Fatigue Failure of a Composite Wind Turbine Blade at Its Root End

Embed Size (px)

Citation preview

  • 8/18/2019 Fatigue Failure of a Composite Wind Turbine Blade at Its Root End

    1/8

    Fatigue failure of a composite wind turbine blade at its root end

    Hak Gu Lee ⇑, Min Gyu Kang, Jisang Park

    Wind Turbine Technology Research Center, Korea Institute of Materials Science, 797 Changwondaero, Changwon, Gyeongnam 641-831, Republic of Korea

    a r t i c l e i n f o

     Article history:

    Available online 7 August 2015

    Keywords:

    Fatigue

    Failure

    Delamination

    Wind turbine blade

    Bumping motion

    a b s t r a c t

    As blade failures at wind farms have increased, the structural safety of composite wind turbine blades is

    ever more important. The recent implementation of considerably larger blades has made the problem

    even more crucial. One of the critical failure modes is the blade root failure, which can result in the bladebeing pulled out from its wind turbine during operation. In this study, we experienced delamination fail-

    ure at the blade root during fatigue testing of a 3 MW full-scale wind turbine blade according to interna-

    tional standard IEC 61400-23: full-scale structural testing of rotor blades. Comparing the measured data

    with the FE analysis results, we simulated the situations the blade had experienced, and then found what

    caused the delamination failure as well as the problem of the conventional design approach. The bump-

    ing motions of the blade shell caused by geometric complexities between the maximum chord and the

    root alter significantly the load distribution at the end of the blade root. Therefore, to enhance the struc-

    tural safety of a large composite wind turbine blade, a more detailed FE analysis on the blade root in the

    design stage is needed.

      2015 Elsevier Ltd. All rights reserved.

    1. Introduction

    With the recent trend toward large slender wind turbine blades,

    questions are being raised regarding their reliability. In order to

    evaluate the static strengths and fatigue lives of these larger

    blades, static and fatigue tests of full-scale prototype blades should

    be conducted according to international standards or equivalent

    guidelines [1–4]. Testing methodologies developed so far are well

    described in the two review papers of Malhotra et al.   [5]   and

    Yang et al.  [6].

    Previous studies pertaining to static strength of a full-scale

    wind turbine blade are as follows. Jensen et al.  [7]  tested a 34 m

    composite wind turbine blade until its structural collapse.

    Debonding of the outer skin was the initial failure mechanism, fol-

    lowed by delamination buckling which led to the blade’s collapse.

     Jensen et al. believed the main root cause was the Braizer effect of 

    the shell structure due to bending. Overgaard et al.   [8–9]  carried

    out a static flapwise bending test of a 25 m wind turbine blade

    to collapse. The Brazier effect had a large influence on the local

    out-of-plane deflection, but its influence on the longitudinal strain

    level in the primary load-carrying laminate was insignificant.

    Overgaard et al. assert that the structural stability of the generic

    wind turbine blade has been governed by buckling and the

    delamination phenomena. Yang et al.  [10] conducted a static test

    of a 40 m wind turbine blade under flapwise loading to collapse.

    Yang et al. concluded the Brazier effect was not the dominant

    failure mechanism, but debonding between the pressure-side and

    the suction-side aerodynamic shells was the initial failure mecha-

    nism followed by its instable propagation which leads to collapse.

    Previous studies pertaining to fatigue of wind turbine blades are

    divided into two categories: a material fatigue behavior and a

    structural fatigue behavior. The material fatigue behavior has

    extensively studied with uni-axial, in-plane loading of balanced

    and symmetrical, relatively thin laminates   [11,12], but they are

    only remotely representative for blade structures [13]. Tests with

    a full-scale wind turbine blade to study the structural fatigue

    behavior are so expensive that few studies have been conducted

    to date. Leeuwen et al. [14] had carried out fatigue tests of 37 wind

    turbine blades 3.4 m in length as well as 35 coupons to compare

    fatigue strength from full-scale blade tests with coupon-based

    predictions. Flapwise failures occurred at the tensile side, but edge-

    wise failures were the result of crack initiation starting in the

    bonding at the trailing edge followed by further crack propagation

    in the laminate. Blade fatigue data compared with coupon data

    fitted reasonably with flapwise tests, but they did not compare

    well for edgewise tests. Marín et al.   [15,16]   inspected fatigue

    damage of a 300 kW wind turbine blade, and then performed a

    FE analysis to reveal the root cause of the fatigue damage. The

    crack initiated at the abrupt geometric-transient region between

    the root zone and the aerodynamic zone had been propagated into

    the laminate. It should be noted that the aforementioned studies

    http://dx.doi.org/10.1016/j.compstruct.2015.08.010

    0263-8223/ 2015 Elsevier Ltd. All rights reserved.

