition of the intermediate channel. This hybrid numerical simulation method first established the corresponding unsteady flow energy equation based on the culvert’s flow state. Subsequently, combined with the results of the three-dimensional numerical simulation of the outflow zone, the flow rate at different times of each outflow zone was quantitatively calculated when the upstream locks filled with water to the intermediate channel. Finally, the flow rate of each outflow zone has been used as the boundary for the two-dimensional numerical simulation. This approach allowed us to determine the navigable flow conditions of the intermediate channel during the filling with water. In addition, we verified the reliability of the hybrid numerical simulation method by physical modeling tests. Additionally, the reliability of the hybrid numerical simulation method was verified. According to the research results, the suggestion of the water conveyance mode of the intermediate channel was put forward.

2. Materials and Methods

2.1. Study Area

A 1:30 physical model of the dispersed outflow intermediate channel, which met the gravity similarity criterion, was constructed, as shown in Figure 2. The upstream and downstream chambers were 180 m long and 23 m wide, with the total area being about 5300 m2. The intermediate channel was about 930 m long and 55 m wide, with the total area being about 46,400 m2. There was a 15 m wide, 4 m high water conveyance culvert at the bottom of the intermediate channel. The distances between the upper boundary of the intermediate channel and the first outflow zone, between the first outflow zone and the second outflow zone, and between the second outflow zone and the third outflow zone were approximately 100 m, 375 m, and 375 m, respectively. This paper described filling the intermediate channel from the upstream lock at a constant 8 min opening rate using a valve. Its working head was 33.73 m, and the initial water depth of the intermediate channel was 11.27 m.

2.2. Hybrid Numerical Simulation Method

In this paper, only the intermediate channel-filling was considered. The physical model test result for filling the intermediate channel with a working head of 33.73 m is shown in Figure 3. Q1, Q2, and Q3 were the culvert flow rate in the previous sections of the first, second, and third outflow zones. Additionally, q1 and q2 were the flow rate in the first and second outflow zones. Notably, the flow rate in the third outflow zone matched the culvert flow in its preceding section. Four flow states were shown in Figure 4: from time 0 to t1, Q 1 , Q 2 , Q 3 , q 1 , q 2 > 0 ; from time t1 to t2, Q 1 , Q 2 , Q 3 , q 2 > 0 and q 1 < 0 ; from time t2 to t3, Q 2 , Q 3 , q 2 > 0 and Q 1 , q 1 < 0 ; and from time t3 to t4, Q 2 , Q 3 > 0 and Q 1 , q 1 , q 2 < 0 . This study focused on hybrid numerical simulations within the time interval from 0 to t4 to simplify the calculation.

2.2.1. Governing Equations

Different split ratios directly affected the magnitude of the resistance loss coefficient [16], and Figure 5 shows six different flow patterns in the outflow zone.
Define the resistance loss coefficient under different flow conditions:
k i . j k l = p i . k p i . l g ρ / v 2 2 g
where k i . j k l is the resistance loss coefficient of water from position k (k = 1, 2, 3) to position l (l = 1, 2, 3) in the flow state of the i outflow zone and the j flow state; p i . k is the total pressure on position k; v is the corresponding speed of the maximum flow rate; g is the acceleration of gravity; and ρ is the density of water.
Prior to deriving the governing equation for intermediate channel filling, some assumptions were made: (1) the water surface fluctuation is neglected in the chamber and intermediate channel during water filling, (2) the sum of all frictional and minor head losses in outflow zone is characterized by loss coefficient, and (3) internal inertia term of each outflow zone is ignored.
Taking the case Q 1 > Q 2 > Q 3 > 0 as an example, as shown in Figure 4a, the other three flow regimes gave the control equations in the Appendix A.
Combining the continuity equation for a culvert,
Q 1 = A 1 A 2 A 1 + A 2 d H d t
where A1 is the area of the upstream chamber, A2 is the area of the intermediate channel, and H is the difference between the upstream chamber and the intermediate channel.
The unsteady flow equations between the upstream lock and the first outflow zone are
H = ξ + ξ t + D 1 + k 1 . a 12 2 g S 2 Q 1 Q 1 + L 1 g S d Q 1 d t .
where ξ and ξ t are the resistance coefficient from the upstream chamber to the first outflow zone and the valve resistance coefficient, respectively; D1 is the resistance coefficient along the first section of the culvert, which is calculated by Darcy’s formula; S is the area of the culvert; and L1 is converted inertia length from the upstream chamber to the first outflow zone.
Combining the unsteady flow energy equations between the culvert section in front of the 1st outflow zone and the culvert section in front of the 2nd outflow zone
H 1 + p 1 ρ g + v 1 2 2 g = H 2 + p 2 ρ g + v 2 2 2 g + k 1 . a 13 v 1 v 1 2 g + L 2 g d v 2 d t + D 2 v 2 v 2 2 g
where H 1 + p 1 ρ g + v 1 2 2 g is the total energy of the culvert section in front of the i (i = 1, 2, 3) outflow zone and v i is the speed of the culvert section in front of the i outflow zone.
Combining the unsteady flow energy equations between the culvert section in front of the 1st outflow zone and the water surface of the intermediate channel
H 1 + p 1 ρ g + v 1 2 2 g = k 1 . a 12 v 1 v 1 2 g + h c
Combining the unsteady flow energy equations between the culvert section in front of the 2nd outflow zone and the water surface of the intermediate channel
H 2 + p 2 ρ g + v 2 2 2 g = k 2 . a 12 v 2 v 2 2 g + h c
Combining the Equations (4)–(6).
k 1 . a 13 k 1 . a 12 2 g S 2 Q 1 Q 1 + L 2 g S d Q 2 d t + D 2 + k 2 . a 12 2 g S 2 Q 2 Q 2 = 0
Likewise, Q 3 can be derived.
k 2 . a 13 k 2 . a 12 2 g S 2 Q 2 Q 2 + L 3 g S d Q 3 d t + D 3 + k 3 . a 12 2 g S 2 Q 3 Q 3 = 0

