You are on page 1of 15

Parameters affecting water hammer in a high-head

hydropower plant with Pelton turbines


U Karadžić1, A Bergant2, P Vukoslavčević1
1 Faculty of Mechanical Engineering, 81000 Podgorica, Montenegro
2 Litostroj E.I. d.o.o., 1000 Ljubljana, Slovenia

ABSTRACT

This paper investigates water hammer effects in a high-head hydropower plant Perućica,
Montenegro. During its first phase of modernisation and revitalisation new distributors
(needle valves) have been installed on the first two Pelton turbine units. Numerical results
using the standard quasi-steady friction model and the convolution based unsteady friction
model for different needle closing laws are compared with results of measurements. The
numerical model with the actual two-speed closing law gives the best match between
computed and measured results for the case of emergency shut-down of the turbine unit.
Inclusion of unsteady friction effects into the numerical models slightly improves
numerical results.

1 NOTATION

The following symbols are used in this kair fluid (air) damping coefficient
paper: L pipe length, length
A pipe area Mair fluid (air) damping torque
Am nozzle area Mfr shaft bearing friction torque
a pressure wave speed Mh hydraulic torque
bov width of the overflow weir Mr rated torque
D pipe diameter, diameter m dimensionless torque (m = M/Mr)
Db shaft bearing diameter mk, nk exponential sum coefficients
Dk wheel diameter N number of computational reaches
Dz spherical valve diameter Nk number of exponential terms
dm nozzle diameter n turbine rotational speed
e pipe wall thickness P power
Fh jet hydraulic force Q discharge
f Darcy-Weisbach friction factor Qm discharge to the turbine wheel
g gravitational acceleration Qu nozzle discharge
H piezometric head (head) Rb resultant force in the bearing
Hd head downstream the nozzle Re Reynolds number (Re = VD/ν )
Hu head upstream the nozzle s needle stroke
J polar moment of inertia Ta mechanical starting time
K constant equal to 4ν /D2 t time
KQ nozzle discharge coefficient tc needle closing time
to needle opening time ω angular velocity
tdef jet deflector operating time ζ in surge tank inflow coefficient
u peripheral velocity ζ out surge tank outflow coefficient
V average flow velocity
Vm jet velocity Subscripts:
v flow velocity def deflector
x distance along the pipe max maximum
yk weighting function component q quasi-steady part
z water level, elevation R reservoir
β momentum correction factor r rated conditions
∆t time step T tunnel
∆x reach length u unsteady part
δ absolute pipe roughness 0 initial or steady state conditions
ϕ relative speed change I,II,III penstock number
ρ water mass density
µ b shaft bearing friction coefficient Abbreviations:
µ ov overflow discharge coefficient QSF quasi-steady friction
CBM convolution-based model
ν kinematic viscosity
HPP hydropower plant
τ dimensionless nozzle opening MOC method of characteristics

2 INTRODUCTION

Planning of new or refurbishment and modernisation of existing hydropower plants (HPPs)


requires detailed water hammer analysis in order to get maximum and minimum pressures
as the most important parameters in the design process of the plant components. Water
hammer induces pressure rise or drop in hydraulic systems, rotational speed variation in
hydraulic turbomachinery (pumps and water turbines) and level fluctuation in surge tanks
and air chambers (1). In hydropower plants water hammer is caused by turbine load
acceptance and reduction, load rejection under governor control, emergency shut-down and
unwanted runaway, and closure and opening of the safety shutoff valve. Unsteady pipe
flow equations are used to describe water hammer phenomena. Boundary conditions for
reaction type water turbines (Francis, Kaplan, bulb) are well defined in the literature (2),
(3), (4). A little is published about detailed modelling of impulse type Pelton turbine (5),
(6). This turbine is usually represented as an end-valve boundary condition.

This paper investigates water hammer effects on operation of high-head Perućica HPP,
Montenegro. In the first part of the paper mathematical tools for solving water hammer
equations are presented. Water hammer is fully described by two hyperbolic partial
differential equations, the equation of continuity and the momentum equation (3), (4).
These equations are solved by the method of characteristics (MOC). Friction losses in
closed conduits are usually estimated by the quasi-steady friction model. In addition, a
convolution based unsteady friction model (7) can be used for calculation of friction factor.
Different closing laws for the distributor (needle valve) of the Pelton turbine are presented.
A novel model for calculation of Pelton turbine rotational speed change during emergency
shut-down is presented in detail. In the second part of the paper comparisons of numerical
and experimental results are made for turbine load rejection cases (load rejection under
governor control, emergency shut-down) from different initial powers. It is shown that
numerical model with actual two-speed closing law gives the best fit with results of
measurements for the case of emergency shut-down of the turbine unit. It is also shown that
unsteady friction effects have small impact on water hammer events in the Perućica HPP
flow-passage system.

