You are on page 1of 23

Accepted Manuscript

Friction Loss and Energy Recovery of a Pelton Turbine for Different Spear
Positions

Heungsu Jeon, Joo Hoon Park, Youhwan Shin, Minsuk Choi

PII: S0960-1481(18)30178-2

DOI: 10.1016/j.renene.2018.02.038

Reference: RENE 9769

To appear in: Renewable Energy

Received Date: 04 February 2017

Revised Date: 29 January 2018

Accepted Date: 06 February 2018

Please cite this article as: Heungsu Jeon, Joo Hoon Park, Youhwan Shin, Minsuk Choi, Friction
Loss and Energy Recovery of a Pelton Turbine for Different Spear Positions, Renewable Energy
(2018), doi: 10.1016/j.renene.2018.02.038

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form.
Please note that during the production process errors may be discovered which could affect the
content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT
1

1 Friction Loss and Energy Recovery of a Pelton Turbine for

2 Different Spear Positions

4 Heungsu Jeon1, Joo Hoon Park2, Youhwan Shin2,* and Minsuk Choi1,*

6 1 Department of Mechanical Engineering, Myongji University, Yongin 17058, South Korea

7 2Center for Urban Energy Research, Korea Institute of Science and Technology, Seoul 02792, South Korea

8 * Corresponding Author: Y. Shin (yhshin@kist.re.kr) & M. Choi (mchoi@mju.ac.kr)

10 Abstract

11 This study is dedicated to find a cause of a critical flow rate in a Pelton turbine operating with a constant

12 runner speed, below which the efficiency of the turbine decreases significantly. A critical flow rate was initially

13 found in the performance test of the Pelton turbine for extracting energy from a PRO (pressure retarded osmosis)

14 pilot plant. For higher flow rates than a critical value, the efficiency of the Pelton turbine was nearly constant

15 independent of flow rates. For lower flow rates than a critical value, however, the efficiency drops with

16 decreasing flow rates. 3D flow simulations were conducted at three different flow rates to investigate effects of

17 flow rates on the performance of the Pelton turbine. It was found in the numerical results that a large friction

18 loss is generated in an injector if the spear is closed too tightly for a low flow rate below a critical value. Head

19 loss coefficients of the injector for three different spear positions were calculated and it was found that the loss

20 is doubled below a critical flow rate. This implies that it is important to include the geometry of an injector and

21 spear in the numerical simulations for Pelton turbines.

22 Keywords: Pelton turbine, Injector, Friction loss, Head loss coefficient, PRO

23

24 1. Introduction

25

26 For a country with energy shortages, it is very important to generate power from renewable energy resources

27 such as solar, hydro, marine and wind energies without any pollution. PRO is a power generation technique

28 based on osmotic energy, with which fresh water passes through a semipermeable membrane to sea water and
ACCEPTED MANUSCRIPT
2

1 increases pressure of sea water consequently. This pressurized water up to 30 bars can be used to generate

2 electricity by a proper energy recovery device.

3 A Pelton turbine is an impulse type turbine and is mainly composed of buckets and injectors. Momentum

4 transfer occurs between two parts when the jet from an injector impacts buckets. The size of Pelton turbines is

5 adjustable depending on power capacity so that the smallest turbine generates a few watts and the largest

6 hundreds of megawatts. A Pelton turbine is suitable for extracting energy from a small amount of high

7 pressurized water in PRO plants.

8 Some researchers have been focusing on the water jet from an injector although there are many research

9 topics related to Pelton turbines. Zhang and Casey [1] measured the velocity of the water jet from an injector

10 with LDV (Laser Doppler Velocimetry) method for analyzing effects of the distributor and bifurcations

11 upstream of the injectors on the water jet. They found that the weak secondary flow structure with two counter-

12 rotating vortices in the jet affects the shape and trajectory of the water jet. Benzon et al. [2] conducted a

13 parametric optimization of impulse turbine injectors using DOE (design of experiment) method and found that

14 larger injector and spear angles are beneficial to reduce losses inside the injectors. Jo et al. [3] investigated

15 effects of the throat angle of the nozzle without a spear and found that the increase of the angle results in the

16 reduction of the cross-sectional area of the flow passing through the nozzle and the increase in power loss due to

17 the recirculation region.

