Francis Turbine 2016

download Francis Turbine 2016

of 12

Transcript of Francis Turbine 2016

  • 8/17/2019 Francis Turbine 2016

    1/12

    Failure analysis of runner blades in a Francis hydraulic turbine —Case study

    Alberto Luna-Ramírez ⁎, Alfonso Campos-Amezcua, Oscar Dorantes-Gómez,Zdzislaw Mazur-Czerwiec, Rodolfo Muñoz-Quezada

    Instituto de Investigaciones Eléctricas, Reforma 113, Col. Palmira, Cuernavaca, Mor., México

    a r t i c l e i n f o a b s t r a c t

     Article history:

    Received 4 June 2015

    Received in revised form 6 October 2015

    Accepted 20 October 2015

    Available online 23 October 2015

    Failure analysis of the moving blades in the runner of a 200 MW Francis type hydraulic turbine

    is presented. Analysis consisted of the determination of the pressure on the blade surface using

    computational  uid dynamics (CFD), the calculation of the stress distribution in the turbine

    runner at different operating conditions using thenite element method, and a simplied fatigue

    analysis. Natural vibration modes in water and Von Kármán's vortex frequencies serve as

    reference. Results showed a large concentration of stresses in the T-joint between the blade and

    crown where the cracks originated. Finally, possible causes of the damage are discussed.

    © 2015 Elsevier Ltd. All rights reserved.

    Keywords:

    Turbine failure

    Cavitation

    Vibration

    Fatigue crack

    Runner blades

    1. Introduction

    The operating regime of a hydraulic turbinecanbe divided into steady stateoperationand transient state operation [1]. During steady

    state, the forces acting on the turbine are: the static weight of the runner including water weight, residual stresses and dynamic forces.

    Residual stresses can be generated not only during construction and assembly of the turbine, but also during the commercial

    operation of the turbine due to weld repair and uneven heating. The effect of residual stress is accentuated on intricate shapes or in

    certain turbine components. Dynamic forces are generated during rotation of the unit as a result of the combination of unbalance and

    misalignment with other transient perturbations, namely intermittent water  ow or foreign objects in the water.

    During steady state,the unit operates at a constanthead, speed, load and constant opening of the guide vanes. The forcesacting during

    this period are constant in magnitude, direction and frequency. However, due to defects such as excess pressure pulsations generated in

    the intake pipe, cavitation or misalignment, random non-periodic forces have different directions, amplitudes and frequencies.

    Transient state operation occurs when there is a change in the head or load or wicket gate opening, i.e. starting, synchronizing,changing load, stopping, load rejections, tripping, turbine failures and overspeed. During these periods, vibrations do not follow a

    single pattern but change in magnitude, direction and frequency; their values increase or decrease depending on the amount of 

    water that enters. Other factors that can induce stresses or a non-steady state are the increase or decrease in load or frequent

    starts or stops. For instance, Casanova  [2]  has reported that the fracture of the draft tube connecting bolts in Francis turbines is

    especially frequent when the machine operates at partial loads.

    During operation, several steady and unsteady forces act on turbine components. Unsteady state forces are superimposed on

    steady state forces, resulting in vibrations. Since several components are fastened to each other, the vibration of a component is

    Engineering Failure Analysis 59 (2016) 314–325

    ⁎  Corresponding author.

    E-mail address: [email protected] (A. Luna-Ramírez).

    http://dx.doi.org/10.1016/j.engfailanal.2015.10.0201350-6307/© 2015 Elsevier Ltd. All rights reserved.

    Contents lists available at ScienceDirect

    Engineering Failure Analysis

     j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / e f a

    http://dx.doi.org/10.1016/j.engfailanal.2015.10.020http://dx.doi.org/10.1016/j.engfailanal.2015.10.020http://dx.doi.org/10.1016/j.engfailanal.2015.10.020mailto:[email protected]://dx.doi.org/10.1016/j.engfailanal.2015.10.020http://www.sciencedirect.com/science/journal/http://www.sciencedirect.com/science/journal/http://localhost/var/www/apps/conversion/tmp/scratch_3/Imprint%20logohttp://dx.doi.org/10.1016/j.engfailanal.2015.10.020http://localhost/var/www/apps/conversion/tmp/scratch_3/Journal%20logomailto:[email protected]://dx.doi.org/10.1016/j.engfailanal.2015.10.020http://crossmark.crossref.org/dialog/?doi=10.1016/j.engfailanal.2015.10.020&domain=pdf

  • 8/17/2019 Francis Turbine 2016

    2/12

    transferred from one to another, so the rigidity of each component must be ensured. These vibrations result in deformation of the

    different components, which in turn causes additional stress on the affected components  [3].