3 THEORETICAL MODELLING

Water hammer equations are applied for calculation of the liquid unsteady pipe flow. A
simplified form of the equations neglecting the convective terms is used for most engineering
applications (3), (4),

∂H 2
∂Q
+ a =0 (1)
∂t gA ∂x

∂H 1 ∂Q f  Q ∣Q∣
+ +
∂x gA ∂ t 2 gDA 2
(2)

where H = piezometric head (head), Q = discharge, a = pressure wave speed, D = pipe


diameter, A = pipe area, g = gravitational acceleration, f = Darcy-Weisbach friction factor,
x = distance, and t = time. The staggered (diamond) grid (4) in applying the method of
characteristics (MOC) is used in this paper. At a boundary (reservoir, Pelton turbine), a
device-specific equation replaces one of the MOC water hammer compatibility equations.

3.1 Modelling friction losses

Friction losses in penstocks of hydropower plants are traditionally estimated by quasi-


steady friction model. This approach does not give good results for fast transients when
numerical results are compared with experiments (8). The friction factor f can be expressed
as the sum of the quasi-steady part fq and the unsteady part fu,

f = f q + fu
(3)

The quasi-steady friction factor is dependent on the Reynolds number and relative pipe
roughness and it is updated every time step. The explicit Halland equation (9) is used in this
paper,
2
1 
 6.9  δ 1.11 
= −1.8 log  +  
fq  
 Re  3.7 D  

(4)

where Re = Reynolds number (Re = VD/ν ), V = average flow velocity, ν = kinematic


viscosity, and δ = absolute pipe roughness. For evaluating of the unsteady friction factor a
convolution based model (CBM) (7) is used in this paper. The unsteady friction factor is
expressed as a finite sum of Nk functions yk(t) (10),

32νA Nk
fu =
DQ Q
∑y (t )
k =1
k

(5)

with

{
yk ( t + 2∆t ) = e − nk K∆t e − nk K∆t yk ( t ) + mk [ Q( t + 2∆t ) − Q( t ) ] }
(6)

where Nk = number of exponential terms (Nk,max = 10), ∆ t = time step, and K = constant equal
to 4ν /D2. Coefficients mk and nk have been developed for Zielke's (7) and Vardy-Brown's
(11), (12) weighting functions and can be found in Vítkovský et al. (10). In addition, a
momentum correction factor β , defined by Eq. (7), is incorporated into the MOC solution
when CBM model is used (13),

βAV 2
= ∫ v 2 dA
A

(7)

where v = flow velocity.

3.2 Pelton turbine boundary condition

3.2.1 Pelton turbine distributor (needle valve)


Pelton turbine distributor (needle valve) is utilized for regulation of discharge and
consequently for regulation of the turbine output. Discharge is controlled by closing or
opening the mouth of the nozzle by means of a needle (Fig. 1) and with appropriate
position of the jet deflector. The discharge through the nozzle is only dependent on the
position of the needle and it is not dependent on the turbine speed (14). Therefore, the water
hammer equations and the dynamic equation of the unit rotating parts can be solved
separately. In this way the head and discharge through the nozzle during the transient event
are calculated by the MOC and these values are used as input in the solution method for the
dynamic equation of the unit rotating parts.

Fig. 1 Pelton turbine distributor (needle valve)

The discharge through the nozzle is defined by the following equation,


(Qu )t = K Q Am 2 g ( H u ,t −H d )
(8)

where KQ = nozzle discharge coefficient, Am = nozzle area (Am = π dm2/4), Hu,t = head
upstream the nozzle, and Hd = const. = head downstream the nozzle. Typical functional
dependency of the discharge coefficient KQ and the ratio of needle stroke s and nozzle
diameter dm is shown in Fig. 2.

The needle closing law is expressed as follows,

s = τ ⋅ smax (9)

where τ = dimensionless nozzle opening, and smax = maximum needle stroke.