2.2.2. Solution of Governing Equations

Firstly, the governing equation was integrated and then solved by computer. The valve resistance coefficient ξ t and the outflow zone resistance coefficient k i . j k l underwent minor changes over shorter Δ t . In order to facilitate the solution, we assumed that the valve resistance coefficient and the outflow zone resistance coefficient were constant over time Δ t . When the unsteady flow energy equation was integrated, these two coefficients were regarded as constants.
The first-order Taylor expansion of H and Q at the time t 0 was carried out.
t t 0 + Δ t H dt = t 0 t 0 + Δ t H t 0 + d H d t t 0 ( t t 0 ) d t = H t 0 Δ t + d H d t t 0 Δ t 2 2
t t 0 + Δ t Q dt = t 0 t 0 + Δ t Q t 0 + d Q d t t 0 ( t t 0 ) d t = Q t 0 Δ t + d Q d t t 0 Δ t 2 2
Wylie highlighted the following relationship in the study of unsteady pipe flow [17].
t t + Δ t Q Q d t = Q t Q t + Δ t Δ t
Equations (2), (3), (7), and (8) were integrated.
H t + Δ t = H t 0 Q 1 t 0 Δ t C + d Q 1 d t t 0 Δ t 2 2 C
Q 1 ( t 0 + Δ t ) = H t 0 Δ t + d H d t t 0 Δ t 2 2 + L 1 g S Q 1 ( t 0 ) δ + ξ + k 1.12 2 g S 2 Q 1 ( t 0 ) Δ t + L S g
Q 2 ( t 0 + Δ t ) = k 1.12 k 1.13 2 g S 2 Q 1 ( t 0 ) Q 1 ( t 0 + Δ t ) Δ t + L 2 g S Q 2 ( t 0 ) L 2 g S + D 2 + k 2.12 2 g S 2 Q 2 ( t 0 ) Δ t
Q 3 ( t 0 + Δ t ) = k 2.12 k 2.13 2 g S 2 Q 2 ( t 0 ) Q 2 ( t 0 + Δ t ) Δ t + L 3 g S Q 3 ( t 0 ) L 3 g S + D 3 + k 3.12 2 g S 2 Q 3 ( t 0 ) Δ t
If we knew the hydraulic parameters at any given moment t , we could subsequently determine the hydraulic parameters for the moment t + Δ t . It was essential to highlight that the magnitude Δ t directly impacts the precision of the computational outcomes. If the Δ t is too large, the previous assumption that the valve resistance coefficient and the outflow zone resistance coefficient were unchanged within time Δ t will become invalid.

