Next Article in Journal
Application Assessment of Pumped Storage and Lithium-Ion Batteries on Electricity Supply Grid
Previous Article in Journal
Produce Low Aromatic Contents with Enhanced Cold Properties of Hydrotreated Renewable Diesel Using Pt/Alumina-Beta-Zeolite: Reaction Path Studied via Monoaromatic Model Compound
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence Mechanism of Geometric Characteristics of Water Conveyance System on Extreme Water Hammer during Load Rejection in Pumped Storage Plants

1
College of Water Conservancy and Hydropower Engineering, Hohai University, Nanjing 210098, China
2
Powerchina Huadong Engineering Corporation Limited, Hangzhou 311122, China
*
Authors to whom correspondence should be addressed.
Energies 2019, 12(15), 2854; https://doi.org/10.3390/en12152854
Submission received: 24 June 2019 / Revised: 19 July 2019 / Accepted: 23 July 2019 / Published: 24 July 2019
(This article belongs to the Section D: Energy Storage and Application)

Abstract

:
Pumped storage plants (PSPs) have achieved rapid development and deployment worldwide since the penetration of intermittent renewable energy sources (RES). Hydraulic transient analysis in the PSP, to obtain the control parameters such as extreme water hammer pressure, is vital to the safe design of water conveyance system. Empirically, simultaneous load rejection (SLR) is commonly accepted as the control condition for extreme water hammer, while it is not completely true for the PSP. Employing theoretical analysis and numerical simulation, this study systematically investigates the effects of geometric characteristics on the extreme water hammer, and reveals the mechanism leading to the maximum spiral case pressure (SCP) during a two-stage load rejection (TLR) process. The results indicate that the extreme water hammer pressure is closely related to geometric characteristics of the water conveyance system, performing the allocation of the water inertia time constant of the main and branch pipelines. When the water inertia time constant in the branch pipe is dominant (η1 > 0.24 for example), the maximum SCP will occur in TLR conditions rather than SLR. Moreover, the maximum SCP is almost the same, providing the water inertia time constants of both the main and branch pipelines are kept constant.

Graphical Abstract

1. Introduction

In recent years, the penetration of renewable energy sources (RES) have increased rapidly worldwide due to the ever-increasing environmental concerns and electricity demand [1,2]. Compared with traditional fossil fuels, RES are regarded as sustainable, clean, low-carbon, and economical energy sources [3]. However, the unpredictable and intermittent nature of some RES—such as wind, solar, and wave energy—brings significant challenges to the grid’s security, stability, reliability, and efficiency [4,5,6]. To solve the above problems caused by the intermittency of RES, energy storage technology is an effective way forward. Among the available technologies, the pumped storage plant (PSP) is proven to be the most mature, cost-effective, and large-scale one, considered to be an essential part of a renewable oriented power system [7]. Therefore, PSPs have broad prospects and are developing rapidly in worldwide.
A PSP usually has the functions of peak shaving, valley filling, frequency modulation, phase modulation, and accident standby in the electric grid system [8,9]. To fulfill these regulatory functions, the PSP working conditions change frequently and the reversible pumped turbines switch rapidly between the pump mode and turbine mode, substantially increasing the probability of load rejection. The following severe water hammer caused by emergency closure of guide vanes may result in serious threat to the safe operation of the water conveyance system. Therefore, hydraulic transient analysis on the water hammer phenomenon of PSPs is an important and essential work, to provide safe and optimal operation, during the design or service stage. This has been extensively studied, both numerically and experimentally. The general theory of water hammer was first developed by Allievi [10]. Further efforts by Wood [11], Rich [12], Parmakian [13], Streeter and Lai [14], and Streeter and Wylie [15] formed the classical mass and momentum equations for one-dimensional water hammer flows, which were then written in numerous classical textbooks for transient flow analysis [16,17]. The above-mentioned governing equations formed a pair of quasi-linear hyperbolic partial differential equations, which could seldom be solved analytically. Therefore, various numerical approaches have been introduced for calculation of the pipeline transients, including the method of characteristics (MOC) [16,17], wave plan method [18], finite volume method [19], finite element method, and finite difference method [20]. Later, these theoretical and numerical approaches were widely applied in hydroelectrical power systems. Yang et al. [21] presented a mathematical model of the hydroelectrical power system based on TOPSYS, as well as its application. Employing MOC, sensitivity analysis of hydraulic transient at runaway and the interaction effect of hydraulic transient flow for two parallel reversible pump turbines (RPTs) have been numerically investigated by Rezghi and Riasi [22,23]. Hu et al. [24] conducted transient simulation and analyzed the transient pressure pulsations characteristics of the RPT under load rejection. Zhou et al. [25] proposed a 3D numerical method based on single-phase and volume of fluid (SP-VOF) to investigate the transient characteristics of a PSP. Fu et al. [26,27] employed a dynamic mesh technology to simulate the flow behavior in an RPT under load rejection, and the energy transformation and flow mechanism were analyzed as well. Different innovative closure schemes are proposed by Kuwabara et al. [28] and Yu et al. [29], respectively, which can effectively mitigate the water hammer due to S-shaped characteristics. Applying the method of internal characteristics, the combined closing scheme of ball valve and wicket gate during load rejection was simulated by Li et al. [30]. Zhang et al. [31] simulated the load rejections of Tongbai PSP in China, and the results agreed well with the field test data. Chen et al. [32] calibrated the boundary conditions based on field tests and gave the numerical prediction of the extreme water hammer pressure under critical load case scenarios.
Reliable prediction of the hydraulic parameters, including maximum spiral case pressure (SCP) and minimum draft tube pressure (DTP), is of great importance for the safe design of the water conservancy. Empirically, simultaneous load rejection (SLR) is usually considered to be the critical load case scenario, in which extreme values of SCP or DTP occur. This is true for conventional hydropower plant with Francis turbines in most cases. For the PSP, however, it is not the case due to the well-known S-shaped characteristics of the RPTs, which is contributed to by the runner design of focusing on pump mode [33]. The hydraulic behavior during load rejection of the PSP is more complicated and distinct from the conventional hydroelectric power plant. Recent studies indicated that extreme low DTP may occur in a specific load case scenario, namely TLR, which was firstly proposed by Yokoyama and Shimmei [34] to study the dynamic characteristics of RPTs in the PSP. TLR usually happens in the system layout of multiple RPTs sharing the same main pipe, resulting from the hydraulic disturbance. The RPT rejecting the load firstly can increase the working head of the other RPTs through the hydraulic connection provided by the bifurcations or trifurcations. Then, the overloaded units may reject the load in turn, defined as TLR (Figure 1) [35]. Currently, TLR is extensively investigated in PSPs with concerns for the minimum DTP. Zhang and Suo [36] first found the minimum pressure of the draft tube may occur during TLR when discussing the setting criterion for tailrace surge tank. Subsequently, Zhang et al. [37,38] conducted a series of studies on TLR, and concluded that the unique S-shaped characteristics of RPT were responsible for the abnormal low DTP, and the downstream pipeline layout had great influence on the magnitude of DTP. By theoretical analysis and numerical simulation, Zeng et al. studied the impact of the pipe diameters on DTP [39], proposed several multiphase closing schemes to mitigate the severe water hammer and rotational speed rise under TLR [40], and explained the physical mechanism of the phenomenon leading to extreme low DTP [41]. In order to control the great pressure drop in the draft tube, several possible countermeasures were proposed by Fang and Koutnik [42].
The above achievements confirm that TLR is the exact critical load case scenario for minimum DTP, which was ascribed to the S-shaped characteristics and the layout of water conveyance system. For maximum SCP, however, TLR can also sometimes be the critical load case rather than SLR [43]. If the extreme water hammer appears in TLR, but is not considered in advance, the unavoidable extreme pressure may result in serious damage to the system. In fact, the extreme water hammer pressure, no matter SCP or DTP, also greatly depends on the geometric characteristic of the water conveyance system, including the diameter and length of the pipelines. Different geometric dimensions of the main and branch pipes may result in the extreme water hammer pressure occurring in different load cases, TLR or SLR. However, the comprehensive influence analysis of geometric characteristics of the water conveyance system on extreme water hammer are obviously limited. How the geometric characteristics affect the extreme water hammer pressure is still unclear. Therefore, an in-depth and systematic study on extreme water hammer during load rejection of different geometric characteristics of water conveyance system is required, especially for the maximum SCP.
The present study investigates the influence mechanism of geometric characteristics of water conveyance system on extreme water hammer during load rejection in the PSP. Rigid water hammer theory and the numerical model for hydraulic transients based on MOC are employed to illustrate how the geometric characteristics affect the extreme water hammer pressure, such as the maximum SCP.

