3
Considerations for Foundation Design and the Necessary Calculations

3.1 Introduction

Offshore wind turbines are complex machines, and the structure is dynamically sensitive. Furthermore, these are located in challenging offshore environment and, therefore, there are several types of design considerations. Before the design considerations are described, it is important to highlight some of the complexities and interdisciplinary nature of the issues. Construction of offshore structures in ultra‐deep water is not new, and there are abundance of references and experience. However, what is new is the large‐scale offshore wind turbine structures where a heavy rotating mass is placed at the top of the slender tower. Figure 3.1 shows the future of offshore wind turbines where heavier turbines are increasingly being placed in taller towers. One of the important design drivers is the dynamics of these systems. Therefore, the next section discusses the modes of vibration of these structures and how they affect the overall design considerations.

Illustration displaying 4 wind turbines such as Siemens SWT-3.6-107, Areva M5000, Siemens SWT-6.0-154, and Vestas V164 with indicated weights placed in towers measuring of 75 m, 85 m, 100 m, and 110 m, respectively.

Figure 3.1 Future of offshore wind turbines.

Schematic of a multiple supported structure with m, M1, M2, and M3 representing the mass of one blade, foundation frame, tower, and onboard machinery, respectively. L represents the height of the tower.

Figure 3.2 Definition of terms for design.

3.2 Modes of Vibrations of Wind Turbine Structures

The modes of vibration depend on the combination of the foundation system (i.e. single foundation such as mono caisson or monopile or a group of piles or a seabed frame supported on multiple shallow foundations) and the superstructure stiffness and mass distribution. The fundamental modes of vibration for these structures can be mainly of two types:

  1. Sway‐bending modes. This consists of flexible modes of the tower together with the top rotor‐nacelle assembly (RNA) mass that is essentially sway‐bending mode of the tower. Effectively in these cases, the foundation is very stiff axially when compared with the tower and the tower vibrates. The foundation provides stiffness and damping to the whole system.
  2. Rocking modes. This occurs when the foundation is axially deformable (less stiff) and is typical of wind turbine generators (WTGs) supported on multiple shallow foundations. Rocking modes can also be coupled with flexible modes of the tower.

The next section describes the modes of vibration through some examples. These aspects were investigated experimentally, numerically, and analytically by Bisoi and Halder (2014), Nikitas et al. (2016, 2017), and Lombardi et al. (2013). In the experimental methods, the modes of vibration were obtained from snap‐back test, also known as free‐vibration tests.

3.2.1 Sway‐Bending Modes of Vibration

Essentially, this form is observed when the foundation is very rigid compared to the superstructure. Wind turbines supported on monopiles and jackets supported on piles will exhibit such kind of modes. Figure 3.3 shows a schematic diagram of modes of vibration for monopile supported wind turbines and Figure 3.4 shows schematic diagram of a jacket supported wind turbine system. It is important to note that the first two modes are quite widely spaced – the typical ratio is about four to six times. One of the important points to note is that the foundation is axially very stiff. These analyses can be easily carried out using standard software.

Image described by caption and surrounding text.

Figure 3.3 Modes of vibration for monopile supported wind turbines.

Image described by caption and surrounding text.

Figure 3.4 Schematic diagram of modes of vibration for jacket structures supported on piles.

It may be worth noting that the first three modes of vibration of a fixed‐based cantilever beam are given by:

3.1equation

where αn is a mode number parameter having the value of 1.875, 4.694, 7.855 for the first, second, and third mode, respectively. EI is the bending stiffness of the beam having length L and M is the mass per unit length of the beam.

3.2.1.1 Example Numerical Application of Modes of Vibration of Jacket Systems

Figures 3.5 and 3.6 shows examples of numerical simulation of a jacket supported offshore wind turbine system where the modes of vibration may be appreciated. This is obtained through eigen solutions. For the twisted jacket (also known as inward battered guided structure [IBGS]), the frequency of first mode and second modes is 0.2678 Hz and the third mode is 1.049 Hz. It may be noted that the first two modes are identical and the task of the designer is to place this in the narrow band of allowable soft‐stiff zone, discussed in Section 2.2 of Chapter 2. Similar observations were also noted for standard jackets on piles, see Figure 3.4. This figure also provides a plan view to visualise the mode shapes. Closed‐form solution for obtaining natural frequencies of jacket is developed in Jalbi and Bhattacharya (2018).

Schematic diagram illustrating the modes of vibration for twisted jacket.

Figure 3.5 Modes of vibration for twisted jacket.

Schematic diagram illustrating the modes of vibration of standard jacket.

Figure 3.6 Modes of vibration of a standard jacket.

3.2.1.2 Estimation of Natural Frequency of Monopile‐Supported Strctures

Following the concept of target natural frequency (Section 2.2 in Chapter 2), in order to place the natural frequency of the whole system in the desired band, iterations are necessary whereby the different design parameters are altered. It is time consuming to carry out the optimization using software, and as a result, closed‐form solutions are effective. Method to calculate natural frequencies are shown in Chapter 5 and worked‐out examples are carried out in Chapter 6.

As an example, the natural frequency of a monopile supported wind turbine system can be estimated following the method developed by Arany et al. (2015a, b, 2016). This simplified methodology builds on the simple cantilever beam formula to estimate the natural frequency of the tower (fixed‐base assumptions), and then applies modifying coefficients to take into account the flexibility of the foundation and the substructure. This is expressed as

3.2equation

where CL and CR are the lateral and rotational foundation flexibility coefficients, fFB is the fixed‐base (cantilever) natural frequency of the tower considering the varying stiffness along the length of the tower and transition piece (TP). Effectively, the formulation shows the flexibility provided by the foundation. The method is described in Chapter 5 and example applications are shown in Chapter 6. Arany et al. (2016) provides a detailed derivation of the method, along with validation and verification with 10 case studies of the wind turbine structure.

3.2.2 Rocking Modes of Vibration

Rocking modes of foundation are typical of wind turbines supported on multiple shallow foundations; see, for example, Figure 3.7 where wind turbine structures are supported on multiple bucket‐type foundations. Figure 3.8 shows three other types of scaled model tests where a seabed frame or jacket is supported on shallow foundations. A schematic representation of the rocking modes of vibration are shown in Figure 3.9 where the tower mode may get coupled with rigid modes of vibration. Essentially, there are two types of vibrations: (i) rocking mode of the jacket, which is very rigid; (ii) flexible modes of the tower. Theoretically, for each rocking mode, there can be two flexible modes of the tower. These were observed through scaled model tests and reported in Bhattacharya et al. (2013a, b, 2017a). The foundation may rock about different planes and is dictated by the orientation of the principle axes i.e. highest difference of second moment of area. Figure 3.9 is a simplified diagram showing the modes of vibration where the tower modes can also interact with the rocking modes i.e. the tower may or may not follow the rocking mode of the foundation. Rocking modes of a foundation can be complex, as they interact with the flexible modes of the tower and it will be shown that these need to be avoided.

