Вы находитесь на странице: 1из 12

Engineering Failure Analysis 59 (2016) 314325

Contents lists available at ScienceDirect

Engineering Failure Analysis


journal homepage: www.elsevier.com/locate/efa

Failure analysis of runner blades in a Francis hydraulic turbine


Case study
Alberto Luna-Ramrez , Alfonso Campos-Amezcua, Oscar Dorantes-Gmez,
Zdzislaw Mazur-Czerwiec, Rodolfo Muoz-Quezada
Instituto de Investigaciones Elctricas, Reforma 113, Col. Palmira, Cuernavaca, Mor., Mxico

a r t i c l e

i n f o

Article history:
Received 4 June 2015
Received in revised form 6 October 2015
Accepted 20 October 2015
Available online 23 October 2015
Keywords:
Turbine failure
Cavitation
Vibration
Fatigue crack
Runner blades

a b s t r a c t
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 the nite element method, and a simplied fatigue
analysis. Natural vibration modes in water and Von Krmn'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.

1. Introduction
The operating regime of a hydraulic turbine can be divided into steady state operation and 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 constant head, speed, load and constant opening of the guide vanes. The forces acting 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
Corresponding author.
E-mail address: aluna@iie.org.mx (A. Luna-Ramrez).

http://dx.doi.org/10.1016/j.engfailanal.2015.10.020
1350-6307/ 2015 Elsevier Ltd. All rights reserved.

A. Luna-Ramrez et al. / Engineering Failure Analysis 59 (2016) 314325

315

Fig. 1. Tailwater level difference.

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 sufcient 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 [46].

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 the
blade, c) magnication of cavitation damage on the underside of the blade.

316

A. Luna-Ramrez et al. / Engineering Failure Analysis 59 (2016) 314325

Table 1
Behavior in the storage reservoir hydropower plant.
Date

(Maximum elevation reached)


(haverage) (m)

Tailwater level
difference h (m)

Approximate running
time of this event (h)

Power output (average)


in (MW) during the event

16 Sep. 2013
17 Sep. 2013
18 Sep. 2013
19 Sep. 2013
20 Sep. 2013

427.6
427.6
423.4
422.2
419.3

427.62418 = 9.62
427.65418 = 9.65
423.44418 = 5.44
422.25418 = 4.25
419.34418 = 1.34

6
11

192.5
192.2
194.7
192.7
194.7

Reference elevation (see Fig. 1).

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 rotorstator interaction [7,8].
As long as the loads are high enough, the startstop 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
Krmn 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 ran
into 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 Krmn 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 20012014 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 2
Operating hours of the three units and number of times they operated at low load (30 MW) from 2001 to 2014.

Hours of operation at low load


Number of times units operated at low loads

Unit 1

Unit 2

Unit 3

75
72

66
78

25
66

A. Luna-Ramrez et al. / Engineering Failure Analysis 59 (2016) 314325

317

Table 3
Nominal operating parameters of the hydraulic turbine.
Net head (m)

Output (MW)

Discharge (m3/s)

Speed (rpm)

101.3
91.2
75.5

214.25
200.5
151.1

227.7
237.9
216.2

128

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 NavierStokes equations were used to model the ow, and the SpalartAllmaras
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 4
Properties of material in a hydraulic turbine runner.
Material property

ASTM A487 Grade CA-6NM

Tensile strength (Su)


Yield strength (Sy)
Density ()
Young's modulus (E)

755 MPa
550 MPa
7695 kg/m3
199.95 GPa

318

A. Luna-Ramrez et al. / Engineering Failure Analysis 59 (2016) 314325

Fig. 3. Geometric model used in the simulation.

weight) and dynamic loads (pressure uctuations caused by the rotorstator 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 displacement 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 (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. 4. Computational grid used in the model.

A. Luna-Ramrez et al. / Engineering Failure Analysis 59 (2016) 314325

319

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

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 the runner that were calculated. This analysis considers only the added water mass; thus, the effects
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 Krmn vortex frequencies
and certain vibration modes by considering the water mass.
3.2. Von Krmn vortex frequencies
The computation of the possible Von Krmn 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 Krmn frequency. The Von Krmn frequency (f) is dened by the following Eq. (1) [18]:
f St C=t v ::

It has been found that the value of St varies between 0.16 and 0.24 [19]. For the following calculations a value of 0.2 is assumed
where:
St
C
T
v

Strouhal number
velocity of water [m/s]
thickness of the blade trailing edge [m]
0.0293 (x/Rex)1/5 = virtual boundary layer thickness [m]

where
x
Rex

blade length or chord length = 1.744 m


Reynolds number = 9.5E7

Table 5
Boundary conditions at rated load (214.25 MW).
Boundary condition

Units

Value

Mass ow
Temperature
Turbulent viscosity
Static pressure at model outlet
Rotational speed

kg/s
C
m2/s
Pa
rpm

227,928
20
0.0001
342,041
128.57

320

A. Luna-Ramrez et al. / Engineering Failure Analysis 59 (2016) 314325

Table 6
Variations in the model at different loads.
No.