    Although the units are designed with a suf cient margin of safety to withstand normal stresses, high vibrations and dynamic

    stresses result in a gradual development of cracks. For this reason, if no precautions are taken to reduce these vibrations, the

    affected component failure may occur without notice. Excessive vibrations and resulting stresses in the affected components

    can cause fatigue and thus component failure. This vibration can also cause wear and fatigue failure of the runner blades, guide

    vane ring, bearings, shafts, runner labyrinth seals and shaft seals, and shear or loosening of various nuts and bolts in all affected

    areas, as well as loosening of bolts and stator core blocks [4–6].

    Fig. 1.  Tailwater level difference.

    Fig. 2. a) Crack located on the trailing edge prole at the junction between the runner blade and crown, b) cavitation pitting damage near the trailing edge of theblade, c) magnication of cavitation damage on the underside of the blade.

    315 A. Luna-Ramírez et al. / Engineering Failure Analysis 59 (2016) 314– 325

    http://localhost/var/www/apps/conversion/tmp/scratch_3/Image%20of%20Fig.%202http://localhost/var/www/apps/conversion/tmp/scratch_3/Image%20of%20&INS%20id=

  • 8/17/2019 Francis Turbine 2016

    3/12

    In Francis turbine runners, fatigue cracks tend to occur early or after decades of operation. The failure mechanism is considered

    a combination of low-cycle fatigue caused by startups and high cycle fatigue caused or induced by rotor–stator interaction [7,8].

    As long as the loads are high enough, the start–stop cycles may propagate fatigue cracks from manufacturing defects or even

    trigger fatigue cracks at areas of stress concentration. With rotational runner speed up to several hundred revolutions per minute,

    the number of vibration cycles or high cycle fatigue due to wakes behind the stationary blades and guide vanes is increased to

    several million per day. Therefore, once a crack has reached a critical size in relation to high cycle fatigue loading, the crack

    may grow and can cause a catastrophic failure in a short period compared to the design life of the turbine runner  [4,8].

    The purpose of this article is to determine the cause or probable causes of failure (cracking) of the runner blades in a Francis

    200 MW type Hydraulic Turbine. Failure analysis is based on the historical unit operation analysis, which determines pressure

    distribution and stress analysis, and compares the natural frequencies in water with the excitation frequencies due to the Von

    Kármán vortex and simplied fatigue analysis, which together serve to determine the cause or causes of blade failure.

    1.1. Damage description

    Given the excess water in the storage reservoir of the hydroelectric power plants, it became necessary to discharge a lot of 

    water. When the three units of the power plant operate under normal conditions, the water level in the landll is 418 m; during

    discharge, the water level rose 10 m above the normal operating level, as it is shown in Fig. 1. The turbines did not stop operating

    during the venting operation, so excess water was used to carry out work on the turbine.

    The discharge lasted several hours and the peak level (427 m.a.s.l.) remained for three days. The temperature of bearings did

    not increase during this special operation and, since their vibration sensors were not implemented on the turbine, it was not

    possible to measure or detect possible episodes of vibration during the reservoir discharge period.

    During the turbine maintenance period, NDT with magnetic particles were conducted on the 13 runner blades in the turbine.

    Large cracks in four consecutive runner blades were detected during the test.  Fig. 2a) shows the size and location of the cracks,

    which began near the junction between the runner blade and the crown of the hydraulic turbine at the trailing edge, then raninto the body of the blade. Fig. 2b) shows cavitation pitting damage near the trailing edge of the blade. The cavitation damage

    indicated in Fig. 1b) may be related to the interblade vortex cavitation or sometimes known as the Von Kármán vortex cavitation

    phenomenon [3,9], which occurs mainly on the trailing edge of the blade, usually when the turbine is operated at partial loads.

    Such cavitation produces structural vibrations at the blade trailing edge  [10]. Fig. 2c) shows a magnication of cavitation damage

    on the bottom surface of the blade.