A scaled model of a tetrapod foundation, with arrows marking the hub, blades, tower, and tetra-pod. Dimensions are indicated.

Figure 3.7 A scaled model of a tetrapod foundation.

Photo displaying different configuration of foundations such as jacket (left), SIWT (middle), and tetrapod (right).

Figure 3.8 Different configuration of foundations.

Image described by caption and surrounding text.

Figure 3.9 Rocking modes of vibration.

Three cases are discussed below:

  1. Wind turbine supported on symmetric tetrapod foundations such as the one shown in Figure 3.7. A simplified model for analysis is illustrated in Figures 3.10 and 3.11. Research by Bhattacharya et al. (2013a,b) shows that even for the same foundations under each support, there will be two closely spaced vibration frequencies. This is due to different vertical stiffness of the foundation associated with variability of the ground. However, after many thousands of cycles of loading and vibration, these closely spaced vibration frequencies may converge to a single peak.
  2. Asymmetric tripod foundation. Example is provided in Figure 3.12 inspired by the concept shown in Figure 1.31 (Chapter 1). Study reported in Bhattacharya et al. (2013a,b) showed that there will two modes of vibration with closely spaced frequencies but with millions of cycles of loading, and these two closely spaced peaks will not converge. This is because the foundation has two different stiffness in two orthogonal planes.
  3. Symmetric tripod foundation. In a bid to understand the modes of vibration for a symmetric tripod, tests were carried out on a triangular foundation shown in Figures 3.13 and 3.14. Free vibration tests were carried out and a typical result is shown in Figure 3.15. The mode is like a ‘beating phenomenon’ well known in physics, which is possible for two very closed spaced vibration frequencies with low damping.
Image described by caption and surrounding text.

Figure 3.10 Rocking modes for a symmetric tetrapod about X−X′ and Y−Y′ plane.

Schematic illustrating the rocking modes about diagonal plane in plan, plane Y-Y', and plane X-X' views.

Figure 3.11 Rocking modes about diagonal plane.

Schematic illustrating the Modes of vibration for symmetric tripod in plan, plane Y-Y', and plane X-X' views. Model of a turbine with symmetric tripod is at the left side.

Figure 3.12 Modes of vibration for symmetric tripod.

Schematic illustrating the Modes of vibration for symmetric foundation in plan, plane Y-Y', and plane X-X' views.

Figure 3.13 Symmetric foundation.

Schematic of planes of vibration depicted 3 triangular foundations each having 3 circles on the vertices labeled K1, K2, and K3 and 2 intersecting lines for planes 1 and 2, planes 3 and 4, and planes 5 and 6 (left-right).

Figure 3.14 Planes of vibration.

Graph illustrating free vibration acceleration response, depicting a sine wave with changing amplitude.

Figure 3.15 Free vibration acceleration response.

Taking into consideration Figures 2.4 and 2.5 (Chapter 2) where the design first natural frequency of the whole system is to be targeted between 1P and 3P, it is important not to have two closely spaced modes of vibration. In practical terms, it is therefore recommended to avoid an asymmetric system. The previous case study also shows that a symmetric tetrapod is better than a symmetric tripod due to higher damping. It may be noted that the beating phenomenon is typical of low damping and two closely spaced modes. Gravity‐based foundation will also exhibit rocking modes of vibration and it may also interact with tower flexible modes. Figures 3.163.18 show a schematic diagram of observed modes of vibration from a small‐scale model test. Figure 3.19 shows the changing of natural frequency of foundations with cycles of loads observed in small scale testing.

Left: Schematic diagram displaying a turbine with small circular based gravity foundation. Right: Graph of power spectral amplitude vs. frequency displaying a curve with 2 peaks labeled Rocking mode and Tower mode.

Figure 3.16 Modes of vibration for a small circular gravity‐based foundation.

Left: Schematic diagram displaying a turbine with rectangular based gravity foundation. Right: Graph of power spectral amplitude vs. frequency displaying a curve with 2 peaks labeled Rocking mode and Tower mode.

Figure 3.17 Modes of vibration in a rectangular‐shaped GBS.

Left: Schematic diagram displaying a turbine with square based gravity foundation. Right: Graph of power spectral amplitude vs. frequency displaying a curve with 2 peaks labeled Rocking mode and Tower mode.

Figure 3.18 Modes of vibration in a square GBS.

Graph of normalised power spectra vs. frequency displaying 3 fluctuating curves representing after 0 cycles, after 40500 cycles, and after 400000 cycles.

Figure 3.19 Changes in modes of vibration of a monopile supported wind turbine with cycles of loading.

3.2.3 Comparison of Modes of Vibration of Monopile/Mono‐Caisson and Multiple Modes of Vibration

Deep foundations such as monopiles will exhibit sway‐bending mode, i.e. the first two vibration modes are widely spaced – typical ratio is four to five. However, multiple pod foundations supported on shallow foundations (such as tetrapod or tripod on suction caisson) will exhibit rocking modes in two principal planes (which are, of course, orthogonal). Figure 3.20 shows the dynamic response of monopile supported wind turbine and tetrapod foundation plotted in the loading spectrum diagram obtained from scaled models tests.

Graph of normalised power spectra vs. frequency displaying 2 fluctuating curves representing monopile foundation and tetrapod foundation.

Figure 3.20 Free vibration response of tetrapod foundations on suction caissons (see Figure 3.8 for the foundation) and monopile.

3.2.4 Why Rocking Must Be Avoided

Foundations (symmetric or asymmetric) on multiple foundations will exhibit two closely spaced natural frequencies corresponding to the rocking modes of vibration in two principal axes. For a soft‐stiff design, these need to be fitted in a narrow gap, as shown in Figure 3.21. Furthermore, as will be discussed in Chapter 5, owing to soil‐structure interaction (SSI), the two spectral peaks change with repeated cycles of loading. Also, for symmetric tetrapods (but not for asymmetric tripods), these two peaks will converge for sandy deposits. From the fatigue design point of view, the two spectral peaks for multipod foundations broaden the range of frequencies that can be excited by the broadband nature of the environmental loading (wind and wave) thereby impacting the extent of motions. Thus, the system lifespan (number of cycles to failure) may effectively increase for symmetric foundations as the two peaks will tend to converge. However, for asymmetric foundations the system life may continue to be affected adversely as the two peaks will not converge. In this sense, designers should prefer symmetric foundations to asymmetric foundations.