18 For the interaction between the water jet and buckets, a stationary configuration with a fixed bucket has been

19 initially adopted because of the difficulties in measurements and simulations with rotating buckets. In particular,

20 Zoppé et al.[4] conducted experiments and numerical simulations for a fixed bucket and nozzle in a Pelton

21 turbine and showed that the pressure distribution on the inner surface of a bucket could be accurately predicted

22 by their simulations. Although most of researches for Pelton turbines are now based on a real configuration with

23 a rotating turbine runner in an aid of developments in measuring techniques and numerical methods, numerical

24 simulations for Pelton turbines are still a challenging task due to many difficulties such as grid generation

25 around a bucket, long transient computing time and two-phase flow of air and water. Perrig et al. [5] measured

26 static pressure on the inner surface of a bucket and compared experimental data to their numerical results. After

27 validation, they described the interaction between the jet and the buckets in detail. Santolin et al. [6] numerically

28 tested two different jet configurations with and without a needle valve in an injector and analyzed the effect of

29 two jets on the performance of a Pelton turbine. They found that the turbine efficiency impinged by a real

30 perturbed jet is lower than by an ideal jet, highlighting the importance of implementing a real jet in numerical
ACCEPTED MANUSCRIPT
3

1 simulations. Stamatelos et al. [7] conducted performance tests of a Pelton turbine with two injectors and

2 estimated the mechanical loss of the Pelton turbine based on their measurements over a wide operating range.

3 Xiao et al. [8] performed numerical simulations for a prototype Pelton turbine with five different discharge

4 conditions and compared their results to experimental data. In their simulations, however, the water jet was

5 assumed to be axisymmetric with a constant velocity profile. Aggidis and Židonis [9] have a performance test

6 facility for Pelton turbines remodeled to generate performance curves automatically. Within 4 hours of testing,

7 they could obtain standard efficiency hill charts of Pelton and Turgo turbines. You et al. [10] calculated the

8 runaway process in a Pelton turbine with changing rotational speeds of the turbine runner. They found that the

9 increase in both the frequency of alternating force and the amplitude of centrifugal force during the runaway

10 process has a negative effect on the safety of the turbine. Benzon et al. [11] applied an optimized injector to

11 Pelton and Turgo turbine runners and obtained improved performance of two turbines. They also found that the

12 nonuniform flow pattern of the jet caused by the spear could generate additional losses in comparison to the

13 ideal uniform jet. Choi et al. [12] simulated the interaction between the water jet and the buckets in a small

14 Pelton turbine and found that the torque generated by a bucket is closely related to the water sheet on the inner

15 surface of the bucket. Židonis et al. [13] applied the FLS (fast Lagrangian solver) and generic algorithm to

16 obtain initially optimized turbine runners and thereafter used two commercial flow solvers to analyze the

17 characteristics of the flow for optimized geometries. Židonis and Aggidis [14] conducted a computational

18 parametric study to investigate the effect of number of buckets, bucket radial position and bucket angular

19 position on the Pelton turbine efficiency and validated their numerical results in their following experiments.

20 Gupta et al. [15] performed numerical simulations of a Pelton turbine with six injectors and investigate the

21 effects of time step, turbulence model and mesh size on the accuracy of their numerical results.

22 By the way, Marongiu et al. [16] applied the SPH (smoothed particle hydrodynamics) method to analyze the

23 jet-bucket interaction, where the water jet is implemented by a discrete number of particles without mesh.

24 Thereafter, they combined the SPH method to the inviscid Euler equation and obtained improved capacity in

25 predicting free-surface flows [17]. Anagnostopoulos and Papantonis [18] applied the FLS to a Pelton turbine

26 and predicted the performance of the turbine accurately. Rosetti et al. [19] also applied the Euler-Lagrangian

27 approach to the simulations for a Pelton turbine. They divided the bucket geometry into six discharging areas

28 and investigated contribution to the bucket total torque, highlighting a performance drop of the bottom part of

29 the bucket at a low flow rate. However, Židonis and Aggidis [20] pointed out in their review paper that

30 Lagrangian methods are attractive and promising for Pelton turbine simulations because of lower computational
ACCEPTED MANUSCRIPT
4

1 cost in comparison to Euler methods but still need to be improved in terms of accuracy.

2 Power generation turbines connected to the national power grid should have the same frequency as its value

3 regardless of operating conditions. With a bipolar induction generator, a rotational speed of the shaft between

4 the turbine and the generator should be kept 3,600 rpm to match the frequency of 60 Hz in South Korea. During

5 designing and testing a Pelton turbine for recovering energy from a PRO pilot plant, it was found in the

6 experiments that the efficiency of a Pelton turbine with various flow rates for the fixed rotational speed is nearly

7 constant over wide discharge conditions but decreases significantly below a critical flow rate. In spite of the

8 many references mentioned above, no work has been reported on this issue because the turbine performance has

9 been generally measured with various rotational speeds for a fixed flow rate. Therefore, this study is dedicated

10 to figure out why a Pelton turbine with a constant runner speed has a critical flow rate, below which the

11 efficiency of the turbine drops significantly.