1.0
0.8
KQ / (KQ)max (-)

0.6
0.4
0.2
0.0
0.0 0.2 0.4 0.6 0.8 1.0
s/dm / (s/dm)max (-)

Fig. 2 Typical discharge coefficient of Pelton turbine nozzle

3.2.2 Rotational speed change during emergency shut-down of the unit


The equation that describes dynamic behaviour of the Pelton turbine unit rotating parts
during emergency shut-down is,


J = M h − ( M fr + M air ) (10)
dt

where J = polar moment of inertia of the turbine unit rotating parts, ω = angular velocity,
Mh = hydraulic torque, Mfr = shaft bearing friction torque, and Mair = fluid damping torque
(ventilation losses in the turbine housing). The turbine is disconnected from the electrical
grid followed by simultaneous gradual full-closure of the needle(s) and rapid activation of
the jet deflector(s) (deflection of the jet from the wheel).

Let us introduce the relative speed change ϕ ,

n − nr n
ϕ= = −1 (11)
nr nr
and the mechanical starting time Ta (2), (3),

π Jn r
Ta = (12)
30 M r

into Eq. (10), after rearrangement it follows,


Ta = mh − m fr − mair (13)
dt

where n = turbine rotational speed (traditionally in rpm), r = rated, Mr = rated torque, and m
= dimensionless torque (m = M/Mr).

The dimensionless hydraulic torque is expressed as follows,

Mh 1 D
mh = = Fh k (14)
Mr Mr 2

where Fh = jet hydraulic force (15),

Fh = 2 ρQm (Vm − u ) (15)

u = peripheral velocity,

Dk πnr
u= (ϕ + 1)
60
(16)

Dk = wheel diameter, Qm = discharge to the turbine wheel (Qm ≡ Qu), and Vm = jet velocity
(Vm = Qm / Am).

The dimensionless shaft bearing friction torque is,

1 D
m fr = µb b ( R Ab + RBb ) (17)
Mr 2

where Db = shaft bearing diameter, and µ b = shaft bearing friction coefficient. The resultant
forces in the shaft bearings RAb, RBb of the horizontal-shaft unit are due to hydraulic force,
weight of the wheels, weight of the shaft and weight of the generator.

The dimensionless fluid damping torque is expressed by,

1
k air nr2 (ϕ +1)
2
mair =
Mr
(18)
where kair = fluid (air) damping coefficient.

The hydraulic torque affects the turbine wheel until the jet deflector deflects all the water
into the tailrace (t=tdef). During t ≤ tdef the hydraulic torque is much greater than the
dissipation torques. Neglecting the dissipation torques in Eq. (13) gives the following
solution for the speed change during t ≤ tdef,

 60Qu   D 2πn ρQ t 
ϕ =  − 1 1 − exp − k r u 
 Dk Amπnr   60 M rTa 
(19)

At t > tdef the hydraulic torque is set to zero and the dissipation torques mfr and mair are
considered in Eq. (13). Now, the relative speed is given by,

µb Db RAb  n  
ϕ= tan  r µb Db RAb kair (t − t ) + atan 

nr2 kair
(ϕdef + 1)  − 1
 µb Db RAb
2 def
nr kair  Ta M r 
(20)

where ϕ def = relative speed at t = tdef.

The solution method that describes dynamic behaviour of the Pelton turbine unit rotating
parts during load rejection under governor control has been developed in a similar way. The
turbine is disconnected from the electrical grid followed by simultaneous gradual closure of
the needle(s) to the speed-no load position and controlled manouvre of the jet deflector(s)
i.e. rapid activation at the first instant followed by gradual adjustment of the deflector to the
speed-no load position.

4 PERUĆICA HPP FLOW-PASSAGE SYSTEM

Perućica HPP is a complex system comprised of a concrete tunnel (LT = 3335 m, DT =


4.8m), surge tank and three parallel steel penstocks with horizontal-shaft Pelton turbines
built at their downstream ends (Fig. 3). The length of each penstock is about 2000 m
whereby penstock I feeds two turbine units (A1 and A2) with rated unit power of 39 MW,
penstock II feeds three turbine units (A3, A4 and A5) of 39 MW each and penstock III
feeds two units (A6 and A7) of 59 MW each. A new turbine unit (A8) with a rated power of
59 MW is to be installed in the near future. The maximum water level at the intake is 613 m
and the minimum one is 602.5 m.
Fig. 3 Layout of Perućica HPP, Montenegro

The surge tank is of cylindrical type with an expansion and overflow (Fig. 4). Width of the
overflow weir is bov = 7.98m and the discharge coefficient is µ ov = 0.4. At the surge tank
intake there is a non-symmetrical orifice with head loss coefficients ζ in = 1.65 and ζ out =
2.48 during inflow and outflow, respectively.