Top: Graph of forcing power vs. frequency with 2 curves with 4 graphs at the top for Froya Wind, Pierson & Moskowitz wave, etc. Bottom: Graph of system gain vs. frequency with 2 curves for tetrapod or tripod and monopile.

Figure 3.21 Fitting of two peaks in a narrow gap.

3.3 Effect of Resonance: A Study of an Equivalent Problem

Offshore wind turbines are a new type of structures, and it is important to learn from other disciplines. In this section, an example from helicopters is taken to show the importance of avoiding certain frequencies. Figure 3.22 shows still photographs from the well‐known helicopter resonance problem known as ground resonance, the video can be accessed from YouTube. Effectively, due to the imbalance in the helicopter rotor the rotation‐induced oscillations get in phase with the rocking frequency of the helicopter on its landing gears. This leads to collapse; the experiment is schematically shown in Figure 3.23. The helicopter starts rocking about the two landing pads (skids) until the stresses induced through resonance exceed the strength of the materials and connections, causing failure. There is a similarity between the helicopter ground resonance and jacket‐supported offshore wind turbines, as shown schematically in Figure 3.24, in the sense that both support a heavy rotating top mass. As the motion under consideration is rocking, the vertical stiffness of the supports is a governing parameter. For a jacket structure, at the onset the vertical stiffness (kN m−1) may not be identical. Therefore, they are shown as K1 and K2. It is clear that resonance must be avoided, and this emphasises the importance of understanding the subtle aspects of the dynamic behaviour of jacket‐supported wind turbines. Resonance not only affects the FLS (fatigue limit state) and SLS (serviceability limit state) but will also impact O&M (operation and maintenance).

Image described by caption and surrounding text.

Figure 3.22 Ground resonance of a helicopter.

Schematic illustrating the rocking motion of a helicopter getting tuned with the RPM of helicopter rotor, with center of gravity depicted by rotor blades (left) and clockwise arrows (middle, right).

Figure 3.23 Rocking motion of a helicopter getting tuned with the RPM of helicopter rotor.

Schematics of a helicopter and offshore wind turbine structure with Euler-Bernoulli beam and circular arrows for excitation forces. Both schematics support a heavy rotating top mass.

Figure 3.24 Similarities between a helicopter and offshore wind turbine structure.

3.3.1 Observed Resonance in German North Sea Wind Turbines

Hu et al. (2014) reported resonance for wind turbines. This is discussed further in Chapter 5.

3.3.2 Damping of Structural Vibrations of Offshore Wind Turbines

As the natural frequency of offshore wind turbines is close to forcing frequencies, damping is critical to restrict damage accumulation and avoid premature maintenance. Therefore, discussion is warranted for issues related to damping, and for practical design purposes can be idealised in fore−aft and side‐to‐side vibrations of offshore wind turbines. The main difference between the sway‐bending (or rocking) vibrations about two axes (X and Y) is that in the along‐wind direction, higher damping is expected due to the high aerodynamic damping caused by the rotating blades interacting with the airflow. On the other hand, for side‐to‐side direction, the aerodynamic damping is orders of magnitudes lower.

A nonoperational (parked or idling) OWT has similar aerodynamic damping in the fore−aft as in the side‐to‐side direction. Since the wind load is acting in the along‐wind direction, the highest load amplitudes are expected in fore−aft motion. This is because for most wind turbines in water depths less than 30 m, wind loading is the dominant load, while for very large diameter monopiles in medium to deep water, wave loading is expected to have equal or higher magnitude. It is worth noting that due to the yaw mechanism of the wind turbine, the along‐wind and cross‐wind directions are dynamically moving and are not fixed; therefore, the foundation is subjected to both cross‐wind and along‐wind loading in all directions during the lifetime of the OWT. One can conclude that analysing vibrations in both directions is important.

Studies considering the damping of the first bending mode either empirically or theoretically include Camp et al. (2004); Tarp‐johansen et al. (2009); Versteijlen et al. (2011); Damgaard and Andersen (2012); Damgaard et al. (2013); Shirzadeh et al. (2013). Based on these studies and other estimates and in the absence of other data, the following assessment of damping ratio contributions is recommended:

  • Structural damping: 0.15−1.5%. The value of structural damping depends on the connections in the structure (such as welded connections, grouted connections, etc) in addition to material damping (usually steel) through energy dissipation in the form of heat (hysteretic damping).
  • Soil damping: 0.444−1%. The sources of damping resulting from soil‐structure interaction (SSI) include hysteretic (material) damping of the soil, wave radiation damping (geometric dissipation) and, to a much lesser extent, pore fluid induced damping. Wave radiation damping and pore fluid induced damping are negligible for excitations below 1 Hz, and therefore hysteretic damping is dominant for the purposes of this study. The soil damping depends on the type of soil and the strain level.
  • Hydrodynamic damping: 0.07−0.23%. Results from wave radiation and viscous damping due to hydrodynamic drag. In the low frequency vibration of wind turbines the relative velocity of the substructure is low and therefore viscous damping, which is proportional to the square of the velocity, is typically very low. The larger contribution results from wave radiation damping, which is proportional to the relative velocity.
  • Aerodynamic damping: in the fore−aft direction for an operational turbine 1−6%, for a parking turbine or in the crosswind direction 0.06−0.23%. Aerodynamic damping is the result of the relative velocity between the wind turbine structure and the surrounding air. Aerodynamic damping depends on the particular wind turbine, and is inherent in the popular blade element momentum (BEM) theory for aeroelastic analysis of wind turbine rotors. The magnitude for a particular wind turbine also depends on the rotational speed of the turbine.

The total damping of the first mode of vibration is typically between 1−4% in side‐to‐side vibration, or parked or stopped or idling turbine. On the other hand, the total damping is between 2−8% for an operational wind turbine in the fore−aft direction.