2. Theoretical Analysis and Mathematical Model

2.1. Theoretical Analysis

The schematic diagram for a typical PSP with two RPT units is shown in Figure 2. RPT 1 and 2 are symmetrically arranged to obtain the best flow conditions, and the initial outputs are the same. As mentioned previously, RPT 1 and RPT 2 share the same headrace tunnel (number 1), thereby RPT 1 rejecting the full load may result in the subsequent load rejection of RPT 2. Rigid water column theory [16,17] is employed for the following quantitative analysis.
With reference to Figure 2, the momentum equations [16,17] of this hydraulic system are given by
L 1 g f 1 d Q T d t = Z u α 1 Q T 2 H F u
L 2 g f 2 d Q 1 d t = H F u α 2 Q 1 2 H t u 1
L 3 g f 3 d Q 2 d t = H F u α 3 Q 2 2 H t u 2
where Li and fi (i = 1–3) are length and cross-sectional area of the i-th pipeline reach, respectively; αi (i = 1–3) is head loss coefficient of the i-th reach; Zu is the upstream reservoir water level; Htuj (j = 1–2) is the head at the spiral case of the j-th RPT, respectively; HFu is the head at upstream bifurcation node; QT is headrace tunnel discharge; Q1 and Q2 are the demand discharges of RPT 1 and RPT 2, respectively; t is time; and g is gravitational acceleration.
The continuity equation [16,17] is
Q T = Q 1 + Q 2
Taking Equations (1) and (2) simultaneously gives
L 1 g f 1 d Q T d t + L 2 g f 2 d Q 1 d t = Z u α 1 Q T 2 α 2 Q 1 2 H t u 1 .
After transformation, Equation (5) takes the form
H t u 1 = Z u α 1 Q T 2 α 2 Q 1 2 ( L 1 g f 1 d Q T d t + L 2 g f 2 d Q 1 d t ) .
Substitution of Equation (4) into Equation (6) yields
H t u 1 = Z u α 1 Q T 2 α 2 Q 1 2 ( L 1 g f 1 + L 2 g f 2 ) d Q 1 d t L 1 g f 1 d Q 2 d t .
In the same manner for RPT 2
H t u 2 = Z u α 1 Q T 2 α 3 Q 2 2 ( L 1 g f 1 d Q T d t + L 3 g f 3 d Q 2 d t ) = Z u α 1 Q T 2 α 3 Q 2 2 ( L 1 g f 1 + L 3 g f 3 ) d Q 2 d t L 1 g f 1 d Q 1 d t
Equations (7) and (8) indicate that the SCP (Htu1 and Htu2) are affected by the following parameters: The upstream reservoir water level (Zu), discharge of headrace tunnel (QT), discharge of corresponding branch pipelines (Q1 or Q2), geometric characteristics of upstream water conveyance system (Li and fi, i = 1–3), head loss of each reaches (α1QT2, α2Q12, and α3Q22), and the rate of discharge change ( d Q 1 d t and d Q 2 d t ). The SCP of RPT 1 (or RPT 2) is related not only to the characteristics of the branch pipe RPT 1 (or RPT 2) location, but is also related to that of the common shared main pipe. Obviously, the critical terms are L 1 g f 1 d Q T d t and L 3 g f 3 d Q 2 d t .
During SLR, Equation (8) can be written as
H t u 2    SLR = Z u | SLR α 1 Q T 2 | SLR α 3 Q 2 2 | SLR L 1 g f 1 d Q T d t | SLR L 3 g f 3 d Q 2 d t | SLR .
As for TSL, assuming RPT 1 firstly rejects full load by emergency closure, then RPT 2 commits full load rejection after a certain interval of time (similarly hereinafter). Equation (8) takes the form
H t u 2    TLR = Z u | TLR α 1 Q T 2 | TLR α 3 Q 2 2 | TLR L 1 g f 1 d Q T d t | TLR L 3 g f 3 d Q 2 d t | TLR .
Due to the same initial condition of a load case,
Z u | TLR = Z u | TLR    ,    α 1 Q T 2 | SLR = α 1 Q T 2 | TLR    ,    α 3 Q 2 2 | SLR = α 3 Q 2 2 | TLR .
By use of Equation (10) minus Equation (9) to eliminate duplicate parameters, it becomes
H t u 2    TLR H t u 2    SLR = L 1 g f 1 ( d Q T d t | TLR + d Q T d t | SLR ) + L 3 g f 3 ( d Q 2 d t | TLR + d Q 2 d t | SLR ) .
Defining the following dimensionless parameters
h = H H r ,    z = Z H r    ,    q = Q Q r
where the subscript r indicates the rated quantities.
Then, the Equations (8) and (11) are transformed into
h t u 2 = z u α 1 Q r 2 H r q T 2 α 3 Q r 2 H r q 2 2 ( T w m , u d q T d t + T w b , u d q 2 d t ) = z u α 1 Q r 2 H r q T 2 α 3 Q r 2 H r q 2 2 T w , u d q 2 d t T w m , u d q 1 d t
h t u 2    TLR h t u 2    SLR = T w m , u ( d q T d t | TLR + d q T d t | SLR ) + T w b , u ( d q 2 d t | TLR + d q 2 d t | SLR )
where T w m , u = L 1 Q r g f 1 H r and T w b , u = L 2 Q r g f 2 H r are water inertia time constant of upstream main pipe and upstream branch pipes, respectively; T w , u = T w m , u + T w b , u is water inertia time constant of upstream pipelines.
For Equation (14), if h t u 2    TLR h t u 2    SLR < 0 , the extreme SCP will occur in SLR. Otherwise, it will occur during TLR. Figure 3 gives the general law of the wicket gates opening, and discharge process of main and branch pipes during SLR and TLR, respectively. As shown in Figure 3, the total closure time of wicket gates for SLR is t4, while that for TLR is t5 (= t4 + tc). That is to say, TLR is equivalent to extending the total closure time of the system compared with SLR. Finally, the times for the main pipe discharge dropping to the first valley under SLR and TLR are t1 and t2, respectively, which results in the decrease of the flow gradient in main pipeline. It is obvious that d Q T d t | SLR < 0 when t < t1, and d Q T d t | TLR < 0 when t < t2. Additionally, the instant for maximum SCP during SLR is less than t1, and that during TLR is less than t2. Thus, there exists ( d Q T d t | TLR + d Q T d t | SLR ) < 0 or ( d q T d t | TLR + d q T d t | SLR ) < 0 for main pipe. This factor makes the maximum SCP under TLR less severe than SLR, and also explains why TLR has been paid little attention during transient calculation. While for the branch pipe numbered 3, it also takes t1 for the discharge to drop to the first valley during SLR. However, the load rejection of RPT 1 inevitably increases the demand discharge of RPT 2, which will be dropped to the first valley at a greater rate of discharge change in a shorter time (t3tc). As a result, the flow gradient of the branch pipe of RPT 2 is much greater than SLR, namely ( d Q 2 d t | TLR + d Q 2 d t | SLR ) > 0 or ( d q 2 d t | TLR + d q 2 d t | SLR ) > 0 . Therefore, Equation (14) is not necessarily negative as it is on the surface. That is, it is not always the case that the maximum SCP during TLR is less than that during SLR.
In fact, the sign of Equation (14) also greatly depends on the geometric characteristics of upstream water conveyance system, such as Li and fi (i = 1–3). If the headrace tunnel is long enough and the area is appropriate, the term of L1/f1 (or Twm,u) will be large enough. The water inertia in the main pipeline becomes the dominant factor, thereby maximum SCP will occur during the load case of SLR. Conversely, if the proportion of the branch pipe length is larger and the corresponding area is relatively small to make L3/f3 large enough, namely Twb,u large enough, the water inertia of the branch pipe will be the dominant factor. At this point, the maximum SCP following TLR will be greater than that of SLR.
In summary, the maximum SCP during load rejection is greatly affected by the geometric characteristics of the water conveyance system. When the water inertia time constant in the branch pipeline is dominant, the extreme water hammer will occur under TLR condition. In addition, ignoring the influence of head loss, if the water inertia time constants of both main and branch pipelines are kept constant by changing the lengths and diameters simultaneously, the discharge change rate ( d q T d t , d q 1 d t and d q 2 d t ) will remain unchanged, and the extreme water hammer will as well.