    ⇑ Corresponding author. Tel.: +82 55 280 3261; fax: +82 55 280 3498.

    E-mail address: [email protected] (H.G. Lee).

    Composite Structures 133 (2015) 878–885

    Contents lists available at  ScienceDirect

    Composite Structures

    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 / c o m p s t r u c t

    http://dx.doi.org/10.1016/j.compstruct.2015.08.010mailto:[email protected]://dx.doi.org/10.1016/j.compstruct.2015.08.010http://www.sciencedirect.com/science/journal/02638223http://www.elsevier.com/locate/compstructhttp://www.elsevier.com/locate/compstructhttp://www.sciencedirect.com/science/journal/02638223http://dx.doi.org/10.1016/j.compstruct.2015.08.010mailto:[email protected]://dx.doi.org/10.1016/j.compstruct.2015.08.010http://crossmark.crossref.org/dialog/?doi=10.1016/j.compstruct.2015.08.010&domain=pdf

  • 8/18/2019 Fatigue Failure of a Composite Wind Turbine Blade at Its Root End

    2/8

    dealt with small wind turbine blades that have relatively higher

    fatigue margins than the large slender wind turbine blades

    presently in development.

    Another approach to enhance the reliability of a wind turbine

    blade is an FE fatigue simulation for the blade. Kong et al.  [17,18]

    designed a 750 kW wind turbine blade, factoring in its fatigue life

    of 20 years based on the well-known S–N linear damage equation,

    the load spectrum, and Spera’s empirical equations. Shokrieh et al.[19] performed a case study with a 23 m wind turbine blade. Using

    its FE shell model with a stochastic approach on fatigue loads, the

    fatigue life was bounded between 18.66 years and 24 years as

    lower and upper limits. Toft et al.  [20] estimated the reliability of 

    a wind turbine blade for a single failure mode, considering statisti-

    cal uncertainties.

    Despite such contributions by many researchers, about 30 blade

    failures are occurring per year throughout the world, and the num-

    ber of blade failure are increasing over time [21]. There have been

    blade failures not observed in previous studies which have

    occurred in the field at Eclipse wind farm and Ocotillo wind farm

    in 2013. In these instances, the wind turbine blades were pulled

    out from the wind turbines due to delamination at the root. The

    authors of this paper have also experienced a similar phenomenon

    during fatigue testing of a 3 MW full-scale wind turbine blade. To

    find a root cause of the phenomenon, loading conditions calculated

    by its FE shell model were applied to the more detailed FE solid

    model of the root subcomponent. Comparing the analysis results

    with measured strain data of the T-bolt, we adjusted loading con-

    ditions of the FE solid model to simulate deformations and stress

    distributions of the blade root. Based on the simulation results, this

    study has found one of the plausible root causes able to incur

    delamination at the root of a wind turbine blade.

    2. Test blade and its failure during fatigue testing 

    The test blade is a 3 MW glass/epoxy composite blade as shown

    in Fig. 1. Blade length and weight are 56 m and 14.5 ton, respec-

    tively. It has been developed as a result of a R&D project fundedby Korean government. After mounting the test blade on a stand

    fixture like a horizontal cantilever beam, the fatigue test setups

    including an aerodynamic fairing, two additional masses, a flap-

    wise exciter, and an edgewise exciter were attached on the test

    blade. Then we carried out a dual-axis resonance fatigue test of 

    the blade using two different resonance frequencies according to

    international standard IEC 61400-23 [1]. The tip motion of the test

    blade during dual-axis resonance fatigue testing is shown in Fig. 2.