3.4 Allowable Rotation and Deflection of a Wind Turbine Structure

Allowable rotation and deflection fall under SLSs i.e. operational tolerances imposed on the turbine system. These limits specify the total deflection, instantaneous deflection, and differential settlement/accumulation of rotation allowable throughout the lifetime of the turbine system. Excessive deflections or settlements can have an adverse effect on nonstructural components such as the generator and gearbox. This section of the chapter highlights these considerations. Possible reasons for the strict requirements from the point of view of the safe operation of the turbine, can be identified and are listed in Table 3.1:

  1. One of the important issues with excessive rotation is the risk of tower strike by the blades and the difficulties with the yaw system of control to provide a constant power. Due to the pre‐bend of the blades caused by mudline tilt, the blade‐tower clearance may reduce, resulting in an increased risk of tower strike; see Figure 3.25. The capacity of the yaw motor may not allow the turbine to yaw into the wind when the rotor‐nacelle assembly has to be turned ‘uphill’. Furthermore, the yaw brakes may not be able to keep the RNA in the upward position.
  2. There are also related issues such as reduced‐power production, increased bending moment on the support structure, reduced lifetime of the bearings, and problems with cooling fluid levels or movement. These issues are becoming less importance with design improvements.

Table 3.1 Possible reasons for strict verticality requirements.

Possible problem area Description of the problem
Blade – tower collision The tilt of the turbine may cause reduction of blade‐tower clearance due to initial deflection of the blade. These effects for an upwind machine may increase the risk of tower strike at the extreme deflection of a blade in the downward pointing position as shown in Figure 3.25. Blade‐tower collision is a serious risk that needs to be taken into account. In order to avoid tower strike, wind turbine manufacturers apply tilt angles to the main shaft of the rotor in the tune of 5–6°.
Reduced energy production The tilt at mudline causes the wind to hit the rotor ‘at an angle’ and reduces the total energy production by the wind turbine.
Yaw motors The limited yaw motor capacity may stop the turbine from yawing into the wind when the rotor has to be turned upwards against gravity.
Yaw brake The yaw brake may not be able to keep the rotor in the upwards pointing position.
Yaw, pitch, and main bearings The tilt at the nacelle level changes the direction and magnitude of the loading of bearings which may reduce their fatigue life or affect movement criteria.
Fluid levels and cooling fluid movements Tilt at the nacelle level might interfere with the cooling system of the turbine.
Increased bending moments The tilt of the turbine causes higher bending moments in the tower, grouted connection of the monopile and the transition piece, and the monopile itself.
Schematics illustrating mechanical issues with higher rotation at the mudline (Part 1), with reduced blade-tower clearance (left) and Yaw motor and Yaw brake (right).

Figure 3.25 Mechanical issues with higher rotation at the mudline (Part 1).

3.4.1 Current Limits on the Rotation at Mudline Level

In terms of SLS criteria, some specific requirements are to be met and the most important being the maximum deflection and rotation (tilt) at the foundation level/pile head (mudline) and at nacelle level. When assessing the total rotation, the initial tilt as well as the accumulated permanent rotation resulting from cyclic and dynamic loading throughout the design lifetime of the offshore wind turbine (OWT) should be considered.

The Det Norske Verita (DNV) code (DNV 2010a,b) gives typical limits of 0.25° and 0.5° for allowable rotation and states that ‘the deformation tolerances are typically derived from visual requirements and requirements for the operation of the wind turbine’.

Serviceability criteria appear to be ‘turbine manufacturer requirements’ for the operation of the wind turbine, which originates from onshore wind technology and is very similar to a typical tall structure (such as tall buildings). The very low tilt angle requirement for monopile supported structures appears to be especially overcautious in light of floating OWT technology, where tilt angles up to 7° are acceptable.

3.5 Internationals Standards and Codes of Practices

The considerations necessary for foundation design depend on its type (grounded or floating system), site location (wind and wave climate and subsurface conditions), availability and expertise of marine contractors, economics, type of contract and investors (governmental or private), and type of economy (developed or developing). However, the aim of this chapter is to discuss the scientific considerations with emphasis on dynamics requirements. Where applicable, general considerations for offshore installation and constructions will be briefly discussed and further reading will be suggested (Figure 3.26).

3 Schematics illustrating mechanical issues with higher rotation at the mudline (Part 2), with labels “Reduced energy production?” (left), “Increased foundation loads?” (middle), and “Reduced bearing lifetime?” (right).

Figure 3.26 Mechanical issues with higher rotation at the mudline (Part 2).

The design criteria and considerations are typically established based on the following:

  1. Design codes. The most important ones are: International Electrotechnical Commission (IEC) regulations, IEC61400‐3 and IEC61400‐22; DNV – guidelines on ‘Design of Offshore Wind Turbine Structures’ (DNV 2014), Germanischer Lloyd (GL) Windenergie's ‘Guideline for the Certification of Offshore Wind Turbines’, ISO standards, American Petroleum Institute (API) codes, Eurocodes, British Standards (BSI), and BSH (German) codes.
    1. The British Standards Institute (BSI) is a business service provider originating out of London, England. In addition to providing standards for the majority of UK industry, it has also produced a series of publications for use in the design and construction of offshore wind turbines. These regulations are BS EN 61400‐3 and BS EN 61400‐22, and they draw heavily from and elaborate on the IEC regulations already mentioned (BSI – British Standards Institution 2009; BSI – British Standards Institution 2011).
    2. DNVs is a Norwegian classification society dating back to 1864 that has produced numerous standards and guidelines for the offshore industry. With the emergence of offshore wind power in the North Sea, DNV started producing standards for the renewable sector, such as DNV‐OS‐J101 and DNV‐OS‐J201 (Det Norske Veritas 2009, 2013).
    3. GL is a German classification society based out of Hamburg. Similar to DNV, GL also produces a series of regulations providing guidance for the design and construction of offshore wind turbines (Germanischer Lloyd 2012). Recently, DNV and GL merged to form a classification society; however, to date the standards produced by each organisation remain separate.
    4. The API is a trade association that has provided regulatory services to the oil and gas industry. Despite no specific regulations regarding the design of offshore wind turbines, their guidance for conventional offshore programs is worth noting (American Petroleum Institute 2007).
  2. Certification body. Typically, a certification body allows for departure from the design guidelines if the design is supported by sound engineering and sufficient evidence/test results.
  3. Client. Occasionally, the client may pose additional requirements based on appointed consultants.
  4. Turbine manufacturer. The manufacturer of the wind turbines typically imposes strict SLS requirements or natural frequency requirements. In addition, the expected hub height is also a requirement for the turbine type and the site. The tower dimensions are also often inputs to foundation design and are normally provided by the turbine supplier.
  5. For specialised design such as seismic, ISO standard and Eurocode 8 may be used. The readers are also referred to text book titled Seismic Design of Foundation: Concepts and Applications by Bhattacharya et al (2018) for guidance on seismic design.