2.3. The Calculation of the Outflow Zone Resistance Coefficient

2.3.1. Governing Equations

With the development of computer technology, computational fluid dynamics (CFD) has been considered a suitable alternative or auxiliary method [18,19,20]. In this paper, the value k i . j k l was determined through the three-dimensional numerical simulation. The controlling continuity equations were the mass conservation equation for incompressible flows and the momentum conservation equation in the Cartesian coordinate system:
u = 0
u t + u u = P ρ + v 2 u + f b + u , ¯ u ,
where u is velocity, P is pressure, f b is gravitational body force and external body force, u , is turbulent velocity fluctuation, and v is dynamic viscosity.
The k ω SST turbulence model was a hybrid model that combines the characteristics of the original k ω model near the wall and the characteristics of the k ε model away from the wall. In addition, the k ω STT model could be likened to the convergence of the k ω model [21]. The k ω SST turbulent model more accurately described the motion of the fluid in numerical simulations of large curvature [22,23,24]. In this study, the outflow zone features a sharp curvature of 90 degrees. Consequently, we employed the k ω SST turbulence model for this simulation.

2.3.2. Computational Domains and Mesh

As shown in Figure 6, the outflow zone was axisymmetric. Only half of the outflow zone was considered for numerical simulation to simplify the model. Each outflow zone was 89.6 m in length and 18 m in width. Two branch holes, spaced 4 m apart, were incorporated every 5.6 m along the zone. These branch holes were 4 m in length, 0.3 m in width, and 1.5 m in height. A total of 32 branch holes, each with an energy dissipation cover plate, were present in every outflow zone.
Three structured grid sizes were used to check the grid irrelevance of the computational domain model dx = 0.15 m, 0.20 m, and 0.25 m; the total grid sizes were about 1.07 million, 1.94 million, and 4.81 million, respectively. The grid independence was tested under the condition that the velocity of inlet 1 was 2.4 m/s and the velocity of inlet 3 was −3 m/s. As shown in Figure 7, the velocity along the culvert no longer varied significantly when dx 0.2 . Therefore, the dx = 0.2 m was selected as the final grid size for model discretization for the precision and time of response.

2.3.3. Boundary Condition

Two velocity inlet boundaries and a pressure outlet boundary controlled different flow regimes in the outflow zone with different splitting ratios (see Figure 6). The wall surface was a no-slip boundary with a viscous bottom standard wall function. The SIMPLEC algorithm and the second-order upwind discrete format were used to solve the momentum equations. The commercial software ANSYS fluent 17.0 was used to help discretize the solution domain and the set of equations to find a more reliable approximate solution.

2.3.4. Simulation Validation

The empirical equation of the single branch hole physical model was chosen to validate the three-dimensional numerical model. The single branch hole outflow model was similar to Figure 5a.
As shown in Figure 8, the results presented through the numerical simulation were in good agreement with the results of the physical experiments in the single branch hole, which indicated that the branch hole resistance coefficient was reflected by the three-dimensional model. Therefore, we were confident in accurately calculating the resistance coefficient for the outflow zone through the three-dimensional model.

2.4. Numerical Simulation of Intermediate Channel

2.4.1. Governing Equations

For the intermediate channel, the one-dimensional hydrodynamic model was difficult to reflect fully the required navigable flow conditions. The two-dimensional hydrodynamic model was well-developed and has been widely used in the numerical simulation of rivers [26,27]. In this study, the horizontal scale was larger than the vertical depth scale. Therefore, we ignored the impact of the Coriolis force and wind. The two-dimensional shallow water equation was selected as the governing equation, which was shown in Equations (19)–(21).
U t + F U = S
U = h h u h v , F = F x , F y , F x = h u h u 2 + g h 2 / 2 h u v , F y = h u h u v h u 2 + g h 2 / 2
S = 0 g h S 0 x S f x g h S 0 y S f y
where u and v are the velocities in x and y directions, respectively; S 0 x = Z x and S 0 y = Z y are the bottom slope terms of the x and y directions, respectively; Z is the bottom elevation; S f x = n 2 u u 2 + v 2 h 4 / 3 and S f y = n 2 v u 2 + v 2 h 4 / 3 are the friction slopes of the x and y directions, respectively; and n is the Manning coefficient.

