5

Developments in materials for offshore wind turbine blades

R. Nijssen,  and G.D. de Winkel     Knowledge Center Wind Turbine Materials and Constructions (WMC), Wieringerwerf, The Netherlands

Abstract

Rotor blades are large-scale integral hybrid structures. For a large part they are made of composites, with bolted blade–hub connections and integrated lightning protection, and including various types of composite structural materials and elements. During their life in an offshore wind farm they suffer from environmental loads, extreme loads and fatigue.

This chapter discusses the role of materials in creating structures capable of dealing with these loads and highlights current and future developments in (composite) materials, manufacturing methods and design philosophy.

A case study is included, where a blade is designed using integrated wind turbine design software, further exemplifying the relationship between the main design drivers, material choice and structural freedom.

Keywords

Fatigue; Loads; Manufacturing; Materials; Rotor blades; Simulation; Testing

5.1. Key requirements for blade materials

5.1.1. Loads on rotor blades

Wind turbine rotor blades suffer from a tremendous amount of fatigue cycles (in the order of 1089 cycles, where the existence of a fatigue limit in some metallic materials has not been proven). There are two main causes: one is the wind shear load cycle (flapwise load is lower at the downward position than the upward position due to vertical wind shear), the other main source of fatigue loading is the gravitational loading of the blades by their own mass (load reverses from one horizontal position to the other).
The development of rotor blade materials cannot be seen separately from the development of wind turbine design. With regards to the state-of-the-art wind turbines (three-bladed upwind turbines of increasing size), various developments can be identified, such as inclusion of lighter materials (to decrease gravitational fatigue loading), and the application of stiffer materials and structures (to keep the blade from hitting the tower at high winds). Joints sometimes need to be reconsidered in order to maintain load transfer between sections that become relatively smaller as blades become more slender.
For offshore wind turbines, higher tip speed allowables may lead to employing two-bladed turbines in a downwind rotor configuration, see Fig. 5.1, leading to new demands with regards to the material. Higher tip speeds require better erosion-resistant leading-edge materials. Increasing size will increase the requirement of light-weight design and/or trailing edge reinforcement. Trailing edge reinforcement helps to keep the blade's flap and lead-lag natural frequencies apart in longer, slenderer blades. Nevertheless, if the remainder of the turbine structure allows, turbine structural stiffness and hence natural frequencies may go down, allowing for more cost-effective material usage (ie, avoidance of carbon).
Moving to offshore potentially increases the influence of humidity (eg, [1]) and salinity.

5.1.2. Structural elements of rotor blades

A typical blade structural layout is shown in Fig. 5.2. Similar information can be found in [2]. The rotor blade is a hollow structure. Near the root, blade–root connector bolts are embedded in a tubular, monolithic, usually multi-axial laminate, ie, with dominant fibre directions both in tip direction as well as oriented at, eg, ±45 degree with respect to the tip direction. The connector bolt configuration can be either as T-bolts or as embedded studs.
image
Figure 5.1 Development of offshore turbines may lead to shifts in material requirements.
image
Figure 5.2 Typical blade layout. P. Brøndsted, H. Lilholt, A. Lystrup, Composite materials for wind power turbine blades, Annu. Rev. Mater. Res. 35 (1) (2005) 505–538.
Further along the blade length, usually around the location where the tubular blade root transforms into an aerodynamically shaped cross-section, an internal spar starts in the structure, running until it is close to the blade tip. Although this is an integral structure, it is composed of components with their own functionality; the spar caps are monolithic slabs of mostly unidirectional laminate (fibres running from root to tip), whereas the shear webs are typically sandwich structures. A blade may have up to three shear webs at heavily loaded sections.