3.6 Definition of Limit States

A limit state is a condition beyond which the structure‐foundation assembly will no longer satisfy the specified performance requirements. For offshore wind turbines applications, they are:

(i) the Ultimate Limit State (ULS); (ii) the Serviceability Limit State (SLS); (iii) the Fatigue Limit State (FLS); and (iv) the Accidental Limit State (ALS).

3.6.1 Ultimate Limit State (ULS)

The ULS is related to the maximum load‐carrying resistance, and can be reached for several reasons: (i) excessive yielding and/or buckling (i.e. loss of structural resistance); (ii) a failure of a component (e.g. brittle fracture of connections); and (iii) a loss of static equilibrium of the structure (whole or part) with a consequent mechanism (e.g. rigid body behaviour, overturning, and capsizing).

As the main aim of a foundation is to transfer all the loads, during its design life, from the wind turbine structure to the ground safely and within the allowable deformations. The design calculations should ensure that the maximum loads on the foundations are much lower than the capacity of the chosen foundation. This calculation is most dependent on the ultimate strength of the soil, i.e. this is a strength type calculation. The first step in design is to estimate the maximum loads on the foundations (predominantly overturning moment, lateral load and the vertical load) due to all possible design load cases – this is covered in Chapter 2 . Subsequently, the loads need to be compared with the capacity of the chosen foundation. This calculation is necessary to avoid failure of foundation; see, for example, Figure 3.27a,b, which shows two cases of ULS failure for monopiles. In the case of Figure 3.27a, the foundation fails due to soil failure around the foundation and eventually through uprooting of the foundation. On the other hand, Figure 3.27b shows the case where the pile fails by forming a plastic hinge where the overturning moment in the monopile exceeds the plastic moment carrying capacity of the pile. For a jacket on small‐diameter piles, this is also very similar. Figure 3.28 shows some of the ULS cases for anchor piles supporting a floating wind turbine, as discussed in Chapter 2 . On the other hand, Figure 3.29 shows the ULS cases for anchor piles where two types of failure mechanisms (overturning i.e. rigid body rotation and sliding/translation) are depicted.

Schematics illustrating ULS failure through exceeding the foundation's ultimate lateral capacity (left) and through plastic hinge of the pile (middle) and SLS failure when tilt angle exceeds allowable value (right).

Figure 3.27 Examples of ULS and SLS failure.

Schematics illustrating ULS loads for spar type floating wind turbine structure with arrows depicting reaction forces, restoring moment, wave load, and wind load,

Figure 3.28 ULS loads for spar type floating wind turbine structure.

Schematics illustrating ULS cases for design of anchor foundations for floating systems: rigid body rotation (left) and horizontal translation (right).

Figure 3.29 ULS cases for design of anchor foundations for floating systems.

3.6.2 Serviceability Limit State (SLS)

The SLS corresponds to tolerance criteria associated with the regular and normal use of the wind turbine, including: (i) excessive deflection leading to the second order effects modifying the distribution of loads between supported and supporting structures; (ii) excessive vibration jeopardising the functioning of the nonstructural components; (iii) displacements that exceed the limitation of the equipment; (iv) differential settlements of foundation and soil causing intolerable tilt of the wind turbine; and (v) temperature‐induced excessive deformations.

SLS is also directly linked to target natural frequency (eigen frequency) as discussed in Chapter 2 due to possible amplification due to dynamics. For these calculations, prediction of the natural frequency of the whole system (eigen frequency) is necessary. As natural frequency is concerned with very small amplitude vibrations, linear eigen value analysis will suffice and prediction of initial foundation stiffness is necessary. Therefore, determination of stiffness of the foundation is an important design step.

3.6.3 Fatigue Limit State (FLS)

The FLS is related to cumulative damage due to repeated loads. For the case of monopile, this would require predicting the fatigue life of the monopile, as well as effects of long‐term cyclic loading on the foundation. Again, this step requires stiffness of the foundation.

3.6.4 Accidental Limit States (ALS)

ALS considers potential accidental or unexpected loads (e.g. vessel impact) that can lead to loss of global or local structural integrity.

3.7 Other Design Considerations Affecting the Limit States

This section discusses design considerations affecting limit states, including scour, corrosion, and marine growth.

3.7.1 Scour

Scour, the manifestation of which is a local lowering of seabed around foundations, is essentially sediment transport or erosion of soil and is caused by waves and current. Scour can affect the structural stability of foundations by making the foundations unsupported up to the scour depth and is an important design consideration. Scour is usually quantified in terms of its equilibrium depth S and the scour extension Ls as shown in Figure 3.30.

Schematic illustration of scour presenting a shaded vertical bar with double-headed arrows labeled Ls and s. A rightward arrow (current) points to the bar. The waves at the upper portion of the bar are labeled.

Figure 3.30 Definition of scour terminology.

The scour depth depends on the current speed and there is a threshold current speed at which undisturbed sand far from the pile just starts to move. If the current speed is lower than the threshold, scour does not develop and the seabed remains stable. However, if the current speed increases scour starts to develop. Table 3.2 summarises empirical formulas on scour depth based on literature and design standards. Caution must be exercised while using these expressions for large diameter monopiles as they may not have been calibrated.

Table 3.2 Formulas for scour depth.

Publication Flow condition Relationships
Sumer et al. (1992) and adopted in DNV‐OS‐J101 (2014) Steady Current
KC → ∞
images
Sumer et al. (1992) adopted in DNV‐OS‐J101 (2014) Waves
KC ≥ 6
images

The equations given in Table 3.2 are governed by Keulegan‐Carpenter (KC) number given by the following equation:

equation

where Umax is the maximum value of the orbital velocity at the bed, T is wave period, and DP is the pile diameter. DNV code recommends using linear theory to find the orbital velocity at the bed.

KC number is a nondimensional number representing the relative importance of drag forces over inertia forces for objects in an oscillatory fluid flow. Scour protection measured are often taken to avoid unplanned maintenance.

3.7.2 Corrosion

Monopiles located in offshore marine environments are susceptible to corrosion, which has implication on long‐term performance and fatigue life. Current design standards such as the DNV and literature suggest different methods to account for in design:

  1. Corrosion allowance, i.e. extra wall thickness, added during design to compensate for any reduction in wall thickness during design life. The corrosion allowance is based on the corrosion rate, which is a minimum of 0.1 mm per year following DNV.
  2. Use a coating system such as zinc paint.
  3. Cathodic protection is a technique to prevent the corrosion of steel surface by making the surface to act as a cathode of an electrochemical cell.