2.2. Mathematical Model

Since MOC is a reliable and effective method for hydraulic transient analysis [35,44], it is employed to simulate hydraulic transients during load rejection. The basic equations of motion and continuity for the pressurized conduits could be referred to in the literature [16,17], as well as the boundary conditions, such as the reservoirs, series connections, junctions, and surge tanks. The key boundary condition of the RPT and the transformation of the RPT characteristics are given herein.

2.2.1. Basic Equations of RPT

An RPT is the vital component for the PSP system. The basic equations include the head balance equation and speed change equation. Figure 4 gives the sketch of the boundary condition of the RPT, and the head balance equation can be expressed as
H = ( H t u + | Q | Q 2 g f s c 2 ) ( H t d + | Q | Q 2 g f d t 2 )
in which fsc is the area at spiral case end; fdt is the area at draft tube inlet; and Q is demand discharge of the RPT.
The speed change in terms of the power produced by the RPT is
P P G = I ω d ω d t
I = W R g 2 / g
where P is the power generated by the RPT; PG is the power absorbed by the generator; ω is the angular velocity; dω/dt is the angular acceleration; I is the polar moment of inertia of rotating fluid and mechanical parts; Rg is gyration radius; W is the weight.
The RPT unit will be disconnected from the grid as soon as load rejection occurs, thereby Pg = 0. The guide vanes are subsequently closed emergently by the given closure law to prevent overspeed of the rotational speed. By integration and Taylor expansion, Equation (16) is transferred to algebraic equations, given by
n = n 0 + Δ t 2 T a ( β + β 0 )
in which n (= N/Nr) and β (= M/Mr) are dimensionless rotational speed and dimensionless torque, respectively; n0 and β0 are the calculation results of a previous time step of n and β, respectively; Ta is inertia time constant of the RPT units, defined as T a = G D 2 | N r | 374.7 M r , with Nr rated rotational speed and Mr rated torque; and Δt is time step.

2.2.2. Modified Suter Transformation for RPT Characteristics

The hydraulic behavior of an RPT is usually described with two four-quadrant characteristics curves, including a series of guide vane opening curves (as shown in Figure 5). The characteristics curves are depicted by the unit rotational speed n11, unit discharge Q11, unit torque M11, and guide vane opening α0, which are defined as follows
n 11 = n D H Q 11 = Q D 2 H M 11 = M D 3 H
where D is runner diameter of the RPT.
On the characteristics curves of RPTs, there is a zone known as S-shape region, where three unit flows or torques exist at a given unit speed. Such a multi-valued nature will pose great difficulty in the interpolation of these curves during the transient simulation. Herein, a modified Suter transform [45] is applied for the mathematic model to overcome this challenge.
W H ( x , y ) = 1 ( Q 11 / Q 11 r + c ) 2 + ( n 11 / n 11 r ) 2
W M ( x , y ) = M 11 / M 11 r
{ x = tan 1 [ ( Q 11 / Q 11 r + c ) / ( n 11 / n 11 r ) ] , n 11 / n 11 r 0 x = π + tan 1 [ ( Q 11 / Q 11 r + c ) / ( n 11 / n 11 r ) ] , n 11 / n 11 r < 0
in which WH is the discharge function; WB is the torque function; x is the polar angle, determined by Equation (21); y is the dimensionless guide vane opening; and c is a constant, usually selected between 1.0 to 1.5. The subscript r denotes the rated values. Through the use of the modified Suter transformation, the characteristics of RPT can be turned as the WH and WB curves, which are single-valued with the polar angle x.

2.2.3. Technical Parameters

One PSP located in China is used for numerical simulation. Figure 2 depicts its simplified layout. The installed capacity of the PSP is 1200 MW, with four identical RPTs. Each RPT can develop 306.1 MW in rated turbine operation. There are two hydraulic units, in which two RPTs share a common headrace tunnel. The normal water levels of upstream and downstream reservoirs are 815.5 m and 413.5 m, respectively. The dead water levels of upstream and downstream reservoirs are 782.0 m and 383.0 m, respectively. The rated discharge 2 × 86.68 m3/s, rated speed is 375.0 r/m, rated head is 400.0 m, maximum head is 432.47 m, and minimum head is 361.08 m. As shown in Figure 5, the pump-turbines have pronounced S-shaped region. The parameters of the conduits are listed in Table 1. The guide vanes are closed down linearly in 26 s.
The technical route of this study can be referred to Figure 6.

