You are on page 1of 10

Engineering Failure Analysis 97 (2019) 579–588

Contents lists available at ScienceDirect

Engineering Failure Analysis


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

Detection of damage in steam turbine blades caused by low cycle


T
and strain cycling fatigue
S. Canoa, J.A. Rodrígueza, , J.M. Rodríguezb, J.C. Garcíaa, F.Z. Sierraa, S.R. Casolcoa,

M. Herreraa
a
Centro de Investigación en Ingeniería y Ciencias Aplicadas, Av. Universidad 1001, Col. Chamilpa, 62209 Cuernavaca, Morelos, Mexico
b
Centro Nacional de Investigación y Desarrollo Tecnológico, Interior Internado Palmira S/N, Col. Palmira, C.P. 62490, Cuernavaca, Morelos, Mexico

ARTICLE INFO ABSTRACT

Keywords: The blades of steam turbines are exposed to low cycle fatigue, as well as to strain cycling fatigue,
Damage of turbine blade which may cause their structural deterioration after a low number of cycles. This represents the
Failure analysis blade damage in steady condition of operation, which may lead to failure. In this work, the blades
Low cycle fatigue of the last stage (L-0) from a 110 MW output steam turbine were analyzed. Samples of material
Life estimation
AISI 410 were investigated to define the number of cycles against the stress to failure. As a result,
the magnitude of the constants of a numerical model were obtained. In addition, numerical and
analytical models predicted the blade damage based on mean stress correction, Strain-life,
Morrow and Smith Watson Topper (SWT). The results showed that centrifugal load caused da-
mage in blades and likely, it transforms into crack initiation due to low cycles fatigue. Steam load
produced stresses low level in magnitude. However these can be dangerous for blade structure
subject to non-linear and resonance conditions, which mainly occur during transient periods.
Important differences in prediction of life estimation upon the method emerged for high levels of
stress magnitude, which should be taken into account during blade design.

1. Introduction

Fatigue Damage is cumulative and occurs when a structure is subject to variable and/or periodic loads. The repeating maximum
values of stresses during a cycling load induces micro-cracks, which could lead to the failure [1]. Some fractures of steam turbine
blades are caused by low cycle fatigue (LCF) during start-up or shut-down of the steam turbines because of variations of centrifugal
and steam loads together with resonance phenomenon. The blades of the last stage are affected by centrifugal forces which cause
alternating stresses during start-up and shut-down stages. This situation can conduct to early cracks, generating failures leading to
shut-down of steam turbines. The stresses in the blades are caused by a combination of two loads: tension because of centrifugal
forces and bending due to steam forces. High destructive dynamic forces appear if the natural frequencies of blades are in resonance
with frequencies harmonics. When resonance occurs in combination with LCF, the yield strength of the material is reached and some
regions are exposed to plastic deformations. The repetitive plastic deformations are the main cause of failure in LCF. Blade failures by
LCF are a common problem in steam turbines of power plants [2]. Different experimental tests and numerical models were performed
where damage accumulation was obtained in LCF. Their approach was based on the elastic and the elastic-plastic analysis [3–5].
Radhakrishnan et al. presented an analysis of cumulative damage in LCF based on absorbed plastic energy [6]. Kumar et al.


Corresponding author.
E-mail address: jarr@uaem.mx (J.A. Rodríguez).

https://doi.org/10.1016/j.engfailanal.2019.01.015
Received 12 August 2018; Received in revised form 12 December 2018; Accepted 2 January 2019
Available online 02 January 2019
1350-6307/ © 2019 Elsevier Ltd. All rights reserved.
S. Cano et al. Engineering Failure Analysis 97 (2019) 579–588

