Next Article in Journal
Uncertainty Quantification of the PHEBUS FPT-1 Test Modelling Results
Previous Article in Journal
Macromodeling High-Speed Circuit Data Using Rational Krylov Fitting Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Vibration Characteristics of a Hydroelectric Generating System with Different Hydraulic-Mechanical-Electric Parameters in a Sudden Load Increasing Process

1
Institute of Water Resources and Hydropower Research, Northwest A&F University, Xianyang 712100, China
2
Key Laboratory of Agricultural Soil and Water Engineering in Arid and Semiarid Areas, Ministry of Education, Northwest A&F University, Xianyang 712100, China
*
Author to whom correspondence should be addressed.
Energies 2021, 14(21), 7319; https://doi.org/10.3390/en14217319
Submission received: 12 October 2021 / Revised: 27 October 2021 / Accepted: 1 November 2021 / Published: 4 November 2021

Abstract

:
In a sudden load increasing process (SLIP), the hydroelectric generating system (HGS) experiences a severe vibration response due to the sudden change of the hydraulic-mechanical-electric parameters (HMEPs). The instability of HGS limits the ability of sudden load increase, and its flexibility and reliability are reduced. Thus, in this study, a new transient nonlinear coupling model of HGS is proposed, which couples the hydro-turbine governing system (HTGS) and the hydro-turbine generator shafting system (HGSS) with the hydraulic-mechanical-electric coupling force, rotating speed, flow rate, hydro-turbine torque, electromagnetic torque, and guide vane opening. By using numerical simulation, the influences of different HMEPs on the vibration characteristics of HGS in SLIP are analyzed. The result shows that, compared with stable operating conditions, the vibration amplitude of HGS increases sharply in SLIP. The increase of the sudden load increasing amount, blade exit flow angle, mass eccentricity and excitation current, and the decrease in guide bearing stiffness and average air gap between the stator and rotor cause abnormal vibration of different degrees in the HGS. Hydraulic factors have the greatest influence on the nonlinear dynamic behavior of HGS. The maximum vibration amplitude of HGS in SLIP is increased by 70.46%, compared with that under stable operating conditions. This study provides reasonable reference for the analysis of the nonlinear dynamic behavior of HGS in SLIP under the multiple vibration sources.

1. Introduction

Hydropower is an environmentally friendly renewable energy, which has been widely used all over the world, and is of great significance to sustainable development [1,2,3]. HGS has the benefits of a quick startup and shutdown, flexible scheduling, and low operating costs [4,5,6]. When HGSs regulate the peak and frequency of the power system, it is impossible to avoid a series of transient processes, such as load increase and decrease [7]. Although the duration of the transient process is short, the parameters of the hydro-turbine, governor, and pipe are abnormal due to the sudden load change [8]. Compared with the large power system, the sudden load change has a greater impact on single-machine single-load power systems (SMSLPSs), which aggravates the impact of HMEPs on HGS, and causes severe vibration of the HGS [9,10]. For example, the water hammer superimposition occurred in Tianshengqiao-II HGS in Guangxi, China, due to the poor guide vane opening and closing strategies, which resulted in damage to the hydro-turbine runner and the surge tank [11].
The vibration of HGS is mainly caused by three factors: hydraulic, mechanical, and electric [12,13]. For hydraulic factors, Zhuang et al. proposed a hydraulic unbalance force model, and concluded that hydraulic factors determine the changing trend of HGSS dynamic behaviors [14]. Favrel et al. introduced a methodology to assess the vortex behavior, the pressure fluctuation, and the dynamic strain, and investigated the impact of precessing cavitation vortices on runner blades [15]. For mechanical factors, Al-Hussain et al. established a coupling model of lateral vibration and torsional vibration, and concluded that parallel misalignment is a source of both vibrations [16]. Yong-Wei T et al. established a coupling model of rub-induced force and temperature distribution, and found that the vibration and temperature increase when the rotor is running below its critical speed [17]. For electric factors, Calleecharan et al. derived an unbalanced magnetic pull (UMP) model, and found that the sensitivity of UMP to forcing frequency increases with the decrease in HGSS stiffness [18]. Xu et al. established a finite element rotor model and a UMP model, and concluded that UMP is limited by the saturation effects as the field current increases [19]. The above studies mainly focus on the vibration characteristics of HGS under stable operating conditions, and their application is limited. The transient process, as an important factor, effectively reflects the overall vibration characteristics of HGS.
In transient processes, the study of HGS vibration characteristics inevitably involves the time variant of parameters and the coupling of multi vibration sources due to the complex flow behavior [20,21]. Thus, the vibration characteristics of HGS in transient processes have not yet been further investigated. Zhang et al. established a hydraulic excitation model induced by a draft tube vortex, and analyzed the vibration characteristics of HGS in a sudden load increase of 10% [22]. Wu et al. proposed a coupling structure model of HGSS and the powerhouse, and researched the vibration characteristics of HGSS under mechanical and electromagnetic factors under start-up conditions [23]. The above studies only analyzed single or double vibration sources, but did not fully study the comprehensive influences of HMEPs on HGS vibration characteristics. Guo et al. established an HGSS model by using the dual-disc rotor system, and explored the vibration characteristics of the generator rotor and the hydro-turbine runner during the load rejection [24]. Although the simplified dual-disc rotor system reflects the vibration characteristics of the generator rotor and hydro-turbine runner, the vibration characteristics of other key parts of the HGSS, such as the guide bearing, are not obtained. In addition, the simplification of the model causes calculation errors.
Thus, this study proposes a new transient nonlinear coupling model of HGS, which couples HTGS and HGSS. The model is modularized, based on the real HGS parameters of China. The sub-models include: the water diversion system (WDS) model established by the method of characteristics (MOC), the generator and load model, the PID governor model, the hydro-turbine model and HGSS model established by the Riccati transfer matrix method (RTMM), and the Wilson-θ numerical integration method (WNIM). Through numerical simulation, the influences of different HMEPs on the vibration characteristics of HGS in SLIP under SMSLPS are investigated.
The innovations of this study lie in three main aspects: (1) a new coupling method is proposed, which adds HGSs, a hydro-turbine characteristic curve, and a PID governor to the HGS model; (2) the vibration characteristics of HGSS in SLIP are studied, which have not yet been further investigated; (3) considering the hydraulic-mechanical-electric coupling vibration source, the influences of different HMEPs on the vibration characteristics of HGS in SLIP are studied.
The structure of this paper is as follows: in Section 2, the transient nonlinear coupling model of HGS is established and verified. Section 3 analyzes the influences of different HMEPs on the vibration characteristics of HGSS in SLIP under SMSLPS. Finally, in Section 4, the main results are summarized.

2. Transient Nonlinear Coupling Model of HGS

The HGS is composed of an upstream reservoir, a diversion tunnel, a surge tank, a pipe, an HGSS, a draft tube, and a downstream reservoir. The schematic diagram of the HGS is shown in Figure 1. The transient nonlinear coupling model of the HGS is established through the coupling of a hydro-turbine and the WDS, and the coupling of HTGS and HGSS. The schematic diagram of the HGS coupling model is shown in Figure 2.