5.1.2.1. Requirements for materials

As described above, rotor blade materials are part of a light-weight, highly loaded integral structure. The excessive load requirements need to be balanced against minimal cost of manufacturing, operation and maintenance, and decommissioning. In the paragraphs below, the requirements for blades in terms of manufacturing and operation are discussed, as well as a short note on cost.
In manufacturing
Most rotor blades are manufactured in a vacuum-infusion process, see Fig. 5.3. In this process, prefabricated parts, dry fabrics, core materials and other components, such as the lightning protection system, are positioned in a mould, typically in the shape of the pressure side or suction side of the blade profile (other modular divisions of the blade are common practice at various manufacturers). Prefabricated parts may include the blade root (eg, with embedded connector bolts), and the spar caps, due to their large thickness and resulting poorer infusibility. Auxiliary materials, such as flow media, are added to the stack of dry fabrics, and a vacuum foil is applied over the complete stack. Release media (spray-on, film, or textile) are typically included on the mould surface and between the laminate stack and auxiliary materials. The result is the future blade shell, surrounded by a rigid mould on the side of the future aerodynamic surface, and a flexible mould on the future inside of the blade. This cavity is then infused with uncured resin, using vacuum assistance, which is subsequently cured, resulting in a rigid polymer with embedded reinforcement.
image
Figure 5.3 Blade manufacturing set-up (schematic).
The success of the infusion step determines to a very large extent the quality of the final product. The speed of infusion and post-cure is an important driver of manufacturing time and cost.
The speed of the infusion depends in part on the infusion strategy, which comprises the selection of resin channels and other flow media, and selection of resins and hardener systems and the associated temperature and pressure profiles with the aim of maximising infusion speed, fabric impregnation, quality, and fibre volume fraction. Simulation software incorporating most or all of these parameters is available and can be used to optimise the infusion strategy (Fig. 5.4[3].
From a materials perspective, the resin cure-kinetics and textile permeabilities (the latter preferably in three directions) play an important role in the design of the infusion strategy. Most resin formulations can be tailored on-site to a faster or slower infusion through the manipulation of temperature, and choice and relative content of accelerators, catalysts, and curing agents. During curing, temperatures increase (resin curing is an exothermic process), and the resin gradually increases viscosity from a water-like viscosity (which is lower for higher starting temperatures) to the polymerised state [4].
Given a certain textile permeability, this means that the faster the resin curing rate is, the shorter the infusion path can be, necessitating the use of a larger number of resin injections as well as vacuum ports. Thus, a fast-curing resin requires more preparatory activities. Faster curing is typically also associated with a higher peak exotherm, which can be self-reinforced in areas of high resin volume during cure, such as in thick laminates (this can be a reason to prefabricate these parts). Higher peak exotherms mean larger temperature trajectories during cooling and potentially more severe shrinkage/residual stress effects. Conversely, slow-curing resins allow for less aggressive curing, reduced mould preparation, but have the disadvantage of longer cycle times.
image
Figure 5.4 Filling time simulation of a rotor blade. Polyworx, http://www.mathfem.it/eng/software-simulation-composite-infusion-RTM-Worx-project-lm-glasfiber.html, (accessed 29.01.16), with permission.
A recent development is found in ‘latent’ resin systems. These are resin systems that do not start curing directly after the final formulation is generated from the resin constituents, but that require a threshold temperature to start the polymerisation. Theoretically, this allows for combining the advantages of a slow resin with those of fast resins, with the added requirement of controllable mould temperature [5].
As for the textiles, good permeability is a prerequisite for successful infusion. Generally, textile permeability is best for non-crimp (non-woven) fabrics, in the direction of the fibres, and in the plane of the textile. High permeability is associated with a ‘loose’ coherence, which is contrary to the strength and stiffness requirement of the high-fibre volume fraction in the final laminate.
An important non-technical requirement, that nevertheless has a significant influence on rotor blade design, is imposed by many manufacturers on the availability of materials. To ensure unimpeded production and repair, a material will only be included in the design if it can be delivered by two or more independent suppliers and be used interchangeably, ie, materials from different suppliers can be used in the production of the same part. Without doubt, this strategy ensures (for the material buyer) favourable competition between suppliers. On the other hand, it may hamper innovation as newly developed improved materials may take some time before they are included in rotor blades.
Specific properties
As blades become longer, to capture more energy at a given wind turbine site, the specific properties (strength and stiffness with respect to mass) become more influential in material selection, favouring higher specific strength and stiffness over ‘heavier’ materials. This is driven by the loading of the blade by its own mass, which generates cyclic stresses with maxima near the leading and trailing edges when the blade is in the horizontal position. Although basic scaling rules dictate that the root bending moment scales with rotor radius R3 + 1 (mass scales according to the square cube law, moment arm scales proportionally to rotor radius), the actual stresses in the blade root only scale proportionally according to these same scaling rules. In practice, this might mean less-than-proportional scaling, as improvements in blade design have resulted in a historic mass increase which is below cubic [6]. On the other hand, the blade root radius is limited by hub- and pitch-bearing design considerations, as a result of which, eg, (partially) embedded root bolt connections are becoming increasingly popular compared to T-bolt connections that require more installation space [7,8].
Cost
Apart from market aspects, such as variations in supplier cost due to international trade policy or currency fluctuations, the business case for blade manufacturing depends on cost of materials, tooling, manufacturing facilities and cost of labour, management and organisation. A relevant effect for the wind industry is that investment in reduction of handling of materials does not lead to significant reductions in blade cost – if labour cost is only 5% of blade cost, then a 20% reduction will lead to a total cost reduction of only 1%. In such a case, striving towards reduction of cycle time has been advocated by [9] as improving the cost reduction more effectively. They asserted that an added advantage of shorter cycle times is improved quality, as production teams will work on the same product at a higher frequency, and faults can be spotted and resolved inherently faster. These effects are equally valid for high-labour cost situations.
Shorter cycle times can be achieved by shorter mould times, which can, eg, be facilitated through the development of faster resin systems, see Section 2.2.1.1, higher material deposition rates, larger ply thicknesses, co-depositing of structural and auxiliary materials.