    After flapwise 510,000 cycles under the equivalent amplitude of 

    5352 kNm and the mean of 5970 kNm and edgewise 780,000

    cycles under the equivalent amplitude of 4454 kNm and the mean

    of 0 kNm at the end of the blade root, fatigue failure was found asshown in Fig. 3.

    Strain values of the T-bolt, located as shown in Fig. 4, were mea-

    sured during the fatigue testing. The shape and specification of the

    T-bolt are represented in Fig. 5 and Table 1. The graphs in Fig. 6 are

    the measured strain values of the two different strain gages

    attached at the same cross section of the T-bolt. In each graph

    the amplitude of tensile strains larger than that of compressive

    strains means separation of the T-bolt joint has occurred.

    Fig. 1.   3 MW test blade 56 m in length.

    Fig. 2.   Tip motion during dual-axis resonance fatigue testing.

    Fig. 3.  Failure at the end of the blade root.

    H.G. Lee et al. / Composite Structures 133 (2015) 878–885   879

    http://-/?-http://-/?-

  • 8/18/2019 Fatigue Failure of a Composite Wind Turbine Blade at Its Root End

    3/8

    Furthermore, the tensile or compressive amplitude of strain gage 2

    being larger than that of strain gage 1 means bending of the T-bolt.

    Separation and bending of the T-bolt joint are unexpected

    phenomena that must be avoided in the design stage. Before going

    into the detailed root cause analysis, the conventional approach

    for a blade root design will be explained in the next section to

    understand why the approach failed to predict the separation

    and bending observed.

    3. Conventional blade root design

    Fig. 7   represents a schematic diagram of a blade root part

    including a pitch bearing assembled by T-bolts and nuts. The blade

    root is a very thick composite laminate able to enclose T-bolts and

    cross nuts, its thickness being about 100 mm. Thus, blade designers

    have believed that, compared with blade shell sandwich structures

    whose thickness are less than about 30 mm, a blade root is so stiff 

    that the stress distribution in it is similar to that of a hollow circu-

    lar cylinder structure when subjected to bending. Based on this

    presumption, the distributions of local moments and axial stresses

    or forces have been calculated as shown in Fig. 7. This conventional

    approach for the test blade gives 14 kNm for the maximum local

    moment and 267 kN for the maximum axial force at each T-bolt

     joint. The axial force is much smaller than the pretension 340 kN

    for the T-bolts in Table 1. Thus the separation and bending of the

    T-bolt cannot be observed during the fatigue testing based on this

    calculation. The presumption regarding the blade root had worked

    well before wind turbine blades became larger and more slender.

    Fig. 4.   Location of the T-bolt where bolt strains were measured.

    Fig. 5.   Schematic diagram of the M36 T-bolt used in this study.

     Table 1

    Specifications of the M36 T-bolt used in this study.

    Grade Min. diameter

    [mm]

    Pretension

    [kN]

    Prestress

    [MPa]

    Min. yield strength

    [MPa]

    10.9 28 340 552 940

    Fig. 6.   Strain values measured by two different strain gages attached at the same cross section of the T-bolt: (a) strain gage 1 and (b) strain gage 2.

    880   H.G. Lee et al. / Composite Structures 133 (2015) 878–885

  • 8/18/2019 Fatigue Failure of a Composite Wind Turbine Blade at Its Root End

    4/8

    However, the current trend is requiring a more detailed analysis on

    the blade root part.

    4. FE simulation for root cause analysis

    A flowchart of the root cause analysis carried out in this study is

    represented in Fig. 8. Using the test setup and the loading condi-

    tions aforementioned in Section   2, a static analysis with the FEshell model for the test blade was conducted to calculate the

    plausible amplitudes of an axial force and a local moment at each

    T-bolt. The calculated values do not represent the real situation

    during the fatigue testing because the clamped boundary

    conditions used in the FE shell model does not match with the

    separation of the T-bolt joint observed. The axial force and the local

    moment were applied to the subcomponent FE solid model for the

    test blade root, and then we modified them, comparing the calcu-

    lated T-bolt strains with the measured strains during the fatigue

    testing. After several modifications, stress distributions at the

    end of the blade root were able to explain the observed delamina-

    tion that occurred.