2.1. Model of WDS

The WDS is composed of a reservoir, a diversion tunnel, a surge tank, and a pipe. MOC is used to transform the partial differential equation of the unsteady flow into an ordinary differential equation, and then the ordinary differential equation is approximately transformed into a differential equation for hydraulic transient analysis [25,26]. The hydraulic transient motion equation L1 and the continuous equation L2 are:
{ L 1 = H X + V g V X + 1 g V t + f | V | V 2 g D = 0 L 2 = H t + V H X + a w 2 g V X + V sin σ = 0
where H is the water head; X is the pipe length; V is the flow velocity; g is the gravitational acceleration; t is the time; f is the Darcy–Weisbach friction coefficient; D is the pipe diameter; aw is the pressure wave velocity; and σ is the angle of the pipe inclined with the horizontal.
Equation (1) is solved by MOC combined with boundary conditions. The schematic diagram of MOC is shown in Figure 3. The flow rate and the water head at point p are calculated through the flow rate and the water head at point d and u, which are shown in Equations (2) and (3). More detailed descriptions of MOC are obtained in [27].
{ C + : Q p = C p C a H p C : Q p = C n + C a H p
{ C a = g A a w C p = Q d + g A a w H d f Δ t 2 D A Q d | Q d | + Q d Δ t g a w 2 sin σ C n = Q u g A a w H u f Δ t 2 D A Q u | Q u | Q u Δ t g a w 2 sin σ
where Qp, Qd, Qu, Hp, Hd, and Hu are the flow rate and the water head at point p, d, and u, respectively; Ca, Cp, and Cn are intermediate variables; A is the cross-sectional area of the pipe; Δt is the time step.

2.2. The Boundary Conditions of WDS

When the WDS model established in 2.1 is used to calculate the flow rate and the water head at each point of the pipe, the corresponding boundary conditions are required. The schematic diagram of boundary conditions for each node of the WDS is shown in Figure 4.
(1)
Upstream and downstream reservoirs
Considering the large area of the reservoir, it is assumed that the water levels of the upstream and downstream reservoirs are constant during the transient flow, and the pipe inlet head loss is ignored. The boundary conditions of the upstream and downstream reservoirs are [28]:
{ H p 1 = H c o n s t 1 Q p 1 = C n 1 + C a 1 H p 1
{ H p 2 = H c o n s t 2 Q p 2 = C p 2 C a 2 H p 2
where Qp1 and Qp2 are the flow rate at points p1 and p2, respectively; Hp1 and Hp2 are the water head at p1 and p2, respectively; Hconst1 and Hconst2 are the water head of upstream reservoir and downstream reservoir, respectively; and Cn1, Cp2, Ca1, and Ca2 are intermediate variables.
(2)
Surge tank
To reduce the water hammer pressure of the pipe in the transient process, a surge tank is built at the connection between the pressure diversion channel and the penstock for long WDS. The mathematical model is described in [29]. The boundary condition of surge tank is:
{ Q d 3 = C p 3 C a 31 H d 3 Q u 3 = C n 3 + C a 32 H u 3 Q d 3 = Q j + Q u 3 H d 3 = H u 3 = H p 3 H p 3 = H j + R j Q j | Q j | A w j 2 H j = H j , Δ t + ( Q j + Q j , Δ t ) Δ t 2 A j
where Qd3 and Qu3 are the flow rate at points d3 and u3, respectively; Hd3 and Hu3 are the water head at points d3 and u3, respectively; Cp3, Cn3, Ca31, and Ca32 are intermediate variables; Qj and Qjt are the surge tank flow rate at this moment and the last moment, respectively; Hj and Hjt are the surge tank water level at this moment and the last moment, respectively; Hp3 is the water head at point p3; Rj is the hydraulic loss coefficient of surge tank inlet impedance; Awj is the cross-sectional area of the impedance hole; and Aj is the cross-sectional area of surge tank.
(3)
Elbow pipe and bifurcated pipe
The front and rear sections of the elbow pipe satisfy the characteristic equation. The elbow pipe also satisfies the flow rate and water head balance equation, as shown in Equation (7). The boundary condition of elbow pipe is [29]:
{ Q d 4 = C p 4 C a 41 H d 4 Q u 41 = C n 4 + C a 42 H u 41 Q u 41 = Q u 42 H u 41 = H u 42 + ξ V u 42 2 2 g
where Qd4, Qu41 and Qu42 are the flow rate at points d4, u41, and u42, respectively; Hd4, Hu41, and Hu42 are the water head at points d4, u41, and u42, respectively; Cp4, Cn4, Ca41, and Ca42 are intermediate variables; ξ is the local head loss coefficient; and Vu42 is the flow velocity at point u42.
The front and rear sections of the bifurcated pipe satisfy the characteristic equation. The head loss at the joint is ignored. According to the energy equation and continuity equation, the boundary condition of the bifurcated pipe is [29]:
{ Q d 5 = C p 5 C a 5 H d 5 Q u 51 = C n 51 + C a 51 H u 51 Q u 52 = C n 52 + C a 52 H u 52 Q u 53 = C n 53 + C a 53 H u 53 H d 5 = H u 51 = H u 52 = H u 53 = H p Q d 5 = Q u 41 + Q u 52 + Q u 53
where Qd5, Qu51, Qu52, and Qu53 are the flow rate at points d5, u51, u52, and u53, respectively; Hd5, Hu51, Hu52, and Hu53 are the water head at points d5, u51, u52, and u53, respectively; Cp5, Cn51, Cn52, Cn53, Ca5, Ca51, Ca52, and Ca53 are intermediate variables.
(4)
Hydro-turbine
The water head of the hydro-turbine is the difference between the inlet of the volute and the outlet of the draft tube. The flow loss of the hydro-turbine is ignored. The boundary condition of the hydro-turbine is [24]:
{ Q d 6 = Q u 6 = Q t Q d 6 = C p 6 C a 61 H d 6 Q u 6 = C n 6 + C a 62 H u 6 H t = H d 6 H u 6
where Qd6 and Qu6 are the flow rate at points d6 and u6, respectively; Hd6 and Hu6 are the water head at points d6 and u2, respectively; Cp6, Cn6, Ca61, and Ca62 are intermediate variables; and Qt is the flow rate of the hydro-turbine, which is obtained from the hydro-turbine model.

2.3. Model of Hydro-Turbine

The modeling process of the hydro-turbine depends on the characteristic curve of hydro-turbine, as the boundary of the pipe, its characteristics need to be. When the guide vane opening, rotating speed, and water head are known, the characteristic line equation of the pipe is solved simultaneously. The hydro-turbine characteristic data are transformed into characteristic curves through the polynomial interpolation method, as shown in Figure 5. The calculation formulas of the flow rate, torque, and speed are [30]:
{ Q 11 = f ( n 11 , a ) , Q t = Q 11 D 1 2 H t M 11 = f ( n 11 , a ) , M t = M 11 D 1 3 H t n t = n 11 H t D 1
where Q11 is the unit flow rate; M11 is the unit torque; n11 is the unit rotating speed; and a is the guide vane opening.

2.4. Model of Generator and Load

In this study, the parallel operation of the HGSS is regarded as a rotating rigid body, and electromagnetic power, torque angle, and the change of the field winding flux are not considered. Thus, the first-order generator model is selected. In addition, HGSS still bears the grid load, and the generator electromagnetic torque is not zero in SLIP. The model of generator and load [31] is:
{ M t M g = J d ω d t m g = C g + A g + e g ( n n r ) n t = n ( t Δ t ) + Δ t T a [ 1.5 ( m t ( t Δ t ) m g ( t Δ t ) ) 0.5 ( m t ( t 2 Δ t ) m g ( t 2 Δ t ) ) ]
where Mt is the driving torque; Mg is the electromagnetic torque; ω is the angular velocity; J is the moment of inertia; mg is the relative electromagnetic torque; Cg is the relative load torque; Ag is the step value of load torque; nr is the relative rated speed; eg is the load self-regulation coefficient; n is the relative rotating speed; and Ta is the inertia time constant.

