You are on page 1of 8

Composite Structures 133 (2015) 878885

Contents lists available at ScienceDirect

Composite Structures
journal homepage: www.elsevier.com/locate/compstruct

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 blade
being pulled out from its wind turbine during operation. In this study, we experienced delamination failure at the blade root during fatigue testing of a 3 MW full-scale wind turbine blade according to international 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 bumping 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 structural 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 [14]. 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, followed by delamination buckling which led to the blades collapse.
Jensen et al. believed the main root cause was the Braizer effect of
the shell structure due to bending. Overgaard et al. [89] 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
Corresponding author. Tel.: +82 55 280 3261; fax: +82 55 280 3498.
E-mail address: hakgulee@kims.re.kr (H.G. Lee).
http://dx.doi.org/10.1016/j.compstruct.2015.08.010
0263-8223/ 2015 Elsevier Ltd. All rights reserved.

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 mechanism 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 edgewise 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. Marn 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

H.G. Lee et al. / Composite Structures 133 (2015) 878885

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 SN linear damage equation,
the load spectrum, and Speras 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 statistical uncertainties.
Despite such contributions by many researchers, about 30 blade
failures are occurring per year throughout the world, and the number 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 conditions 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.

879

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

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, respectively. It has been developed as a result of a R&D project funded
by 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 flapwise 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 as
shown in Fig. 3.
Strain values of the T-bolt, located as shown in Fig. 4, were measured 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. 3. Failure at the end of the blade root.

880

H.G. Lee et al. / Composite Structures 133 (2015) 878885

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

understand why the approach failed to predict the separation


and bending observed.
3. Conventional blade root design
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

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

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 circular 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. 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.

H.G. Lee et al. / Composite Structures 133 (2015) 878885

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 conditions aforementioned in Section 2, a static analysis with the FE
shell 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 calculated T-bolt strains with the measured strains during the fatigue

881

testing. After several modifications, stress distributions at the


end of the blade root were able to explain the observed delamination 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 commercial FE solver, ABAQUS 6.13, and its 4 node shell element, S4R.
The number of the elements is 57,969. The boundary condition
was the clamped condition at the end of the blade root, and the
loading conditions were the flapwise and the edgewise test bending 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: E1 of 40.14 GPa, E2 of 12.30 GPa, v12 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.

882

H.G. Lee et al. / Composite Structures 133 (2015) 878885


Table 3
Subcomponent FE solid model for the test blade root.

Table 2
FE shell model for the test blade.
Element type

No. of the
elements

Boundary condition

ABAQUS
6.13

4 node shell element


(S4R)

57,969

Clamped condition at
the root

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 v23 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 Poissons ratios v12 and
v23. Then G23 can be calculated from Eq. (1), resulting in
4.44 GPa, and the other properties E3, v13, and G13 are the same
as E2, v12, and G12, respectively.

G23

E2
21 v 23

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)

the FE shell model, which were applied to the cross section of the
blade root part along the spanwise direction and the hoop direction, 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 laminated plate theory (CLPT) cannot calculate whole equivalent orthotropic properties because even interlaminar stresses at the
interface of two laminae are discontinuous [23]. CLPT gives us only
in-plane laminate properties such as Ex, Ey, vxy, and Gxy and
through-thickness Poissons ratios such as vxz and vyz, which are
calculated from Eqs. (2)(4) [22].

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

FE solver

FE solver

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

v xz
v yz
Fi

1
1
A1
11 F 1 A12 F 2 A16 F 6

2HA1
11
1
1
A1
21 F 1 A22 F 2 A26 F 6

2HA1
22

N
X
 k k
 
S Q

k  k k  k
i 1; 2; 6
13 1i S23 Q 2i S36 Q 6i t k

k1

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


where Aij , 
Skij , Q
ij
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 total
thickness 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, Ez, Gxz, and Gyz 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.
Table 4
Properties of the glass NCF/epoxy unidirectional lamina.
E1
[GPa]

E2
[GPa]

E3
[GPa]

v12

v13

v23

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.