Table 1 shows geometrical characteristics of the three penstocks (Fig. 3).

Table 1. Geometrical characteristics of penstocks


Penstock I Penstock II Penstock III
Pipe Pipe wall Pipe Pipe wall Pipe Pipe wall
Section Lenght L
diameter D thickness e diameter D thickness e diameter D thickness e
(m)
(mm) (mm) (mm) (mm) (mm) (mm)
T1-T2 75.0 2200 10 2200 10 2650 12
T2-T3 61.0 2200 10 2200 10 2650 12
T3-T4 330.5 2200-2100 10 2200 10-16 2650 12-13
T4-T5 318.0 2100-2000 16-25 2200 17-23 2650 13-21
T5-T6 123.0 2000 26-29 2200 24-27 2650 21-24
T6-T7 672.0 1900 27.5 2200-2100 26 2500 23
T7-T8 238.0 1800 27-39 2100 26-34 2500 24-29
I: 53.0
T8-T9 II: 99.8 1800 39 2100 34 2500 29
III: 146.6
Fig. 4 Perućica HPP surge tank

Basic characteristics of the Pelton turbine units, built at the downstream end of the
penstocks, are presented in Table 2.

Table 2. Characteristics of Pelton turbine units


Rated unit power Rated net head Rated speed
Turbine unit
Pr (MW) Hr (m) n (min-1)
A1,A2,A3,A4 39 526 375
A5 39 526 375
A6,A7 59 526 428
The polar moment of
Number of runners per Number of needles per
inertia of the unit
Turbine unit turbine unit turbine runner
rotating parts J (tm2)
A1,A2,A3,A4 2 168.75 1
A5 2 168.75 1
A6,A7 2 200 2
Closing time of the Opening time of the
Stroke of the needle
Turbine unit needle needle
smax (mm)
tc (s) to (s)
A1,A2,A3,A4 150 85 30
A5 195 80 30
A6,A7 166 80 50

The runner diameter of turbine units A1 to A5 is Dk= 2400 mm and for turbine units A6
and A7 is Dk = 2100 mm. The turbine inlet spherical valves diameters are Dz = 1000 mm
and Dz = 1200 mm, respectively. The valves are equipped with a passive actuator
comprised of a hydraulic servomotor. These valves are fully opened during turbine load
rejection tests.

The following quantities have been continuously measured during transient regimes:
pressure at the upstream and the downstream end of the turbine inlet valve (spherical
valve), stroke of the needle, stroke of the jet deflector and turbine rotational speed.
Pressures at the upstream and downstream end of the turbine inlet valve are the same. All
measurements have been carried out on turbine units A1 and A2. Absolute pressures have
been measured by high-pressure piezoelectric transducers Cerabar T PMP 131-A1101A70
Endress+Hauser (pressure range 0 to 100 bar, uncertainty in measurement ± 0.5 %). The
needle stroke and the stroke of the jet deflector have been measured by discplacement
transducers Balluff BTL5-S112-M0175-B-532 and Balluff BTL5-S112-M0275-B-532,
respectively. Uncertainty of these sensors is ± 0.03 mm. The turbine rotational speed has
been measured using inductive sensor Balluff BES M18MI-PSC50B-S04K (uncertainty in
measurement ± 0.03 %).

5 COMPARISONS OF NUMERICAL AND FIELD TEST RESULTS

Various operating regimes have been performed in the plant during commissioning of the
turbine units A1 and A2, including the unit start-up, load acceptance and reduction, load
rejection under governor control and emergency shut-down, and closure of turbine safety
valve against the discharge. Water hammer model includes intake, tunnel, surge tank and
three parallel penstocks with Pelton units at their downstream end (see Fig. 3). Numerical
results from standard quasi-steady friction model (QSF) and convolution based unsteady
friction model (CBM) for different needle's closing laws are compared with results of
measurements. The following results of measurements and corresponding numerical
simulations are presented:
1. Emergency shut-down of turbine unit A1 from initial power P0 = 39.5 MW (Test
A1P39.5MW)
2. Simultaneous load rejection under governor control of turbine units A1 and A2
from initial power of P0 = 42 MW each (Test A1&A2P42MW)
Tables below present the main initial parameters for these two tests. Flow in penstock I is
turbulent with a large Reynolds number.