12

13 2. Test configuration

14

15 2.1 Test facility for Pelton turbine

16 3D numerical flow simulations were conducted to analyze the internal flow in a Pelton turbine, which had

17 been designed to extract electrical energy from the high-pressurized water in a PRO pilot plant. Figures 1 and 2

18 show the shapes of a bucket and a spear-nozzle designed and manufactured by KIST (Korea Institute of Science

19 and Technology) and the assembled test facility to measure the performance of the turbine respectively. The

20 working fluid of the performance test facility is water, which circulates in a closed loop where it passes through

21 a high-pressure pump, accelerates through the spear nozzle, and returns to the water tank. There is a bypass line

22 in the middle of the pipe to control the flow rate while maintaining the pressure. A flow meter was installed in

23 the main pipe, and a pressure transducer and a pressure gauge were installed near the spear valve. A torque

24 transducer was attached between the Pelton turbine and the generator shaft. A proximity sensor was installed on

25 one side of the generator shaft, and the output power of the generator was measured using a power meter. The

26 system was configured to monitor and save all electrical signals from the flow meter, pressure transducer, torque

27 transducer and power meter in real time. The maximum power output from the generator is about 10 kW.

28 The Pelton turbine consists of 25 buckets and one injector due to a small amount of the water. A step motor

29 outside the injector can move the spear back and forth to adjust the volumetric flow rate. All parts of the test

30 facility were made of the rustproof stainless steel 316L. In the tests, the total pressure of the water and the
ACCEPTED MANUSCRIPT
5

1 rotational speed of the bucket runner were fixed to be 30 bar (Gauge) and 3,600 rpm respectively, and only the

2 spear position was changed to adjust the volumetric flow rate from 100 L/min to 300 L/min. For reference, the

3 design volumetric flow rate of the Pelton turbine is 277.8 L/min and the corresponding specific speed (ns) is

4 2.79, which was calculated by the following equation.


5/4
5 𝑛𝑠 = 0.2626 𝜔 𝑃/𝐻 (1)

6 where ω is the runner speed (rpm), P is the power at the design condition (kW) and H is the water head (m). The

7 factor 0.2626 is required when the specific speed is to be converted to English units. Referring to Munson et al.

8 [21], the impulse turbines are best at low specific speeds of ns < 10.

10

11

12 Fig. 1 Manufactured Pleton bucket and spear nozzle

13

14

15 Fig. 2 Assembled test facility with Pleton turbine

16

17
ACCEPTED MANUSCRIPT
6

1 2.2 Computational Domain and Mesh

2 The geometry of the test facility has been reconstructed for unsteady flow simulations as shown in Fig. 3. The

3 computational domain is mainly composed of three parts, namely, turbine runner and bucket domain (R1),

4 nozzle domain (N1) and casing domain (C1). The fluid region around a bucket was filled with an unstructured

5 mesh of tetras because of its complex shape. In particular, it is very difficult to generate hexahedral mesh around

6 the cut-out. In other two domains of N1 and C1, a multi-block hexahedral mesh was generated as shown in Fig.

7 4. For grid independency tests, as listed in Table 1, three grids with different resolutions have been generated

8 and tested in the simulations.

10

11 Fig. 3 Reconstructed computational domain and geometry

12

13

14

15 Fig. 4 Computational grids near injector and on bucket surface

16
ACCEPTED MANUSCRIPT
7

1 Table 1 Number of nodes for three different resolutions

Grid No. Bucket Casing Nozzle Total

1 1,457,550 582,734 18,810 2,059,094

2 2,027,025 1,027,504 42,336 3,096,865

3 2,243,674 1,592,564 92,862 3,929,101

3 2.3 Numerical methods and boundary conditions

4 In this study, the flow field inside the Pelton turbine has been calculated using a commercial flow solver,

5 ANSYS CFX-16. The three-dimensional Reynolds-averaged Navier-Stokes equations, including the continuity,

6 momentum, and energy equations, were solved with constant densities of air and water. Two-phase flow was

7 resolved by the VOF (Volume of Fluid) method, and the turbulent viscosity was calculated by the k-ε model. A

8 high-resolution scheme was used for the convection terms, while the spatial derivatives of the diffusion terms

9 were calculated by the shape functions based on the finite element method. For time derivatives, the second-

10 order implicit time marching was employed to obtain unsteady flow fields with a time step equal to 1/40 of a

11 bucket pitch.