    The information on the FE shell model for the test blade is in

    Fig. 9 and  Table 2. The shell model reflects the shape of the test

    blade and laminating sequences of composites. We used a com-

    mercial FE solver, ABAQUS 6.13, and its 4 node shell element,  S4R.

    The number of the elements is 57,969. The boundary conditionwas the clamped condition at the end of the blade root, and the

    loading conditions were the flapwise and the edgewise test bend-

    ing moment distributions along the positive   y- and   x-directions,

    where the positive  y   means a chord direction toward the trailing

    edge in the pitch zero section and the positive  x  means the cross

    product of the positive  y  with the pitch axis.

    Properties of unidirectional glass NCF/epoxy composites used in

    this study are in   Table 4. Four properties were measured from

    coupon tests:   E 1   of 40.14 GPa,   E 2   of 12.30 GPa,   v 12   of 0.26, and

    Fig. 7.   Schematic diagram of a blade root part and a pitch bearing.

    Fig. 8.   Flowchart of the root cause analysis taking the fatigue testing conditions into account.

    H.G. Lee et al. / Composite Structures 133 (2015) 878–885   881

    http://-/?-http://-/?-http://-/?-

  • 8/18/2019 Fatigue Failure of a Composite Wind Turbine Blade at Its Root End

    5/8

    G12   of 3.40 GPa, where the subscripts 1, 2, 3 mean the fiber

    direction, the transverse direction, and the thickness direction,

    respectively. A transversely isotropic material needs 5 independent

    material properties, so we assumed   v 23   of glass NCF/epoxy

    composites as 0.38 compared with S-2 glass/epoxy composites in

    Ref.   [22], the ratio of 1.48 between two Poisson’s ratios   v 12   and

    v 23. Then   G23   can be calculated from Eq.   (1), resulting in

    4.44 GPa, and the other properties   E 3,   v 13, and   G13   are the same

    as E 2,   v 12, and  G12, respectively.

    G23  ¼  E 2

    2ð1 þ v 23Þ  ð1Þ

    The information on the subcomponent FE solid model for the

    test blade root is in Fig. 10 and Table 3. ABAQUS 6.13 was also used

    as the FE solver for the subcomponent model, which was

    constructed with the 8 node solid element,  C3D8I . The number of 

    the element was 53,629. The boundary conditions used in the

    model were the axially symmetric and symmetric condition along

    the hoop direction, the fixed condition along the blade length

    (spanwise) direction on the positions of two bearing ball arrays,

    and the contact conditions with the friction coefficient of 0.3 on

    the several contact surfaces of the T-bolt and the nuts. The loading

    conditions were the axial force and the local moment calculated by

    the FE shell model, which were applied to the cross section of the

    blade root part along the spanwise direction and the hoop direc-

    tion, respectively.

    Equivalent orthotropic properties of the blade root laminate

    whose stacking sequence is [45/0/45]n  were generated for the

    convenience of FE modeling with solid elements. Classical lami-

    nated plate theory (CLPT) cannot calculate whole equivalent ortho-

    tropic properties because even interlaminar stresses at the

    interface of two laminae are discontinuous [23]. CLPT gives us only

    in-plane laminate properties such as   E  x,   E y,   v  xy, and   G xy   and

    through-thickness Poisson’s ratios such as   v  xz   and   v  yz , which are

    calculated from Eqs. (2)–(4) [22].

    v  xz  ¼ð A

    111F 1 þ A

    112F 2 þ A

    116F 6Þ

    2HA111

    ð2Þ

    v  yz  ¼ð A

    121F 1 þ A

    122F 2 þ A

    126F 6Þ

    2HA122

    ð3Þ

    F i  ¼XN 

    k¼1

    S k13Q k1i þ

    S k23Q k2i þ

    S k36Q k6i

    t k

      ði ¼ 1;2;6Þ ð4Þ

    where Aij, S kij ,

     Q kij, t k, and 2H  are the i, j component of  A matrix, the i, j

    component of the transformed compliance matrix in the kth lamina,

    the   i,   j  component of the transformed reduced stiffness matrix in

    the   kth lamina, the thickness of the   kth lamina, and the totalthickness of the laminate, respectively. The six properties were