Table 3. Initial discharges through tunnel and penstocks


Test QI (m3/s) ReI × 106 QII (m3/s) QIII (m3/s) QT (m3/s)
A1P39.5MW 8.4 5.5 0 22.0 30.4
A1&A2P42MW 18.6 12.1 0 6.4 25.0

Table 4. Steady friction factors and momentum correction factors


Penstock I Penstock II Penstock III Tunnel
Test
f0 β f0 β f0 β f0 β
A1P39.5MW 0.0126 1.0123 0.0125 1.0122 0.0126 1.0123 0.0146 1.0142
A1&A2P42MW 0.0125 1.0122 0.0125 1.0122 0.0129 1.0126 0.0146 1.0142

Table 5. Closure time, intake level and initial opening of the nozzle
Test zR (m) tc (s) s0 (mm) tdef (s)
A1P39.5MW 605.9 55.3 117 1.6
A1&A2P42MW 608.5 82.7 146 2.0

Pressure wave speeds are as follows, aT = 1345 m/s , aI = 1000 m/s, aII = 983 m/s and
aIII=1007 m/s.

5.1 Comparison of numerical and measured head at the turbine inlet

Transient head and discharge have been computed using a staggered grid MOC code. Basic
time step was ∆ t = 0.040 s. Computed and measured results are shown in Figs. 5 and 6.

Numerical and measured head at the turbine inlet (HI) for emergency shut-down of the unit
A1 are compared in Fig. 5 (Test A1P39.5MW). The computed and the measured total
needle closure time are the same (tc = 55.3 s - see Fig. 5a). The closure time is much larger
than the water hammer reflection time of 2LI/aI = 3.84 s. Maximum measured head of
557.7 m occurs when the nozzle is fully closed. Head rise for this case is 24.5 m. Maximum
calculated heads obtained by QSF are 555.9 m (one-speed closure) and 556.4 m (two-speed
closure; the cushioning stroke is 2.5 %), and by CBM are 556.4 m (one-speed closure) and
556.8 m (two-speed closure). All maximum calculated heads match the measured one.
Calculated and measured heads are much lower than the maximum permissible head of 602
m. Numerical results agree well with measured results during the needle closure period.
After this, a phase shift for all numerical models is evident from the third pressure pulse on.
A model with two-speed closure better attenuates pressure pulses compared to the model
with one-speed closure. It means that at the end of the closure the needle actually moves a
little bit slower (natural damping). Friction losses are described slightly better by the CBM
model (Fig. 5d). The Perućica flow-passage system is not unsteady friction dominant
system during water hammer events.

20 580
Measurement Measurement QSF
570
15 One-speed closure CBM
s / smax (%)

560
HI (m)

Two-speed closure
10 550
5 540
530
0 One-speed closure
520
45 50 55 60 0 20 40 60 80 100 120
a) Time (s) b) Time (s)
580 560
Measurement QSF Two-speed closure
570
CBM 550
560
HI (m)
HI (m)

550 540
540
530 QSF
530
Two-speed closure CBM
520 520
0 20 40 60 80 100 120 60 80 100 120 140 160
c) Time (s) d) Time (s)

Fig. 5 Comparison of needle stroke (s) and head at the turbine inlet (HI; datum level z
= 65.8 m; time step ∆ t=0.040 s). Emergency shut-down of A1 from P0 = 39.5 MW
Numerical and measured heads at the turbine inlet for simultaneous load rejection under
governor control of turbine units A1 and A2 are compared in Fig. 6. Test A1&A2P42MW
induces maximum measured head of 573.9 m with head rise of 57.0 m. The maximum
calculated heads obtained by QSF are 571.5 m (one-speed closure) and 572.7 m (two-speed
closure), and by CBM are 572.4 m (one-speed closure) and 571.9 m (two-speed closure). In
this case the needle is closed to its speed-no load position (3.8 %). Process is governed by
the turbine control system. Up to this time all numerical models show good agreement with
results of measurement and all maximum heads are well below the maximum permissible
head of 602 m. All models produce similar results after the closure period. In this case there
is no cushioning effect for the two-speed closure because the needle speed-no load position
is larger (3.8 %) than the cushioning one (2.1 %). Consequently, the two-speed closure is
actually the one-speed closure with slightly shorter closure time.