2.4.2. Computational Domains and Mesh

The two-dimensional numerical simulation area was the entire intermediate channel. The unstructured mesh was used to divide the simulation domain. This mesh comprises a total of 5931 elements and 3192 nodes. Enhanced refinement was applied to the mesh at the two ends of the intermediate channel, and the grid and the computational domain, as shown in Figure 9.
The result derived from the numerical simulation in Section 2.2 was used as the flow rate of the two-dimensional numerical simulation outlet zone. The total flow change curve of centralized outflow was consistent with the decentralized outflow, but all the flow had the first outflow partition zone, as shown in Figure 10. Feature points were set at 35 m intervals along the middle channel and 3 m from the convex bank, with a total of 26.

2.5. Simulation Method Validation

To validate the reliability of the simulation method, we extracted the experiment results of the water level of the upstream chamber, the average water level of the intermediate channel, each culver flow rate, and the water level in different positions of the intermediate channel (see Figure 11 and Figure 12). The statistical indicator of Kling Gupta efficiency (KGE) was used to evaluate the preference for the hybrid numerical simulation method [28,29]. The equation for KGE was as follows:
K G E = 1 R 1 2 + β 1 2 + γ 1 2 β = C s i m ¯ C exp ¯ , γ = σ s i m / C d . s i m ¯ σ exp / C d . exp ¯ R = i = 1 n C exp . i C exp ¯ × C s i m . i C s i m ¯ i = 1 n C exp . i C exp ¯ i = 1 n C s i m . i C s i m ¯
where n is the total number of data, C exp is the experimented value, C s i m is the simulated value, C exp ¯ is the mean value of the experimented value, C s i m ¯ is the mean value of the simulated value, σ exp is the standard deviation of the experimented value, σ s i m is the standard deviation of the simulated value, and R is the correlation coefficient.
The results presented through the hybrid numerical simulation were in good agreement with the results of the physical experiments (the KGE of all parameters exceeded 0.95), which indicated this numerical method was reliable.

3. Results

As was seen in Figure 13, the maximum water surface slope in the intermediate channel with a centralized outflow was significantly worse than that in the intermediate channel with a decentralized outflow. The maximum water surface slope of the intermediate channel with a concentrated outflow was 2.65‰ at 300 s, which seriously threatened the ship’s berthing and navigation safety. The maximum water surface slope of the intermediate channel with a dispersed outflow changed more slowly, and the maximum of only 0.62‰ appeared in 300 s.
Additionally, we paid attention to the water surface slope at different locations of the intermediate channel. Consequently, the water surface slope data were extracted for the upstream, middle, and downstream regions in the intermediate channel (see Figure 14). It was quite obvious that the intermediate channel with a centralized outflow was larger than the water surface slope with a decentralized outflow. However, there was little difference in the water surface slope between the middle and downstream regions. A large amount of water for a short period flowed from the upstream chamber to the intermediate channel when the upstream lock emptied. All water flowed out in the first outflow zone for the concentrated outflow mode, which made the water level near the outflow zone higher than other areas. A propulsive wave was formed above the water surface and propagated downstream, which made the water surface slope drop in the upstream region by a much larger amount than that in the middle and downstream regions. However, water flowed out of each outflow zone in a dispersed outflow mode, thereby alleviating the excessive local water surface slope in the upstream region.
The change in the maximum flow velocity in the intermediate channel during water filling is shown in Figure 15, with only the outflow mode changed. The maximum flow velocity of the intermediate channel with a concentrated outflow was larger, and the flow velocity change was more intense. The maximum flow velocity of the intermediate channel with a concentrated outflow was 1.45 m/s, which appeared in 420 s. The maximum flow velocity of the intermediate channel with a dispersed outflow was 0.56 m/s, which appeared in 300 s.
We were also interested in the flow state in the intermediate channel with different outflow modes at 6 min (see Figure 16). Water flowed from the only outflow zone into the intermediate channel and diffused downstream in the concentrated outflow water conveyance mode. Water could flow into the channel from multiple outflow zones in the intermediate channel with a dispersed outflow, meaning it no longer only flowed downstream. Changing the direction of water flow diffusion helps to decrease the discharge for unit width in the intermediate channel. The water level fluctuation was ignored relative to the water depth. The flow velocity was directly related to the discharge for unit width. From the perspective of flow velocity, the choice of a dispersed outflow was better than that of the concentrated outflow, especially in the upper area of the intermediate channel with a concentrated outflow, and the navigable flow conditions were particularly harsh.