studied the evolution low cycle fatigue damage in turbine blade using a finite element approach. They found that LCF causes
sufficient stiffness degradation of blades [7]. Keum suggested that the energy density of plastic strain was a damage parameter and
showed good correlation in low-cycle fatigue for grey cast iron [8].
Chen et al. presented the models of modified linear damage summation method and the modified strain range partitioning method
to conduct life prediction of turbine blades [9]. Mestánek conducted a low cycle fatigue analysis of low-pressure steam turbine blades.
The blades were cyclically loaded by the centrifugal force because of the repeated start-ups of the turbine [10]. Suhas and Anwar
investigated fatigue and creep in blades of the last stage of a steam turbine. They developed a method to assess low cycle fatigue life
of the blades and to determine the number of start-ups and shut-down cycles to failure [11]. Shlyannikov et al. presented a study on
the fatigue life of cracked steam turbine blades which operated under cyclic loading conditions. They also presented an experimental
study using round bar samples of the steam turbine blades with single edge notch. The samples were subject to low cycle fatigue tests
using a loading test cycle similar to the start-stop steam turbine cycle. [12]. Tulsidas et al. carried out life estimation of a steam
turbine blade using linear elastic FEA and elastic plastic FEA [13]. Yu et al. presented a new local stress-strain range approach based
on elasto-plastic finite element analysis and Neuber rule. This approach and the accumulated damage rule were applied to compute
the LCF life of steam turbine long blades [14]. Ince proposed a fatigue damaged model based on the distortional strain energy energy
to cmpute fatigue predictions for Incoloy 901 superalloy, 120–90-02 ductile cast iron and 7075-T651 aluminum alloy [15]. Ba-
naszkiewicz used the equivalent strain energy density rule to online to compute and monitor low cycle fatigue of a steam turbine
rotor [16]. Also various numerical models based on probabilistic methods have been used to compute LCF life prediction of turbine
components [17–21].
In this work, a technique to assess fatigue life of the blades and to determine the number of cycles to the crack initiation was
developed. An experimental and numerical fatigue analysis to the last stage blades of a 110 MW steam turbine was done under the
condition of LCF. The blades useful life was computed based on Strain-life, Morrow and Smith Watson Topper (SWT) methods. In
these last two methods, the mean stress in the blades is considered. The results revealed that centrifugal loads, due to the high level of
stress generated could cause the initiation and propagation of cracks. The number of load cycles required to appear cracks were
calculated. The steam loads generated low stresses in zones where many failures are presented. The results have shown that SWT
present better results to estimate useful life under different load conditions.

2. Experimental and numerical analysis

2.1. Experimental setup

The material of turbine blades was stainless steel AISI 410 and the mechanical properties of this steel are shown in Table 1. An
experimental fatigue analysis was carried out to obtain the SeN curve of this material under normal conditions at the temperature of
25 °C. ASTM E739–91 and manual of model RBF 200 were used for experimental analysis [22–24].
A rotating bending fatigue testing machine was used to determine the strength of materials under the action of fatigue loads.
Dimensions of the specimen are shown in Fig. 1 and the Eq. (1) shows the relation between the moment (Mmax) and the stress (σ)
[24].

Mmax = D
32 (1)
The moment may be determined for different stresses. D is the diameter of the specimen. The loads were calculated with Eq. (2),
where P is the load and dAB is the distance between bearings with value = 0.225. The fatigue strength of a material was calculated
with different tests in rotating beam, using a frequency of 20 Hz and loads between 15 and 50 N.
2Mmax
P=
dAB (2)
The fatigue strength of a material was calculated with different tests in rotating beam fatigue. Basquin's relationship defines two
zones, the fatigue in brittle fracture and the fatigue due to yielding. The results of stress-life (SeN) data were plotted using a log–log
scale (Fig. 2). The fatigue failure surface was used to find the fatigue properties, b and σ'f, which are the fatigue strength exponent
and fatigue strength coefficient respectively [25,26]. Regarding the fatigue empirical constants ɛ'f and c are used as proposed by Rao
[27] (Table 2).

Table 1
Mechanical properties AISI 410 stainless steel [23].
Property Magnitude

Density, ρ (kg/m3) 7800


Elastic Module, E (GPa) 200
Tensile strength, Sut (MPa) 480
Creep strength, Sy (MPa) 275
Elongation, EI (%) 20
Reduction area, RA (%) 45
Hardness, HB 223

580
S. Cano et al. Engineering Failure Analysis 97 (2019) 579–588

Fig. 1. Specimen dimensions in mm.

Fig. 2. S-N diagram AISI 410 stainless steel.

Table 2
Fatigue properties AISI 410 stainless steel [23].
Property Magnitude

Fatigue strength coefficient, σ'f (MPa) 825


Fatigue ductility coefficient, ɛ'f 0.5978
Fatigue strength exponent, b −0.085
Fatigue ductility exponent, c −0.62
Cyclic strain hardening exponent, n' 0.1371
Cyclic strength coefficient, K' (MPa) 885.28

2.2. Numerical setup

The LCF analysis for the last stage of steam turbine blades was carried out using the finite element approach in the program
Ansys®, employing the fatigue module. This steam turbine stage has 55 nozzles and 110 blades. The blades form a wheel attached to
the shaft in groups of 10 blades. Each blade has a length of 0.60 m. In Fig. 3 a turbine blade and its geometrical model for finite
element analysis are shown. The geometrical model was meshed using SOLID185 element type.