2.5. Model of PID Governor

In SLIP, the control mode of HGS is changed to frequency control under SMSLPS in order to ensure that the HGS tracks the load change automatically, and maintains the frequency near the rated frequency [32]. Where KP, KI, and KD are the proportional, integral, and differential coefficients, respectively; x1 is the intermediate variable; x2 is the output variable; Ty is the servomotor response time; and bp is the permanent difference coefficient. The equations are shown in Equations (12) and (13) [33]. The schematic diagram of the PID governor is shown in Figure 6.
{ x ˙ 1 = K I n t b p e x ˙ 2 = 1 T y x 2 + 1 T y e
e = x 1 K P n t K D n ˙ t
When transformed in Equations (12) and (13), the state equation of the PID governor model is:
[ x ˙ 1 x ˙ 2 ] = [ b p 0 1 / T y 1 / T y ] [ x 1 x 2 ] + [ b p K P K I b p K D K P / T y K D / T y ] [ n t n ˙ t ]

2.6. Model of HGSS

The HGSS is composed of runners, turbine shafts, generator shafts, generator rotors, and guide bearings. This study discretizes the vertical HGSS with a continuous mass distribution, and transforms it into a multi degree of freedom nodes with the centralized mass and the centralized moment of inertia [34]. These nodes are generally selected in the generator rotors, bearings, runners, and shaft diameter changes. In this section, the structural model of HGSS is established through the combination of RTMM and WNIM. The HGSS is divided into 18 nodes and 17 shaft segments for modeling. The schematic diagram of the HGSS is shown in Figure 7.
(1)
WNIM
WNIM is a stepwise integration method. The displacements of each node at time t + θΔt and the displacements, velocities, and accelerations of each node at time t are known to obtain the displacements, velocities, and accelerations of each node at time t + Δt [35]. The expression of WNIM is
{ x ¨ t + Δ t = 6 θ 3 Δ t 2 ( x t + θ Δ t x t ) 6 θ 2 Δ t x ˙ t + ( 1 3 θ ) x ¨ t x ˙ t + Δ t = x ˙ t + Δ t 2 ( x ¨ t + Δ t + 2 x ¨ t ) x t + Δ t = x t + x ˙ t Δ t + Δ t 2 6 ( x ¨ t + Δ t + 2 x ¨ t )
where x is the displacement. When θ ≥ 1.37, WNIM is unconditionally stable.
(2)
Differential equations of motion:
{ m i { x ¨ y ¨ } i = [ K x x K x y K y x K y y ] i { x y } i [ C x x C x y C y x C y y ] i { x ˙ y ˙ } i + { P x L P x R P y L P y R } i + { F x ( t ) F y ( t ) } i J d i { φ ¨ ψ ¨ } i = ω [ 0 J p J p 0 ] i { φ ˙ ψ ˙ } i + { M x R M x L M y R M y L } i + { L a x ( t ) L a y ( t ) } i
where subscripts i, x, y, d, and p are the ith node, x-direction, y-direction, radial direction and axial direction, respectively; superscripts L and R are the left and right sections of each node; m is the mass; x, y, ψ, and φ are the displacements; K is the stiffness coefficient; C is the damping coefficient; P and M are the shear force and the shear moment, respectively; and Fa and La are additional forces and additional moments, respectively.
(3)
RTMM
The state vectors of each node are divided into f = {Mx, Qx, My, Qy}T and e = {x, y, ψ, φ}T. Combined with the differential equation of motion, the transfer matrix at time t + θΔt is:
{ f e } i + 1 = [ B 11 B 12 B 21 B 22 ] [ T 11 T 12 T 21 T 22 ] { f e } i + { T 13 T 23 } i = [ u 11 u 12 u 21 u 22 ] { f e } i + { F f F e } i
[ u 22 ] i = [ B 21 ] i [ T 12 ] i + [ B 22 ] i , { F f } i = [ B 11 ] i { T 13 } i , { F e } i = [ B 21 ] i { T 13 } i , [ T 23 ] = { 0 0 0 0 } T , [ T 11 ] i = [ T 22 ] i = [ 1 0 0 0 0 1 0 0 0 0 1 0 0 0 0 1 ] , [ T 12 ] i = [ 0 K 2 0 K 3 W 1 0 W 2 0 0 K 3 0 K 2 W 3 0 W 4 0 ] , [ B 11 ] i = [ B 22 ] i = [ u 11 ] i = [ 1 l 0 0 0 1 0 0 0 0 1 l 0 0 0 1 ] , [ u 12 ] i = [ l W 1 K 2 l W 2 K 3 W 1 0 W 2 0 l W 3 K 3 l W 4 K 2 W 3 0 W 4 0 ] , [ B 21 ] i = [ u 21 ] i = [ a 2 a 3 0 0 a 1 a 2 0 0 0 0 a 2 a 3 0 0 a 1 a 2 ] , [ T 13 ] = { K 2 φ t 6 J d i θ Δ t φ ˙ t 2 J d i φ ¨ t K 3 ψ t 2 J p i ω ψ ˙ t 1 2 J p i ω θ Δ t ψ ¨ t L x ( t + θ Δ t ) ( 2 m + 1 2 θ Δ t C x x ) x ¨ t + 1 2 θ Δ t C x y y ¨ t + ( 6 m θ Δ t + 2 C x x ) x ˙ t + 2 C x y y ˙ t + ( 6 m ( θ Δ t ) 2 + 3 C x x θ Δ t ) x t + 3 θ Δ t C x y y t + F x ( t + θ Δ t ) K 2 ψ t 6 J d i θ Δ t ψ ˙ t 2 J d i ψ ¨ t + K 3 φ t + 2 J p i ω φ ˙ t + 1 2 J p i ω θ Δ t φ ¨ t L y ( t + θ Δ t ) ( 2 m + 1 2 θ Δ t C y y ) y ¨ t + 1 2 θ Δ t C y x x ¨ t + ( 6 m θ Δ t + 2 C y y ) y ˙ t + 2 C y x x ˙ t + ( 6 m ( θ Δ t ) 2 + 3 C y y θ Δ t ) y t + 3 θ Δ t C y x x t + F y ( t + θ Δ t ) } ,
K 1 = 6 m i ( θ Δ t ) 2 , K 2 = 6 J d i ( θ Δ t ) 2 , K 3 = 3 J p i ω θ Δ t , W 1 = K x x 3 θ Δ t C x x 6 m ( θ Δ t ) 2 , W 2 = K x y 3 θ Δ t C x y , W 3 = K y x 3 θ Δ t C y x , W 4 = K y y 3 θ Δ t C y y 6 m ( θ Δ t ) 2 , a 1 = l E J , a 2 = l 2 2 E J , a 3 = l 3 6 E J .
where l is the length of the shaft segment; and EJ is the bending stiffness.
The Riccati transformation is:
{ f } i = [ S ] i { e } i + { P } i
where [S]i and {p}i are the undetermined matrices.
It is obtained from Equations (17) and (20):
{ [ S ] i + 1 = [ u 11 S + u 12 ] i [ u 21 S + u 22 ] i 1 { P } i + 1 = { u 11 P + F f } i [ S ] i + 1 [ u 21 P + F e ] i
From the boundary conditions {f}1 = 0 and {e}1 ≠ 0, [S]1 = 0 and {p}1 = 0 is obtained. In addition, from Equation (21), [S]i and {p}i are obtained. The boundary condition {f}N + 1 = 0 and Equation (21) are combined to obtain:
{ e } N + 1 = [ S ] N + 1 1 { P } N + 1
Thus, the displacements of each node at time t + θΔt are obtained:
{ e } i = [ u 21 S + u 22 ] i 1 ( { e } i + 1 [ u 21 P + F e ] i )