5.2. Role of testing materials and structures in the blade design process

There are several reasons for testing materials, components, and full-scale structures. One of the main reasons is validation of design. Design is based on assumptions and simplified models. The development of these models themselves requires an ample experimental basis for scientific observation. However, even if models are so well developed that they accurately and reliably predict structural behaviour in all foreseeable situations, the models themselves need to be fed input data for calibrating material factors and properties. Finally, in the development of appropriate test methods, the accuracy of the material models determines to a large extent the test's capability to accurately represent the operational conditions. For example, the test load in a rotor blade test is derived from the design load using fatigue life prediction methods that are derived from coupon tests.
In practice, rotor blade design starts with material selection, and to this end, candidate materials may be inventorised based on performance in manufacturing as well as in operation, availability, and cost. A selection of these materials is then subjected to standardised material testing. For a typical blade, ca. 5–10 materials may be submitted to such a test programme, with subjecting the materials to several different quasi-static test types (eg, tension, compression, in-plane shear, interlaminar shear), each of which is standardised to yield a prescribed description of material parameters (typically a stress–strain diagram and Poisson contraction). Each configuration is tested 6–10 times under identical conditions to obtain an impression of inherent scatter in the material or properties. Information regarding scatter is used in a later stage to obtain either a pre-defined safety factor or, in probabilistic design, a targeted reliability. From the best results, a few materials are then often tested in fatigue, where typically three R-ratios are investigated up to 107 load cycles (the R-ratio is defined as the ratio of minimum to maximum cyclic load applied in a fatigue test; typically R = 0.1 (all tension), R = 1 (reversed loading), and R = 10 (all compression) are used). In all, a material testing programme performed in the framework of a blade certification can amount to 500 individual tests and last a few months.
A systems engineering insight that has been prevalent in (notably) aerospace engineering and is being investigated for wind turbine design is the concept of the verification escalade. This entails the reduction of uncertainties in design by testing at the material, subcomponent and full-scale levels, consecutively, for a schematic see Fig. 5.5.
image
Figure 5.5 The blade verification escalade.
In design practice up to the early 2000s, it was common practice to essentially base a rotor blade design on knowledge of the ‘bulk’ material properties, ie, material properties obtained from laboratory tests on standard coupons. This approach to a large extent disregards the interaction between different materials (composite or hybrid combinations of composites and metals) that exist in and around bondlines and connections, and in sandwich panels. This has led to designs where design details rather than laminates were critical. In European projects such as UPWIND [10], and the more recent IRPWIND [11], the notion of subcomponent testing development was elaborated [12]. A subcomponent test is a mechanical experiment on an isolated part of the structure where failure mechanisms are expected (or known to exist from full-scale blade experience), with the aim of inferring these failure mechanisms and learning about their behaviour. Typically, these are structural details for which (numerical) models cannot accurately predict their behaviour.
As such, the subcomponent testing level seems to occupy a highly necessary intermediate step between material and full-scale testing. A full-scale test raises many questions (such as: “What is the predictive power of the result of a single full-scale blade with regards to the entire production?”), and in practice fails at the critical design detail (as opposed to failing everywhere simultaneously in ideal optimal design). In this light it is striking that the development of subcomponent testing has lagged behind with respect to the aerospace industry, and has only recently been targeted for inclusion in some of the design guidelines [13]. A possible explanation for this relatively late introduction is the fact that a wind turbine rotor blade is in many ways a more integral structure than, eg, an airframe, due to the structural layout and the materials used. An aircraft can, at least from a structural point of view, be separated into a large amount of the structural parts, consisting of isotropic materials, which are each connected in a distinct fashion, often through mechanical fasteners such as bolts and rivets. This simplifies to a certain extent the division in representative test specimens, as well as the appropriate representation of boundary conditions (a riveted connection to the rest of the structure can be replaced by a bolted or riveted connection with the test machine). Other reasons may be the fast-changing designs and design scales (rendering previous test programmes on structural details less useful for next-generation designs). Furthermore, standardisation is of significant influence on the introduction of subcomponent testing in common design practice. A certain degree of standardisation of subcomponent testing is necessary for useful inclusion in the blade certification process, but the intellectual property represented in structural details specific to certain blade designs makes them unsuitable for publication in a standard, whereas generic subcomponent specimen designs and test set-ups can only be relevant for a limited range of design details.