2.3. Damage calculation in steam turbine blades

2.3.1. Boundary conditions


Two load cases were considered: a) the blades were under a load of steam forces and b) the centrifugal forces were taken into
account acting during the start-up and shut-down of the steam turbine blades. The steam forces on blades under operating conditions
were calculated using a CFD computation in a previous work [28]. The steam forces behave as a harmonic function of time (Fig. 4).
The maximum and minimum of this harmonic function are presented in Table 3. The steam forces were applied in the centroid of
each blade through local coordinates (See Fig. 5).
The centrifugal load and the centrifugal stress during a cycle of start-up and shut-down is represented as in Fig. 6. So a centrifugal
load history for the steam blades after a series of start-ups and shut-downs over the useful life of a steam turbine could be represented
as a harmonic function. As observed in Fig. 6, the centrifugal stress changes proportionally to the angular speeds of the turbine. Data

581
S. Cano et al. Engineering Failure Analysis 97 (2019) 579–588

Fig. 3. (a) Steam Turbine blade, (b) geometrical model.

Fig. 4. History of load due to steam forces.

Table 3
Maximum and minimum steam forces in components [26].
Maximum force (N) Minimum force (N)

Total force, FR 231.8 220.30


Tangential force, Ft 186.22 175.65
Axial force, Fa 131.80 125.85
Radial force, Fr 41.38 43.12

from the centrifugal load on the steam blades were considered in the fatigue analysis.

2.3.2. Damage calculation


The damage was calculated according to Palmgren-Miner Rule, Eq. (3).
i
ni
D=
i=0
Nf , i (3)

where D is the total damage, Nf is the estimated life and ni is the number of loading cycles. D can take values between 0, corre-
sponding to the startup of cyclic loading; and 1 when a failure occurs [29]. Three methods were used for fatigue damage analysis,
Strain-Life, Eq. (4). Morrow, Eq. (5) and Smith-Watson-Topper, SWT, Eq. (6). It is important to have a comparative parameter when
used three different methods to evaluated fatigue damage. The comparative parameter is ni = 108 cycles, infinite life. [30–32]. These
methods were used because their results are more precise with respect to other methods and equations.
,
f ,
= (2Nf ) b + (2Nf )c
2 E f (4)

582
S. Cano et al. Engineering Failure Analysis 97 (2019) 579–588

Fig. 5. Nodal forces in blades.

Fig. 6. Centrifugal load and stresses at steam turbine blades during start-up and shut-down.

,
f m ,
= (2Nf ) b + (2Nf )c
2 E f (5)
,2
f , ,
= (2Nf ) 2b + (2Nf )b + c
max
2 E f f (6)
LCF was analyzed considering steam loads and centrifugal forces for angular speeds since 0 to 3600 RPM. Statics stresses were
computed and showed for 500, 1000, 1500, 2000, 2500, 3000 and 3600 RPM, where the last one is the operation speed of the
turbine.
The present analysis of estimating the life of the blades by deformation-life equation requires reducing the state of triaxial stresses.
The Von Mises' theory was used for this reduction. Table 4 shows the settings for the Ansys® fatigue module [33].

Table 4
Fatigue tools settings.
Group of 10 blades

Steam forces Centrifugal forces

Type of analysis Deformation-life


Type of load R = 0.9462 Zero-base
Mean stress correction None
Morrow
SWT
Stress Von-Mises

583
S. Cano et al. Engineering Failure Analysis 97 (2019) 579–588

Fig. 7. Maximum stress due steam forces.

3. Results

3.1. Stresses by steam forces

The effect of the steam forces on a group of ten blades is shown in Figs. 7 and 8. The effect of these forces are low stresses
compared with the effect of centrifugal load but these loads can be considered for useful life estimated and damage. The steam forces
on the steam turbine blades vary in time as harmonic function during normal operation. Maximum and minimum stresses occur when
the maximum and minimum steam loads are applied, as observed in Fig. 7 and Fig. 8, respectively. In both cases, maximum stresses
are located near to the blade root.