2.7. Additional Forces on HGSS

(1)
Nonlinear oil film force
Most of the guide bearings on HGSS are oil-lubricated sliding bearings with tilting pads. The oil film force of the bearing has a non-linear relationship with the speed and displacement of the journal. The expression of nonlinear oil film force [36] is:
{ F x F y } = σ { f x f y } = [ ( x 2 y ) 2 + ( y + 2 x ) 2 ] 1 / 2 1 x 2 y 2 { 3 x V ( x , y , α ) sin α G ( x , y , α ) 2 cos α S ( x , y , α ) 3 y V ( x , y , α ) + cos α G ( x , y , α ) 2 sin α S ( x , y , α ) }
V ( x , y , α ) = 2 + ( y cos α x sin α ) G ( x , y , α ) 1 x 2 y 2
G ( x , y , α ) = 2 ( 1 x 2 y 2 ) 1 / 2 [ π 2 + arctan y cos α x sin α ( 1 x 2 y 2 ) 1 / 2 ]
S ( x , y , α ) = x cos α + y sin α 1 ( x cos α + y sin α ) 2
α = arctan y + 2 x x 2 y π 2 sgn ( y + 2 x x 2 y ) π 2 sgn ( y + 2 x )
σ = μ ω R b L b ( R b C b ) 2 ( L b D b ) 2
where μ is the lubricating oil viscosity; Rb is the guide bearing radius; Lb is the guide bearing length; Cb is the guide bearing clearance; and Db is the guide bearing diameter.
(2)
Mechanical unbalance force
Due to uneven mass distribution and installation accuracy, mass eccentricity occurs at the generator rotor and the hydro-turbine runner, which causes the rotation center to shift. The expression of mechanical unbalance force [37] is:
F m u f = m e ω 2
where Fmuf is the mechanical unbalance force; and e is the eccentricity.
(3)
UMP
When the air gap between the stator and the rotor is uneven due to manufacturing or installation errors, the rotor produces a lateral unbalanced magnetic tension. The expression of UMP [38] is:
{ F x u m p = R r L r π 4 μ 0 F j 2 ( 2 Λ 0 Λ 1 + Λ 1 Λ 2 + Λ 2 Λ 3 ) cos η F y u m p = R r L r π 4 μ 0 F j 2 ( 2 Λ 0 Λ 1 + Λ 1 Λ 2 + Λ 2 Λ 3 ) sin η
Λ n = { μ 0 δ 0 1 1 ε 2 ( n = 0 ) 2 μ 0 δ 0 1 1 ε 2 ( 1 1 ε 2 ε ) n ( n > 0 )
where Fx-ump and Fy-ump are the unbalanced magnetic pulls in x-direction and y-direction, respectively; Rr is the generator rotor radius; Lr is the generator rotor length; μ0 is the air permeability; Fj is the fundamental magnetomotive force of the excitation current; ε is the relative eccentricity; δ0 is the average air gap between stator and rotor; and η is the rotating angle.
(4)
Hydraulic excitation force
When the hydro-turbine performs energy conversion, it also brings complicated hydraulic vibration problems to the HGSS. The reasons for this include many aspects, such as operation under non-design conditions, draft tube pressure pulsation, unequal sealing gaps, and excessive axial water thrust, etc. Zhuang [14] proposed an excitation force model on a single blade of the hydro-turbine runner. Superposition of the forces on all the blades of the hydro-turbine runner is the hydraulic excitation force, and its expression is
{ F x h e f = n b γ C y Q 2 F b cos ( β m λ ) 8 g cos λ [ 1 ( s 1 π D 1 b 0 sin β 1 ) 2 + 1 ( s 2 π D 2 2 sin β 2 ) 2 + 2 cos ( β 2 β 1 ) s 1 s 2 π 2 D 1 D 2 2 b 0 sin β 1 sin β 2 ] cos η F y h e f = n b γ C y Q 2 F b cos ( β m λ ) 8 g cos λ [ 1 ( s 1 π D 1 b 0 sin β 1 ) 2 + 1 ( s 2 π D 2 2 sin β 2 ) 2 + 2 cos ( β 2 β 1 ) s 1 s 2 π 2 D 1 D 2 2 b 0 sin β 1 sin β 2 ] sin η
β m = arcsin ( 2 ( s 1 D 1 b 0 + s 2 D 2 2 ) ( s 2 D 2 2 sin β 1 ) 2 + ( s 1 D 1 b 0 D 2 2 sin β 2 ) 2 + 2 s 1 s 2 D 1 D 2 2 b 0 cos ( β 2 β 1 ) sin β 1 sin β 2 )
β 1 = arctan ( 2 Q tan α 1 π D 1 2 ω s 1 b 0 tan α 1 2 Q )
where Fx-hef and Fy-hef are the hydraulic excitation forces in x-direction and y-direction, respectively; nb is the number of runner blades; γ is the unit weight; λ = arctan(Cx/Cy); Cy is the lift coefficient; Cx is the drag coefficient; Fb is the maximum area of the blade; βm is the direction angle of the relative average speed; s1 and s2 are the entrance excretion coefficient and exit excretion coefficient, respectively; D1 and D2 are the entrance diameter and exit diameter of hydro-turbine runner, respectively; b0 is the height of guide vane; β1 and β2 are the blade entrance flow angle and the blade exit flow angle, respectively; and α1 is the guide vane opening angle.

2.8. Model Validation

This study selects the real HGS parameters of Qinghai Province, China, to establish the model, as well as the comprehensive verification based on the operation characteristics and vibration characteristics of HGS. The HGS parameters of Qinghai Province in China are shown in Table 1.
First, the operation characteristics of HGSS #1, #2 and #3 in the simultaneous load rejection condition are simulated, and the simulation results are compared with the experimental results from [39]. Table 2 shows the comparison of the simulation results and the experiment results.
From Table 2, it can be seen that the maximum error of the maximum and minimum pressure at the volute are 3.9% and 2.79%, respectively; the maximum error of the maximum speed rise rate is 0.15%; and the error of the maximum and minimum surge in surge tank are 0.13% and 0.31%, respectively. The above comparisons show that the maximum error between the simulation and the experiment do not exceed 4%, and indicate that the established model is correct and reasonable in operation characteristics.
Second, 50%-, 75%-, and 100%-rated load conditions in an operable load area are selected to simulate the vibration characteristics of the HGS, and the simulation results are compared with the National Standard of China “Specification installation of hydraulic turbine generator units. GB/T 5864-2003” [40]. For the HGSS in this study, the allowable absolute vibration amplitudes of the upper guide bearing, the lower guide bearing, the flange, and the water guide bearing are 0.022 mm, 0.076 mm, 0.106 mm, and 0.2 mm, respectively. The vibration characteristics of different guide bearings and the flange are shown in Figure 8.
The vibration characteristics of HGSS #1, #2, and #3 are similar, and HGSS #1 is used as an example. As shown in Figure 6, the upper guide bearing and the lower guide bearing are mainly affected by the mechanical unbalance force and UMP at the generator rotor, and the vibration amplitudes are the largest under the 100%-rated load condition, which are 0.0096 mm and 0.008 mm, respectively. The flange and the water guide bearing are mainly affected by the hydraulic excitation force, and the vibration amplitudes are the largest under the 50%-rated load condition, which are 0.017 mm and 0.098 mm, respectively. The above results show that the vibration amplitudes of HGS do not exceed the allowable vibration amplitude in National Standard of China, and indicate that the established model is correct and reasonable in vibration characteristics.