4. Discussion

The most common method was to treat all the branch holes as a single entity, and the decentralized filling–emptying system was simplified into a centralized filling–emptying system for calculation [30]. However, this method could not calculate the flow rate in each branch hole, much less understand the flow conditions inside the lock intermediate channel. The hybrid numerical simulation method is the most important innovation in this paper, whose core is to calculate the flow rate of each outflow zone based on the flow patterns in the culvert.
Poor water flow conditions in the upstream region are unavoidable in the intermediate channel with a high head and concentrated outflow mode. The method generally adopted is to increase the resistance coefficient of the filling–emptying system and thus reduce the water flow rate, but this will undoubtedly increase the filling time and reduce the navigable efficiency. Hybrid numerical simulation results show that the intermediate channel with decentralized water conveyance has better navigable flow conditions than the centralized outflow, and so it can consider the safety of the ship’s navigation and berthing and navigable efficiency. These results are similar to the superiority of decentralized filling–emptying systems over centralized filling–emptying in high-head ship locks [31]. However, the water conveyance mode of the dispersed outflow is more complicated in its design. It is necessary to consider the problem of the uneven outflow in each outflow zone due to the inertia of the long water conveyance culvert.
Similar to many studies, this study encountered limitations. This paper discusses two different discharge modes of the intermediate channel in the upstream locks operating alone. The layout of the outflow zone within the intermediate channel, operation modes (the filling and emptying can be carried out separately, simultaneously, or at different times), and others have not been thoroughly discussed. In the future, we are looking for a more efficient and safe operation and arrangement of decentralized locks to provide a reference for constructing high-head, large-scale intermediate channels.

5. Conclusions

This paper established a hybrid numerical simulation method to comprehensively evaluate the influence of the outflow mode of the intermediate channel in decentralized cascade locks on its navigable flow conditions. The maximum water surface slope and the maximum flow velocity were evaluation indices of the navigable flow condition. The hybrid numerical simulations were conducted for the intermediate channel filling with a working head of 33.73 m. The main conclusions are as follows.
1. The flow in each culvert of the chamber’s water level, the average water level of the intermediate channel, and the water level in different positions of the intermediate channel were used as standards. The hybrid numerical simulation method in this paper, to quantify the individual hydraulic parameters of the intermediate channel, was reliable by the 1:30 physical model test.
2. When centralized water conveyance is arranged, the intermediate channel with a high head is more worrying. Both indicators of the centralized outflow are much worse than the decentralized outflow for the intermediate channel. The maximum water surface slope and the maximum flow velocity increased by 327.4% and 158.9% in the numerical simulation, respectively. The navigable flow conditions are extremely poor, especially in the upstream region of the centralized outflow.
3. We recommend that the intermediate channel of the high-head decentralized cascade locks adopt a decentralized outflow mode to safeguard the navigable flow conditions in the channel.

Author Contributions

L.M.: Data collection, Methodology, Visualization, Writing—original draft; J.A.: Conceptualization, Writing—review and editing; Z.L.: Writing—review and editing; S.Z.: Data collection; L.D.: Data collection. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data available on request due to restrictions eg privacy or ethical. The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors thank the reviewers for their constructive comments and suggestions.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Flow state of Figure 4b governing equations:
Q 1 = C 1 d H d t
H = ξ + ξ t + D 1 2 g S 2 Q 1 Q 1 + k 1 . b 13 k 1 . b 23 2 g S 2 Q 2 Q 2 + + L 1 g S d Q 1 d t
L 2 g S d Q 2 d t + k 1 . b 23 + D 2 + k 2 . a 12 2 g S 2 Q 2 Q 2 = 0
k 2 . a 13 k 2 . a 12 2 g S 2 Q 2 Q 2 + L 3 g S d Q 3 d t + D 3 + k 3 . a 12 2 g S 2 Q 3 Q 3 = 0
Flow state of Figure 4c governing equations:
Q 1 = C 1 d H d t
H = ξ + ξ t 2 g S 2 Q 1 Q 1 + k 1 . c 21 2 g ω 2 Q 1 Q 2 Q 1 Q 2 + L 1 g S d Q 1 d t
L 2 g S d Q 2 d t k 1 . c 23 2 g ω 2 Q 1 Q 2 Q 1 Q 2 + D 2 + k 2 . a 12 2 g S 2 Q 2 Q 2 = 0
k 2 . a 13 k 2 . a 12 2 g S 2 Q 2 Q 2 + L 3 g S d Q 3 d t + D 3 + k 3 . a 12 2 g S 2 Q 3 Q 3 = 0
Flow state of Figure 4d governing equations:
Q 1 = C 1 d H d t
H = ξ + ξ t 2 g S 2 Q 1 Q 1 + k 1 . c 21 2 g ω 2 Q 1 Q 2 Q 1 Q 2 + L 1 g S d Q 1 d t
L 2 g S d Q 2 d t k 1 . c 23 2 g ω 2 Q 1 Q 2 Q 1 Q 2 + k 2 . b 13 k 2 . b 23 2 g S 2 Q 3 Q 3 + D 2 2 g S 2 Q 2 Q 2 = 0
L 3 g S d Q 3 d t + k 2 . b 23 + D 3 + k 3 . a 12 2 g S 2 Q 3 Q 3 = 0