    1.2. Turbine operation history

    Before commencing failure analysis of the unit, operating conditions of the three power units in the power plant are presented

    in order to compare the modes of operation of each turbine. Table 1 summarizes the extraordinary conditions of operation for the

    increased tailwater level. Table 2 outlines the operating hours and the number of times that units operated at low loads (30 MW)

    in the 2001–2014 period.

    2. Fluid ow analysis

     2.1. Hydraulic turbine characteristics

    The dimensional characteristics and nominal operation parameters of the analyzed hydraulic turbine are: number of runner

    blades 13, number of guide vanes 24, number of stay vanes 12, and turbine ratings, which are presented in  Table 3. The runner

     Table 1

    Behavior in the storage reservoir hydropower plant.

    Date (Maximum elevation reached)

    (haverage) (m)

    Tailwater level

    difference Δh (m)

    Approximate running

    time of this event (h)

    Power output (average)

    in (MW) during the event

    16 Sep. 2013 427.6 427.62–418⁎ = 9.62 6 192.5

    17 Sep. 2013 427.6 427.65–418 = 9.65 11 192.2

    18 Sep. 2013 423.4 423.44–418 = 5.44   –   194.7

    19 Sep. 2013 422.2 422.25–418 = 4.25   –   192.7

    20 Sep. 2013 419.3 419.34–418 = 1.34   –   194.7

    ⁎   Reference elevation (see Fig. 1).

     Table 2

    Operating hours of the three units and number of times they operated at low load (30 MW) from 2001 to 2014.

    Unit 1 Unit 2 Unit 3

    Hours of operation at low load 75 66 25

    Number of times units operated at low loads 72 78 66

    316   A. Luna-Ramírez et al. / Engineering Failure Analysis 59 (2016) 314– 325

  • 8/17/2019 Francis Turbine 2016

    4/12

    material is 13Cr/4Ni (ASTM A487 Grade CA-6NM) martensitic stainless steel. The properties of the runner material, used for all

    the analyses reported in this paper, are described in  Table 4.

    Fluid analysis through the turbine runner is required to determine the hydraulic load on the surface of the runner blade. The

    computational model domain comprises the guide vanes and turbine runner. Since both the blade distributor and runner blade

    are periodic, a guiding vane and runner blade were modeled in addition to the curves dening the crown and band of the

    impeller. A 3D model, using periodicity conditions, was created.

    Fig. 3(a) shows the guide vane, runner blade and diffuser section, whereas  Fig. 3(b) shows the runner full model using

    periodicity conditions. Analyses were conducted at steady state operation for different loads, considering the working  uid in

    three dimensions as a continuous medium. The Navier–Stokes equations were used to model the  ow, and the Spalart–Allmaras

    turbulence model was used to complete the equation system. Flow analysis was performed using the Computational Fluid

    Dynamics (CFD) technique using NUMECA  [11]  commercial software.

    The computational mesh is an unstructured grid of almost 3 million of hexahedral cells, with a higher density of nodes in areas

    where the  uid  ow was expected to considerably change, as at the leading edge and trailing edges of guide vanes and runner

    blades, where a  ner grid resolution was assumed to be benecial for achieving an accurate resolution, Fig. 4.

    A preliminary mesh was worked out including a more rened boundary-layer, for which the  ow solution no longer changed

    with subsequent rening. A detail of blade to blade mesh can be seen in  Fig. 5.

    Boundary conditions to be imposed on the model for normal operating conditions of the turbine are shown in  Table 5. The

    angle at which the  ow enters the model (inlet to the guide vanes) corresponds to the output angle of the stay vanes.

    When the level increased in the lower reservoir of the Power Plant to 9.65 m, the outlet head raised to 436,695 Pa, that is,

    there was an increase of 94,655 Pa compared to the nominal head. Computational runs at partial loads and overload were

    performed to determine the pressure on the surface of the blades, and subsequently calculate stress. The boundary conditions

    of these models are the same as in  Table 6, except for mass  ow. The values of the variables (load, angle and throat) that are

    modied in the model for each analysis are shown in  Table 6.

     2.2. CFD results

    Pressure distribution for the pressure and suction sides of the runner blades for different load conditions is shown in  Fig. 6.