3. Numerical Simulation and Result Analysis

3.1. Hydraulic Factors

(1)
Sudden load increasing amount
For HGS in this study, 50–100%-rated load area is the operable area. Thus, in this subsection, the vibration characteristics of HGSS #1, #2, and #3 which collectively suddenly increase 30%, 40%, and 50%-rated load from the 50%-rated load (ΔP = 30%, ΔP = 40%, and ΔP = 50%) are analyzed. The operation and vibration characteristics of the hydro-turbine runner under different sudden load increasing amounts are shown in Figure 9, Figure 10 and Figure 11, respectively.
From Figure 9, the PID governor tracks the load change automatically, and opens the guide vane quickly. The rotating speed decreases by 11.16%, 12.87%, and 15.21%, and the water head decreases by 6.67%, 7.52%, and 8.46%, with the increase of sudden load increasing amount, respectively.
From Figure 10, it can be seen that Area I and Area III are the stable operating areas; Area II is the operating area of SLIP. When t = 16 s, the maximum vibration response occurs at the hydro-turbine runner. When ΔP = 50%, the vibration amplitudes of the hydro-turbine runner in x-direction and y-direction in Area II are the largest, which increase by 69.89% and 69.95%, compared with that of Area I, respectively. When ΔP = 30%, the vibration amplitudes of the hydro-turbine runner in x-direction and y-direction in Area III are the largest, which increase by 33.79% and 34.1%, compared with that of Area I, respectively. Thus, the hydraulic excitation force causes the vibration amplitude under a partial load to be greater than that under a rated load. In addition, as demonstrated in Figure 11, the greater the sudden load increasing amount, the greater the axis trajectory and vibration amplitude in Area II, and the smaller axis trajectory and vibration amplitude in Area III.
(2)
Blade exit flow angle
When the Francis HGSS was operating in the non-optimal load condition area, the direction of the water flow from the blade exit had a circumferential component, which caused the vortex. Thus, in this subsection, the vibration characteristics of HGS under ΔP = 50% and the blade exit flow angles, taken as 0.05 rad, 0.1 rad, and 0.15 rad, are analyzed, respectively. The vibration characteristics of the hydro-turbine runner under different blade exit flow angles are shown in Figure 12 and Figure 13.
From Figure 12 and Figure 13, when t = 15 s, the maximum vibration response occurs at the hydro-turbine runner. When β2 = 0.15 rad, the vibration amplitudes of the hydro-turbine runner in x-direction and y-direction are the largest, which increase by 67.36% and 66.54%, compared with that of Area I, respectively. The greater the blade exit flow angle, the greater the axis trajectory and vibration amplitude of the hydro-turbine runner.

3.2. Mechanical Factors

(1)
Mass eccentricity
Uneven mass distribution of the rotating parts caused by poor materials and insufficient installation accuracy is common, which causes the vibration of the HGS. Thus, in this subsection, the vibration characteristics of the HGS under ΔP = 50% and the mass eccentricities, taken as 0.4 mm, 0.6 mm, and 0.8 mm, are analyzed, respectively. The vibration characteristics of the generator rotor and the hydro-turbine runner under different mass eccentricities are shown in Figure 14 and Figure 15.
From Figure 14 and Figure 15, when t = 18 s and t = 15 s, the maximum vibration response occurs at the generator rotor and the hydro-turbine runner, respectively. When e = 0.8 mm, the vibration amplitudes of the generator rotor and the hydro-turbine runner in x-direction and y-direction are the largest, which increase by 40.71%, 37.25%, 65.3%, and 64.95%, compared with that of Area I, respectively. The axis trajectory and vibration amplitude of the HGS increase with mass eccentricity increase. In addition, the mass eccentricity has a great impact on the generator rotor, but a lesser impact on the hydro-turbine runner, as the hydro-turbine runner is greatly affected by hydraulic factors.
(2)
Guide bearing stiffness
The guide bearing is used to bear the radial force during the operation, and limit the vibration of HGSS. Thus, the unreasonable guide bearing stiffness leads to the severe vibration of HGS. In this subsection, the vibration characteristics of the HGS under ΔP = 50% and different stiffnesses of the upper guide bearing, lower guide bearing, and water guide bearing are analyzed, respectively. The guide bearing stiffness is taken as 1.0 × 108 N × m−1, 3.0 × 108 N × m−1, and 5.0 × 108 N × m−1, respectively. The vibration characteristics of the guide bearing under different stiffnesses are shown in Figure 16 and Figure 17.
As demonstrated in Figure 16 and Figure 17, the maximum vibration response of upper guide bearing, lower guide bearing, and water guide bearing occurs at t = 22 s, 17 s, and 15 s, respectively. When K = 1.0 × 108 N × m−1, the vibration amplitudes of the upper guide bearing, lower guide bearing, and water guide bearing in x-direction and y-direction are the largest, which increase by 41.24%, 36.35%, 63.54%, 60.01%, 70.46%, and 70.17%, compared with that of Area I, respectively. In conclusion, the stiffness of the water guide bearing has the greatest influence on vibration response, and the stiffness of the lower guide bearing has less influence on the vibration response than that of the upper guide bearing. The axis trajectory and vibration amplitude at guide bearing increase as guide bearing stiffness decreases.

3.3. Electric Factors

(1)
Excitation current
The change of excitation current causes the change of UMP, which leads to the vibration of the HGS. Thus, in this subsection, the vibration characteristics of HGS under ΔP = 50% and the excitation currents, taken as 50% Ij, 75% Ij, and 100% Ij, are analyzed, respectively. The vibration characteristics of the generator rotor under different excitation currents are shown in Figure 18 and Figure 19.
As can be observed in Figure 18 and Figure 19, when t = 21 s, the maximum vibration response occurs at the generator rotor. When 75% excitation current is conducted, the vibration amplitudes of the generator rotor in x-direction and y-direction are the largest, which increase by 40% and 34.67%, compared with that of Area I, respectively. The axis trajectory and vibration amplitude at the generator rotor increase as the excitation current increases.
(2)
Average air gap between stator and rotor
An uneven air gap between the stator and rotor produces greater UMP, and aggravates the vibration of the HGSS. Thus, in this subsection, the vibration characteristics of HGS under ΔP = 50% and the average air gaps between the stator and rotor, taken as δ0 = 20 mm, δ0 = 30 mm, and δ0 = 40 mm, are analyzed, respectively. The vibration characteristics of the generator rotor under different average air gaps between the stator and rotor are shown in Figure 20 and Figure 21.
As demonstrated in Figure 20 and Figure 21, when t = 21 s, the maximum vibration response occurs at the generator rotor. When δ0 = 20 mm, the vibration amplitudes of the generator rotor in x-direction and y-direction are the largest, which increase by 40% and 34.54%, compared with that of Area I, respectively. The axis trajectory and vibration amplitude at the generator rotor increase as the average air gap between the stator and rotor decreases.