3. Numerical Simulation and Results

To further validate the above conclusions, a numerical simulation was conducted herein. The numerical simulation for the influence of the geometric characteristics of the upstream pipelines on the maximum SCP can be carried out through following three scenarios. Scenario 1, keeping the total length of the upstream pipelines to be constant, and allocating the main pipe length and branch pipe length. Scenario 2, only altering the branch pipe area, and the rest of the parameters are kept unchanged. Scenario 3, changing the length and area of both main and branch pipes simultaneously to keep Tw,u of each scheme the same as the original one. For convenient analysis, define η1 = Twb,u/Tw,u, characterizing the proportion of the water inertia of the branch pipe to that of the upstream pipelines.
The procedure of the TLR is described as follows. RPT 1 rejects full load at t = 0 as the first load rejection turbine, and after some interval of time, RPT 2 rejects full load caused by violent hydraulic disturbance as the subsequent load rejection turbine. The worst interval time for extreme high SCP can be obtained by repeated trial calculation.

3.1. Scenario 1: Changing Length of Both Main and Branch Pipes

For Scenario 1, the total length of the diversion pipelines was kept at a constant of 1320 m, and diameters of the main and branch pipes remained at 6.8 m and 2.8 m, respectively. The downstream water conveyance system was consistent with Table 1. The other parameters of different schemes are shown in Table 2, as well as the results of the maximum SCP.
As shown in Table 2, on the premise of constant total length, main pipe areas, and branch pipe areas, the water inertia time constant of the main pipe (Twm,u) increased with the increasing main pipe length. While the decreasing branch pipe length led to the decrease in Twb,u. Eventually, the ratio of the water inertia of the branch pipe to that of the upstream pipelines (η1) decreased. The results illustrate that the maximum SCP decreased with decreasing η1, which is consistent with conventional theory. The difference is that at the larger ratio of η1 (η1 ≥ 0.37), the maximum SCP occured in TLR rather than SLR. Figure 7 gives the maximum SCP with different time intervals for each η1 in Scenario 1. For the four schemes of η1 = 0.71, η1 = 0.66, η1 = 0.55, and η1 = 0.37, the worst interval time was 4.8 s, 4.8 s, 4.8 s, and 1.8 s, respectively. The maximum SCPs increased with increasing time intervals and attained peak value of 673.10 m, 666.61 m, 652.35 m, and 634.03 m at corresponding worst interval times. The variation of the SCP at worst interval time for different η1 can be referred to Figure 8. IN particular, the difference between maximum SCP during SLR and TLR at 4.8 s reached 21.9 m for η1 = 0.71. If the safety margin is not reserved enough, the unavoidable extreme pressure increase during TLR may cause the rupture of spiral case, which will lead to serious damage to the system. As for η1 = 0.18, with the increase of time intervals, the maximum SCP presented a downward trend. That is, the maximum SCP appeared during SLR. The unexpected phenomenon herein is different from the conventional theory that maximum SCP occurs in SLR, but greatly related to the water inertia of both main and branch pipes.
Taking the typical schemes of η1 = 0.71 and η1 = 0.18 for further investigation as they presented the different results. Figure 8 and Figure 9 give the time history of discharge of the main and branch pipes during SLR and TLR for η1 = 0.71 and η1 = 0.18, respectively. The total closure time for SLR was 26 s, while it was (26 + tc) s for TSL. That is, TLR actually extended the total closure time of the system compared with SLR, which was consistent with analysis in Section 2. The flow gradient in the main pipe was definitely decreased under TLR compared with SLR. Besides, d Q T d t was negative before the discharge first drops to the valley. Thus d Q T d t | TLR + d Q T d t | SLR < 0 is always valid. The change rate of discharge of main and branch pipes at the instant for maximum SCP are listed in Table 3, and can be seen in Figure 9 and Figure 10 as well. For η1 = 0.71, ( d Q T d t | TLR + d Q T d t | SLR ) = 16.16 − 27.75 = −11.59 m2/s2 < 0, while for η1 = 0.18, ( d Q T d t | TLR + d Q T d t | SLR ) = 22.4 − 32.27 = −9.87 m2/s2 < 0.
The smaller the water inertia in main pipe, the severer change in flow gradient between TLR and SLR. When SLR occurred, the discharge of RPT 2 dropped to the first valley at a greater rate of discharge change in a shorter time for the worst time interval. The flow gradient of the branch pipe of RPT 2 was much greater than SLR. Thus d Q 2 d t | TLR + d Q 2 d t | SLR > 0 is always valid. Similarly, for η1 = 0.71, there exists ( d Q 2 d t | TLR + d Q 2 d t | SLR ) = 20.08−13.66 = 6.42 m2/s2 > 0, while for η1 = 0.18, ( d Q 2 d t | TLR + d Q 2 d t | SLR ) = 45.69 − 20.71 = 24.98 m2/s2 > 0. The smaller the water inertia in branch pipe, the severer change in flow gradient between TLR and SLR. The results are consistent with the theoretical analysis.
According to Equation (11) or Equation (14),
( h t u 2    TLR h t u 2    SLR ) η 1 = 0.71 = T w m , u ( d q T d t | TLR + d q T d t | SLR ) + T w b , u ( d q 2 d t | TLR + d q 2 d t | SLR ) = 0.57 × ( 0.134 ) + 1.38 × 0.074 = 0.026 > 0
Equations (22) and (23) and Table 3 indicate the critical load case for maximum SCP was closed related to the water inertia of the branch and main pipes, which was determined by the geometric dimension of the upstream water conveyance system. The maximum SCP occurred in TLR for larger η1 (≥ 0.37), while it came to SLR for smaller η1 (≤ 0.18). Additionally, the worst interval times, changing from 4.8 s to 1.8 s and finally 0 s, also explain the effect of the different η1 on the SCP. In conclusion, the allocation of the water inertia in main and branch pipes was the critical factor determining in which load cases the maximum SCP will occur. When η1 is at a large value, the water inertia in branch pipe is the dominant effect for inducing maximum SCP during TLR. If the water inertia in the main pipe becomes the dominant factor (η1 ≤ 0.18), maximum SCP will occur in SLR condition.
Additionally, Figure 11 and Figure 12 present the time histories of the discharge of the main pipe and branch pipe for different η1, respectively. For the schemes of η1 = 0.71, η1 = 0.66 and η1 = 0.55, the worst interval times were all 4.8 s. The discharge variation curve clusters of the main pipe for these three schemes had almost the same variation trend before the instant for maximum SCP. However, obvious differences appeared for the branch pipe discharge curve clusters after the load rejection of RPT 2, which indicates that the geometric dimension change of the branch pipe has greater influence on the maximum SCP.