3.2. Centrifugal stresses

The centrifugal stresses were computed for different rotation speeds since 0 to 3600 RPM, tabulated some values in Table 5. For
these conditions, maximum centrifugal stresses were obtained for a maximum speed of 3600 RPM. The stress distribution in the group
due to centrifugal load is shown in Fig. 9. The maximum centrifugal stress is located approximately on 3/4 of the blade length.

3.3. Effects of steam forces on the useful life

The results of damage and useful life to a group of blades using equation strain-life, Morrow and SWT correction are presented in
Table 6. In all cases, life's cycles are very high and the damage is extremely small. Therefore, the effect of steam forces on the turbine
blades analyzed is very small.

Fig. 8. Minimum stress due steam forces.

584
S. Cano et al. Engineering Failure Analysis 97 (2019) 579–588

Table 5
Von-Mises stresses obtained at different speeds of operation.
Rotational velocity (RPM) Maximum Von-Mises stresses (Pa)

500 1.4655 × 107


1000 5.8618 × 107
1500 1.3189 × 108
2000 2.3447 × 108
2500 3.6636 × 108
3000 5.2756 × 108
3600 7.5969 × 108

Fig. 9. Centrifugal stresses in blades (σVM = 76 MPa).

Table 6
Estimated life and damage due to steam forces.
Strain-life Morrow SWT

Cycles Damage Cycles Damage Cycles Damage

39 −31 39 −31 30
3.759E10 2.660 × 10 3.1552 × 10 3.169 × 10 2.1653 × 10 4.618 × 10−22

3.4. Effects of centrifugal force on the useful life

The effect of centrifugal force on Nf life was got with the strain-life relationship, Morrow, and SWT correction and also the
corresponding damage was calculated (Table 7). It is shown that while the RPM is increased, Nf life is decreased and the damage is
increased. The Morrow correction estimated 22,270 cycles of life and a comparative damage of 44,903 was obtained, while the SWT
correction computed 14,320 cycles and damage of 69,832 for the same speed.

Table 7
Estimated life and damage due to effects of centrifugal forces.
Velocity Strain-life Morrow SWT

RPM Cycles Damage Cycles Damage Cycles Damage

500 6.83 × 1023 1.46 × 10−15 6.15 × 1023 1.63 × 10−15 1.16 × 1022 8.64 × 10−14
1000 5.64 × 1016 1.77 × 10−8 3.686 × 1016 2.71 × 10−8 9.56 × 1016 1.045 × 10−6
1500 4.056 × 1012 2.47 × 10−4 1.52 × 1012 6.56 × 10−4 6.91 × 1010 1.45 × 10−2
2000 4.68 × 109 0.214 8.52 × 108 1.1738 9.286 × 107 10.769
2500 2.603 × 107 38.42 4.24 × 106 235.67 1.304 × 106 767.18
3000 5.43 × 105 1842.6 2.032 × 105 4922.5 1.030 × 105 9708
3600 31,839 31,408 22,270 44,903 14,320 69,832

585
S. Cano et al. Engineering Failure Analysis 97 (2019) 579–588

Fig. 10. Minimum life in blades. Life at 3600 RPM in blades using SWT correlation.

Fig. 11. Maximum damage in blades. Damage at 3600 RPM in blades using SWT correlation.

In Figs. 10 and 11, the locations of minimum life and maximum damage in the group of blades are shown. The maximum damage
is located among the bonding wires and it can be identified as the most critical point of the blades.
In Fig. 12, the graph of stress vs. estimated life is showed. The magnitude of the stresses vs. estimated life was got with the

1.00E+09

Morrow
Stress/ [Pa]

SWT
1.00E+08

1.00E+07
1.00E+00 1.00E+23
Life / Nf [Cycles]
Fig. 12. Stress vs estimated life.

586
S. Cano et al. Engineering Failure Analysis 97 (2019) 579–588

8.10E+08

7.10E+08

Morrow 6.10E+08

Stress/ [Pa]
SWT 5.10E+08

4.10E+08

3.10E+08

2.10E+08

1.10E+08

1.00E+07
1.00E+00 Damage 2.00E+04

Fig. 13. Stress vs damage of blades.

theories of strain-life, Morrow, and SWT. In this Fig., the tendency of estimated life by the applied load is shown with the dotted line.
At low levels of stresses, the estimated Nf life computed with the three theories mentioned has a great difference. While at high levels
of stresses, the difference between estimated life seems too small. In Fig. 13, the stresses versus damage of blades are shown. The
dotted line marks the tendency of the damage depending on the level of stress applied to the blades. In this case, the computed
damage differs at high-stress levels. The SWT correction computed the bigger damaged for a certain level of stress.