5.3. Case study on material selection and blade design

In this chapter, the dependence on material characteristics and blade design is illustrated. In this study, full-glass reinforced load-carrying blade elements, the spar caps, were compared to full-carbon reinforced spar caps as well as glass-carbon hybrid spar caps. On one side, the relative advantages of using glass spars are the low cost and high material availability in non-crimp reinforcement configuration. For large blades, materials with a higher specific strength and/or stiffness are required, and although carbon fibre has been pointed out as the most likely candidate for decades, it has not been until the latest generations of designs that larger blades consistently exhibit full-carbon structural elements. The potential downsides of this solution are the high cost (approximately a factor of three more costly than glass), poorer infusability due to the smaller fibre diameter and resulting denser fibre packing, and more pronounced sensitivity in compression loading (in fibre direction) to geometrical aberrations such as wrinkles and waviness. Glass–carbon hybrid laminates constitute an interesting compromise between the two extremes, where the glass layers double as infusion paths for the resin, offer improved support for the carbon layers in compression, and limit the overall cost of the structural element.
In a recent study described in Ref. [14], the experimentally determined quasi-static and fatigue characteristics of three different laminates are used, for each laminate, to optimise the internal structure of a reference blade. The three laminates are a full-glass reinforced epoxy; a full-carbon reinforced epoxy; a hybrid laminate consisting of both glass and carbon layers. The one-million cycle stresses measured in fatigue tests in three R-ratios are shown in Fig. 5.6.
image
Figure 5.6 Fatigue performance at one-million cycles for three materials (quasi-static stresses indicated on abscissa) [14].
Several design limits and boundary conditions are taken into account in the blade design. For optimisation, the masses of the blade's load-carrying elements are converted to cost (applying a straightforward multiplier), giving a general impression of the potential of using alternative laminates.
An important boundary condition in this study is that only the internal structure of the blade could be modified, ie, the outside geometry was fixed to an available reference geometry. This is analogous to a design dictated by aerodynamic performance requirements. In integrated design, it is advisable to optimise the blade more globally, including a more thorough interaction between aero-elastic design and structural design.
Basic design limits were taken into account, ie:
Tower clearance
a limit on blade bending during the maximum operational gust.
Flapwise and edgewise fatigue
a limit on fatigue damage in the main spar flanges and trailing edge reinforcements (if any).
Buckling
avoidance of buckling in the spar flanges – buckling of skin panels was not taken into account.
The optimisation of blade design for this study consisted of finding the best distribution of spar cap thickness, ie, resulting in a minimum amount of material required to resist the design loads (see Fig. 5.7).
The resulting mass and cost from the spar cap thickness distribution optimisation are shown in Fig. 5.8. Using carbon, due to its higher stiffness and better fatigue properties, the spar cap thickness can be reduced by approximately 50% by going to a hybrid laminate. Further ‘removal’ of glass to obtain an all-carbon spar cap leads to a relatively smaller decrease in mass (ca. 35%), which can be attributed partly to buckling constraints. In terms of cost, the reductions are significantly smaller than in mass, due to the higher material cost. Note that in this study a reference blade design is used, but since the hybrid and carbon versions contain significant ±45 degree layers to improve coupon test performance and infusability, the glass spar cap was modified in the simulation software optimisations to also contain ±45 degree layers. This approach facilitates comparing apples to apples, but the ‘optimised’ glass spar cap (‘Glass blade’ in Fig. 5.7) was actually 2.7 tonnes heavier than the reference (all-UD) spar cap (‘Reference blade’).
image
Figure 5.7 Optimised blade spar cap thickness for different materials.
image
Figure 5.8 Mass and cost comparison of all-glass, all-carbon, and hybrid materials.
In this study, only one parameter was optimised at a time. More detailed routines might include multi-parameter optimisation, including spar cap width and laminate configuration. A reasonable amount of ‘engineering judgement’ was used in the partly manual optimisation executed for this single study. However, in a design department the more extensive optimisation routines available in some integrated design packages may be employed to maximise optimisation of automisation. The study discussed here reflects the state of the art in the sense that blade designs are only to a certain extent ‘integrated’. In most blade designs, if not all, separate elements fulfil separate tasks. As design as well as manufacturing automation might become more prevalent, further integration of functions and optimisation may occur.
More detailed parameterisation of (parts of) rotor blades for optimisation can easily result in a very large number of parameters. Even a relatively simple thickness distribution of a UD spar as shown in Fig. 5.9 has some locations where the thickness needs to be defined, even with the simplification that individual ply-drops are not taken into account.
In addition, parametrisation of an existing blade design may be hard and time-consuming and the blade definition may need to be simplified to end with a reasonable number of parameters in the optimisation. For some recent optimisation projects where layups with a complex thickness distribution had to be parametrised, a novel parametrisation method was implemented in the FOCUS6 wind turbine design software [15]. Instead of a direct parameterisation of the layup, a shape modifier was applied on the thickness distribution of the various layups in the blade. The used method is similar to the ‘Adjust Color Curves’ function as commonly found in photo-editing applications such as Photoshop and Gimp (Fig. 5.10).
image
Figure 5.9 Thickness distribution of a typical UD spar.
With a few degrees of freedom (the markers in the graph), the distribution can be changed. The shape modifier is applied to the existing thickness distribution resulting in a new thickness distribution. The advantage of this is that the starting point of the optimisation can be a complex thickness distribution, while the shape modifier has only a few degrees of freedom, thus less parameters need to be included in the optimisation routine.
image
Figure 5.10 Adjust Color Curves dialogue screen of Gimp.
For the shape function, a third-degree dimensionless polynomial function is used, that runs from j = 0 to 1 in the domain x = 0 to 1:

j(x,α1,α2)=x(1+272(x1)[α1(x23)α2(x13)])

image

where α1, α2 alter the curve shape. These parameters can be adjusted during the optimisation. The parameter x is the relative position in spanwise direction within a layup. The factors α1 and α2 are defined here at the 1/3 and 2/3 relative to spanwise coordinate locations, but can be defined at other locations as well.
If the thickness distribution in the spanwise direction is defined as t(s), the thickness distribution with the shape function applied is

t(s,αh,α1,α2)=(1+αh)t(sj(x,α1,α2))

image

The function j modifies the spanwise distribution of the thickness, while αh scales the thicknesses.
Figs 5.115.13 demonstrate graphically how the thickness distribution can be altered by changing the curve of the shape modifier by adjusting α1 and α2 (upper left graph) and the adjustment of the scaling (αh). In case α1 = α2 = αh, the original thickness distribution is obtained. For reasonable distributions α1 and α2 must be kept between 1/3 and +1/3. The method is not limited to the third order polynomial function as used here.
image
Figure 5.11 Thickness distribution ‘squeezed’.
image
Figure 5.12 Thickness distribution ‘shifted’.

5.4. Future trends

5.4.1. Material and structural damping

For larger blades and more slender blades (shorter chord length, smaller profile thickness), improved knowledge on damping properties will be required during design to provide appropriate stability checks, see eg, [16].

5.4.2. Multi-axial testing and damage progression

In isotropic material characterisation, the test is often designed so as to induce a uni-axial stress or strain in the material. In the case of multi-axial stress states, the performance in a uni-axial situation is converted to the multi-axial situation using eg, von Mises-type formulations. In laminated composites, however, the stress state is, more or less by definition, multi-axial; lamina with differing stiffness and different directions transfer loads mutually through both shear and normal stress. Sandwich composite panels must resist in-plane loads originating from edgewise, as well as flatwise, blade loads. In thick laminates, a through-the-thickness component may exist. Moreover, in details employing adhesive bondlines, two- or three-axial stress states occur; due to geometry and manufacturing not all bondlines are loaded in pure shear (which is usually common in adhesives).
image
Figure 5.13 Thickness distribution ‘scaled’.
Already, there is increased attention for the effects of stress multi-axiality in design details, eg, [17], but the development of appropriate testing methods on the material level requires further efforts. In design guidelines, the topic has not been addressed to date, although most guidelines have been in operation employing multi-axial failure criteria for the quasi-static assessment of composites, such as Puck [18] or the maximum strain or modified Tsai-Wu criterion [1,19].
Up to a certain point, the initiation and propagation of damage is a related topic, as damage often initiates in complex design details, that are relatively difficult to tackle in testing and design guidelines. The result is that structural failure is currently based on data that reflect full loss of load-bearing capacity. In the case of more damage-tolerant designs, as eg, discussed in Section 5.4, the behaviour of materials and structures in the presence of damage is much more relevant and testing and (laboratory) monitoring methods will need to cater for the quantification of damage progression.