3.2. Scenario 2: Changing Diameter of Branch Pipe

Since the change of the branch pipe diameter can greatly influence proportion of the water inertia of the branch pipe to that of the upstream pipelines, herein only the diameter of the branch pipe is changed for Scenario 2. The rest parameters are kept constant, as well as the downstream water conveyance system. The other parameters of different schemes and results are listed in Table 4.
As shown in Table 4 and Figure 13 and Figure 14, the results present the same law to that of Scenario 1. With the decrease of the branch pipe diameter, the water inertia of the upstream pipelines increased, thereby increasing the maximum SCP. For the schemes of η1 = 0.16 and η1 = 0.18, the maximum SCP both appeared in LSR because of the relative low proportion of water inertia in the branch pipe. When η1 rose to 0.24, the maximum SCP began to appear in TLR. At this point, the worst interval time was 0.6 s, and the difference in the maximum SCP between TLR and SLR was merely 0.66 m. After that, the maximum SCP still occured in TLR and the difference in the maximum SCP between TLR and SLR was increasing with the increasing η1, as well as the worst interval time. The results of Scenario 2 also indicate that the allocation change of the water inertia between the main and branch pipes may result in the critical load case transfer for maximum SCP. Figure 15 and Figure 16 give the discharge variations of main and branch pipes of different η1 for Scenario 2. The diameter changes in the branch pipe also influence the discharge variation of both main and branch pipes.

3.3. Scenario 3: Changing Length and Area of Both Main and Branch Pipes Simultaneously

To further reveal the effect of water inertia on extreme water hammer, the water inertia time constant of both branch and main pipe were kept constant in Scenario 3. The length and area of both main and branch pipes were simultaneously changed to keep Twm,u and Twb,u of each scheme the same as the original one. The rest of the parameters were kept constant, as well as the downstream water conveyance system. Table 5 lists the other parameters of different schemes and corresponding results.
For the five schemes in Scenario 3, the values of Twm,u and Twb,u were kept constant as 0.73 and 0.43, respectively, by changing the lengths and areas of the main and branch pipes simultaneously. Thereby η1 = 0.37 remained unchanged. As shown in Table 5, the maximum SCP of each scheme was almost the same no matter if it was the SLR or TLR. The maximum difference between the five schemes was within 0.6 m. Figure 17 presents the maximum SCP with different time intervals for each scheme of Scenario 3, which had almost the same variation trend. Additionally, the worst interval time for each scheme under TLR was about 1.9 s, and also can be regarded to be the same. As for the SCP for each scheme of Scenario 3, although the initial pressure was not the same due to the different head loss, nearly the same extreme SCP was reached, meaning the curves varied in the same way (seen in Figure 18). Moreover, the discharge variation curve clusters of the main and branch pipes are almost overlapped, which is shown in Figure 19. The above results verify the conclusion that the discharge change rate remained little changed, thereby the extreme SCPs were almost the same providing the water inertia time constants of both main and branch pipelines were kept constant. It should be noted that the slight difference in discharge was due to the influence of head loss. Although η1 is kept constant for each scheme, there is no guarantee that head loss is the same. Except from this factor, the difference of pressure change curve clusters was also related to the influence of the velocity head.

4. Discussion

For conventional hydroelectrical power plants, SLT is generally accepted as the critical load case for extreme water hammer. However, it is not completely applied for the PSPs because of the unique S-shaped characteristics. The above theoretical analysis and numerical simulation show that the maximum SCP during load rejection is closely related to the geometric dimensions of the upstream water diversion system, including the area and length of both branch and main pipes. The changes of the geometric characteristics of the upstream pipelines actually cause the changes in the water inertia. When the water inertia time constant in the branch pipe is dominant, the extreme water hammer will occur under TLR condition. For the numerical simulation herein, if η1 > 0.24 (η1 denotes the ratio of the water inertia of the branch pipe to that of the upstream pipelines), then the maximum SCP would appear in TLR rather than SLR (as shown in Figure 20), which is not consistent with the conventional knowledge. Moreover, with the continuous increase of η1, the differences of the maximum SCP between SLR and TLR increased greatly. Especially, for η1 = 0.71, the maximum SCP during TLR was nearly 22 m greater than that during SLR, which is inconceivable and catastrophic for the PSP system. Therefore, in the design stage of a PSP, the importance of this phenomenon should be fully recognized, and the geometric characteristics of branch pipes should be carefully examined to obtain a reasonable and relatively small η1, so as to ensure the maximum SCP occurring at the load case of SLR.
In addition, if the pipeline area and length are changed at the same time to keep constant of η1, the maximum SCP is almost the same, verifying the inference from theoretical analysis. The achievements also provide another idea on the optimal design of the water conveyance system of the PSP.

5. Conclusions

The maximum SCP is one of the control parameters for the water conveyance system design of a PSP. As a result, the determination of the critical load case becomes the primary task. In this study, the effect of geometric characteristics of upstream pipelines on extreme water hammer during load rejection was examined. Rigid water hammer theory was introduced for quantitative analysis. Based on MOC and modified Suter transformations, the dynamic transient numerical model was developed and applied to a practical engineering of a numerical simulation. The fact that TLR can be the critical load case for maximum SCP is presented, and the influence mechanism of the geometric dimension on the maximum SCP was revealed.
In conclusion, the extreme water hammer pressure is closely related to geometric characteristics of the water conveyance system, namely the allocation of the water inertia time constant of the main and branch pipelines. If the water inertia in the branch pipeline is dominant, the maximum SCP will occur under TLR condition, and vice versa. The ratio of the water inertia of the branch pipe to that of the upstream pipelines η1 can serve as the index to measure whether the maximum SCP is occurring in SLR or TLR. With the increase of η1, the differences of the maximum SCP between SLR and TLR increase greatly. Moreover, the discharge change rate remains little changed, providing the water inertia time constants of both main and branch pipelines are kept constant (a constant of η1), eventually resulting in the unchanged maximum SCP.

Author Contributions

Conceptualization, S.C., J.Z., and X.Y.; methodology, S.C. and G.L.; validation, J.Z.; writing—original draft, S.C.; writing—review and editing, G.L. and X.Y.

Funding

This research was supported by the National Natural Science Foundation of China (Grant No. 51709087 and No. 51839008), and the Fundamental Research Funds for the Central Universities (Grant No. 2018B55514).