    calculated using Table 4 and the stacking sequence [45/0/45]s, in

    which the thicknesses of 45/45 lamina and 0 lamina were

    0.15 mm and 0.60 mm, respectively. The large staking number of 

    the unsymmetric laminate [45/0/45]n   makes its properties

    converge into those of the symmetric laminate [45/0/45]s, so we

    used the symmetric stacking sequence instead of the unsymmetric

    one. The remaining three properties,  E z, G xz , and G yz  were calculated

    from a FE cube model whose stacking sequence is [45/0/45]10s. By

    applying normal forces or shear forces to the cube surfaces, we

    obtained a pertinent deformation value at each case. From the

    deformation value and loading conditions, the cube stiffness was

    calculated at each case.   Table 5  shows the calculated equivalent

    orthotropic properties of the blade root laminate.

    Fig. 9.   FE shell model used in this study.

     Table 2

    FE shell model for the test blade.

    FE solver Element type No. of the

    elements

    Boundary condition

    ABAQUS

    6.13

    4 node shell element

    (S4R)

    57,969 Clampe d c ondition at

    the root

    Fig. 10.  FE solid model used in this study.

     Table 3

    Subcomponent FE solid model for the test blade root.

    FE solver Element type No. of the

    elements

    Boundary conditions

    ABAQUS 6.13 8 node solid

    element(C3D8I)

    53,629 Axially symmetric condition

    Symmetric condition

    Fixed condition

    Contact condition (l = 0.3)

     Table 4

    Properties of the glass NCF/epoxy unidirectional lamina.

    E 1[GPa]

    E 2[GPa]

    E 3[GPa]

    v 12   v 13   v 23   G12[GPa]

    G13[GPa]

    G23[GPa]

    40.14 12.30 12.30 0.26 0 .26 0 .38 3 .40 3.40 4.44

     Table 5

    Properties of the [45/0/45]n  laminate.

    E  x[GPa]

    E  y[GPa]

    E  z [GPa]

    v  xy   v  xz    v  yz    G xy[GPa]

    G xz [GPa]

    G yz [GPa]

    30.69 13.44 12.70 0.42 0 .21 0 .34 6 .26 3.55 4.21

    882   H.G. Lee et al. / Composite Structures 133 (2015) 878–885

  • 8/18/2019 Fatigue Failure of a Composite Wind Turbine Blade at Its Root End

    6/8

    5. Analysis results

    The analysis result of the FE shell model shows that the

    maximum values of the axial force and the local moment per each

    T-bolt are greater than those of the conventional model. As shown

    in Fig. 11(a), the blade root end near 110 degrees from the leading

    edge receives severe tension, 387 kN, even larger than pretension

    of the T-bolt, 340 kN, resulting in the separation of the T-bolt joint.

    The local moment and the axial force in the FE shell model greaterthan those in the conventional model in   Fig. 11   come from the

    inward bumping motion in Fig. 12, and the lower values come from

    the outward bumping motion. The strangest location for these

    differences is the trailing edge. The location is expected to move

    inward because the flapwise and the edgewise bending moment

    are applied along the positive   y- and the positive   x-direction.

    However, it moved outward, as shown in   Fig. 12, resulting in a

    small axial force and a small local moment. This opposite moving

    direction at the trailing edge may be caused by the geometric com-

    plexities from the maximum chord of the blade to the root. In the

    same cross section of the blade, the alleviation of the axial force

    and the local moment in some location incurs an increase of the

    axial force and the local moment in another location. Thus the

    bumping motions of the blade are thought to be the main reasonof the unfavorable load distribution at the end of the blade root.

    Fig. 11.   Comparison between the results of the FE shell model and those of the conventional approach: (a) axial force distribution and (b) local moment distribution.

    Fig. 12.   Bumping motion of the blade shell during the fatigue testing.

    Fig. 13.   Comparison between the measured and the calculated strain ranges.

    H.G. Lee et al. / Composite Structures 133 (2015) 878–885   883

    http://-/?-http://-/?-http://-/?-

  • 8/18/2019 Fatigue Failure of a Composite Wind Turbine Blade at Its Root End

    7/8

    Fig. 14.   Partial separation between the blade root and the pitch bearing.

    Fig. 15.   Stress distributions that incur delamination at the end of the blade root.

    Fig. 16.   Schematic diagram of the blade root failure caused by delamination followed by crack propagation.