    Pressure differences for the pressure and suction sides are higher at higher loads (Fig. 6a5) and b5)).The increasing pressure

    on the blades at the leading edge for all load conditions on both sides is clearly visible.

    Under all load conditions, pressure is observed to gradually decrease in the direction of the  ow, starting at the suction side of 

    the blade. The low pressure zone (blue) may be viewed after half the length of blade, at the bottom of the blade trailing edge. For

    example, at 100% load, maximum pressure is 980 kPa at the leading edge and at least 2 kPa at the trailing edge. This low pressure

    area is susceptible to cavitation. The pressure reduction on the runner blade suction side near the trailing edge region may be

    explained by the higher magnitudes of the velocity speed at the trailing edge  [12]. In all cases or load conditions, the pressure

    differences on both sides of the blade (suction and pressure) are high; these differences result in a torque on the shaft. Similar

    results have been documented in other studies [13].

    3. Stress analysis in turbine runner 

    The stress analysis of the runner subjected to different loading conditions is an important issue in the analysis of failure since it

    provides information on the structural integrity of the turbine. In this study, analysis was performed using ANSYS, which has

    proven to be an effective tool for modeling stresses in hydraulic turbine components  [5,12,13,14].

    Mechanisms that can damage the hydraulic turbine runner are a combination of loads acting on the runner, and are divided

    into two categories: stationary loads (the  uid pressure on the surface of the blades, the centrifugal force and the runner's own

     Table 3

    Nominal operating parameters of the hydraulic turbine.

    Net head (m) Output (MW) Discharge (m3/s) Speed (rpm)

    101.3 214.25 227.7 128

    91.2 200.5 237.9

    75.5 151.1 216.2

     Table 4

    Properties of material in a hydraulic turbine runner.

    Material property ASTM A487 Grade CA-6NM

    Tensile strength (Su) 755 MPa

    Yield strength (Sy) 550 MPa

    Density (ρ) 7695 kg/m3

    Young's modulus (E) 199.95 GPa

    317 A. Luna-Ramírez et al. / Engineering Failure Analysis 59 (2016) 314– 325

  • 8/17/2019 Francis Turbine 2016

    5/12

    weight) and dynamic loads (pressure uctuations caused by the rotor–stator interaction, as well as phenomena such as cavitation

    and water hammer). In the Francis turbine, stresses caused by centrifugal force during normal operation condition are much

    smaller than those caused by water pressure  [12].

    The geometric model with the established boundary conditions is shown in  Fig. 7. Letter A corresponds to the zero displace-

    ment condition (UX = UY = UZ = 0), which is represented as a  xed support in the area of connection to the shaft  ange. Letter

    B represents the rotational clockwise speed, and letter C represents gravity acceleration.

    Fig. 8 shows the results of stresses calculations at different loading conditions (15, 30, 50, 75, 100, 110%) on the runner blades,

    considering the pressure distribution exerted by the water  ow on the surface of the runner blades and centrifugal forces. The

    stress distribution shown in Fig. 8 is expressed in terms of equivalent stress (Von Mises stress).

    At 100% load, maximum stress (Fig. 8e)) was around 224.86 MPa, and at 110% load, maximum stress was 254.30 MPa

    (Fig. 8g)); both values represent slightly less than half the material yield stress (σ y = 555 MPa). For all loading conditions,

    maximum stress started at the trailing edge at the junction between blade and crown. Maximum stress obtained for a load of 

    100% but considering output pressure of 436.7 Pa (corresponding to 9.65 m increased level in the lower reservoir vent of the

    Power Plant) was 241.55 MPa. This value is approximately 7% larger than the obtained at a nominal load of 100% and 5% less

    than the 110% load (Fig. 8f)).

    It is important to note that the location of the higher stresses in all cases coincides with the location of the cracks observed in

    the runner blades. This suggests that although the stresses alone do not cause blade failure, they are the source of crack induction

    and growth.

    The literature [4,5,12,13,15,16] mentions that in a Francis turbine runner, the region near the trailing edge at the blade crown

    intersection is identied as the region subjected to high stresses, and is therefore a critical area for cracks to occur due to fatigue.

    From a numerical analysis standpoint, it can be concluded that the maximum calculated stresses are located at the transition

    between blade and crown on the trailing edge, which correspond to the location of the cracks observed during maintenance of 

    the unit.

    Fig. 3.  Geometric model used in the simulation.