12 The R1 domain has been solved in the rotating frame and other two domains, N1 and C1 domains, in the

13 stationary frame. As shown in Fig. 3, the interface between R1 and C1 was treated using the sliding plane,

14 where the fluxes in mass, momentum, and energy were conservative through the interface. The nozzle domain

15 was attached to the casing domain using the GGI (general grid interface), through which the connection is made

16 automatically between the mutually overlapping surfaces and all fluxes are preserved.

17 At the nozzle inlet in the simulation, the total pressure of 30 bar (Gauge) was given. At the outlet of the

18 casing domain, the ambient static pressure was specified and only air could pass through the boundary freely.

19 The rotational speed of the turbine runner was fixed to be 3,600 rpm and the position of the spear in the injectors

20 was changed to adjust volumetric flow rate. Other surfaces on the buckets, spear, injector, and casing were

21 treated as solid wall so that the no-slip condition was applied and the standard wall function was employed to

22 obtain the velocity on the first node from the wall. Since only half of the whole Pelton turbine was simulated for

23 minimizing the calculation time, the symmetric condition was applied.

24 During unsteady simulations, in order to determine whether a computational result reaches a periodic solution

25 or not, the torque time history has been checked. The torque time history showed the cyclic motions after 15

26 pitches from the start of an unsteady simulation, implying that the computational results reached a quasi-steady
ACCEPTED MANUSCRIPT
8

1 state.

3 3. Results and discussion

5 3.1. Grid independency test

6 A grid independency test has been performed for three grids with different grid resolutions listed in Table 1.

7 As shown in Fig. 5, the time-averaged torque during 1 pitch period of each grid has been compared with one

8 another for Q = 219.7 L/min. The torque of grid 1 is much smaller than the torques of the other two grids, but

9 the torque of grids 2 and 3 has a similar value. It seems that no significant change in torque is expected from a

10 further mesh refinement. Therefore, grid 2 has been selected for further analysis.

11

12

13 Fig. 5 Grid dependency of time-averaged torque

14

15 Figure 6 shows the interaction of the water jet to the buckets with grid 2 at the given flow rate, where the

16 water volume fraction (Vw) is equal to 0.85. When the water jet impinges the turbine runner, splashes of water

17 droplets are generated in the casing. It is confirmed that the water jet diffuses at the interface to make the jet

18 diameter decrease slightly. The cross-section of the water jet is perfectly circular because only straight part of

19 the injector is used in the simulation. The water jet moves in the straight line before impinging to a bucket, but

20 turns its direction with smaller velocity after the impingement. This figure shows that grid 2 has a sufficient

21 resolution to analyze the interaction between the water jet and the buckets and to predict the torque and power of

22 the Pelton turbine.

23
ACCEPTED MANUSCRIPT
9

2 Fig. 6 Iso-surface of water volume fraction at Q2 = 219.7 L/min (VW=0.85)

4 3.2 Effects of Volumetric Flow Rate

5 For the fixed rotational speed of the turbine runner and the inlet total pressure, only volumetric flow rate (Q)

6 has been changed by three different positions of the spear inside the injector, as shown in Fig. 7. The

7 corresponding flow rates for the given spear positions are listed in Table 2.

9 (a) Q1 (b) Q2 (c) Q3

10 Fig. 7 Three spear positions in injector

11

12 Table 2 Volumetric flow rates for different spear positions

No. Flow Rate (L/min)

Q1 288.4

Q2 219.7

Q3 120.1

13
ACCEPTED MANUSCRIPT
10

1 The time-averaged torque and efficiency have been predicted numerically for these three injector

2 configurations and compared with the experimentally measured data. For the given runner speed and inlet total

3 pressure, the torque on the shaft of the turbine was measured with a torque transducer in the test facility. Shaft

4 power is the product of torque and rotational speed and the measured value is in the range of 3kW and 11kW,

5 depending on the volumetric flow rate. As shown in Fig. 8, the volumetric flow rate and the torque have been

6 non-dimensionalized to the flow coefficient and the power coefficient respectively by the following equations.
Q
7 𝐶𝑄 = 3 (2)
𝜔𝑅

𝜏𝜔
8 𝐶𝑃 = 3 5 (3)
𝜌𝜔 𝑅

9 where all variables are metric. It seems that the power coefficient is linearly proportional to the flow coefficient

10 and the experimental data can be expressed by two linear lines having different slopes at Point A with CQ = 465.

11 The corresponding flow rate at Point A is 130 L/min, which could be a critical flow rate. The efficiency was

12 calculated by the following equation.