Load [%]

Throat [mm]

Angle []

Mass ow [kg/s]

1
2
3
4
5
6

15 (30 MW)
30
50
75
100 (200 MW)
110

63
126
210
315
420
462

82.8
77.1
69.4
59.9
50.4
46.5

34,189
68,378
113,964
170,946
227,928
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 load
conditions.

A. Luna-Ramrez et al. / Engineering Failure Analysis 59 (2016) 314325

321

Fig. 6 (continued).

Table 8 shows the results of frequency calculation by Von Krmn vortex for different condition loads.
Analysis of the results in Table 8 shows that at 30% load, the Von Krmn 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 Krmn 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 Krmn vortex shedding frequency and the natural frequency
of the blades [16,2023].

322

A. Luna-Ramrez et al. / Engineering Failure Analysis 59 (2016) 314325

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

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 t r f r :::::::::::::::::

where [24]:
Nf
tr
fr

resistance fatigue [cycles]


time in resonance [s]
resonance frequency
For steel, Nf = 107 cycles are considered.
Solving tr in Eq. (2), time (tr) to failure (cracking) is calculated when the component operates in resonance; see Eq. (3):
t r N f =f r :::::::::::::

When Eq. (2) includes the fatigue resistance of the steel Nf = 107 cycles and the natural frequencies calculated at the runner,
fr1 = 30.92 Hz (rst mode) and fr2 = 116.67 (mode) are obtained:
7

t r1 10 =30:92 89:8 h; t r2 10 =116:67 23:8 h


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.

A. Luna-Ramrez et al. / Engineering Failure Analysis 59 (2016) 314325

323

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

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

324

A. Luna-Ramrez et al. / Engineering Failure Analysis 59 (2016) 314325

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)

ND1
ND2
ND3
ND4
ND5

In water

In air

30.92
71.12
88.76
116.67
135.37

43.01
83.39
142.51
158.06
194.71

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 phenomenon 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.

Table 8
Results of frequency calculation by Von Krmn vortex.
Load

Velocity of water

Thickness of blade trailing edge

Von Krmn vortex

[m]

C (m/s)

t (m)

f (Hz)

0.00129

6.7
13.4
22.3
33.4
44.6

0.0219

57.7
115.51
192.2
287.8
384.4

15
30
50
75
100

A. Luna-Ramrez et al. / Engineering Failure Analysis 59 (2016) 314325

325

The Von Krmn 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) 22022208.
[3] U. Dorji, R. Ghomashchi, Hydro turbine failure mechanisms: an overview, Eng. Fail. Anal. 44 (2014) 136147.
[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. Rodrguez, A fatigue analysis of a hydraulic Francis turbine runner, World J. Mech. 2 (2012) 2834.
[6] M. Gagnon, T. Antoine, B.P. Thibault, On the fatigue reliability of hydroelectric Francis runners, Procedia Eng. 66 (2013) 565574.
[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 turbinesa review, J. Renew. Sustain. Energy 14 (2010) 374383.
[10] S. Khurana, H. Singh, Effect of cavitation on hydraulic turbines a review, Int. J. Curr. Eng. Technol. 2 (1) (2012) 172177.
[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) 12451255.
[13] R. Negru, S. Muntean, L. Marsavina, R. Susan-Resiga, N. Pasca, Computation of stress distribution in a Francis turbine runner induced by fluid flow, Comput. Mater.
Sci. 64 (2012) 253259.
[14] M.O. Popoviciu, I. Bordeasu, Analyses of fissures in bulb turbine shafts, Mach. Des. 3 (4) (2011) 233240.
[15] X. Ruofu, W. Zhengwei, L. Yongyao, Dynamic stresses in a Francis turbine runner based on fluidstructure interaction analysis, Tsinghua Sci. Technol. 13 (5)
(2008) 587592.
[16] A. Mirasso, L. Milln, Aplicacin del MEF en la determinacin de las causas de fisuracin de un rodete de turbina Francis, Mecnica Computacional Vol. XX11
ENIEF 2003XIII Congreso sobre Mtodos Numricos y sus Aplicaciones, 2003 (In spanish).
[17] Q.W. Liang, C.G. Rodrguez, E. Egusquiza, X. Escaler, M. Farhat, F. Avellan, Numerical simulation of fluid added mass effect on a Francis turbine runner, Comput.
Fluids 36 (2007) 11061118.
[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) 409427.
[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. 543547 (2014) 268271.
[21] L.R. Malik, M.L. Chawla, Cracking of hydro-turbine runner blades, Proceeding Forty Third Annual Research Session, Vol IVB-Power 1973, pp. 107117 (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.

Вам также может понравиться