    Fig. 4.  Computational grid used in the model.

    318   A. Luna-Ramírez et al. / Engineering Failure Analysis 59 (2016) 314– 325

    http://localhost/var/www/apps/conversion/tmp/scratch_3/Image%20of%20Fig.%204http://localhost/var/www/apps/conversion/tmp/scratch_3/Image%20of%20&INS%20id=

  • 8/17/2019 Francis Turbine 2016

    6/12

     3.1. Determination of natural frequencies of the runner in water and air 

    Since the runners work while submerged in water, dynamic behavior could be strongly affected by the presence of this heavy

    uid [17]; therefore, the study on turbine runner structures considering water effect is of highly practical interest. A numerical

    model was therefore developed, including a  nite volume representing an added mass of water.  Fig. 9 shows the numerical

    model and natural frequencies of therunner that were calculated. This analysis considers only the added watermass; thus, theeffects

    of vibration in the water surrounding the runner are not taken into account.  Table 7 shows the mode and natural frequency of 

    vibration of the runner in water and air at 128.57 rpm.

    When the obtained result in the runner's simulations in air and water are compared, it was observed that a decrease in the

    natural frequencies does exist, because of the water that surrounds the structure. The ratio of frequencies reduction varies from

    0.147 to 0.377 depending on the frequencies. The results presented in Table 7 were used to correlate Von Kármán vortex frequencies

    and certain vibration modes by considering the water mass.

     3.2. Von Kármán vortex frequencies

    The computation of the possible Von Kármán excitation frequency ( f ) must be done for every load condition. These frequencies

    are compared with the runner natural frequencies in water in order to evaluate the chance to have a natural mode shape excited

    by any Von Kármán frequency. The Von Kármán frequency ( f ) is dened by the following Eq.  (1)  [18]:

     f   ¼ S t    C =t  þ  δvð Þ………::   ð1Þ

    It has been found that the value of  S t  varies between 0.16 and 0.24 [19]. For the following calculations a value of 0.2 is assumed

    where:

    S t    Strouhal number

    C velocity of water [m/s]

    T thickness of the blade trailing edge [m]

    δv   0.0293 (x/Rex)1/5 = virtual boundary layer thickness [m]

    where

    x blade length or chord length = 1.744 m

    Rex   Reynolds number = 9.5E7

    Fig. 5.  Fragment of the guide blade to blade mesh.

     Table 5

    Boundary conditions at rated load (214.25 MW).

    Boundary condition Units Value

    Mass ow kg/s 227,928

    Temperature °C 20

    Turbulent viscosity m2/s 0.0001

    Static pressure at model outlet Pa 342,041

    Rotational speed rpm 128.57

    319 A. Luna-Ramírez et al. / Engineering Failure Analysis 59 (2016) 314– 325

    http://localhost/var/www/apps/conversion/tmp/scratch_3/Image%20of%20Fig.%205

  • 8/17/2019 Francis Turbine 2016

    7/12

     Table 6

    Variations in the model at different loads.

    No. Load [%] Throat [mm] Angle [°] Mass  ow [kg/s]

    1 15 (30 MW) 63 82.8 34,189

    2 30 126 77.1 68,378

    3 50 210 69.4 113,964

    4 75 315 59.9 170,946

    5 100 (200 MW) 420 50.4 227,928

    6 110 462 46.5 250,721

    Fig. 6.  Solutions for pressure on the blade; a) pressure on the pressure side of the blade, b) pressure on the suction side of the blade, under different loadconditions.

    320   A. Luna-Ramírez et al. / Engineering Failure Analysis 59 (2016) 314– 325

    http://localhost/var/www/apps/conversion/tmp/scratch_3/Image%20of%20Fig.%206

  • 8/17/2019 Francis Turbine 2016

    8/12

    Table 8 shows the results of frequency calculation by Von Kármán vortex for different condition loads.

    Analysis of the results in Table 8 shows that at 30% load, the Von Kármán frequency (115.51 Hz) is very close to the natural

    frequency (116.7 Hz) of the fourth vibration mode of the runner in water. Furthermore, by varying the turbine load from 15 to

    30% (30 to 60 MW), the Von Kármán vortex frequencies range from 57.7 to 115.5 Hz, and they can result in the excitation of 

    the runner natural frequencies in water in second (71.1 Hz) and third mode (88.7 Hz) vibration; see  Table 7.