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

Ex
[GPa]

Ey
[GPa]

Ez
[GPa]

vxy

vxz

vyz

Gxy
[GPa]

Gxz
[GPa]

Gyz
[GPa]

30.69

13.44

12.70

0.42

0.21

0.34

6.26

3.55

4.21

H.G. Lee et al. / Composite Structures 133 (2015) 878885

883

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.

5. Analysis results

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

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 greater
than 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 complexities 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 reason
of the unfavorable load distribution at the end of the blade root.

884

H.G. Lee et al. / Composite Structures 133 (2015) 878885

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.

H.G. Lee et al. / Composite Structures 133 (2015) 878885

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 different strain gages. The simulation shows us partial separation
between the blade root and the pitch bearing as shown in
Fig. 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 subcomponent 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).

885

References
[1] International Electrotechnical Commission. International standard IEC 6140023 wind turbines part 23: full-scale structural testing of rotor blades. 2014.
[2] Det Norske Vertias. Standard DNV-DS-J102 Design and manufacture of wind
turbine blades, offshore and onshore wind turbines. 2010.
[3] Germanischer Lloyd. Rules and guidelines IV industrial services 1 guideline for
the 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 Sustainable
Energy Rev 2012;16:28492.
[6] Yang B, Sun D. Testing, inspecting and monitoring technologies for wind
turbine blades: a survey. Renewable Sustainable Energy Rev 2013;22:51526.
[7] Jensen FM, Falzon BG, Ankersen J, Stang H. Structural testing and numerical
simulation of 34 m composite wind turbine blade. Compos Struct
2006;76:5261.
[8] Overgaard LCT, Lund E, Thomsen OT. Structural collapse of a wind turbine
blade. Part A: static test and equivalent single layered models. Compos Part A
2010;41:25770.
[9] Overgaard LCT, Lund E. Structural collapse of a wind turbine blade. Part B:
progressive interlaminar failure models. Compos Part A 2010;41:27183.
[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-scale
static tests. Compos Struct 2013;97:1529.
[11] Kensche CW. Fatigue of composites for wind turbines. Int J Fatigue
2006;28:136374.
[12] Mandell JF, Samborsky DD, Agastra P, Sears AT, Wilson TJ. Analysis of SNL/
MSU/DOE fatigue database trends for wind turbine blade materials. Sandia
National Laboratory report: SAND2010-7052. 2010.
[13] Nijssen RPL. Fatigue life prediction and strength degradation of wind turbine
rotor blade composites. Sandia National Laboratory report: SAND2006-7810P.
2006.
[14] van Leeuwen H, van Delft D, Heijdra J, Braam H, Jrgensen ER, Lekou D, et al.
Comparing fatigue strength from full scale blade tests with coupon-based
predictions. Trans ASME 2002;124:40411.
[15] Marn JC, Barroso A, Pars F, Caas J. Study of damage and repair of blades a 300
kW wind turbine. Energy 2008;33:106883.
[16] Marn JC, Barroso A, Pars F, Caas J. Study of fatigue damage in wind turbine
blades. Eng Fail Anal 2009;16:65668.
[17] Kong C, Bang J, Sugiyama Y. Structural investigation of composite wind turbine
blade considering various load cases and fatigue life. Energy
2005;30:210114.
[18] Kong C, Kim T, Han D, Sugiyama Y. Investigation of fatigue life for a medium
scale composite wind turbine blade. Int J Fatigue 2006;28:13828.
[19] Shokrieh MM, Rafiee R. Simulation of fatigue failure in a full composite wind
turbine blade. Compos Struct 2006;74:33242.
[20] Toft HS, Srensen JD. Reliability-based design of wind turbine blades. Struct
Saf 2011;33:33342.
[21] <http://www.caithnesswindfarms.co.uk/AccidentStatistics.htm>, 2015.
[22] Herakovich CT. Mechanics of fibrous composites. John Wiley & Sons: New
York; 1998. p. 14, 112183.
[23] Reddy JN. Mechanics of laminated composite plates theory and analysis. New
York: CRC Press Inc; 1997. p. 651657.

You might also like