4. Conclusions

This study proposes a new transient nonlinear coupling model of HGS to analyze the influences of different HMEPs on the vibration characteristics of HGS in SLIP under SMSLPS. The main conclusions obtained are as follows:
(1)
For hydraulic factors, the greater the sudden load increasing amount and blade exit flow angle, the greater the vibration of the hydro-turbine runner. The maximum vibration amplitude is increased by 69.95%, compared with that under stable operating conditions. Moreover, the vibration under a partial load is greater than that under a rated load. Hydraulic factors determine the tendency of vibration at hydro-turbine runner.
(2)
For mechanical factors, the vibration of HGSS increases as mass eccentricity increases and guide bearing stiffness decreases. The maximum vibration amplitude is increased by 70.46%, compared with that under stable operating conditions. The mass eccentricity has a great impact on the generator rotor, but a lesser impact on the hydro-turbine runner. The influence of the water guide bearing stiffness on the vibration response is greater than that of the upper guide bearing and the lower guide bearing.
(3)
For electric factors, the vibration of the generator rotor increases as the excitation current increases and the average air gap between the stator and rotor decreases. The maximum vibration amplitude is increased by 40%, compared with that under stable operating conditions.
More importantly, this study provides reasonable reference for the analysis of the nonlinear dynamic behavior of HGS in SLIP, and the stable operation of the hydropower station.

Author Contributions

Y.G.: conceptualization, methodology, software, writing—original draft, writing—reviewing and editing. X.L.: software, investigation, data curation, writing—review and editing. Z.N.: software, formal analysis, and data curation. Z.C.: visualization, writing—review and editing. L.L.: conceptualization, methodology and data curation. H.X.: software, writing—review and editing. D.C.: funding acquisition, supervision, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the scientific research foundation of the Natural Science Foundation of Shaanxi Province of China (2019JLP-24) and the Technology Innovation Team (2020TD-025).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

SLIPSudden Load Increasing Process
HMEPsHydraulic-Mechanical-Electric Parameter
HGSSHydro-turbine Generator Shafting System
UMPUnbalanced Magnetic Pull
MOCMethod Of Characteristics
WNIMWilson-θ Numerical Integration Method
HGSHydroelectric Generating System
HTGSHydro-Turbine Governing System
SMSLPSSingle-Machine Single-Load Power System
WDSWater Diversion System
RTMMRiccati Transfer Matrix Method