    This excitation can generate high levels of vibration, which in turn can produce high cycle fatigue failure on mechanical

    components, especially on turbine runner blades [13]. Cases have been reported of cracking of runner blades attributed to the

    occurrence of vibrations on account of resonance between the Von Kármán vortex shedding frequency and the natural frequency

    of the blades [16,20–23].

    Fig. 6   (continued).

    321 A. Luna-Ramírez et al. / Engineering Failure Analysis 59 (2016) 314– 325

    http://localhost/var/www/apps/conversion/tmp/scratch_3/Image%20of%20Fig.%206

  • 8/17/2019 Francis Turbine 2016

    9/12

    4. Simplied fatigue analysis

    The fatigue analysis serves to characterize the ability of a material subjected to cyclic loading, such as the hydraulic turbine

    runner. Startups, shutdowns and load changes produce low cycle fatigue, whereas cyclic loads induced by vibration,   ow

    uctuations, recirculation and resonance produce high cycle fatigue. The fatigue strength of steel is dened by the following

    Eq. (2):

    N  f   ¼ t r   f r ð Þ:::::::::::::::::   ð2Þ

    where [24]:

    Nf    resistance fatigue [cycles]

    tr   time in resonance [s]

    f r   resonance frequency

    For steel,  N  f  = 107 cycles are considered.

    Solving  t r  in Eq. (2), time (t r ) to failure (cracking) is calculated when the component operates in resonance; see Eq.  (3):

    t r  ¼ N  f  = f r :::::::::::::   ð3Þ

    When Eq. (2) includes the fatigue resistance of the steel  N  f  = 107 cycles and the natural frequencies calculated at the runner,

     f r1  = 30.92 Hz (rst mode) and  f r2  = 116.67 (mode) are obtained:

    t r1  ¼  107 = 30:92 ¼ 89:8 h½ ; t r2  ¼  10

    7 =116:67  ¼  23:8 h½

    The results indicate that if the runner operates in resonance with the  rst or fourth mode of vibration, cracks in the runner

    would appear within 24 to 90 h of operation. In case of resonance at higher modes of the runner, cracking would occur in less

    time.

    On the other hand, it is known that the runner has been operated for 26 years without cracking. This means that runner failure

    could not be induced during normal operating conditions (constant nominal load). It is thus suspected that some other temporary

    transients occurred, causing the excitation in one of the natural frequencies of the runner, resulting in high cycle fatigue failure.

    It is known that normal fatigue life is around 90% the total life for crack initiation and 10% for the propagation phase  [25]. This

    implies that no signs of fatigue are noticed on the runner blade surface for a signicant period of time even if fatigue conditions occur

    during operation (excitation, resonance). Fatigue damage is cumulative and becomes visible following a considerable number of fatigue cycles.

    Fig. 7.  Boundary conditions used in the  nite element analysis.

    322   A. Luna-Ramírez et al. / Engineering Failure Analysis 59 (2016) 314– 325

    http://localhost/var/www/apps/conversion/tmp/scratch_3/Image%20of%20Fig.%207

  • 8/17/2019 Francis Turbine 2016

    10/12

    5. Conclusions

    Taking into account the results of the calculated stresses at maximum operating loads, the stresses are below the yieldstrength (σ y  = 550 MPa). Therefore, these stresses alone could not have caused blade failure.

    Fig. 8.  Stress distribution on the surface of the runner blades under different load conditions.

    323 A. Luna-Ramírez et al. / Engineering Failure Analysis 59 (2016) 314– 325

    http://localhost/var/www/apps/conversion/tmp/scratch_3/Image%20of%20Fig.%208

  • 8/17/2019 Francis Turbine 2016

    11/12

    The maximum stresses obtained for 100% load with additional 94,655 Pa outlet pressure were approximately 7% higher than

    those obtained at 100% nominal load. On the other hand, these stresses were 5% lower than at 110% load, which also indicates

    that this level of stress alone (without any other phenomena) could not be the direct cause of the   “sudden” failure of the blades.

    Namely, the rise in water level at the turbine outlet caused the stresses to increase, which in combination with another phenom-

    enon could have accelerated blade failure.