References

  1. Liang, S.; Yang, X.; Bi, F.; Ye, C. Vessel traffic scheduling method for the controlled waterways in the upper Yangtze River. Ocean Eng. 2019, 172, 96–104. [Google Scholar] [CrossRef]
  2. Zhou, D.Q.; Huang, F.; Wang, Q.W.; Liu, X. The role of structure change in driving CO2 emissions from China’s waterway transport sector. Resour. Conserv. Recycl. 2021, 171, 105627. [Google Scholar] [CrossRef]
  3. Hecht, J.S.; Lacombe, G.; Arias, M.E.; Dang, T.D.; Piman, T. Hydropower dams of the Mekong River basin: A review of their hydrological impacts. J. Hydrol. 2019, 568, 285–300. [Google Scholar] [CrossRef]
  4. Ma, H.Q.; Chi, F.D. Technical Progress on Researches for the Safety of High Concrete-Faced Rockfill Dams. Engineering 2016, 2, 332–339. [Google Scholar] [CrossRef]
  5. Wan, H.; Xie, Y.; Li, B.; Cai, Y.; Yang, Z. An integrated method to identify and evaluate the impact of hydropower development on terrestrial ecosystem. Environ. Impact Assess. Rev. 2023, 99, 107042. [Google Scholar] [CrossRef]
  6. Meng, X. Study on Unsteady Flow in the Approach Channel of Ship Lock. Ph.D. Thesis, Tianjin University, Tianjin, China, 2010. [Google Scholar]
  7. Kiri, U.; Leng, X.; Chanson, H. Positive surge propagating in an asymmetrical canal. J. Hydro-Environ. Res. 2020, 31, 41–47. [Google Scholar] [CrossRef]
  8. Reichstetter, M.; Chanson, H. Negative Surges in Open Channels: Physical and Numerical Modeling. J. Hydraul. Eng.-Asce 2013, 139, 341–346. [Google Scholar] [CrossRef]
  9. Zheng, F. Characteristics of Ship-Lock-Induced Surges in an Intermediate Navigation Channel. Ph.D. Thesis, Hohai University, Nanjing, China, 2019. [Google Scholar]
  10. Wan, Z.Y.; Li, Y.; Wang, X.G.; An, J.F.; Cheng, L.; Liao, Y.P. Effect of ship-lock-induced surges on navigation safety in a branched lower approach channel system. J. Hydroinform. 2022, 24, 481–496. [Google Scholar] [CrossRef]
  11. Chen, Y.; Hu, Y.; Li, Z. Research on Influence from Ship Navigating in the Intermediate Channel between Ship Lifts on Hydraulic Characteristics. In Proceedings of the 10th PIANC Smart Rivers Conference, Smart Rivers 2022, Nanjing, China, 18–21 October 2022; pp. 599–610. [Google Scholar]
  12. Liu, J.K.; Hu, Y.A.; Li, Z.H.; Xue, S. Risk Analysis and Optimization of Water Surface Deviation from Shafts in the Filling-Emptying System of a Mega-Scale Hydro-Floating Ship Lift. Water 2021, 13, 1377. [Google Scholar] [CrossRef]
  13. Wu, B.; Hu, Y.-A.; Xue, S. Study on the flow field uniformity of hydro-floating ship lift combined type hydraulic-driven system based on the residual energy theory. J. Hydrodyn. 2022, 34, 1156–1161. [Google Scholar] [CrossRef]
  14. Scheffermann, J.M.; Stockstill, R.L. Pressures and discharges during operation of a lock valve, physical and numerical model results. J. Hydraul. Res. 2009, 47, 137–141. [Google Scholar] [CrossRef]
  15. Badano Nicolas, D.; Menendez Angel, N. Numerical modeling of Reynolds scale effects for filling/emptying system of Panama Canal locks. Water Sci. Eng. 2021, 14, 237–245. [Google Scholar] [CrossRef]
  16. Yin, Y.F.; Li, A.G.; Wu, D.M.; Wen, X.Q.; Li, J.X.; Guo, J.N.; Ma, Y.Q. Comparison and analysis of energy loss and flow characteristics of T-junctions via secondary flow and entropy production. Build. Environ. 2022, 225, 109635. [Google Scholar] [CrossRef]
  17. Wylie, E.B. The Microcomputer and Pipeline Transients. J. Hydraul. Eng. 1983, 109, 1723–1739. [Google Scholar] [CrossRef]