τω
13 η = 𝜌𝑔𝐻𝑄 (4)

14 The measured efficiency is nearly constant when the flow coefficient is larger than the critical value, while

15 the efficiency decreases suddenly below the critical value. The numerically predicted torque and efficiency

16 agree well with the experimental values. In particular, two efficiencies at Q1 and Q2 are similar to each other but

17 the efficiency at Q3 is smaller than other two values. Although there are some differences in absolute value

18 between experiments and simulations, the predicted trends in torque and efficiency are consistent with the

19 measured data. According to the previous researches of Santolin et al. [6] and Choi et al. [12], the difference

20 between the experimental and computational data might be caused by the interface between the turbine runner

21 domain (R1) and the casing domain (C1), through which the water jet diffuses slightly.

22
ACCEPTED MANUSCRIPT
11

2 (a) Power coefficient

4 (b) Efficiency

5 Fig. 8 Perofmrnace of Pelton turbine

7 Torque variation of a bucket by the water jet for three different flow rates can be extracted during unsteady

8 simulations and the results are shown in Fig. 9. Each bucket of the Pelton turbine is influenced by the water jet

9 from the injector during four pitch periods. The water jet impingement on each bucket starts at the same angular

10 position where the cut-out of a bucket is in contact with the jet outer boundary, and the torque increases

11 continuously and rapidly thereafter. The maximum torques are achieved at about 2 pitch periods after the

12 initiation of the water jet impingement, where the bucket is nearly vertical to the jet, and the torque decreases

13 quickly thereafter. As the volumetric flow rate increases, torque at each instant increases simultaneously.
ACCEPTED MANUSCRIPT
12

2 Fig. 9 Torque variation of a bucket by water jet

4 In the previous work, Choi et al. [12] investigated the relationship between the time-variation of the torque

5 and the wet area on the bucket inner surface. They found that torque of a bucket has a maximum value when the

6 water sheet fills the inner surface completely and the water jet is vertical to the splitter. Figure 10 shows the wet

7 area of the inner surface of the bucket. The water impingement on a bucket starts at the cut-out in the upward-

8 diagonal direction, from the viewpoint of the bucket. As the turbine runner rotates, the incidence angle of the

9 water jet is increasing and the impinging point on the bucket surface is moving towards the center. When the

10 bucket is vertical to the jet, the water jet is divided by the splitter evenly and the water sheet is moving in the

11 horizontal direction, where the torque of a bucket reaches the maximum value. After the vertical position, the

12 incidence angle increases further, the impinging point moves back towards the cut-out and the torque decreases

13 gradually. At 2 pitch periods, the torque of each bucket has the maximum value and the wet area on the inner

14 surface reaches the maximum area. For two higher flow rates, the pattern of the wet area at position 2 is the

15 same as each other, although the wet area for Q2 is smaller than the area for Q1 at position 1. For the lowest flow

16 rate of Q3, the wet area is much smaller than other two cases at all instants.
ACCEPTED MANUSCRIPT
13

3 Fig. 10 Water volume fraction on bucket inner surface

5 Figure 11 shows the water jet from the injector for analyzing effects of the flow rate on the jet thickness on

6 the symmetric plane, where the boundary of the water is set to be the region of Vw>0.5. The white arrows

7 indicate the thickness of the water jet in each case and especially the thickness for Q1 is presented by the white

8 dotted line on all images for comparison to one another. The water jet becomes thinner as the volumetric flow

9 rate decreases. As shown in Fig. 11 (a) and (b), however, the amount of the change in the thickness is relatively

10 small above the critical flow rate of 130 L/min in spite of different flow rates. Below the critical flow rate, the

11 thickness of the water jet decreases significantly.

12

13

14 (a) Q1 (b) Q2 (c) Q3

15 Fig. 11 Water jet thickness for three different flow rates

16
ACCEPTED MANUSCRIPT
14

1 The effects of volumetric flow rates on the performance of a Pelton turbine might be explained theoretically

2 by the following equation, which is derived by applying Euler turbine equation to a stationary Pelton turbine

3 [21]. Although these equations assumed that the water jet always impinges the bucket center in the vertical

4 direction and leaves after turning of angle β, some physical interpretation is still possible based on them.