4. Conclusions

The fatigue life of steam turbine blades and useful life was evaluated in condition of LCF. Steam forces do not generate high
stresses or enough damage in order to represent a serious risk of crack in the steam turbine blades. However, this work revealed that
stresses due to centrifugal forces might contribute to the initiation and propagation of cracks. A minimum number of cycles required
to appear cracks in blades were calculated. The results showed that maximum damage in the group of blades was located in the
second blade between bonding wires, approximately at 40 cm from the root when the yield strength of the material is achieved and
the plastic deformation are presented. The damage located at 15 cm from the root of the blades was very high in according to results
reported in previous studies [34–37]. Therefore, cracks are expected to appear in that region. The model SWT has a better prediction
in steam turbine blades for detecting damage due to this model contains more variables of analysis. The results do not correlate with
any other property of the material.

References

[1] A. Fatemi, L.F. Yangt, Cumulative fatigue damage and life prediction theories: a survey of the state of the art for homogeneous materials, Int. J. Fatigue 20 (1)
(1998) 9–34.
[2] M. Zhang, Collections on aeroengine service life and speeding-up simulated trial operation, Liaoning Science Technology Press, Shenyang, 1991.
[3] Benudhar Sahoo, R.K. Satpathy, S.K. Panigrahi, Analysis of a turbine blade failure in a military turbojet engine, Int. J. Turbo Jet Eng. 33 (2) (2015) 1–9.
[4] S.M. Beden, S. Abdullah, A.K. Ariffin, N.A. Al-Asady, M.M. Rahman, Fatigue life assessment of different steel-based shell materials under variable amplitude
loading, Eur. J. Sci. Res. 29 (2) (2009) 157–169.
[5] Cyrus B. Meher-Homji, Gas turbine axial compressor fouling: a unified treatment of its effects, detection, and control, Int. J. Turbo Jet Eng. 9 (1992) 331–334.
[6] V.M. Radhakrishnan, Cumulative damage in low-cycle fatigue, Exp. Mech. 18 (8) (1978) 292–296.
[7] S. Kumar, N. Roy, R. Ganguli, Monitoring low cycle fatigue damage in turbine blade using vibration characteristics, Mech. Syst. Signal Process. 26 (2005)
480–501.
[8] L. Keum, L. Soon, Definition of damage parameter in low-cycle fatigue of gray cast iron, Key Eng. Mater. 345–346 (2007) 367–370.
[9] L. Chen, Y. Liu, L. Xie, Power-exponent function model for low-cycle fatigue life prediction and its applications – Part II: Life prediction of turbine blades under
creep–fatigue interaction, Int. J. Fatigue 29 (1) (2007) 10–19.
[10] P. Mestanek, Low cycle fatigue analysis of a last stage steam turbine blade, Appl. Comput. Mech (2008) 71–82.
[11] B. Suhas, K. Anwar, Fatigue and creep interaction in steam turbine bladed disk, Int. J. Innov. Res. Sci. Eng. Technol. (2014) 13597–13603.
[12] V.N. Shlyannikov, R.R. Yarullin, A.P. Zakharov, Fatigue of steam turbine blades with damage on the leading edge, Procedia Mater. Sci. 3 (2014) 1792–1797.
[13] D. Tulsidas, M. Shantharaja, V.G. Bharath, Life estimation of a steam turbine blade using low cycle fatigue analysis, Procedia Mater. Sci. 5 (2014) 2392–2401.
[14] D. Yu, A. He, H. Yang, J. Yang, A LCF Life assessment method for steam turbine long blade based on elastoplastic analysis and local strain approach, ASME Turbo
Expo, Turbine Technical Conference and Exposition, Montreal, Quebec, Canada, June 15–19, 2015, Paper No. GT2015-43955, ASME, New York, 2015.
[15] A. Ince, A generalized mean stress correction model based on distortional strain energy, Int. J. Fatigue 104 (2017) 273–282.
[16] M. Banaszkiewicz, The low-cycle fatigue life assessment method for online monitoring of steam turbine rotors, Int. J. Fatigue 113 (2018) 311–323.
[17] L. Chen, Y. Liu, L. Xie, Power-exponent function model for low-cycle fatigue life prediction and its applications—part I: models and validations, Int. J. Fatigue 29
(1) (2007) 1–9.
[18] S.P. Zhu, H.Z. Huang, R. Smith, et al., Bayesian framework for probabilistic low cycle fatigue life prediction and uncertainty modeling of aircraft turbine disk
alloys, Probab. Eng. Mech. 34 (2013) 114–122.
[19] S.P. Zhu, H.Z. Huang, W. Peng, et al., Probabilistic physics of failure-based framework for fatigue life prediction of aircraft gas turbine discs under un-certainty,
Reliab. Eng. Syst. Saf. 146 (2016) 1–12.
[20] H. Gao, G. Bai, Y. Gao, et al., Reliability analysis for aeroengine turbine disc fatigue life with multiple random variables based on distributed collaborative
response surface method, J. Cent. S. Univ. Technol. 22 (12) (2015) 4693–4701.
[21] H. Gao, A. Wang, G. Bai, C. Wei, C. Fei, Substructure-based distributed collaborative probabilistic analysis method for low-cycle fatigue damage assessment of
turbine blade–disk, Aerosp. Sci. Technol. 79 (2018) 636–646.
[22] ASTM E739-91, Standard practice for statistical analysis of linearized stress life (S-N) and strain life (ε-N) Fatigue data, (2004).
[23] J.E. Bringas, The Metals Black Book, 2nd ed., Casti Publishing Inc, Edmonton, Canada, 1997.
[24] Instruction Manual, Model RBF-200, Rotating Beam Fatigue Testing Machine, Fatigue Dynamics, Inc. P.O. Box 2533, Dearborn, MI, 48123, (2000).
[25] M. Clemente, Experimental and Numerical Analysis in Cracks of Stainless Steel in Corrosive Environment with Application in Steam Turbine: In Spanish, Ph.D.