5.4.3. Micromechanical modelling and interaction with condition monitoring

The ample diversity in available reinforcements, resins, core materials, and resulting composites has the advantage of far-reaching possibilities for tailoring design. Especially the long-term mechanical characteristics of laminates are increasingly thought to depend on reinforcement architecture, performance of the interphase where the matrix interacts with fibres, and environmental influences such as water uptake. Initiation and propagation of cracks inside the laminate lead to macroscopic changes in some of the properties and could be quantified as macroscopic damage, which accumulates during the product's lifetime, eventually leading to failure.
There are certain contrasts with current design practice, which use ‘bulk properties’ of lamina and laminate macroscopic performance, and relies on the prediction of failure rather than damage accumulation. At the moment this is probably satisfactory; as was described earlier in this chapter, macroscopic structural elements such as bondlines and mechanical connections are the most critical parts in design, and can be improved through a combination of structural modelling and subcomponent experimental evaluation.
However, as design details become more optimal, the focus of the designer will once more shift to the ‘bulk’ material. The quasi-static and fatigue characterisation of candidate materials effectively prohibit optimal laminate design.
In various material research disciplines, micromechanical modelling is a methodology aiming to predict macroscopic behaviour based on the micromechanical structure and material properties. In composites, analytical work dates back several decades, eg, [20], and recent efforts have focused on employing numerical models. Inclusion of fatigue damage initiation and propagation is one of the most recent additions to the research found in current literature, eg, [21]. Application in design tools, however, is still limited. In wind energy, as well as in many other applications of light-weight composite design, the potential of accurate life predictions based on the initial sub-ply-level structure of a laminate is enormous, in terms of optimising material test programmes and integrating the reinforcement and matrix preparation (for example weaving and stitching) with manufacturing parameters.
A next step in the development would be the integration of monitoring techniques, where the loads are monitored and in addition the instantaneous material state (including damage) is fed into a micromechanical numerical model. This allows basing predictions of residual life on the actual damage state combined with the actually occurring loads, instead of an assumed load history. Such a strategy does require enhanced understanding of the link between potential measurands (crack density and type, local hardness or stiffness, local state of the resin such as glass-transition temperature and moisture content, etc.).

5.4.4. Coatings, erosion protection

For performance reasons, a high tip-speed ratio (ratio of tip speed to wind speed) is favourable for the blades and drive train. As one of the limiting factors (tip noise) becomes less important for large offshore turbine blades, these turbines employ higher absolute tip speeds. This is associated with increased abrasion through particles (sand) or water (rain erosion), which increases the need for experimental validation of coatings, notably in rain erosion tests, as the eroding effect of rain droplets has been shown to be proportional to the relative velocity to the 5th power [22].
There are two main types of rain erosion test set-ups. One is the droplet erosion set-up, where a single drop is (periodically) propelled onto the substrate. This type of set-up is appropriate for more fundamental research on the interaction of droplets with a surface. A more commonly used test is the ‘helicopter’ test. In this configuration, one, or a couple of samples are fixed to a fast rotating ‘helicopter rotor’, which is located under a rain simulator. There are many test institutes who offer this type of test, but the standardisation of this test in terms of velocities, droplet size, flow rates, is still under development, so it is difficult to compare results from different laboratories. The advantage of the strong correlation of erosion effects with relative velocity is that rain erosion tests can be highly accelerated. Thus, several candidates for erosion protection can be relatively quickly compared by increasing the test rotational speed. Both methods have qualitative or relative estimates of erosion mechanisms and resistance, but they do not give accurate quantitative results that can be linked directly to the behaviour in actual operation.
Erosion protection is often a strip of thermoplastic-based adhesive tape which can be applied to the leading edge in the outboard blade section. Recent rain erosion results show significant development in the longevity of the erosion protection [23]. The advantages of such protection are longer potential service intervals, and reduced chance of damage to the laminate. Erosion damage to the coating is certainly an aerodynamic nuisance; if the laminate below is damaged costly repairs are required, causing significant downtime.