5 τ =‒ 𝑚𝑅(𝑈 ‒ 𝑉𝑗)(1 ‒ 𝑐𝑜𝑠𝛽) =‒ 𝜌𝑄𝑅(𝑈 ‒ 𝑉𝑗)(1 ‒ 𝑐𝑜𝑠𝛽) (5)

𝑈
6 η =‒ 𝑔𝐻(1 ‒ 𝑐𝑜𝑠𝛽)(𝑈 ‒ 𝑉𝑗) (6)

7 Here, the torque is proportional to Q(U-Vj) and the efficiency U-Vj only. In order to quantify the jet velocity, Vj,

8 the velocity magnitude of the water jet has been extracted from the numerical results, where the velocity is the

9 value at the center of the water jet just before impinging on the bucket. As shown in Fig. 12, the velocity of the

10 water jet decreases as the flow rate decreases so that the lower flow rate could cause lower efficiency of Pelton

11 turbines. However, the variation of the jet velocity is confined in a narrow range (74~77m/s) in spite of the large

12 variation of the flow rate. Since U-Vj might be considered as a constant regardless of the flow rate, the torque

13 should be directly proportional to the flow rate and the efficiency should be constant according to Eqs. (5) and

14 (6). This implies that there is another cause of the critical mass flow rate indicated by point A in Fig. 8, where

15 the slope of torque variation and the efficiency are significantly changed.

16

17

18 Fig. 12 Velocity at the center of the water jet just before impingement

19
ACCEPTED MANUSCRIPT
15

2 Fig. 13 Interaction of water jet and buckets (VW=0.85)

4 The similar velocity magnitude of the water jet for three different flow rates makes the interaction between

5 the jet and the buckets similar as shown in Fig. 13. For all cases, the cut-out starts cutting the water jet at the

6 position 0. Thereafter, the water jet impinges onto the bucket at the position 1 with an incidence angle less than

7 90° and the jet is blocked completely by the bucket. The water jet segment cut by the bucket exists between two

8 positions of 1 and 2 and is still pushing the bucket at the position 2 with the right angle. In this figure, only

9 difference for different flow rates is the splash of water droplets generated by the interaction. As the flow rate

10 increases, more water droplets are generated after the water jet impingement on the buckets. However, the effect

11 of the droplets on the torque and the power of the Pelton turbine is limited.

12

13 3.3 Loss in injector

14 Figure 14 shows total pressure distribution of the water jet for different flow rates. At the first glance, the

15 wake of the spear has lower total pressure than the given inlet total pressure of 30 bar and the wake region

16 becomes thicker as the flow rate decreases. For Q3, in particular, the wake region accounts for most of the water

17 jet because the jet thickness decreases while the wake region increases with decreasing flow rate. Lower total

18 pressure in the wake region implies higher hydraulic loss inside the injector. As shown in Fig. 14(c), the flow

19 accelerates quickly through the narrow nozzle area and the fast flow causes high friction loss rather than other

20 two cases. This result corresponds to the numerical results of Benzon et al. [2], where the losses in an injector

21 increase at lower flow rates when the flow is restricted by the spear.

22
ACCEPTED MANUSCRIPT
16

1
2 (a) Q1

3
4 (b) Q2

5
6 (c) Q3

7
8 Fig. 14 Total pressure distribution of water jet on symmetric plane
9
10 In order to quantify the hydraulic loss through the injector, total pressure distribution across the water jet was

11 extracted at the position downstream of the spear, which is indicated black line in Fig. 14, from the numerical

12 results. As shown in Fig. 15, the spear wake exists at the center of the water jet at three flow rates. The total

13 pressure at the center decreases as the flow rate decreases. For Q2, total pressure distribution across the jet is

14 similar to the distribution of Q1 in spite of the difference in the flow rate. As mentioned above, the water jet for

15 Q2 has slightly thin thickness less than the water jet for Q1. For these two flow rates, total pressure reaches about

16 30 bar in all other regions except for the center. For Q3, however, the water jet becomes thinner significantly in

17 comparison to the other cases and the maximum value of total pressure is also smaller than 30 bar. This implies

18 that there is considerable friction loss through the injector when the nozzle area is smaller than a critical value.

19
ACCEPTED MANUSCRIPT
17

2 Fig. 15 Total pressure distribution across water jet

4 The head loss in the injector has been usually ignored in analytical methods and the water jet is considered to

5 be uniform. However, it could be easily included in the analysis by assuming the loss as a minor loss caused by

6 an injector geometry. The head loss coefficient has been calculated by the following equation and the results are

7 shown in Fig. 16.