  18. Battiston, C.C.; Bombardelli, F.A.; Schettini, E.B.C.; Marques, M.G. Mean flow and turbulence statistics through a sluice gate in a navigation lock system: A numerical study. Eur. J. Mech.-B/Fluids 2020, 84, 155–163. [Google Scholar] [CrossRef]
  19. Paternina-Verona, D.A.; Coronado-Hernandez, O.E.; Espinoza-Roman, H.G.; Besharat, M.; Fuertes-Miquel, V.S.; Ramos, H.M. Three-Dimensional Analysis of Air-Admission Orifices in Pipelines during Hydraulic Drainage Events. Sustainability 2022, 14, 14600. [Google Scholar] [CrossRef]
  20. Daneshfaraz, R.; Norouzi, R.; Abbaszadeh, H.; Kuriqi, A.; Di Francesco, S. Influence of Sill on the Hydraulic Regime in Sluice Gates: An Experimental and Numerical Analysis. Fluids 2022, 7, 244. [Google Scholar] [CrossRef]
  21. Menter, F.R. Two-equation eddy-viscosity turbulence models for engineering applications. AIAA J. 1994, 32, 12149. [Google Scholar] [CrossRef]
  22. Yu, Y.F.; Liu, X.R.; Wu, J.H.; Meng, H.B. Comparison among Turbulence Models for a Novel Static Circulating Jet Mixer. Appl. Mech. Mater. 2010, 26–28, 382–385. [Google Scholar] [CrossRef]
  23. Pang, A.L.J.; Skote, M.; Lim, S.Y. Modelling high Re flow around a 2D cylindrical bluff body using the k-omega (SST) turbulence model. Prog. Comput. Fluid Dyn. 2016, 16, 48–57. [Google Scholar] [CrossRef]
  24. You, Y.; Seibold, F.; Wang, S.; Weigand, B.; Gross, U. URANS of turbulent flow and heat transfer in divergent swirl tubes using the k-ω SST turbulence model with curvature correction. Int. J. Heat Mass Transf. 2020, 159, 120088. [Google Scholar] [CrossRef]
  25. Zhang, R. Hydraulic Study on Unsteady Flow in Ship Locks with Composite Culverts. Master’s Thesis, Nanjing Hydraulic Research Institute, Nanjing, China, 1982. [Google Scholar]
  26. Zhang, S.; Jing, Z.; Li, W.; Wang, L.; Liu, D.; Wang, T. Navigation risk assessment method based on flow conditions: A case study of the river reach between the Three Gorges Dam and the Gezhouba Dam. Ocean Eng. 2019, 175, 71–79. [Google Scholar] [CrossRef]
  27. Timbadiya, P.V.; Krishnamraju, K.M. A 2D hydrodynamic model for river flood prediction in a coastal floodplain. Nat. Hazards 2023, 115, 1143–1165. [Google Scholar] [CrossRef]
  28. Daneshfaraz, R.; Norouzi, R.; Abbaszadeh, H.; Azamathulla, H.M. Theoretical and experimental analysis of applicability of sill with different widths on the gate discharge coefficients. Water Supply 2022, 22, 7767–7781. [Google Scholar] [CrossRef]
  29. Knoben, W.J.M.; Freer, J.E.; Woods, R.A. Technical note: Inherent benchmark or not? Comparing Nash-Sutcliffe and Kling-Gupta efficiency scores. Hydrol. Earth Syst. Sci. 2019, 23, 4323–4331. [Google Scholar] [CrossRef]
  30. Davis, J.P. Hydraulic Design of Navigation Locks; US Army Engineer Waterways Experiment Station: Vicksburg, MI, USA, 1989. [Google Scholar]
  31. Ma, X.; Hu, Y.; Li, Z. Impact of an energy dissipator on mooring conditions in ultra-high head locks with a comprehensive uniform-strength evaluation. Ocean Eng. 2022, 266, 112806. [Google Scholar] [CrossRef]
Figure 1. Decentralized cascade locks schematic diagram.
Figure 1. Decentralized cascade locks schematic diagram.
Water 15 03394 g001
Figure 2. Physical test model of the decentralized cascade locks (scale 1:30).
Figure 2. Physical test model of the decentralized cascade locks (scale 1:30).
Water 15 03394 g002
Figure 3. The physical test for process lines of flows.
Figure 3. The physical test for process lines of flows.
Water 15 03394 g003
Figure 4. Four flow states of the intermediate channel at different times.
Figure 4. Four flow states of the intermediate channel at different times.
Water 15 03394 g004