    884   H.G. Lee et al. / Composite Structures 133 (2015) 878–885

  • 8/18/2019 Fatigue Failure of a Composite Wind Turbine Blade at Its Root End

    8/8

    The subcomponent FE solid model simulated well the T-bolt

     joint when applying 100% of the axial force and 73% of the local

    moment at the location of 90 degrees. The measured and

    calculated strain values at the top and bottom positions of the

    T-bolt are shown in Fig. 13; the measured strain value at the top

    position was extrapolated based on the two measured data of dif-

    ferent strain gages. The simulation shows us partial separation

    between the blade root and the pitch bearing as shown inFig. 14. Furthermore, the interlaminar shear stress,   r23, and the

    peel stress, r33, are very severe when the partial separation occurs.

    The positions of the two severe stresses are well matched with the

    observed delamination positions as shown in Fig. 15. Therefore, we

    conclude that the delamination at the end of the blade root due to

    partial separation followed by crack propagation into the root

    laminate, as shown in   Fig. 16, would bring about the failures in

    which blades are pulled out from their wind turbines.

    6. Conclusion

    In this study we have experienced delamination failure at the

    end of the blade root during its full-scale fatigue testing. To find

    what caused the failure, FE analyses were carried out using a sub-

    component FE solid model as well as a full-scale FE shell model.

    The analysis results reveal that for a slender and large wind turbine

    blade the real load distribution at the root is very different from

    that calculated by the conventional approach, which assumes the

    blade root has enough stiffness to be modeled as a bending of a

    hollow circular cylinder. The bumping motions of the blade shell

    alter load distribution at the end of the blade root, resulting in

    the alleviation of load in some locations and the increase of load

    in other locations. The severe increase of load incurs partial

    separation of the T-bolt joints followed by delamination at the

    end of the root, which may lead to pulling of the blade out from

    its wind turbine during operation. Therefore, detailed analyses on

    the blade root should be carried out to enhance its structural safety

    especially for a slender and large wind turbine blade.

     Acknowledgements

    This work was supported by the New & Renewable Energy of 

    Korea Institute of Energy Technology Evaluation and Planning

    (KETEP) grant funded Korea government Ministry of Trade,

    Industry and Energy (No. 2012T100201707).

    References

    [1] International Electrotechnical Commission. International standard IEC 61400-23 wind turbines – part 23: full-scale structural testing of rotor blades. 2014.

    [2] Det Norske Vertias. Standard DNV-DS-J102 Design and manufacture of windturbine blades, offshore and onshore wind turbines. 2010.

    [3] Germanischer Lloyd. Rules and guidelines IV industrial services 1 guideline forthe certification of wind turbines. 2010.

    [4] Germanischer Lloyd. Rules and guidelines IV industrial services 2 guideline for

    the certification of offshore wind turbines. 2012.[5]  Malhotra P, Hyers RW, Manwel JF, McGowan JG. A review and design study of blade testing systems for utility-scale wind turbines. Renewable SustainableEnergy Rev 2012;16:284–92.

    [6]   Yang B, Sun D. Testing, inspecting and monitoring technologies for windturbine blades: a survey. Renewable Sustainable Energy Rev 2013;22:515–26.

    [7]  Jensen FM, Falzon BG, Ankersen J, Stang H. Structural testing and numericalsimulation of 34 m composite wind turbine blade. Compos Struct2006;76:52–61.

    [8]   Overgaard LCT, Lund E, Thomsen OT. Structural collapse of a wind turbineblade. Part A: static test and equivalent single layered models. Compos Part A2010;41:257–70.

    [9]   Overgaard LCT, Lund E. Structural collapse of a wind turbine blade. Part B:progressive interlaminar failure models. Compos Part A 2010;41:271–83.

    [10]   Yang J, Peng C, Xiao J, Zeng J, Xing S, Jin J, et al. Structural investigation of composite wind turbine blade considering structural collapse in full-scalestatic tests. Compos Struct 2013;97:15–29.

    [11]   Kensche CW. Fatigue of composites for wind turbines. Int J Fatigue2006;28:1363–74.