5.4.5. Design philosophies – safe life or damage tolerance?

Wind turbine rotor blades are designed for 100% availability during a predetermined operational life. This complies with a ‘safe life’ design, which means that a certain reserve/conservatism in design is required, so that materials typically operate at a lower level than their maximum capability, resulting in inherent addition of material (mass) to cover for uncertainties in external loading and material and structural models.
In light-weight design, it is therefore logical to employ damage-tolerant design. Here, damage due to operation is taken into account in the design phase through knowledge of the initiation and propagation of damage. Damage-tolerant design is based on the design of inspection intervals and associated repair procedures, and as such makes it unattractive to the design of wind turbines and specifically rotor blades, as inspection as well as repair are associated with prohibitive downtime and cost. A commercial airliner may spend an estimated 4% of its life in scheduled inspection and maintenance (A–D checks), where notably the C and D checks require specific working conditions, whereas an offshore wind turbine needs to achieve availabilities of around 98%, and blade inspection and repair are preferably done on-site.

5.4.5.1. Inspection, maintenance and repair

In any blade, regardless of design philosophy, inspection, maintenance and repair play an important role in maintaining operational integrity. In most blades online condition monitoring/structural health monitoring systems are limited or absent, so periodic inspection is valuable for the operator to find structural damage and/or damage that influences aerodynamic performance. Maintenance can consist of cleaning operations, and application or replacement of leading-edge protective tape and aerodynamic add-ons, such as vortex generators. Repair is still limited to cosmetic repairs. An overview of structural repair is given in [24]. Restoration of fatigue life to its original intended value is unrealistic [25]. Although the scarce publicly available reports [26] of in-field damage are most likely not statistically representative, most of the structural damage, however, seems to occur due to impact and erosion near the blade tip, where relative velocity is highest, and due to lightning, which is most likely to strike near the blade tip. Thus, in many cases cosmetic repairs will suffice.

5.4.5.2. Role of condition monitoring

In the life of a blade, the actual loads and other external influences will differ from those used as the basis for design. This eventually results in overloading or underloading of a rotor blade. In blade design, this uncertainty with regards to the ratio of operational and design loads is compensated for by partial factors. More optimal design would be facilitated by the application of condition monitoring, which would allow the (continuous) quantification of operational loads and comparison against design loads. Thus, ‘condition monitoring’ refers to the evaluation of load (and potentially humidity, icing, etc.) conditions during operation.
A further step would be to use this information in the assessment of residual life.

5.4.5.3. Life re-assessment and extension

Extending the use of information from, eg, strain sensors to monitor operational loads (condition monitoring) to an evaluation of cumulative life consumption or damage growth (structural health monitoring, SHM) requires linking the operational load measurements to a blade design model.
Such SHM techniques can be implemented continuously, but in current practice there are a few reasons why this is not common practice. A practical reason is that, in many turbine locations or wind farms, ‘repowering’ is performed, which means that, subject to a local planning and permit organisation, a wind turbine is replaced with a larger one. The second-hand turbine is then often shipped to other parts of the world with different prerequisites regarding warranty and certification. Thus, the turbine is replaced and moved years before the end-of-life, and intensive structural health monitoring may not be urgent. A second consideration is that early failures (eg, due to poor design or installation of the blade–root connection) are not captured in most structural health systems.
Nevertheless, application of SHM offers the possibility of estimating expended life versus age of the rotor blade for a specific location. This can be the basis for a warrant for life extension, or, alternatively, for condition-based operation, where certain extreme or cyclic loads are avoided for the remainder of the turbine's operation, thus extending the amount of wind energy that can be captured until decommissioning.

Abbreviations and nomenclature

R-ratioRatio of minimum to maximum cyclic value (valley-to-peak) in fatigue load
UDUnidirectional (refers to fibre direction in composite laminates)
t (s)Thickness distribution in spanwise direction
t′(s)Thickness distribution transformed with shape function
jThickness distribution shape function
α1, α2Shape parameters in thickness distribution shape function
αhScale parameter in thickness distribution shape function
SHMStructural Health Monitoring
CrimpThe tortuosity of a fibre or roving due to the weaving pattern

References

[1] Sayer F, Bürkner F, Buchholz B, Strobel M, van Wingerde A.M, Busmann H, Seifert H. Influence of a wind turbine service life on the mechanical properties of the material and the blade. Wind Energy. 2013;16:163–174.

[2] Brøndsted P, Nijssen R.P.L, eds. Advances in Wind Turbine Blade Design and Materials. Woodhead Publishing; 2013.

[3] Koorevaar A. Simulation of liquid injection molding. In: Sampe-JEC Conference, Paris. 2002.

[4] Teuwen J. Thermoplastic Composite Wind Turbine Blades – Kinetics and Processability Dissertation. Delft University of Technology; 2011: 978-94-6186-010-1.