References

  1. Dehghani, M.; Riahi-Madvar, H.; Hooshyaripor, F.; Mosavi, A.; Shamshirband, S.; Zavadskas, E.K.; Chau, K.-W. Prediction of Hydropower Generation Using Grey Wolf Optimization Adaptive Neuro-Fuzzy Inference System. Energies 2019, 12, 289. [Google Scholar] [CrossRef] [Green Version]
  2. Guo, W.C. A review of the hydraulic transient and dynamic behavior of hydropower plants with sloping ceiling tailrace tunnels. Energies 2019, 12, 3220. [Google Scholar] [CrossRef] [Green Version]
  3. Sipahutar, R.; Bernas, S.M.; Imanuddin, M.S. Renewable energy and hydropower utilization tendency worldwide. Renew. Sustain. Energy Rev. 2013, 17, 13–15. [Google Scholar]
  4. Singh, V.K.; Singal, S.K. Operation of hydro power plants-a review. Renew. Sustain. Energy Rev. 2017, 69, 610–619. [Google Scholar] [CrossRef]
  5. Chazarra, M.; Garcia-Gonzalez, J.; Perez-Diaz, J.I.; Arteseros, M. Stochastic optimization model for the weekly scheduling of a hy-dropower system in day-ahead and secondary regulation reserve markets. Electr. Power Syst. Res. 2016, 130, 67–77. [Google Scholar] [CrossRef]
  6. Trivedi, C. Study of pressure pulsations in a Francis turbine designed for frequent start-stop. J. Energy Resour. Technol. 2021, 143, 1–24. [Google Scholar] [CrossRef]
  7. Guo, P.C.; Zhang, H.; Gou, D.M. Dynamic characteristics of a hydro-turbine governing system considering draft tube pressure pulsation. IET Renew. Power Gener. 2020, 14, 1210–1218. [Google Scholar] [CrossRef]
  8. Li, H.; Chen, D.; Tolo, S.; Xu, B.; Patelli, E. Hamiltonian Formulation and Analysis for Transient Dynamics of Multi-Unit Hydropower System. J. Comput. Nonlinear Dyn. 2018, 13, 101004. [Google Scholar] [CrossRef]
  9. Pan, Z.; Dai, Y. The influence of the overspeed protection control for the isolated grid stability. In Proceedings of the International Conference on Power System Technology, Hangzhou, China, 24–28 October 2010; pp. 1–6. [Google Scholar]
  10. Yan, D.L.; Wang, W.Y.; Chen, Q.J. Nonlinear modeling and dynamic analyses of the hydro-turbine governing system in the load shedding transient regime. Energies 2018, 11, 1244. [Google Scholar] [CrossRef] [Green Version]
  11. Liao, C.W.; Zhang, J.S. Analysis of the breast wall collapse of the surge shaft of Tianshengqiao-II hydropower station. Water Power 1994, 10, 24–26. (In Chinese) [Google Scholar]
  12. Ishida, Y.; Liu, J.; Inoue, T.; Suzuki, A. Vibrations of an Asymmetrical Shaft with Gravity and Nonlinear Spring Characteristics (Isolated Resonances and Internal Resonances). J. Vib. Acoust. Trans. ASME 2008, 130, 041004. [Google Scholar] [CrossRef]
  13. Xu, B.; Chen, D.; Behrens, P.; Ye, W.; Guo, P.; Luo, X. Modeling oscillation modal interaction in a hydroelectric generating system. Energy Convers. Manag. 2018, 174, 208–217. [Google Scholar] [CrossRef]
  14. Zhuang, K.; Gao, C.; Li, Z.; Yan, D.; Fu, X. Dynamic Analyses of the Hydro-Turbine Generator Shafting System Considering the Hydraulic Instability. Energies 2018, 11, 2862. [Google Scholar] [CrossRef] [Green Version]
  15. Favrel, A.; Junior, J.G.P.; Müller, A.; Landry, C.; Yamamoto, K.; Avellan, F. Swirl number based transposition of flow-induced mechanical stresses from reduced scale to full-size Francis turbine runners. J. Fluids Struct. 2020, 94, 102956. [Google Scholar] [CrossRef]
  16. Al-Hussain, K.M.; Redmond, I. Dynamic response of two rotors connected by rigid mechanical coupling with parallel misalignment. J. Sound Vib. 2002, 249, 483–498. [Google Scholar] [CrossRef]
  17. Yong-Wei, T.; Jian-Gang, Y. Research on vibration induced by the coupled heat and force due to rotor-to-stator rub. J. Vib. Control 2011, 17, 549–566. [Google Scholar] [CrossRef]
  18. Calleecharan, Y.; Aidanpaa, J.O. Dynamics of a hydropower generator subjected to unbalanced magnetic pull. Proc. Inst. Mech. Eng. Part C J. Mech. Eng. Sci. 2011, 225, 2076–2088. [Google Scholar] [CrossRef]
  19. Xu, Y.; Li, Z.H. Computational model for investigating the influence of unbalanced magnetic pull on the radial vibration of large hydro-turbine generators. J. Vib. Acoust. Trans. ASME 2012, 134, 051013. [Google Scholar] [CrossRef]
  20. Li, H.; Chen, D.; Zhang, X.; Wu, Y. Dynamic analysis and modelling of a Francis hydro-energy generation system in the load rejection transient. IET Renew. Power Gener. 2016, 10, 1140–1148. [Google Scholar] [CrossRef]
  21. Zeng, Y.; Zhang, L.; Guo, Y.; Qian, J.; Zhang, C. The generalized Hamiltonian model for the shafting transient analysis of the hydro turbine generating sets. Nonlinear Dyn. 2014, 76, 1921–1933. [Google Scholar] [CrossRef] [Green Version]
  22. Zhang, L.; Wu, Q.; Ma, Z.; Wang, X. Transient vibration analysis of unit-plant structure for hydropower station in sudden load increasing process. Mech. Syst. Signal Process. 2019, 120, 486–504. [Google Scholar] [CrossRef]
  23. Wu, Q.Q.; Zhang, L.K.; Ma, Z.Y. A model establishment and numerical simulation of dynamic coupled hydraulic–mechanical–electric–structural system for hydropower station. Nonlinear Dyn. 2016, 87, 459–474. [Google Scholar] [CrossRef]
  24. Guo, B.Q.; Xu, B.B.; Chen, D.Y.; Li, F.; Behrens, P.; Egusquiza, E. Vibration characteristics of the hydroelectric generating systems during load rejection transient process. J. Comput. Nonlinear Dyn. 2019, 14, 071006. [Google Scholar] [CrossRef]
  25. 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]
  26. Chaudhry, M.H. Characteristics and finite-difference methods. Appl. Hydraul. Transients 2014, 94, 65–113. [Google Scholar]
  27. Riasi, A.; Tazraei, P. Numerical analysis of the hydraulic transient response in the presence of surge tanks and relief valves. Renew. Energy 2017, 107, 138–146. [Google Scholar] [CrossRef] [Green Version]
  28. Karlsson, M.; Aidanpää, J.-O.; Perers, R.; Leijon, M. Rotor Dynamic Analysis of an Eccentric Hydropower Generator with Damper Winding for Reactive Load. J. Appl. Mech. 2007, 74, 1178–1186. [Google Scholar] [CrossRef]
  29. Rezghi, A.; Riaasi, A.; Tazraei, P. Multi-objective optimization of hydraulic transient condition in a pump-turbine hydropower considering the wicket-gates closing law and the surge tank position. Renew. Energy 2020, 148, 478–491. [Google Scholar] [CrossRef]
  30. Yang, W.; Yang, J.; Guo, W.; 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]
  31. Guo, W.C.; Yang, J.D. Stability performance for primary frequency regulation of hydro-turbine governing system with surge tank. Appl. Math. Model. 2017, 54, 446–466. [Google Scholar] [CrossRef]
  32. Hou, J.; Li, C.; Tian, Z.; Xu, Y.; Lai, X.; Zhang, N.; Zheng, T.; Wu, W. Multi-Objective Optimization of Start-up Strategy for Pumped Storage Units. Energies 2018, 11, 1141. [Google Scholar] [CrossRef] [Green Version]
  33. Li, C.; Zhang, N.; Lai, X.; Zhou, J.; Xu, Y. Design of a fractional-order PID controller for a pumped storage unit using a gravitational search algorithm based on the Cauchy and Gaussian mutation. Inf. Sci. 2017, 396, 162–181. [Google Scholar] [CrossRef]
  34. Huang, S.; Chi, X.; Xu, Y.; Sun, Y. Dynamic characteristics of coupling shaft system in tufting machine based on the Riccati whole transfer matrix method. Proc. Inst. Mech. Eng. Part C J. Mech. Eng. Sci. 2016, 231, 356–371. [Google Scholar] [CrossRef]
  35. Yin, S.H. Vibration of a simple beam subjected to a moving sprung mass with initial velocity and constant acceleration. Int. J. Struct. Stab. Dyn. 2016, 16, 176–185. [Google Scholar] [CrossRef]
  36. Cui, Y.; Liu, Z.; Wang, Y.; Ye, J. Nonlinear Dynamic of a Geared Rotor System with Nonlinear Oil Film Force and Nonlinear Mesh Force. J. Vib. Acoust. 2012, 134, 041001. [Google Scholar] [CrossRef]
  37. Gao, H.; Song, W. Analysis on unbalance vibration and compensation for maglev flywheel. Int. J. Appl. Electromagn. Mech. 2016, 50, 177–187. [Google Scholar] [CrossRef]
  38. Zhang, L.; Ma, Z.; Wu, Q.; Wang, X. Vibration analysis of coupled bending-torsional rotor-bearing system for hydraulic generating set with rub-impact under electromagnetic excitation. Arch. Appl. Mech. 2016, 86, 1665–1679. [Google Scholar] [CrossRef]
  39. Wang, M.H. Study on Calculation of Transition Process of Long-Loaded Hydropower Station. Master’s Thesis, North China University of Water Resources and Electric Power, Zhengzhou, China, 2019. [Google Scholar]