20 600
Measurement Measurement QSF
15 One-speed closure 580 CBM
s / smax (%)

Two-speed closure HI (m)


560
10
540
5 520
One-speed closure
0 500
70 75 80 85 90 95 100 0 20 40 60 80 100 120 140
a) Time (s) b) Time (s)
600 580
Measurement QSF 570 CBM
580 CBM
560
HI (m)

560
HI (m)

550
540 540
530
520 520
One-speed closure
Two-speed closure Two-speed closure
500 510
0 20 40 60 80 100 120 140 80 100 120 140 160
c) Time (s) d) Time (s)

Fig. 6 Comparison of needle stroke (s) and head at the turbine inlet (HI; datum level z
= 65.8 m; time step ∆ t=0.040 s). Simultaneous load rejection under governor control
of A1 and A2 from P0 = 42 MW

5.1.1 Statement on convergence and stability


Convergence relates to the behaviour of the solution as ∆ x and ∆ t approach zero, whereas
stability is concerned with the growth of round-off error. Analysis of computational results
obtained with different numbers of reaches (see Table 6) reveals that the magnitude and
timing of the pressure pulses predicted by QSF and CBM converge to practically the same
solution. As an example, Fig. 7 shows heads at the turbine inlet (HI) for the case of the
emergency shut-down of turbine unit A1 from initial power P0 = 39.5 MW (Test
A1P39.5MW). The QSF model (Fig. 7a) and the CBM model (Fig. 7b) results are
compared for two different time steps ∆ t = {0.040; 0.005} s.
Table 6. Number of reaches for tunnel and penstocks
N (∆ t=0.04 s) N (∆ t=0.02 s) N (∆ t=0.01 s) N (∆ t=0.005 s)
Tunnel 62 124 248 496
Penstock I 48 96 192 384
Penstock II 50 100 200 400
Penstock III 50 100 200 400

580 580
QSF: ∆ t =0.040 s CBM: ∆ t =0.040 s
570 QSF: ∆ t =0.005 s
570 CBM: ∆ t =0.005 s

HI (m)
560 560
HI (m)

550 550
540 540
530 Two-speed closure 530 Two-speed closure
520 520
0 20 40 60 80 100 120 0 20 40 60 80 100 120
a) b)
Time (s) Time (s)

Fig. 7 Comparison of head at the turbine inlet (HI; datum level z = 65.8 m) for
different time steps. Emergency shut-down of A1 from P0 = 39.5 MW

5.2 Comparison of computed and measured turbine rotational speed

Turbine rotational speed during (1) emergency shut-down of turbine unit A1 from initial
power P0 = 39.5 MW (Test A1P39.5MW) and (2) during simultaneous load rejection under
governor control of turbine units A1 and A2 from initial power of P0 = 42 MW each (Test
A1&A2P42MW) has been calculated using adequate solution method of the dynamic
equation of the unit rotating parts (Eq. (10)). The input head and discharge through the
nozzle during the transient event have been previously calculated by the MOC (see Section
5.1). Fig. 8 shows comparison between computed and measured rotational speed for both
case studies.

120 120
110 Measurement Measurement
n / n0 (%)

Computation 110 Computation


n / n0 (%)

100
90 100
80
90
70
60 80
0 20 40 60 80 100 120 140 160 0 20 40 60 80 100 120 140 160
Time (s) Time (s)
a) Emergency shut-down of A1 from P = 39.5 MW b) Load rejection of A1&A2 from P = 42 MW

Fig. 8 Rotational speed change (n0 = 375 min-1)

The maximum measured turbine speed rise for Test A1P39.5MW of 8.1 % occurs at time t
= tdef. The maximum computed turbine speed rise of 8.0 % agrees well with measured one
(Fig. 8a). After jet deflector deflects all the water into the tailrace, the computed turbine
speed decrease reasonably agree with measured one. There is a good agreement between
the maximum measured and calculated turbine speed rise for Test A1&A2P42MW of 11.2
% and 11.1 %, respectively (Fig. 8b). There are some discrepancies in the phase of speed
decrease due to complex flow behaviour in the turbine housing. In both investigated cases
the maximum speed rise is well below the permissible speed rise of 25%.