    All maximum stresses calculated on the runner were located near or at the transition radius between band and crown on the

    trailing edge of the blades. Their locations coincide with blade failure, which indicates that this stress concentration was where

    the crack started and spread. Static pressure values on model surfaces at 100% load, obtained by  ow analysis, indicate that

    cavitation prone areas were found, which coincided with cavitation damage recorded during inspection of the units 1 and 3.

    Turbine operation at low loads may excite the natural frequency of the blades in three of the  rst four modes of vibration,

    resulting in high cycle fatigue of the runner blades. The most likely cause of failure in the turbine runner blades was turbine

    operation at low loads, accelerated by detected cavitation, and high water levels in the upper reservoir and tailwater.

    Fig. 9.  Numerical model runner submerged in water mass.

     Table 7

    Vibration frequency for different vibration modes in water and air.

    Vibration mode Vibration frequency (simulation) (Hz)

    In water In air

    ND1 30.92 43.01

    ND2 71.12 83.39

    ND3 88.76 142.51ND4 116.67 158.06

    ND5 135.37 194.71

     Table 8

    Results of frequency calculation by Von Kármán vortex.

    Load   δv   Velocity of water Thickness of blade trailing edge Von Kármán vortex

    % [m] C (m/s) t (m) f (Hz)

    15

    0.00129

    6.7

    0.0219

    57.7

    30 13.4 115.51

    50 22.3 192.2

    75 33.4 287.8

    100 44.6 384.4

    324   A. Luna-Ramírez et al. / Engineering Failure Analysis 59 (2016) 314– 325

    http://localhost/var/www/apps/conversion/tmp/scratch_3/Image%20of%20Fig.%209

  • 8/17/2019 Francis Turbine 2016

    12/12

    The Von Kármán vortex shedding phenomenon may explain the appearance of cracks in the runner blades, since some

    excitation frequencies are similar or close to the natural frequencies of the runner in water, contributing to damage by high

    cycle fatigue. However, more studies are needed to corroborate this conclusion.

    References

    [1] Er R.K. Aggarwal, Metal fatigue due to excess vibration and dynamic stresses on a hydro power plantRetd. Member (Power) BBMB. (Power Consultant, Certified

    Energy Auditor BEE & Empanelled Energy Auditor PCRA)  https://cdm.unfccc.int, 2013.[2]   F. Casanova, Failure analysis of the draft tube connecting bolts of a Francis-type hydroelectric power plant, Eng. Fail. Anal. 16 (7) (2009) 2202–2208.[3]   U. Dorji, R. Ghomashchi, Hydro turbine failure mechanisms: an overview, Eng. Fail. Anal. 44 (2014) 136–147.[4]   D. Frunzaverde, et al., Failure analysis of a Francis turbine runner, 25th IAHR Symposium on Hydraulic Machinery and Systems, IOP Conf. Series: Earth and

    Enviromental Science, 12 2010, p. 012115.[5]   M. Flores, G. Urquiza, J.M. Rodríguez, A fatigue analysis of a hydraulic Francis turbine runner, World J. Mech. 2 (2012) 28–34.[6]   M. Gagnon, T. Antoine, B.P. Thibault, On the fatigue reliability of hydroelectric Francis runners, Procedia Eng. 66 (2013) 565–574.[7]   B. Hermond, Design, perfomance and maintenance of Francis turbine, Glob. J. Res. Eng. 13 (2013).[8]   H.-J. Huth, Fatigue design of the hydraulic turbine runners(PhD Thesis) Norwegian University of Science and Technology, 2005.[9]   P. Kumar, R.P. Saini, Study of cavitation in hydro turbines—a review, J. Renew. Sustain. Energy 14 (2010) 374–383.

    [10]   S. Khurana, H. Singh, Effect of cavitation on hydraulic turbines  —  a review, Int. J. Curr. Eng. Technol. 2 (1) (2012) 172–177.[11] FINE™/Turbo v9.1 Flow Integrated Environment, NUMECA International, Belgium.[12]   R.A. Saeed, A.N. Galybin, V. Popov, Modelling of flow-induced stresses in a Francis turbine runner, Adv. Eng. Softw. 41 (2010) 1245–1255.[13]   R. Negru,S. Muntean, L. Marsavina, R. Susan-Resiga,N. Pasca, Computation of stress distribution in a Francis turbine runnerinduced by fluid flow, Comput. Mater.