Figure 5. (af) Six flow patterns of the outflow zone. (The arrow represents the flow direction, and the digital number represents the position number).
Figure 5. (af) Six flow patterns of the outflow zone. (The arrow represents the flow direction, and the digital number represents the position number).
Water 15 03394 g005
Figure 6. Schematic diagram of the simulation computational domain and mesh division.
Figure 6. Schematic diagram of the simulation computational domain and mesh division.
Water 15 03394 g006
Figure 7. Effect of different grid densities on culvert velocity.
Figure 7. Effect of different grid densities on culvert velocity.
Water 15 03394 g007
Figure 8. Comparison of numerical simulation and actual physical test in single branch hole resistance coefficient [25].
Figure 8. Comparison of numerical simulation and actual physical test in single branch hole resistance coefficient [25].
Water 15 03394 g008
Figure 9. Computation domain and mesh generation.
Figure 9. Computation domain and mesh generation.
Water 15 03394 g009
Figure 10. Process lines of flow rate over time: (a) dispersed outflow; (b) concentrated outflow.
Figure 10. Process lines of flow rate over time: (a) dispersed outflow; (b) concentrated outflow.
Water 15 03394 g010
Figure 11. Comparison of hybrid numerical simulation and the physical test: (a) Process lines of water level over time; (b) Process lines of flow rate over time.
Figure 11. Comparison of hybrid numerical simulation and the physical test: (a) Process lines of water level over time; (b) Process lines of flow rate over time.
Water 15 03394 g011
Figure 12. Comparison of 2D numerical simulation and actual physical testing of water level in intermediate channel: (a) 1st monitoring point; (b) 13th monitoring point; (c) 26th monitoring point.
Figure 12. Comparison of 2D numerical simulation and actual physical testing of water level in intermediate channel: (a) 1st monitoring point; (b) 13th monitoring point; (c) 26th monitoring point.
Water 15 03394 g012
Figure 13. The maximum water surface slope over time.
Figure 13. The maximum water surface slope over time.
Water 15 03394 g013
Figure 14. Comparison of water surface slope at different regions of the intermediate channel: (a) the upstream region; (b) the middle region; (c) the downstream region.
Figure 14. Comparison of water surface slope at different regions of the intermediate channel: (a) the upstream region; (b) the middle region; (c) the downstream region.
Water 15 03394 g014
Figure 15. The maximum flow velocity over time.
Figure 15. The maximum flow velocity over time.
Water 15 03394 g015
Figure 16. The Velocity field cloud image at 6 min, (a) dispersed outflow; (b) concentrated outflow.
Figure 16. The Velocity field cloud image at 6 min, (a) dispersed outflow; (b) concentrated outflow.
Water 15 03394 g016
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mei, L.; An, J.; Li, Z.; Zhang, S.; Ding, L. Study of Navigable Flow Conditions in the Intermediate Channel of Decentralized Cascade Locks. Water 2023, 15, 3394. https://doi.org/10.3390/w15193394

AMA Style

Mei L, An J, Li Z, Zhang S, Ding L. Study of Navigable Flow Conditions in the Intermediate Channel of Decentralized Cascade Locks. Water. 2023; 15(19):3394. https://doi.org/10.3390/w15193394

Chicago/Turabian Style

Mei, Lingqin, Jianfeng An, Zhonghua Li, Shouyuan Zhang, and Leyu Ding. 2023. "Study of Navigable Flow Conditions in the Intermediate Channel of Decentralized Cascade Locks" Water 15, no. 19: 3394. https://doi.org/10.3390/w15193394

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

Article Metrics

Back to TopTop