  40. General Administration of Quality Supervision. Specification Installation of Hydraulic Turbine Generator Units; GB/T 8564-2003; China Standards Press: Beijing, China, 2004. (In Chinese) [Google Scholar]
Figure 1. The schematic diagram of the HGS.
Figure 1. The schematic diagram of the HGS.
Energies 14 07319 g001
Figure 2. The schematic diagram of the HGS transient nonlinear coupling model.
Figure 2. The schematic diagram of the HGS transient nonlinear coupling model.
Energies 14 07319 g002
Figure 3. The schematic diagram of MOC.
Figure 3. The schematic diagram of MOC.
Energies 14 07319 g003
Figure 4. The schematic diagram of boundary conditions for each node of the WDS.
Figure 4. The schematic diagram of boundary conditions for each node of the WDS.
Energies 14 07319 g004
Figure 5. The characteristics of hydro-turbine. (a) The flow rate; (b) the torque.
Figure 5. The characteristics of hydro-turbine. (a) The flow rate; (b) the torque.
Energies 14 07319 g005
Figure 6. The schematic diagram of the PID governor.
Figure 6. The schematic diagram of the PID governor.
Energies 14 07319 g006
Figure 7. The schematic diagram of the HGSS.
Figure 7. The schematic diagram of the HGSS.
Energies 14 07319 g007
Figure 8. Vibration characteristics of different guide bearings and the flange. (a) The upper guide bearing; (b) the lower guide bearing; (c) the flange; (d) the water guide bearing.
Figure 8. Vibration characteristics of different guide bearings and the flange. (a) The upper guide bearing; (b) the lower guide bearing; (c) the flange; (d) the water guide bearing.
Energies 14 07319 g008
Figure 9. The operation characteristics of the hydro-turbine runner under different sudden load increasing amounts. (a) Guide vane opening; (b) rotating speed; (c) water head.
Figure 9. The operation characteristics of the hydro-turbine runner under different sudden load increasing amounts. (a) Guide vane opening; (b) rotating speed; (c) water head.
Energies 14 07319 g009
Figure 10. The time-domain figure of the hydro-turbine runner under different sudden load increasing amounts. (ac) In x-direction; (df) in y-direction.
Figure 10. The time-domain figure of the hydro-turbine runner under different sudden load increasing amounts. (ac) In x-direction; (df) in y-direction.
Energies 14 07319 g010
Figure 11. The axis trajectory of the hydro-turbine runner under different sudden load increasing amounts. (a) ΔP = 30%; (b) ΔP = 40%; (c) ΔP = 50%.
Figure 11. The axis trajectory of the hydro-turbine runner under different sudden load increasing amounts. (a) ΔP = 30%; (b) ΔP = 40%; (c) ΔP = 50%.
Energies 14 07319 g011
Figure 12. The time-domain figure of the hydro-turbine runner under different blade exit flow angles. (a) In x-direction; (b) in y-direction.
Figure 12. The time-domain figure of the hydro-turbine runner under different blade exit flow angles. (a) In x-direction; (b) in y-direction.
Energies 14 07319 g012
Figure 13. The axis trajectory of the hydro-turbine runner under different blade exit flow angles. (a) β2 = 0.05 rad; (b) β2 = 0.1 rad; (c) β2 = 0.15 rad.
Figure 13. The axis trajectory of the hydro-turbine runner under different blade exit flow angles. (a) β2 = 0.05 rad; (b) β2 = 0.1 rad; (c) β2 = 0.15 rad.
Energies 14 07319 g013
Figure 14. The time-domain figure of and under different mass eccentricities. (a,b) Generator rotor; (c,d) hydro-turbine runner.
Figure 14. The time-domain figure of and under different mass eccentricities. (a,b) Generator rotor; (c,d) hydro-turbine runner.
Energies 14 07319 g014
Figure 15. The axis trajectory of the generator rotor and the hydro-turbine runner under different mass eccentricities. (ac) Generator rotor; (df) hydro-turbine runner.
Figure 15. The axis trajectory of the generator rotor and the hydro-turbine runner under different mass eccentricities. (ac) Generator rotor; (df) hydro-turbine runner.
Energies 14 07319 g015
Figure 16. The time-domain figure under different stiffnesses. (a,d) Upper guide bearing; (b,e) lower guide bearing; (c,f) water guide bearing.
Figure 16. The time-domain figure under different stiffnesses. (a,d) Upper guide bearing; (b,e) lower guide bearing; (c,f) water guide bearing.
Energies 14 07319 g016
Figure 17. The axis trajectory under different stiffnesses. (ac) Upper guide bearing; (df) lower guide bearing; (gi) water guide bearing.
Figure 17. The axis trajectory under different stiffnesses. (ac) Upper guide bearing; (df) lower guide bearing; (gi) water guide bearing.
Energies 14 07319 g017
Figure 18. The time-domain figure of the generator rotor under different excitation currents. (a) In x-direction; (b) in y-direction.
Figure 18. The time-domain figure of the generator rotor under different excitation currents. (a) In x-direction; (b) in y-direction.
Energies 14 07319 g018
Figure 19. The axis trajectory of the generator rotor under different excitation currents. (a) 50% Ij; (b) 75% Ij; (c) 100% Ij.
Figure 19. The axis trajectory of the generator rotor under different excitation currents. (a) 50% Ij; (b) 75% Ij; (c) 100% Ij.
Energies 14 07319 g019
Figure 20. The time-domain figure of the generator rotor under different average air gaps between the stator and rotor. (a) In x-direction; (b) in y-direction.
Figure 20. The time-domain figure of the generator rotor under different average air gaps between the stator and rotor. (a) In x-direction; (b) in y-direction.
Energies 14 07319 g020
Figure 21. The axis trajectory of the generator rotor under different average air gaps between the stator and rotor. (a) δ0 = 40 mm; (b) δ0 = 30 mm; (c) δ0 = 20 mm.
Figure 21. The axis trajectory of the generator rotor under different average air gaps between the stator and rotor. (a) δ0 = 40 mm; (b) δ0 = 30 mm; (c) δ0 = 20 mm.
Energies 14 07319 g021
Table 1. The HGS parameters of Qinghai province in China.
Table 1. The HGS parameters of Qinghai province in China.
Model ParameterSymbolValueUnit
Length of the tunnelTL5693m
Length of the surge tankSTL2.3m
Length of the penstock behind the surge tankPBSTL84m
Length of the sloping penstockSPL58.97m
Length of the penstock behind the sloping penstockPBSPL562.44m
Length of pipe 1/2/3PL1/PL2/PL3360.6m
Diameter of the tunnelTD3.2m
Diameter of the surge tankSTD6m
Diameter of the surge tank impedanceSTID2.3m
Diameter of the penstock behind the surge tankPBSTD3.2m
Diameter of the sloping penstockSPD3m
Diameter of the penstock behind the sloping penstockPBSPD3m
Diameter of pipe 1/2/3PD1/PD2/PD31.73m
Upstream water levelHstart3456m
Downstream water levelHend3252.05m
Rated outputPr11MW
Rated flow ratenr600rpm
Rated water headQr6.8m3/s
Runner diameterD11.4m
Generator rotor diameterD23.144m
Guide bearing stiffnessK1 × 108N/m
Generator rotor mass eccentricitye11 × 10−4m
Runner mass eccentricitye21 × 10−4m
Uniform air gap thicknessδ02 × 10−2m
Air permeabilityμ04π × 10−7H/m
Table 2. The comparison of the simulation results and the experiment results.
Table 2. The comparison of the simulation results and the experiment results.
ResultUpper and Lower Water Level (m)HGSSThe Maximum Pressure Head at the Volute (m)The Minimum Pressure Head at the Volute (m)The Maximum Speed Rise Rate (%)The Maximum and Minimum Surge Level in Surge Tank (m)
The simulation3456.00
3252.05
#1252.75180.0158.983471.7
3453.0
#2252.76180.0358.96
#3252.74180.0158.95
The experiment3456.00
3252.05
#1243.45185.1859.073476.29
3442.43
#2243.12184.9258.97
#3243.36184.9459.03
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Guo, Y.; Liang, X.; Niu, Z.; Cao, Z.; Lei, L.; Xiong, H.; Chen, D. Vibration Characteristics of a Hydroelectric Generating System with Different Hydraulic-Mechanical-Electric Parameters in a Sudden Load Increasing Process. Energies 2021, 14, 7319. https://doi.org/10.3390/en14217319

AMA Style

Guo Y, Liang X, Niu Z, Cao Z, Lei L, Xiong H, Chen D. Vibration Characteristics of a Hydroelectric Generating System with Different Hydraulic-Mechanical-Electric Parameters in a Sudden Load Increasing Process. Energies. 2021; 14(21):7319. https://doi.org/10.3390/en14217319

Chicago/Turabian Style

Guo, Yixuan, Xiao Liang, Ziyu Niu, Zezhou Cao, Liuwei Lei, Hualin Xiong, and Diyi Chen. 2021. "Vibration Characteristics of a Hydroelectric Generating System with Different Hydraulic-Mechanical-Electric Parameters in a Sudden Load Increasing Process" Energies 14, no. 21: 7319. https://doi.org/10.3390/en14217319

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