    Sci. 64 (2012) 253–259.[14]   M.O. Popoviciu, I. Bordeasu, Analyses of fissures in bulb turbine shafts, Mach. Des. 3 (4) (2011) 233–240.[15]   X. Ruofu, W. Zhengwei, L. Yongyao, Dynamic stresses in a Francis turbine runner based on fluid–structure interaction analysis, Tsinghua Sci. Technol. 13 (5)

    (2008) 587–592.[16]   A. Mirasso, L. Millán, Aplicación del MEF en la determinación de las causas de fisuración de un rodete de turbina Francis, Mecánica Computacional Vol. XX11

    ENIEF 2003—XIII Congreso sobre Métodos Numéricos y sus Aplicaciones, 2003 (In spanish).[17]   Q.W. Liang, C.G. Rodríguez, E. Egusquiza, X. Escaler, M. Farhat, F. Avellan, Numerical simulation of fluid added mass effect on a Francis turbine runner, Comput.

    Fluids 36 (2007) 1106–1118.[18]   C.A. Gonwer, A study of vanes singing in the water, Trans. ASME 74 (1952).[19]   G.P. Hans, J.C. Zanutto, W.P. Ferreira, Hydraulic power plant machine dynamic diagnosis, Shock. Vib. 13 (2006) 409–427.[20]   Z. Lv, W. Xu, G. Tang, Z. Wu, J. Wang, The generation of cracks in hydraulic turbine blades and the instances, Appl. Mech. Mater. 543–547 (2014) 268–271.[21]   L.R. Malik, M.L. Chawla, Cracking of hydro-turbine runner blades, Proceeding Forty Third Annual Research Session, Vol IV —B-Power 1973, pp. 107–117 (New

    Delhi).[22]   R.K. Fisher, U. Seidel, G. Grosse, W. Gfeller, R. Klinger, A case study in resonant hydroelastic vibration: the causes of runner cracks and the solutions implemented

    for the Xiaolangdi hydroelectric project, Proceedings of the XXI IAHR Symposium on Hydraulic Machinery and Systems, Lausanne, Switzerland, 2002 .[23]   Q. Shi, Abnormal noise and runner cracks caused by Von Karman vortex shedding: a case study in Dachaoshan hydroelectric project, 22nd IAHR Symposium on

    Hydraulic Machinery and System, Stockholm, Sweden, 2004.[24]   D.R. Askeland, The Science and Engineering of Materials, Sixth edition SI PWS Publishers, USA, 1985.[25] Cordes T. Predicting total fatigue life (crack initiation and crack propagation). Presented to the SAE FD & E Committee at Cincinnati, OH on 15 April 2013.

    325 A. Luna-Ramírez et al. / Engineering Failure Analysis 59 (2016) 314– 325

    https://cdm.unfccc.int/http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0005http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0005http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0005http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0010http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0010http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0010http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0015http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0015http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0020http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0020http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0020http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0025http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0025http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0025http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0030http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0035http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0040http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0040http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0040http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0040http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0040http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0045http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0045http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0045http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0045http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0045http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0050http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0050http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0050http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0055http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0055http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0055http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0055http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0060http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0060http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0060http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0065http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0065http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0065http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0065http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0065http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0065http://refhub.elsevier.com/S1350-6307(15)30120-5/rf1235http://refhub.elsevier.com/S1350-6307(15)30120-5/rf1235http://refhub.elsevier.com/S1350-6307(15)30120-5/rf1235http://refhub.elsevier.com/S1350-6307(15)30120-5/rf1235http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0070http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0070http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0070http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0070http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0075http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0080http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0080http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0080http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0085http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0085http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0085http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0085http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0085http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0090http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0090http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0090http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0090http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0090http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0090http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0095http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0095http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0100http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0100http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0100http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0105http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0105http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0100http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0100http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0095http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0095http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0090http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0090http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0085http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0080http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0075http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0070http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0070http://refhub.elsevier.com/S1350-6307(15)30120-5/rf1235http://refhub.elsevier.com/S1350-6307(15)30120-5/rf1235http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0065http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0065http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0060http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0055http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0055http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0050http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0045http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0040http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0035http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0030http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0025http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0020http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0015http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0015http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0010http://refhub.elsevier.com/S1350-6307(15)30120-5/rf0005https://cdm.unfccc.int/