Δ𝐻
8 ℎ𝐿 = 2 (7)
𝑉 /2𝑔

9 where ΔH is the head difference between the injector inlet and outlet and V is the area-averaged velocity at the

10 injector outlet. The head loss coefficient increases as the spear moves forward because more friction loss is

11 generated in a narrow nozzle area. The head loss coefficient is about 0.06 above the critical flow fate but it

12 doubles to become about 0.13 below the critical flow rate. Referring to Munson et al. [21], these values are

13 similar to the loss coefficients of a fully-opened ball valve and a fully-opened gate valve respectively. It should

14 be noted that the loss coefficients obtained in this study is dependent of the injector geometry and the values

15 could be changed in different injectors. However, it is meaningful that the loss coefficient of the injector

16 increases significantly below a critical flow rate.


ACCEPTED MANUSCRIPT
18

2 Fig. 16 Head loss coefficient at different spear positions

4 4. Conclusion

6 A small sized Pelton turbine for a PRO pilot plant had been designed and tested at a constant runner speed to

7 measure torque and efficiency. In the experimental data, the efficiency of the turbine was nearly constant

8 regardless of volumetric flow rates larger than a critical value, but the efficiency decreased quickly below a

9 critical value. The slope of the power coefficient with respect to the flow coefficient also changed at this critical

10 flow rate. Numerical simulations using a commercial flow solver have been conducted at three different flow

11 rates of the Pelton turbine to figure it out. The following conclusions were drawn from this work.

12

13 1. To predict the performance of Pelton turbines numerically, it is important to include the spear of an

14 injector in flow simulations, because the jet diameter changes unpredictably depending on flow rates and

15 the velocity profile across the water jet is not constant due to the spear wake.

16 2. The water jet diameter is nearly constant above a critical flow rate, while the diameter decreases

17 significantly below a critical flow rate. However, the water jet velocity at the center remains unaffected

18 by the flow rate and the jet diameter.

19 3. Flow rates below a critical value make the wet area on the inner surface of a bucket with the maximum

20 torque become small. The pattern of the wet area remains unchanged by the flow rate above a critical

21 value except for the initial stage of the water impingement on a bucket.

22 4. The nozzle area and the flow rate decrease as the spear position moves forward. When the nozzle area is

23 smaller than a critical value, the narrow gap between the spear and the inner surface of an injector causes
ACCEPTED MANUSCRIPT
19

1 significant friction loss and makes the spear wake thicker, consequently resulting in the decrease of total

2 pressure in the water jet. This is why there exists a critical flow rate for a Pelton turbine operating at a

3 constant runner speed.

4 5. The head loss coefficient of the injector is about 0.06 above a critical flow rate but it is doubled to

5 become about 0.13 below a critical flow rate.

7 Nomenclature

8 CP power coefficient

9 CQ flow coefficient

10 g gravitational acceleration

11 H water head

12 hL head loss coefficient

13 m mass flow rate

14 P power

15 Q volumetric flow rate

16 R radius of turbine runner

17 U bucket speed

18 V water velocity at injector outlet

19 Vj water jet velocity

20 VW volume fraction of water

21 β bucket turning angle

22 η efficiency

23 ρ water density

24 τ torque

25 ω runner rotational speed

26

27 Acknowledgements

28 This research was supported by a grant (code 16IFIP-B065893-04) from Industrial Facilities & Infrastructure

29 Research Program funded by Ministry of Land, Infrastructure and Transport of Korean government.

30
ACCEPTED MANUSCRIPT
20

1 References

2 [1] Z. Zhang and M. Casey, 2007, “Experimental studies of the jet of a Pelton turbine,” Proc IMechE, Part A: J Power

3 Energy, Vol. 221, pp. 1181-1192.

4 [2] D. Benzon, A. Židonis, A. Panagiotopoulos, G. A. Aggidis, J. S. Anagnostopoulos, D. E. Papantonis, 2015, “Impulse

5 turbine injector design improvement using computational fluid dynamics,” J Fluid Eng- T ASME, Vol. 137, p. 041106.

6 [3] I. C. Jo, J. H. Park, J.-W. Kim, Y. Shin and J. T. Chung, 2016, “Jet quality characteristics according to nozzle shape of

7 energy-recovery Pelton turbines in pressure-retarded osmosis,” Desalin Water Treat, Vol. 57, pp. 24626-24635.

8 [4] B. Zoppé, C. Pellone, T. Maitre and P. Leroy, 2006, “Flow analysis inside a Pelton turbine bucket”, J Turbomach, Vol.

9 128, pp. 500-511.

10 [5] A. Perrig, F. Avellan, J.-L.Kueny, M. Farhat and E. Parkinson, 2006, “Flow in a Pelton Turbine Bucket: Numerical and

11 Experimental Investigations,” J Fluid Eng- T ASME, Vol. 128, pp. 350-358.

12 [6] A. Santolin, G. Cavazzini, G. Ardizzon and G. Pavesi, 2009, “Numerical investigation of the interaction between jet and

13 bucket in a Pelton turbine,” Proc IMechE, Part A: J Power Energy, Vol. 223, pp. 721-728.