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

cConstant selected between 1 and 1.5-
DDiameter of runnerm
fiCross-sectional area of i-th pipelinem2
fscArea at spiral case endm2
fdtArea at draft tube inletm2
gGravitational accelerationm/s2
HWorking head of RPTm
HFuPiezometric head at upstream bifurcationm
HrRated headm
hDimensionless head-
htuDimensionless head at spiral case end-
Htu1Piezometric head at the spiral case of RPT 1m
Htu2Piezometric head at the spiral case of RPT 2m
iNumber of pipeline reaches, =1–3-
IInertia polar momentkg·m2
LiLength of i-th pipeline reachm
MShaft torqueN·m
MrRated shaft torqueN·m
M11Unit torqueN·m
NRotational speedr/min
NrRated rotational speedr/min
nDimensionless rotational speedr/min
n0Result of a previous time step of nr/min
n11Unit speedr/min
n11rRated unit speedr/min
PPower generated by RPTkW
PgPower absorbed by generatorkW
QDemand discharge of RPTm3/s
qDimensionless discharge-
q1Dimensionless demand discharge of RPT 1-
Q1Demand discharge of RPT 1m3/s
Q11Unit dischargem3/s
Q11rRated unit dischargem3/s
q2Dimensionless demand discharge of RPT 2-
Q2Demand discharge of RPT 2m3/s
QrRated dischargem3/s
qTDimensionless rated discharge-
QTDischarge of headrace tunnelm3/s
RgGyration radiusm
tTimes
tcWorst interval times
TaMechanical starting times
Tw,uWater inertia time constant of upstream pipeliness
Twm,uWater inertia time constant of upstream main pipes
Twb,uWater inertia time constant of upstream branch pipes
WWeightkg
WBTurbine torque characteristics-
WHTurbine head characteristics-
xPolar anglerad
yDimensionless guide vane opening-
zDimensionless water level-
zuDimensionless upstream reservoir water level-
ZuUpstream reservoir water levelM
αiHead loss coefficient of i-th pipeline reach-
α0Wicket gate opening°
βDimensionless torque-
β0Result of a previous time step of β-
ωAngular velocityrad/s
η1Proportion of water inertia of branch pipe to that of upstream pipelines-
ΔtTime steps
d q 1 d t Dimensionless rate of discharge change of RPT 1S−1
d q 2 d t Dimensionless rate of discharge change of RPT 2s−1
d q T d t Dimensionless rate of discharge change of headrace tunnels−1
d Q 1 d t Rate of discharge change of RPT 1s−1
d Q 2 d t Rate of discharge change of RPT 2s−1
d Q T d t Rate of discharge change of headrace tunnels−1

Abbreviations

DTPDraft tube pressure
MOCMethod of characteristic
PSPPumped storage plant
RESRenewable energy sources
RPTReversible pump turbine
SCPSpiral case pressure
SLRSimultaneous load rejection
TLRTwo-stage load rejection