[5] Sambale H. Latent Hardener for Rotor Manufacturing Kunststoffe.de. August 17, 2009.

[6] Wiser R, Yang Z, Hand M, Hohmeyer O, Infield D, Jensen P.H, Nikolaev V, O’Malley M, Sinden G, Zervos A. Wind energy. In: Edenhofer O, Pichs-Madruga R, Sokona Y, Seyboth K, Matschoss P, Kander S, Zwickel T, Eickemeier P, Hansen G, Schloemer S, von Stechow C, eds. IPCC Special Report on Renewable Energy Sources and Climate Change Mitigation. Cambridge, United Kingdom and New York, NY, United States: Cambridge University Press; 2011 via srren.ipcc. wg3.de/report/IPCC_SRREN_Ch07.pdf.

[7] Lillo P. Static and Fatigue Analysis of Wind Turbine Blades Subject to Cold Weather Conditions Using Finite Element Analysis Dissertation. University of Victoria; 2001.

[8] Montejo R, Saenz J, Nuin I, Ugarte A, Sanz J. Development and validation of an innovative joint system for sectional blades. In: Proc. EWEA. 2013 Paper 299.

[9] van Breugel S. Production at higher quality and lower cost. In: CWP Workshop, Beijing. October 2014.

[10] UPWIND Project Site. http://www.ewea.org/eu-funded-projects/completed-projects/upwind/.

[11] IRPWIND Project Site. http://www.irpwind.eu/.

[12] Stammes E, Westphal T, Nijssen R.P.L. Guidelines for Design, Stress Analysis and Testing of a Structural Blade Detail Deliverable D 3.1.4/WMC-2010-95. February 2011.

[13] Pansart S. Evolution in certification requirements for rotor blades. In: DNVGL Symposium on the Future of Rotor Blades for Wind Turbines, Hamburg. November 4, 2014.

[14] Westphal T, Bortolotti P, Nijssen R.P.L. Carbon glass hybrid materials for wind turbine rotor blades. In: Proc. EWEA 2013. February 2013 Vienna, Paper ID 281.

[15] Dekker H, de Winkel G, Kuiken J, Lindenburg K. Optimization of the blade layup. In: Blade Manufacturing and Composites. May 12, 2015 London, UK.

[16] Resor B, Owens B, Griffith D. Aeroelastic instability of very large wind turbine blades. In: Proc. EWEA. 2012.

[17] van Hemelrijck D, Smits A, Philippidis T.P. Study of the behaviour of a glass fibre reinforced epoxy composite system used for wind turbine rotor blades under biaxial load conditions. In: Proc. EWEC. 2006.

[18] Germanischer Lloyd Industrial Services GmbH. Rules and Guidelines, IV Industrial Services, Part 2 Guideline for the Certification of Offshore Wind Turbines. 2012.

[19] DNV-DS-J102, Design and Manufacture of Wind Turbine Blades, Offshore and Onshore Wind Turbines. Det Norsk Veritas; 2010.

[20] Reifsnider K.L. The critical element model: a modelling philosophy. Eng. Fract. Mech. 1986;25:739–749.

[21] Qian C. Multi-Scale Modelling of Fatigue of Wind Turbine Rotor Blade Composites Dissertation. Delft University of Technology; 2013.

[22] Thiruvengadam A, Rudy S, Gunasekaran M. Experimental and analytical investigations on liquid impact erosion. In: Characterization and Determination of Erosion Resistance, ASTM STP 474. American Society for Testing and Materials; 1970:249–287.

[23] Claus C. Latest Protection Methods Against Leading Edge Erosion for Increased Tip Speeds, WindPower Events – Blade Inspection. Hamburg: Damage & Repair Forum; September 2014.

[24] Katnam K.B, Comer A.J, Roy D, da Silva L.F.M, Young T.M. Composite repair in wind turbine blades: an overview. J. Adhes. January 2, 2015;91(1–2):113–139.

[25] R. Nijssen, A. van Wingerde, (WMC), Fatigue of Long Reference and Repaired MD Specimens. OPTIMAT report #: OB_TG4_R012 rev. 000, doc. no. 10324, from: http://www.wmc.eu/public_docs/10324_000.pdf, accessible through http://www.wmc.eu/optimatbladesdocs.php.

[26] Lutz O. Deficiency, defects and damage spots in rotor blades. In: Proc. Wind Turbine Rotor Blades. Essen, Germany: Haus der Technik; June 2011.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
18.118.139.224