    [12] Mandell JF, Samborsky DD, Agastra P, Sears AT, Wilson TJ. Analysis of SNL/MSU/DOE fatigue database trends for wind turbine blade materials. SandiaNational Laboratory report: SAND2010-7052. 2010.

    [13] Nijssen RPL. Fatigue life prediction and strength degradation of wind turbinerotor blade composites. Sandia National Laboratory report: SAND2006-7810P.2006.

    [14]  van Leeuwen H, van Delft D, Heijdra J, Braam H, Jørgensen ER, Lekou D, et al.Comparing fatigue strength from full scale blade tests with coupon-basedpredictions. Trans ASME 2002;124:404–11.

    [15]   Marín JC, Barroso A, París F, Cañas J. Study of damage and repair of blades a 300kW wind turbine. Energy 2008;33:1068–83.

    [16]   Marín JC, Barroso A, París F, Cañas J. Study of fatigue damage in wind turbineblades. Eng Fail Anal 2009;16:656–68.

    [17]  Kong C, Bang J, Sugiyama Y. Structural investigation of composite wind turbineblade considering various load cases and fatigue life. Energy2005;30:2101–14.

    [18]  Kong C, Kim T, Han D, Sugiyama Y. Investigation of fatigue life for a mediumscale composite wind turbine blade. Int J Fatigue 2006;28:1382–8 .

    [19]   Shokrieh MM, Rafiee R. Simulation of fatigue failure in a full composite windturbine blade. Compos Struct 2006;74:332–42.

    [20]   Toft HS, Sørensen JD. Reliability-based design of wind turbine blades. StructSaf 2011;33:333–42.

    [21] , 2015.[22]   Herakovich CT. Mechanics of fibrous composites. John Wiley & Sons: New

    York; 1998. p. 14, 112–183.[23]  Reddy JN. Mechanics of laminated composite plates theory and analysis. New

    York: CRC Press Inc; 1997. p. 651–657.