References

  1. Dai, H.C.; Xie, X.X.; Xie, Y.; Liu, J.; Masui, T. Green growth: The economic impacts of large-scale renewable energy development in China. Appl. Energy 2016, 162, 435–449. [Google Scholar] [CrossRef]
  2. Ellabban, O.; Abu-Rub, H.; Blaabjerg, F. Renewable energy resources: Current status, future prospects and their enabling technology. Renew. Sustain. Energy Rev. 2014, 39, 748–764. [Google Scholar] [CrossRef]
  3. Bhattacharya, M.; Paramati, S.R.; Ozturk, I.; Bhattacharya, S. The effect of renewable energy consumption on economic growth: Evidence from top 38 countries. Appl. Energy 2016, 162, 733–741. [Google Scholar] [CrossRef]
  4. Rehman, S.; Al-Hadhrami, L.M.; Alam, M.M. Pumped hydro energy storage system: A technological review. Renew. Sustain. Energy Rev. 2015, 44, 586–598. [Google Scholar] [CrossRef]
  5. Zeng, W.; Yang, J.D.; Yang, W.J. Instability analysis of pumped-storage stations under no-load conditions using a parameter-varying model. Renew. Energy 2016, 90, 420–429. [Google Scholar] [CrossRef]
  6. Chaudhary, P.; Rizwan, M. Energy management supporting high penetration of solar photovoltaic generation for smart grid using solar forecasts and pumped hydro storage system. Renew. Energy 2018, 118, 928–946. [Google Scholar] [CrossRef]
  7. Ghasemi, A.; Enayatzare, M. Optimal energy management of a renewable-based isolated microgrid with pumped-storage unit and demand response. Renew. Energy 2018, 123, 460–474. [Google Scholar] [CrossRef]
  8. Zeng, M.; Feng, J.J.; Xue, S.; Wang, Z.J.; Zhu, X.L.; Wang, Y.J. Development of China’s pumped storage plant and related policy analysis. Energy Policy 2013, 61, 104–113. [Google Scholar] [CrossRef]
  9. Hino, T.; Lejeune, A. Pumped storage hydropower developments. Compre. Renew. Energy 2012, 6, 405–434. [Google Scholar] [CrossRef]
  10. Allievi, L. Teoria del Colpo D’ariete; Stabilimenti grafici Stucchi: Milan, Italy, 1913. [Google Scholar]
  11. Wood, F.M. The Application of heaviside’s Operational Calculus to the Solution of Problems in Waterhammer. Trans. ASME 1937, 59, 707–713. [Google Scholar]
  12. Rich, G. Waterhammer Analysis by the Laplace-Mellin Transformations. Trans. ASME 1945, 59, 361–376. [Google Scholar]
  13. Parmakian, J. Water-Hammer Analysis; Prentice-Hall, Inc.: Englewood Cliffs, NJ, USA, 1955. [Google Scholar]
  14. Streeter, V.L.; Lai, C. Waterhammer Analysis Including Fluid Friction. Trans. Am. Soc. Civ. Eng. 1963, 128, 1491–1524. [Google Scholar] [CrossRef]
  15. Streeter, V.L.; Wylie, E.B. Hydraulic Transients; McGraw-Hill: New York, NY, USA, 1967. [Google Scholar]
  16. Chaudhry, M.H. Applied Hydraulic Transients; Van Nostrand Reinhold: New York, NY, USA, 1987. [Google Scholar]
  17. Wylie, E.B.; Streeter, V.L.; Suo, L.S. Fluid Transients in Systems; Prentice-Hall, Inc.: Englewood Cliffs, NJ, USA, 1993. [Google Scholar]
  18. Wood, D.J.; Dorsch, R.G.; Lightnor, C. Wave-Plan Analysis of Unsteady Flow in Closed Conduits. J. Hydraul. Div. 1966, 92, 83–110. [Google Scholar]
  19. Guinot, V. Riemann Solvers for Water Hammer Simulations by Godunov Method. Int. J. Numer. Methods Eng. 2002, 49, 851–870. [Google Scholar] [CrossRef]
  20. Izquierdo, J.; Iglesias, P.L. Mathematical modelling of hydraulic transients in simple systems. Math. Comput. Model. 2002, 35, 801–812. [Google Scholar] [CrossRef]
  21. Yang, W.J.; Yang, J.D.; Guo, W.C.; Zeng, W.; Wang, C.; Saarinen, L.; Norrlund, P. A Mathematical Model and Its Application for Hydro Power Units under Different Operating Conditions. Energies 2015, 8, 10260–10275. [Google Scholar] [CrossRef]
  22. Rezghi, A.; Riasi, A. Sensitivity analysis of transient flow of two parallel pump-turbines operating at runaway. Renew. Energy 2016, 86, 611–622. [Google Scholar] [CrossRef]
  23. Rezghi, A.; Riasi, A. The interaction effect of hydraulic transient conditions of two parallel pump-turbine units in a pumped-storage power plant with considering “S-shaped” instability region: Numerical simulation. Renew. Energy 2018, 118, 896–908. [Google Scholar] [CrossRef]
  24. Hu, J.H.; Yang, J.D.; Zeng, W.; Yang, J.B. Transient Pressure Analysis of a Prototype Pump Turbine: Field Tests and Simulation. J. Fluids Eng. 2018, 140, 071102. [Google Scholar] [CrossRef]
  25. Zhou, D.Q.; Chen, H.X.; Zhang, L.G. Investigation of Pumped Storage Hydropower Power-Off Transient Process Using 3D Numerical Simulation Based on SP-VOF Hybrid Model. Energies 2018, 11, 1020. [Google Scholar] [CrossRef]
  26. Fu, X.; Li, D.; Wang, H.; Zhang, G.; Li, Z.; Wei, X. Analysis of transient flow in a pump-turbine during the load rejection process. J. Mech. Sci. Technol. 2018, 32, 2069–2078. [Google Scholar] [CrossRef]
  27. Fu, X.; Li, D.; Wang, H.; Zhang, G.; Li, Z.; Wei, X.; Qin, D. Energy analysis in a pump-turbine during the load rejection process. J. Fluids Eng. 2018, 140, 101107. [Google Scholar] [CrossRef]
  28. Kuwabara, T.; Katayama, K.; Nakagawa, H.; Hagiwara, H. Improvements of transient performance of pump turbine upon load rejection. In Proceedings of the IEEE Power Engineering Society Summer Meeting, Seattle, WA, USA, 16–20 July 2000; pp. 1783–1788. [Google Scholar] [CrossRef]
  29. Yu, X.; Zhang, J.; Miao, D. Innovative closure law for pump-turbines and field test verification. J. Hydraul. Eng. 2015, 141, 05014010. [Google Scholar] [CrossRef]
  30. Li, X.Q.; Tang, X.L.; Shi, X.Y.; Chen, H.H.; Li, C.S. Load rejection transient with joint closing law of ball-valve and guide vane for two units in pumped storage power station. J. Hydroinform. 2018, 20, 301–315. [Google Scholar] [CrossRef]
  31. Zhang, J.; Wang, D.L.; Hu, J.Y.; Zhou, J.; Fang, J. Study on Field Test and Simulating Calculation Following Load Rejections of Tongbai Pumped Storage Power Station. In Proceedings of the 2009 ASME Fluids Engineering Division Summer Conference, Jacksonville, FL, USA, 10–14 August 2008. [Google Scholar]
  32. Chen, S.; Zhang, J.; Li, G.H.; Yu, X.D. Load rejection test and numerical prediction of critical load case scenarios for pumped storage plant. In Proceedings of the ASME 2017 Fluids Engineering Division Summer Meeting, FEDSM 2017, Waikoloa, HI, USA, 30 July–3 August 2017. [Google Scholar]
  33. Yin, J.L.; Wang, D.Z.; Wei, X.Z.; Wang, L.Q. Hydraulic Improvement to Eliminate S-Shaped Curve in Pump Turbine. ASME. J. Fluids Eng. 2013, 135, 071105. [Google Scholar] [CrossRef]
  34. Yokoyama, T.; Shimmei, K. Dynamic Characteristics of Reversible Pump Turbines in Pumped Storage Plants. In Proceedings of the USDE and EPRI Symposium, Boston, MA, USA, 24–26 August 1984. [Google Scholar]
  35. Chen, S.; Zhang, J.; Wang, X.C. Characterization of surge superposition following 2-stage load rejection in hydroelectric power plant. Can. J. Civil Eng. 2016, 43, 844–850. [Google Scholar] [CrossRef]
  36. Zhang, J.; Suo, L.S. Study on tailrace surge chamber installation and hydraulic transients at pumped-storage plant. Water Res. Power 2008, 26, 83–87. [Google Scholar]
  37. Zhang, J.; Lu, W.H.; Hu, J.Y.; Fan, B.Q. Study on the hydraulic transients of pumped-turbines load successive rejection in pumped storage plant. In Proceedings of the 5th Joint ASME/JSME Fluids Engineering Summer Conference, FEDSM 2007, San Diego, CA, USA, 30 July–2 August 2007; pp. 57–61. [Google Scholar] [CrossRef]
  38. Zhang, J.; Lu, W.H.; Fan, B.Q.; Hu, J.Y. The influence of layout of water conveyance system on the hydraulic transients of pump-turbines load successive rejection in pumped storage station. J. Hydroelectric Eng. 2008, 27, 158–162. [Google Scholar]
  39. Zeng, W.; Yang, J.D. Effects of pipe diameters on the pressures during delayed load rejection in high-head pumped storage power stations. In Proceedings of the 27th IAHR Symposium on Hydraulic Machinery and System, Montreal, QC, Canada, 22–26 September 2014. [Google Scholar]
  40. Zeng, W.; Yang, J.D.; Hu, J.H.; Yang, J.B. Guide-vane closing schemes for pump-turbines based on transient characteristics in S-shaped region. ASME. J. Fluids Eng. 2016, 138, 051302. [Google Scholar] [CrossRef]