3.7.3 Marine Growth

The flora and fauna (such as bacteria, plant, and animal life) surrounding an offshore submerged structure may causes marine growth, see Figure 3.33. This will have implications on the structural design as it adds weight, influences the geometry and surface texture, and affects the corrosion rate. Marine growth can be due to tubeworms, barnacles, corals, and sea anemones, which colonise on a structure shortly after installation. The thickness and rate of growth depend on numerous factors such as salinity, oxygen content, pH, current, and temperature, age, orientation, and depth of the structural component below sea level. Structurally, the extra weight needs to be addressed relative to the total mass the foundations and when performing dynamic analysis of the whole system. The increased thickness will also increase the hydrodynamic loads. In the absence of field data, the DNV recommendations provided in Table 3.3 may be used as a guidance. The values are based on Norwegian and UK waters and are provided to give the reader an indication about the size and impact of marine growth on the structural performance of offshore wind turbine foundations.

Photo displaying marine growth around piles.

Figure 3.33 Marine growth around piles.

[Sources: Foundation Zone]

Table 3.3 Marine growth.

Marine growth thickness (mm)
Depth below mean water level (m) Central and north sea Norwegian sea
−2 to 40 100 60
>40 50 30

3.8 Grouted Connection Considerations for Monopile Type Foundations

Additional considerations must be considered:

  • Connection design between monopile, TP, and the tower
  • Long‐term tilt considerations in light of 30 years of uncertain wind and wave conditions

For monopile type of foundations, transition pieces are meant to transmit high bending moments from the superstructure to the pile. Grout were adopted for this connection for most of the initial European projects for prior oil and gas experience, speedy construction, and apparent cost savings. Subsequently, it has been observed that these grouts have settled, cracked, and failed to deliver the intended performance for monopiles. Back analysis revealed that most of these designs excluded reinforcing shear keys and resulted in expensive retrofitting on many European projects and in some cases legal battles.

In this context, it must be mentioned that for decades oil & gas (O&G) platform jackets used API designed grouted connections. For O&G jackets, the grout connection transfers mainly axial load. In contrast, OWT structures are subjected to cyclic overturning moments apart from the axial loads and the stresses in the grout are therefore a combination of axial, cyclic bending, and shear. Behaviour of grouted connections is still an area of active research and industry best practice, and codes are therefore constantly being updated.

Other connection options that are being considered are:

  • Bolted flange connections (e.g. in Scroby Sands Wind Farm)
  • Use conical TP sections as a solution

3.9 Design Consideration for Jacket‐Supported Foundations

Jackets or seabed frames supported on multiple shallow foundations are currently being installed to support offshore wind turbines in deep waters ranging between 23 and 60 m – see, for example, Borkum Riffgrund 1 (Germany, water depth 23–29 m), Alpha Ventus Offshore (Germany, water depth 28–30 m), Aberdeen Offshore wind farm (Scotland, water depth 20–30 m) (4C Offshore Limited). Typical design is three‐ or four‐legged jackets supported on either deep foundations (piles) or shallow foundations (suction caissons). The height of the jacket currently in use is between 30 and 35 m where height is governed by water depth and wave height. However, it is expected that future offshore developments will see jacket heights up to 65 m to support larger turbines (12–20 MW) in deeper waters. Figure 3.34 shows a schematic of a three‐legged jacket inspired by some recent offshore developments.

Schematic (left) and photo (right) of a three-legged jacket supported on a suction caisson.

Figure 3.34 Schematic of a three‐legged jacket supported on a suction caisson.

[Source: DONG Energy (now known as Orsted) and Carbon Trust Offshore Wind Accelerator Project.]

A jacket needs to be engineered towards a no‐rocking solution by optimising two parameters: (i) ratio of vertical stiffness of the foundation to lateral superstructure stiffness; (ii) aspect ratio of the jacket‐tower geometry. A low value of vertical foundation stiffness values together with a low aspect ratio will promote a rocking mode of vibration. On the other hand, a high vertical stiffness of the foundation with higher aspect ratio (broader base of the tower) will encourage a sway‐bending mode. It can be shown that the transition from rocking to sway‐bending is nonlinear and depends not only on the aspect ratio but also on the ratio of vertical stiffness of the foundation and lateral stiffness of jacket‐tower configuration.

Other aspects are design of the joints for the jacket, together with the fatigue design. Welds must be periodically checked.

3.10 Design Considerations for Floating Turbines

Floating offshore wind farms (FOWFs) are emerging as a solution with many demonstration projects over the last decade. Design considerations include the selection of the anchoring system, which has a significant impact as it influences the mooring radius, seabed penetration, required anchor size, length of the forerunner, as well as the anchoring (transportation and installation, [T&I]) costs. There is a set of design choices for floating wind turbine platforms, with the most important ones being the stabilisation type (spar, semi‐sub, tension‐leg platform, semi‐spar), the mooring system type (catenary, taut, or tension‐leg), and the anchor type (gravity anchors, drag anchors, pile anchors, suction caissons). Considerations that are also important are the forerunner and mooring chain types (chain, rope, cable, etc.).

The most common approach in the mooring system design is to decouple the sea‐keeping analysis (i.e. hydrodynamic motions analysis of the platform, including the mooring system) and the anchor system design. In other words, the expected anchor loads may be obtained and an anchor chosen without accounting for any effect of the anchor on the seakeeping of the platform.

Catenary mooring lines have so far been the most common choice for tested floating wind turbine concepts such as Hywind. In the case of catenary mooring lines, a so‐called theoretical anchor point (TAP) is determined on the seabed through hydrodynamic motions analysis of the platform. The TAP is a point where no motion of the mooring line is expected over the life of the platform. The mooring line tension can be determined for this point, and thus the anchor loads can be calculated and the anchor chosen accordingly. This requires the assessment of the following for various types of anchors:

  • Load transfer process
  • Failure mechanisms
  • Anchor size and weight
  • Required seabed preparation
  • Required seabed area
  • Impact of soil conditions on site and the available knowledge about it in the early phases of the project
  • T&I
  • Forerunner type and length
  • Long‐term behaviour
  • Overall anchoring costs

3.11 Seismic Design

Aspects of seismic design are provided in Chapter 2. However, there are two important considerations:

  1. The response of the ground to the expected strong motion. Ground may respond in one of the two ways: either by amplifying the motion (site amplification) or some layers may liquefy. As a rule of thumb, loose‐ to medium‐dense sandy soil liquefies and clay layers typically amplify the motion.
  2. The change in properties of the supporting soil (strength and stiffness) will alter the response of the wind turbine structure. These need to be evaluated and allowance must be made so as to not have permanent deformation of the structure.

3.12 Installation, Decommission, and Robustness

These considerations will ascertain that the foundation can be installed and taken off the ground and that there are adequate redundancies in the system. Figure 3.35 shows an installed wind farm.