6 CONCLUSIONS

Computed results using the standard quasi-steady friction model (QSF) and the convolution
based unsteady friction model (CBM) for different distributor (needle valve) closing laws
are compared with results of measurements performed in Perućica HPP, Montenegro. Both
the QSF model and the CBM model produce practically the same results. The CBM model
only slightly better captures measured data during the decay period of the transient event.
The effects of unsteady friction on water hammer events in Perućica flow-passage system
are indeed small; however, they might have a strong influence on behaviour close to
resonance and this is a subject of authors’ future studies. The two-speed closing law
considers natural damping effect in the hydraulic servomotor close to its fully closed
position. That is why the computed results using the two-speed closing law for the case of
the unit emergency shut-down fit better the results of measurement compared with the
computed results using the one-speed i.e. linear closing law. For the case of load rejection
under governor control the needle is closed to its speed-no load position; then the two-
speed closing law is actually one-speed law with slightly shorter closure time. The turbine
rotational speed change is calculated separately by a novel model of Pelton turbine
rotational speed change. There is a reasonable agreement between the computed and
measured results.

7 ACKNOWLEDGMENTS

The authors wish to thank ARRS (Slovenian Research Agency) for their generous support of
research on fluid transients. The support of research by ZAMTES (Montenegrian Bureau for
International Scientific, Educational, Cultural and Technical Cooperation) and The Ministry
of Education and Science of Montenegro are gratefully acknowledged as well.

8 REFERENCES

(1) Pejović, S., Boldy, A.P., Obradović, D. (1987). Guidelines to hydraulic transient
analysis, Gower Technical Press Ltd., Aldershot, United Kingdom.
(2) Krivchenko, G.I., Arshenevskij, N.N., Kvjatovskaja, E.V., Klabukov, V.M. (1975).
Gidromehanicheskie perehodnie processi v gidroenergeticheskih ustanovkah,
Energija, Moscow, Russia (in Russian).
(3) Chaudhry, M.H. (1987). Applied hydraulic transients, Van Nostrand Reinhold
Company, New York, USA.
(4) Wylie, E.B., Streeter, V.L. (1993). Fluid transients in systems, Prentice Hall,
Englewood Cliffs, USA.
(5) Fasol, K.H. (1964). Die Berücksichtigung des dynamischen Verhaltens der Pelton
düsen bei Druckstoβ berechnungen, ÖZE - Österreichische Zeitschrift für
Elektrizitätwirtschaft, 17(8), 453-456 (in German).
(6) Evangelisti, G., Boari, M., Guerrini, P., Rossi, R. (1973 & 1974). Some applications
of waterhammer analysis by the method of characteristics, L’Energia Elettrica, 50(1),
1-12 & 51(6), 309-324.
(7) Zielke, W. (1968). Frequency-dependent friction in transient pipe flow, Journal of
Basic Engineering, ASME, 90(1), 109-115.
(8) Bergant, A., Simpson, A.R., Vítkovský, J. (2001). Developments in unsteady pipe
flow friction modeling, Journal of Hydraulic Research, IAHR, 39(3), 249-257.
(9) Haaland, S.E. (1983). Simple and explicit formulas for the friction factor in turbulent
pipe flow, Journal of Fluids Engineering, ASME, 105(3), 89-90.
(10) Vítkovský, J., Stephens, M., Bergant, A., Lambert, M., Simpson, A. (2004). Efficient
and accurate calculation of Zielke and Vardy-Brown unsteady friction in pipe
transients, Proceedings of the 9th International Conference on Pressure Surges, BHR
Group, Chester, UK, 15 pp.
(11) Vardy, A.E., Brown, J.M.B. (2003). Transient turbulent friction in smooth pipe flows,
Journal of Sound and Vibration, 259(5), 1011-1036.
(12) Vardy, A.E., Brown, J.M.B. (2004). Transient turbulent friction in fully rough pipe
flows, Journal of Sound and Vibration, 270(1-2), 233-257.
(13) Bergant, A., Karadžić, U., Vítkovský, J., Vušanović, I., Simpson, A.R. (2005). A
discrete gas-cavity model that considers the frictional effects of unsteady pipe flow,
Strojniški Vestnik – Journal of Mechanical Engineering, 51(11), 692-710.
(14) Benišek, M. (1998). Hidraulichne turbine, Faculty of Mechanical Engineering,
Belgrade, Serbia, (in Serbian).
(15) Nechleba, M. (1957). Hydraulic turbines – their design and equipment, Artia, Prague,
Czech Republic.
(16) Edelj, J.U. (1963). Kovshovie gidroturbini – teorija, issledovanie, raschet, Mashgiz,
Moscow, Russia (in Russian).

You might also like