  41. Zeng, W.; Yang, J.D.; Tang, R.B.; Yang, W.J. Extreme water-hammer pressure during one-after-another load shedding in high-head pumped-storage stations. Renew. Energy 2016, 99, 35–44. [Google Scholar] [CrossRef]
  42. Fang, Y.J.; Koutnik, J. The numerical simulation of the delayed load rejection of a pump-turbine power plant. In Proceedings of the 26th IAHR Symposium on Hydraulic Machinery and System, Beijing, China, 19–23 August 2012. [Google Scholar]
  43. Chen, S. Investigation on 2-Stage Load Rejection and Water Hammer Protection in Hydroelectric Power Plants. Ph.D. Thesis, Hohai University, Nanjing, China, 12 September 2014. [Google Scholar]
  44. Ghidaoui, M.S.; Zhao, M.; McInnis, D.A.; Axworthy, D.H. A Review of Water Hammer Theory and Practice. Appl. Mech. Rev. 2005, 58, 49–76. [Google Scholar] [CrossRef]
  45. Yang, K.L. Hydraulic Transient and Regulation in Hydropower Station and Pump Station, 1st ed.; China Water & Power Press: Beijing, China, 1999. [Google Scholar]
Figure 1. Schematic diagram of two-stage load rejection.
Figure 1. Schematic diagram of two-stage load rejection.
Energies 12 02854 g001
Figure 2. Schematic of a typical pumped storage plant system.
Figure 2. Schematic of a typical pumped storage plant system.
Energies 12 02854 g002
Figure 3. Schematic of wicket gate opening, discharge of the main and branch pipes during simultaneous load rejection (SLR) and two-stage load rejection (TLR).
Figure 3. Schematic of wicket gate opening, discharge of the main and branch pipes during simultaneous load rejection (SLR) and two-stage load rejection (TLR).
Energies 12 02854 g003
Figure 4. Sketch of reversible pump turbines (RPT) in pipeline.
Figure 4. Sketch of reversible pump turbines (RPT) in pipeline.
Energies 12 02854 g004
Figure 5. Four-quadrant characteristics of the RPT: (a) Q11~n11; (b) M11~n11.
Figure 5. Four-quadrant characteristics of the RPT: (a) Q11~n11; (b) M11~n11.
Energies 12 02854 g005aEnergies 12 02854 g005b
Figure 6. Flowchart of the technical route.
Figure 6. Flowchart of the technical route.
Energies 12 02854 g006
Figure 7. Maximum SCP with different time intervals of different η1 for Scenario 1.
Figure 7. Maximum SCP with different time intervals of different η1 for Scenario 1.
Energies 12 02854 g007
Figure 8. Spiral case pressure of different η1 versus time for Scenario 1.
Figure 8. Spiral case pressure of different η1 versus time for Scenario 1.
Energies 12 02854 g008
Figure 9. Discharge of the main and branch pipes versus time during SLR and TLR for η1 = 0.71.
Figure 9. Discharge of the main and branch pipes versus time during SLR and TLR for η1 = 0.71.
Energies 12 02854 g009
Figure 10. Discharge of the main and branch pipes versus time during SLR and TLR for η1 = 0.18.
Figure 10. Discharge of the main and branch pipes versus time during SLR and TLR for η1 = 0.18.
Energies 12 02854 g010
Figure 11. Main pipe discharge of different η1 versus time for Scenario 1.
Figure 11. Main pipe discharge of different η1 versus time for Scenario 1.
Energies 12 02854 g011
Figure 12. Branch pipe discharge of different η1 versus time for Scenario 1.
Figure 12. Branch pipe discharge of different η1 versus time for Scenario 1.
Energies 12 02854 g012
Figure 13. Maximum SCP with different time intervals of different η1 for Scenario 2.
Figure 13. Maximum SCP with different time intervals of different η1 for Scenario 2.
Energies 12 02854 g013
Figure 14. Spiral case pressure of different η1 versus time for Scenario 2.
Figure 14. Spiral case pressure of different η1 versus time for Scenario 2.
Energies 12 02854 g014
Figure 15. Main pipe discharge of different η1 versus time for Scenario 2.
Figure 15. Main pipe discharge of different η1 versus time for Scenario 2.
Energies 12 02854 g015
Figure 16. Branch pipe discharge of different η1 versus time for Scenario 2.
Figure 16. Branch pipe discharge of different η1 versus time for Scenario 2.
Energies 12 02854 g016
Figure 17. Maximum SCP with different time intervals for each scheme of Scenario 3.
Figure 17. Maximum SCP with different time intervals for each scheme of Scenario 3.
Energies 12 02854 g017
Figure 18. Maximum SCP with different time intervals for each scheme of Scenario 3.
Figure 18. Maximum SCP with different time intervals for each scheme of Scenario 3.
Energies 12 02854 g018
Figure 19. Main and branch pipe discharge versus time for each scheme of Scenario 3 during TLR.
Figure 19. Main and branch pipe discharge versus time for each scheme of Scenario 3 during TLR.
Energies 12 02854 g019
Figure 20. Variation of maximum SCP with η1.
Figure 20. Variation of maximum SCP with η1.
Energies 12 02854 g020
Table 1. Parameters of water conveyance system.
Table 1. Parameters of water conveyance system.
Number of ReachesLength L (m)Diameter
D (m)
Head Loss Coefficient
α (×10−6) (/)
Initial Demand Discharge
(m3/s)
112006.8123.24162.42
21202.8781.8081.21
31202.8781.8081.21
42005.0109.3281.21
52005.0109.3281.21
6507.010.02162.42
713007.0104.37162.42
Table 2. Conduits parameters and corresponding results for Scenario 1.
Table 2. Conduits parameters and corresponding results for Scenario 1.
Main PipeL1 (m)934.26997.581096.451200.001271.63
D1 (m)6.8
f1 (m2)36.32
Twm,u (s)0.570.610.670.730.77
Branch PipeL3 (m)385.74322.42223.55120.0048.37
D3 (m)2.8
f3 (m2)6.16
Twb,u (s)1.381.160.800.430.17
Tw,u(s)
η1
1.951.771.471.160.94
0.710.660.550.370.18
Max. SCP during SLR (m)651.17645.25637.63628.02622.28
Max. SCP during TLR (m)673.10666.61652.35634.03622.28
Worst Interval Time (s)4.84.84.81.80 1
1 SLR actually is the special case of TLR. If the maximum spiral case pressure (SCP) of TLR is always equal or less than that of SLR, the worst interval time is 0 s.
Table 3. Change rate of discharge of main and branch pipes at the instant for maximum spiral case pressure (SCP).
Table 3. Change rate of discharge of main and branch pipes at the instant for maximum spiral case pressure (SCP).
η1Load CasesInstant for Maximum SCP
(s)
dQT/dt
(m3/s2)
dQ2/dt
(m3/s2)
h t u 2    TLR h t u 2    SLR Critical Load Case
0.71SLR9.3−27.75−13.66>0TLR
TLR13.2−16.16−20.08
0.18SLR8.3−32.27−20.71<0SLR
TLR10.9−22.4−45.69
Table 4. Conduits parameters and corresponding results for Scenario 2.
Table 4. Conduits parameters and corresponding results for Scenario 2.
Main PipeL1 (m)1200
D1 (m)6.8
f1 (m2)36.32
Twm,u (s)0.73
Branch PipeL3 (m)120.00
D3 (m)5.024.593.822.81.78
f3 (m2)19.7916.5411.476.162.48
Twb,u (s)0.130.160.230.431.07
Tw,u(s)
η1
0.860.890.961.161.80
0.160.180.240.370.59
Max. SCP during SLR (m)616.88618.02621.17628.02646.92
Max. SCP during TLR (m)616.88618.02621.83634.03661.66
Worst Interval Time (s)000.61.84.8
Table 5. Conduits parameters and corresponding results for Scenario 3.
Table 5. Conduits parameters and corresponding results for Scenario 3.
Main Pipe.L1 (m)934.26997.581096.451200.001271.63
D1 (m)6.06.26.56.87.0
f1 (m2)28.2730.1933.1836.3238.48
Twm,u (s)0.73
Branch PipeL3 (m)385.74322.42223.55120.0048.37
D3 (m)5.024.593.822.81.78
f3 (m2)19.7916.5411.476.162.48
Twb,u (s)0.43
Tw,u(s)
η1
1.16
0.37
Max. SCP during SLR (m)627.46627.51627.65628.02628.04
Max. SCP during TLR (m)633.55633.76634.10634.03633.94
Worst Interval Time (s)1.91.91.91.81.9

Share and Cite

MDPI and ACS Style

Chen, S.; Zhang, J.; Li, G.; Yu, X. Influence Mechanism of Geometric Characteristics of Water Conveyance System on Extreme Water Hammer during Load Rejection in Pumped Storage Plants. Energies 2019, 12, 2854. https://doi.org/10.3390/en12152854

AMA Style

Chen S, Zhang J, Li G, Yu X. Influence Mechanism of Geometric Characteristics of Water Conveyance System on Extreme Water Hammer during Load Rejection in Pumped Storage Plants. Energies. 2019; 12(15):2854. https://doi.org/10.3390/en12152854

Chicago/Turabian Style

Chen, Sheng, Jian Zhang, Gaohui Li, and Xiaodong Yu. 2019. "Influence Mechanism of Geometric Characteristics of Water Conveyance System on Extreme Water Hammer during Load Rejection in Pumped Storage Plants" Energies 12, no. 15: 2854. https://doi.org/10.3390/en12152854

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