    H.G. Lee et al. / Composite Structures 133 (2015) 878–885   885

    http://refhub.elsevier.com/S0263-8223(15)00687-X/h0025http://refhub.elsevier.com/S0263-8223(15)00687-X/h0025http://refhub.elsevier.com/S0263-8223(15)00687-X/h0025http://refhub.elsevier.com/S0263-8223(15)00687-X/h0030http://refhub.elsevier.com/S0263-8223(15)00687-X/h0030http://refhub.elsevier.com/S0263-8223(15)00687-X/h0035http://refhub.elsevier.com/S0263-8223(15)00687-X/h0035http://refhub.elsevier.com/S0263-8223(15)00687-X/h0035http://refhub.elsevier.com/S0263-8223(15)00687-X/h0040http://refhub.elsevier.com/S0263-8223(15)00687-X/h0040http://refhub.elsevier.com/S0263-8223(15)00687-X/h0040http://refhub.elsevier.com/S0263-8223(15)00687-X/h0045http://refhub.elsevier.com/S0263-8223(15)00687-X/h0045http://refhub.elsevier.com/S0263-8223(15)00687-X/h0050http://refhub.elsevier.com/S0263-8223(15)00687-X/h0050http://refhub.elsevier.com/S0263-8223(15)00687-X/h0050http://refhub.elsevier.com/S0263-8223(15)00687-X/h0055http://refhub.elsevier.com/S0263-8223(15)00687-X/h0055http://refhub.elsevier.com/S0263-8223(15)00687-X/h0070http://refhub.elsevier.com/S0263-8223(15)00687-X/h0070http://refhub.elsevier.com/S0263-8223(15)00687-X/h0070http://refhub.elsevier.com/S0263-8223(15)00687-X/h0070http://refhub.elsevier.com/S0263-8223(15)00687-X/h0075http://refhub.elsevier.com/S0263-8223(15)00687-X/h0075http://refhub.elsevier.com/S0263-8223(15)00687-X/h0075http://refhub.elsevier.com/S0263-8223(15)00687-X/h0080http://refhub.elsevier.com/S0263-8223(15)00687-X/h0080http://refhub.elsevier.com/S0263-8223(15)00687-X/h0085http://refhub.elsevier.com/S0263-8223(15)00687-X/h0085http://refhub.elsevier.com/S0263-8223(15)00687-X/h0085http://refhub.elsevier.com/S0263-8223(15)00687-X/h0085http://refhub.elsevier.com/S0263-8223(15)00687-X/h0090http://refhub.elsevier.com/S0263-8223(15)00687-X/h0090http://refhub.elsevier.com/S0263-8223(15)00687-X/h0095http://refhub.elsevier.com/S0263-8223(15)00687-X/h0095http://refhub.elsevier.com/S0263-8223(15)00687-X/h0100http://refhub.elsevier.com/S0263-8223(15)00687-X/h0100http://refhub.elsevier.com/S0263-8223(15)00687-X/h0100http://www.caithnesswindfarms.co.uk/AccidentStatistics.htmhttp://refhub.elsevier.com/S0263-8223(15)00687-X/h0110http://refhub.elsevier.com/S0263-8223(15)00687-X/h0110http://refhub.elsevier.com/S0263-8223(15)00687-X/h0110http://refhub.elsevier.com/S0263-8223(15)00687-X/h0115http://refhub.elsevier.com/S0263-8223(15)00687-X/h0115http://refhub.elsevier.com/S0263-8223(15)00687-X/h0115http://refhub.elsevier.com/S0263-8223(15)00687-X/h0115http://refhub.elsevier.com/S0263-8223(15)00687-X/h0115http://refhub.elsevier.com/S0263-8223(15)00687-X/h0110http://refhub.elsevier.com/S0263-8223(15)00687-X/h0110http://www.caithnesswindfarms.co.uk/AccidentStatistics.htmhttp://refhub.elsevier.com/S0263-8223(15)00687-X/h0100http://refhub.elsevier.com/S0263-8223(15)00687-X/h0100http://refhub.elsevier.com/S0263-8223(15)00687-X/h0100http://refhub.elsevier.com/S0263-8223(15)00687-X/h0095http://refhub.elsevier.com/S0263-8223(15)00687-X/h0095http://refhub.elsevier.com/S0263-8223(15)00687-X/h0090http://refhub.elsevier.com/S0263-8223(15)00687-X/h0090http://refhub.elsevier.com/S0263-8223(15)00687-X/h0085http://refhub.elsevier.com/S0263-8223(15)00687-X/h0085http://refhub.elsevier.com/S0263-8223(15)00687-X/h0085http://refhub.elsevier.com/S0263-8223(15)00687-X/h0080http://refhub.elsevier.com/S0263-8223(15)00687-X/h0080http://refhub.elsevier.com/S0263-8223(15)00687-X/h0075http://refhub.elsevier.com/S0263-8223(15)00687-X/h0075http://refhub.elsevier.com/S0263-8223(15)00687-X/h0070http://refhub.elsevier.com/S0263-8223(15)00687-X/h0070http://refhub.elsevier.com/S0263-8223(15)00687-X/h0070http://refhub.elsevier.com/S0263-8223(15)00687-X/h0070http://refhub.elsevier.com/S0263-8223(15)00687-X/h0055http://refhub.elsevier.com/S0263-8223(15)00687-X/h0055http://refhub.elsevier.com/S0263-8223(15)00687-X/h0050http://refhub.elsevier.com/S0263-8223(15)00687-X/h0050http://refhub.elsevier.com/S0263-8223(15)00687-X/h0050http://refhub.elsevier.com/S0263-8223(15)00687-X/h0045http://refhub.elsevier.com/S0263-8223(15)00687-X/h0045http://refhub.elsevier.com/S0263-8223(15)00687-X/h0040http://refhub.elsevier.com/S0263-8223(15)00687-X/h0040http://refhub.elsevier.com/S0263-8223(15)00687-X/h0040http://refhub.elsevier.com/S0263-8223(15)00687-X/h0035http://refhub.elsevier.com/S0263-8223(15)00687-X/h0035http://refhub.elsevier.com/S0263-8223(15)00687-X/h0035http://refhub.elsevier.com/S0263-8223(15)00687-X/h0030http://refhub.elsevier.com/S0263-8223(15)00687-X/h0030http://refhub.elsevier.com/S0263-8223(15)00687-X/h0025http://refhub.elsevier.com/S0263-8223(15)00687-X/h0025http://refhub.elsevier.com/S0263-8223(15)00687-X/h0025http://-/?-http://-/?-