Photo of London array wind farm after installation.

Figure 3.35 London array wind farm after installation.

3.12.1 Installation of Foundations

Installation of large monopiles is currently carried out in the following ways: (i) hydraulic or vibration hammers from above water; (ii) underwater hydraulic hammers; (iii) in different ground conditions, drill‐drive‐drill operation; (iv) drill and grout. In comparison to drilling operations, driving a pile is faster and is relatively less weather sensitive. However, driving may damage the pile head and there may be issues with verticality. As expected, drill‐drive is slower than driving and is used if the driving operation didn't reach the required penetration depth. There are some calculations that engineers must carry out, and these are discussed next (Figure 3.36).

Photo displaying driving of piles in London Array wind farm.

Figure 3.36 Driving of piles in London Array wind farm Courtesy: London Array.

3.12.1.1 Pile Drivability Analysis

One‐dimensional wave equation analysis is used to simulate the pile response to the driving equipment. The analysis is based on the assumption that from the moment of impact the ram starts to transmit its energy to the pile cap and pile head, and an energy wave starts to travel through the pile at high velocity. The wave is assumed to be one‐dimensional, acting longitudinally down the pile axis. The amplitude of the wave depends on the energy transmitted. Due to the energy losses in the whole system of hammer, pile, and soil, the amplitude of the wave decreases during travelling down. If sufficient energy is left once the wave has reached the pile tip, the pile starts to penetrate into the soil. As soon as all the energy is consumed, the pile stops penetrating and a permanent set is reached. This set for one blow can be computed by a program, and this is an iterative process. The program needs an initial guess of soil resistance to driving (SRD). Usually, the static‐bearing capacity data are used to determine the initial guess of SRD. In the next steps of the analysis, SRD is varied, keeping all the remaining pile and hammer parameters constant, and the corresponding permanent sets are computed. When the so‐obtained permanent sets are plotted versus the corresponding SRD, a blowcount resistance curve is obtained. A blowcount resistance curve shows the relation between blowcounts and SRD for certain pile, hammer, and soil conditions. There are many such program (GRLWEAP, CAPWAP) that can carry out the analysis. The output of the analysis is a prediction of driving stresses, hammer performance, and total driving time. These studies allow the designer to optimise the hammer requirements for a certain pile and soil condition.

3.12.1.2 Predicting the Increase in Soil Resistance at the Time of Driving (SRD) Due to Delays (Contingency Planning)

During pile installation, often driving has to be stopped and restarted, for various reasons such as changes of cushion or hammer or the addition of a follower. Such delays typically last from a few hours up to a few days. At the design stage, it is necessary to estimate both the SRD during continuous driving and also the amount of set‐up (increase in pile driving resistance) that may occur during delays, to ensure that the hammers taken offshore are sufficient to meet all eventualities. This is a critical design consideration for offshore pile installation projects. Empirical relationships have been proposed for quantifying set‐up – see Table 3.4 for one such case. Set‐up is primarily dependent on three mechanisms or physical processes – consolidation, stress relaxation, and soil ageing. Theoretically, all these mechanisms/processes start acting as soon as the pile penetrates the ground. It is, however, still uncertain which of the above dominate in a particular soil condition, how long each process continues, and the contribution of each component to the observed overall set‐up. There is widespread evidence showing that the capacity of most driven piles increases with time after driving; see Bhattacharya et al. (2009b).

Table 3.4 Empirical formulas for predicting pile capacity with time.

Reference Equation Type of soils Comments (all times are in days)
Skov and Denver (1988) images where Qt and Q0 are the pile capacities at time t and t0 respectively. Sand and clay For sand: A = 0.2, t0 = 0.5
For clay: A = 0.6, t0 = 1.0
Huang (1988) Qt = QEOD + 0.236{1 + loge(t)(Qmax − QEOD)} Soft ground soil of Shanghai Qt = Pile capacity at time t
Qmax = Maximum pile capacity
QEOD = Pile Capacity at the end of driving (EOD).
Guang‐Yu (1988) Qt = QEOD(0.372St + 1) Piles driven in soft soils. Q14 = pile capacity at 14 days and St is the sensitivity of soil.
Svinkin (1996) Qt = A. QEOD. t0.1 Sand For upper bound; A = 1.4
For lower bound; A = 1.025

3.12.1.3 Buckling Considerations in Pile Design

Piles can never be assumed to be straight and free from residual stresses due to construction processes. Apart from out‐of‐line straightness (see Figure 3.37), a pile may be thin walled and therefore vulnerable to local buckling (see Figure 3.38). This has been observed in a number of cases where offshore piles have collapsed during driving due to progressive closure of the internal dimensions – the initiating mechanism being local buckling. Thus, the design method has also to consider the interaction between local and global buckling. Figure 3.39 shows a typical offshore installation and Figure 3.40 shows the pile stick‐up. Once the pile is in the sleeve, it is important to check the buckling potential of the pile under the action of the lateral forces due to the wave loading and the hammer weight. Often attachments are used for transportation and lifting, see Figure 3.41, and it is necessary to evaluate the impact of such attachments on pile driving. Details of buckling considerations can be found in Bhattacharya et al. (2005) and Aldridge et al. (2005).

3 Graphs of pile penetration or pile length versus deviation from base illustrating imperfections in driven pile. Each graph displays a descending curve. The right graph has the longest curve.

Figure 3.37 Imperfections in driven pile.

Image described by caption.

Figure 3.38 Pile tip buckling. The figure shows changes in cross‐sections leading to progressive pile collapse during driving following an initial deformation of the pile tip.

Schematic of a typical offshore pile installation with arrows depicting the pile, pile guide, grout packer, mudmat, grout, pile sleeve, and sea bed (mudline).

Figure 3.39 A typical offshore pile installation.

Schematic of a pile stick-up (Ø2.134 m) with arrows depicting the pile guide, pile sleeve (Ø2.234 m), wave induced lateral load, and hammer. The distance between the sea level and the sea bed (mudline) is 60 m.

Figure 3.40 Pile stick‐up.

Photo of attachments in pile for transportation and lifting.

Figure 3.41 Attachments in pile for transportation and lifting.

3.12.2 Installation of Suction Caissons

The installation of suction caissons occurs in two distinct stages.