587
S. Cano et al. Engineering Failure Analysis 97 (2019) 579–588

Thesis Universidad Autónoma del Estado de Morelos, México, 2017.


[26] K.S. Lee, J.H. Song, Estimation methods for strain-life fatigue properties from hardness, Int. J. Fatigue 28 (2006) 386–400.
[27] J.S. Rao, Turbine Blade Life Estimation, 1st ed., Alpha Science International Ltd., New Delhi, India, 2000.
[28] J.C. García, Determination of Forces Induced by Steam Flow in Turbines: In Spanish, Ph.D. Thesis Universidad Autonoma del Estado de Morelos, México, 2008.
[29] A. Fatemi, L.F. Yangt, Cumulative fatigue damage and life prediction theories: a survey of the state of the art for homogeneous materials, Int. J. Fatigue 20 (1)
(1998) 9–34.
[30] Swanson Jhons, Ansys® (Version 16.1) [Software], (2016).
[31] N.E. Dowling, Mean stress effects in strain–life fatigue, Fatigue Fract. Eng. Mater. Struct. 32 (2009) 1004–1019.
[32] A. Ince, G. Glinka, A modification of Morrow and Smith–Watson–Topper mean stress correction models, Fatigue Fract. Eng. Mater. Struct. (2011) 854–867.
[33] S. Cano, Experimental and numerical simulation of damage in steam turbine blades under LCF: in Spanish, MSc. Thesis Universidad Autónoma del Estado de
Morelos, México, 2016.
[34] J. Kubiak, R. González, D. Juarez, J. Nebradt, F. Sierra, An investigation on the failure of an L-0 steam turbine blade, J. Fail. Anal. Prev. 4 (2004) 47–51.
[35] J.A. Rodríguez, Y. El Hamzaoui, J.A. Hernández, J.C. García, J.E. Flores, A.L. Tejeda, The use of artificial neural network (ANN) for modeling the useful life of the
failure assessment in blades of steam turbines, Eng. Fail. Anal. 35 (2013) 562–575.
[36] Hamzaoui Y. El, J.A. Rodríguez, J.A. Hernández, V. Salazar, Optimization of operating conditions for steam turbine using an artificial neural network inverse,
Appl Thermal Eng. 75 (2015) 648–657.
[37] J.A. Rodríguez, J.C. Garcia, E. Alonso, Hamzaoui Y. El, J.M. Rodríguez, G. Urquiza, Failure Probability Estimation of Steam Turbine Blades by Enhanced Monte
Carlo Method, 56 (2015), pp. 80–88.

588

You might also like