14 [7] F. G. Stamatelos, J. S. Anagnostopoulos and D. E. Papantonis, 2010, “Performance measurements on a Pelton turbine

15 model,” Proc IMechE, Part A: J Power Energy, Vol. 225, pp. 351-362.

16 [8] Y. Xiao, Z. Wang, J. Zhang, C. Zeng and Z. Yan, 2013, “Numerical and experimental analysis of the hydraulic

17 performance of a prototype Pelton turbine,” Proc IMechE, Part A: J Power Energy, Vol. 228, pp. 46-55.

18 [9] G. A. Aggidis and A. Židonis, 2014, “Hydro turbine prototype testing and generation of performance curves: Fully

19 automated approach,” Renew Energ, Vol. 71, pp. 433-441.

20 [10] J. You, X. Lai, W. Zhou and Y. Cheng, 2015, “3D CFD simulation of the runaway process of a Pelton turbine,” Proc

21 IMechE, Part A: J Power Energy, Vol. 230, pp. 234-244.

22 [11] D. Benzon, A. Židonis, A. Panagiotopoulos, G. A. Aggidis, J. S. Anagnostopoulos and D. E. Papantonis, 2015,

23 “Numerical investigation of the spear valve configuration on the performance of Pelton and Turgo turbine injectors and

24 runners,” J Fluid Eng- T ASME, Vol. 137, p. 111201.

25 [12] M. Choi, Y.-J. Jung and Y. Shin, 2015, “Unsteady flow simulations of Pelton turbine at different rotational speeds,” Adv

26 Mech Eng, Vol. 7, No. 11, pp. 1-9.

27 [13] A. Židonis, A. Panagiotopoulos, G. A. Aggidis, J. S. Anagnostopoulos, D. E. Papantonis, 2015, “Parametric

28 optimization of two Pelton turbine runner designs using CFD,” J Hydrodyn, Ser. B, Vol. 27, pp. 403-412.

29 [14] A. Židonis and G. A. Aggidis, 2016, “Pelton turbine: Identifying the optimum number of buckets using CFD,” J

30 Hydrodyn, Ser. B, Vol. 28, pp. 75-83.

31 [15] V. Gupta, V. Prasad and R. Khare, 2016, “Numerical simulation of six jet Pelton turbine model,” Energy, Vol. 104, pp.

32 24-32.

33 [16] J. C. Marongiu, F. Leboeuf and E. Parkinson, 2007, “Numerical simulation of the flow in a Pelton turbine using the
ACCEPTED MANUSCRIPT
21

1 meshless method smoothed particle hydrodynamics: a new simple solid boundary treatment,” Proc IMechE, Part A: J

2 Power Energy, Vol. 221, pp. 849-856.

3 [17] J. C. Marongiu, F. Leboeuf, J. Caro and E. Parkinson, 2010, “Free surface flows simulations in Pelton turbines using an

4 hybrid SPH-ALE method,” J Hydraul Res, Vol. 48, pp. 40-49.

5 [18] J. S. Anagnostopoulos and D. E. Papantonis, 2012, “A fast Lagrangian simulation method for flow analysis and runner

6 design in Pelton turbines,” J Hydrodyn, Ser. B, Vol. 24, pp. 930-941.

7 [19] A. Rossetti, G. Pavesi, G. Cavazzini, A. Santolin and G. Ardizzon, 2013, “Influence of the bucket geometry on the

8 Pelton performance,” Proc IMechE, Part A: J Power Energy, Vol. 228, pp. 33-45.

9 [20] A. Židonis and G. A. Aggidis, 2015, “State of the art in numerical modeling of Pelton turbines,” Renew Sust Energ Rev,

10 Vol. 45, pp. 135-144.

11 [21] B. R. Munson, T. H. Okiishi, W. W. Huebash and A. P. Rothmayer, 2013, Fluid Mechanics, 7th ed. Singapore: John

12 Wiley & Sons.


ACCEPTED MANUSCRIPT

 A Pelton turbine has been designed to recover energy from a PRO pilot plant.
 In performance tests at a fixed speed, the turbine has a threshold of flow rate.
 The Pelton turbine efficiency drops significantly below a critical flow rate.
 A large friction loss in an injector causes the efficiency drop of the turbine.
 The head loss coefficient of the injector is doubled below a critical flow rate.

You might also like