3.12.2.1 First Stage

In this stage, the buoyant self‐weight of the caisson and attached structure pushes the caisson into the marine bed. This buoyant weight is resisted by a combination of skin friction and annulus‐bearing force. The caisson will continue to sink into the sea bed until a sufficient portion of the caisson skirt is embedded in the sediment, creating an equivalent resistance to the self‐weight of the structure. During this installation, valves are left open in the caisson lid to allow water to flow out of the space encompasses by the caisson. If sufficient water pressure develops in the caisson, outward piping failure can occur in the sediment surrounding the caisson. One of the design considerations is to allow sufficient drainage capacity within the caisson lid to prevent this effective from having more drainage valves.

3.12.2.2 Second Stage

In this stage of installation, pumping the water (contained in the cavity that exists between the caisson lid and the soil plug) is carried out. The act of pumping water out of this void is twofold: the first is to increase the downward force, effectively pulling the caisson into the sea bed. The second is to create an upward hydraulic gradient in the soil mass that loosens the soil around the caisson annulus and on the inside of the caisson skirt. This suction needs to be limited to avoid boiling of the sediment on the inside of the caisson. This will still further install the caisson over that achieved through self‐weight installation. In some instances, caissons may be installed without applying suction (stage one only); this is achieved by using a sufficiently large surcharge. A simplified figure showing the installation process is shown in Figure 3.42.

The major reference for the installation load for suction caisson have come from a series of papers produced by Houlsby and Byrne (2005a,b). These estimations are based on theory and corroborated against experimentally obtained data. Both papers provide a conclusive guide as to caisson installation in sand and clay.

Schematics illustrating stages of suction caisson installation, (left) self-weight, (middle) suction assisted, and (right) installed.

Figure 3.42 Stages of suction caisson installation, (a) self‐weight, (b) suction assisted, and (c) installed.

3.12.3 Assembly of Blades

The blades can be assembled in two ways: (i) each blade is individually attached see Figure 3.43 (ii) star type: All the blades are assembled and lifted and installed. Figure shows from London Array wind farm see Figure 3.44.

Photo of London array.

Figure 3.43 London Array installation.

Image described by caption.

Figure 3.44 UK – Ormond Site [Photo Courtsey: A2SEA].

For jackets or seabed frame supported on multiple piles, there are two methods:

  1. Subsea templates are first laid in the seabed, and piling is done at the correct location using the template (also known as pre‐piling). The jackets are then transported and the jacket leg‐pile connection grouted.
  2. Jackets are transported and then laid in the seabed using mudmats. The piles are then driven.

Figure 3.45 shows installation photograph for Germany's Alpha Ventus offshore Wind farm in deeper waters.

2 Photos displaying installation using a template.

Figure 3.45 Installation using a template.

[Source: Alpha Ventus Offshore Wind Farm Project.]

3.12.4 Decommissioning

Decommissioning of energy infrastructure is increasingly becoming an important design consideration from the view of sustainability. This section of the chapter provides a case study of the decommissioning of Lely Wind Farm (Netherlands) where the monopile is also extracted from the ground. All the self‐explanatory photos are provided. In this context, it is necessary to provide an overview of the current legislation on offshore installations.

For the United Kingdom, decommissioning of offshore structures is regulated by UK law and by the OSPAR Decision 98/3 on the Disposal of Disused Offshore Installations (OSPAR Commission 1998). However, the removal of offshore piles is not required under legislation. Typically, the piles are cut below the seabed, at a suitable distance appropriate, often around three metres below the natural seabed.

3.13 Chapter Summary and Learning Points

Offshore wind turbines are new types of offshore structure characterised by low stiffness (as a result, flexible and having low natural frequency) and therefore sensitive to the dynamic loading imposed on them. The design guidelines available for offshore oil and gas installation foundations cannot be direct extrapolated/interpolated to offshore wind turbine foundation design.

3.13.1 Monopiles

Monopiles have been predominantly used to support WTGs in water depths up to 30 m. However, there are discussions with regard to the use of monopiles in deeper water depths termed as ‘XL’ monopile. Preliminary calculations suggest that 10 m diameter monopiles weighing 1200 t may be suitable for 45 m water depth, of course dependent on ground conditions. However, the use is uncertain due to the following:

  • No codified cyclic design to predict long term tilt;
  • Lack of redundancy in foundation system and therefore chance of single‐point failure;
  • Installation costs and lack of adequate specialised vessels;
  • Connection between foundation, transition piece, and the tower.

Some of these aspects are described below in further details.

  1. Lack of redundancy. Monopiles are ‘overturning moment’ resisting structures, and there are two main components: (i) overturning moment arising from the thrust acting at the hub level; (ii) overturning moment due to the wave loading. Also, these two moments can act in two different planes and will vary constantly, depending on the time of the day and time of the year. Monopiles are rigid piles and the foundation collapse can occur if the soil around the pile fails, i.e. there would be rigid body movement. If the foundation starts to tilt, it is very expensive to rectify.
  2. Cyclic (rather dynamic) design of monopile. The response of monopile under cyclic/dynamic load is not well understood and there is a lack of guidance in codes of practice. If cyclic design is incorrect, monopiles can tilt in the long term. If the tilt is more than the allowable limit, the turbine may need a shutdown. Monopile design is usually (also wrongly) carried out using API design procedure calibrated for flexible pile design, where the pile is expected to fail by plastic hinge.
  3. Issues related to installation of monopiles. Large monopile installation requires suitable vessel availability, as well as specialised heavy lifting equipment. Other issues are noise refusals, buckling of the pile tip, drilling out, and grouted connections. If the site contains weak rock (siltstone/sandstone/mudstone) and where the local geology shows bedrock or hard glacial soils at shallow depths, drive‐drill‐drive techniques may be required, with subsequent increases in cost and schedule. It must be mentioned here that driving reduces the fatigue life.

3.13.2 Jacket on Flexible Piles

There has been a considerable interest in jacket‐type structures for deeper‐water applications but is perceived being expensive due to the amount of steel required. However, jackets supported on piles can be considered as a safe solution due to excellent track record of good performance in offshore O&G industry. Offshore O&G industry has been using long flexible piles (diameters up to 2.4 m), which are easy to drive, and necessary vessels are readily available (relatively as opposed to vessels to install monopiles). This aspect will drive down the TIC (time in construction) costs regarding piling, and also, large vessels are not required for pile installation. However, there are costs associated with jacket installation. One of the requirement is the optimisation of the jacket so as to consume minimum steel. There are two types of jacket – normal jacket or twisted jacket. The advantage of the twisted jacket over normal jacket is fewer number of joints and therefore less fatigue issue.

3.13.3 Jackets on Suction Caissons

Jackets on suction caissons need to be designed so that rocking modes of vibration are avoided.

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

You can't read the all page of ebook, please click here login for view all page.
Reset
18.117.152.250