Sunteți pe pagina 1din 12

Geotechnical Design and Design Optimization of a

Pile-Raft Foundation for Tall Onshore Wind Turbines in


Multilayered Clay
Shweta Shrestha, S.M.ASCE1; Nadarajah Ravichandran, M.ASCE2; and Parishad Rahbari, S.M.ASCE3
Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 01/29/18. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Although the pile-raft foundation is preferred for supporting a tall wind turbine, the geotechnical design and selection of suitable
design parameters are based on a complex procedure. Except the foundation, all the other aboveground components are precast members that
are assembled at the project site to build a wind turbine. Therefore, it is necessary to consider the possible variations in soil properties and wind
speed in the design of the foundation. In this paper, a reliability-based robust design procedure for a pile-raft foundation that supports a 130-m-
tall wind turbine on a layered clayey soil is presented. Upon completion of the geotechnical design for the mean wind speed and undrained shear
strength, a parametric study and Monte Carlo simulation were conducted by varying the wind speed and the undrained cohesion of each layer to
establish a relationship among the design variables (number and length of piles and radius of the raft) and the random variables (wind speed and
undrained cohesion). Finally, a reliability-based robust design was created considering the total cost and robustness as the objectives. The stand-
ard deviation of the response of concern, which is the differential settlement, was considered the measure of robustness. The optimization
yielded a set of preferred designs known as the Pareto front, and the suitable design was selected for a given cost limitation and performance
requirement using the Pareto front. DOI: 10.1061/(ASCE)GM.1943-5622.0001061. © 2017 American Society of Civil Engineers.
Author keywords: Pile raft; Geotechnical design; Design optimization; Wind turbine; Hybrid foundation.

Introduction higher energy at a small additional cost. Lewin (2010) found that an
increase in the turbine height from 80 to 100 m results in a 4.6%
Although it is one of the fastest growing clean and renewable ener- greater wind speed and a 14% increase in power output, and that an
gies in the world, wind energy accounts for only 3.3% of the total increased height from 80 to 120 m results in an 8.5% greater wind
electricity generated worldwide. Nonetheless, according to the speed and a 28% increase in power production. Because the initial
2015 Global Wind Report (GEWC), the global cumulative installed construction cost of a wind farm covers the highest percentage of
wind capacity greatly increased by the end of 2015, up 17% from the total cost of the project, it is logical and cost-efficient to increase
the preceding year (GEWC 2015). By the end of 2015, the United the wind energy production by building taller towers that can gener-
States added 4,000 new wind turbines contributing approximately ate additional power from the same number of wind turbines.
8,598 MW of energy, which increased the total installed capacity, Although it increases the wind energy production of a single
from the end of 2014, by 13% (GWEC 2015). Although a signifi- wind turbine, building taller towers poses significant challenges to
cant number have been installed in the United States, wind turbines the geotechnical engineer in designing and selecting the most cost-
account for only 4.7% of the total electricity produced nationwide. efficient foundation for the given subsurface and wind conditions.
The selection of suitable locations for onshore wind farms depends A taller wind turbine tower not only increases the vertical dead load
on factors such as wind speed, soil condition, availability of con- but also significantly increases the lateral load and bending moment
struction material, environmental impacts, and other limitations at the base of the tower. Larger design loads, especially the moment,
imposed by local and federal agencies. not only make the foundation design more complex but also make it
The energy output of individual wind turbines can be increased larger, demanding that a significant amount of resources be allotted
by building taller towers to access higher and steadier wind. It has into foundation design and construction to meet safety and service-
been shown that wind energy is directly proportional to the third
ability requirements. Because a significant percentage of the total
power of wind speed, so taller towers can produce significantly
cost of installing a wind turbine is allocated for the design and
construction of the foundation, it is necessary to develop new
1
Graduate Student, Glenn Dept. of Civil Engineering, Clemson methodologies to design and select the most cost-efficient foun-
Univ., 123 Lowry Hall, Clemson, SC 29634. E-mail: shwetas@ dation for a given set of geotechnical and wind conditions.
clemson.edu Typically, the mat (raft) foundation, the pile group foundation,
2
Associate Professor, Glenn Dept. of Civil Engineering, Clemson and the pile-raft foundation are used to support wind turbines,
Univ., 202 Lowry Hall, Clemson, SC 29634 (corresponding author). depending on the subsurface condition, tower height, and wind
E-mail: nravic@clemson.edu speed at the site. The raft foundation is easy to construct and pro-
3
Graduate Student, Glenn Dept. of Civil Engineering, Clemson Univ., vides significant bearing capacity because of its larger footprint,
123 Lowry Hall, Clemson, SC 29634. E-mail: prahbar@clemson.edu
but its design is controlled by differential and total settlements,
Note. This manuscript was submitted on March 15, 2017; approved on
August 18, 2017; published online on November 28, 2017. Discussion pe- especially when subjected to larger loads. In large-load situations,
riod open until April 28, 2018; separate discussions must be submitted for deep foundations are added to the raft foundation to create what is
individual papers. This paper is part of the International Journal of known as a hybrid or pile-raft foundation, which is economical
Geomechanics, © ASCE, ISSN 1532-3641. for supporting tall wind turbines.

© ASCE 04017143-1 Int. J. Geomech.

Int. J. Geomech., 2018, 18(2): 04017143


It is unfortunate that the mobilized capacities of the piles and raft Site Condition and Design Loads
vary with the amount of settlement because the variation in the
mobilized capacities greatly complicates the design of a pile-raft
Wind Farm Site and Soil Properties
foundation. Also, because of the large number of design variables
(radius of raft, number of piles, length of piles, etc.), many designs A site in the city of Charleston, South Carolina, along the east coast
can be produced, leaving the design engineer to pick the design ran- of the United State was selected in this study. The soil profile and
domly or with little knowledge. Therefore, a proper methodology other necessary subsurface soil properties required for the design
must be developed to help the engineer select the most appropriate were obtained from a geotechnical report for a location in
foundation for the given variations in the wind speed and soil condi- Charleston, South Carolina (WPC 2010). It is noteworthy that this
tions. This paper details the efforts of the authors to develop such a geotechnical report was produced for the construction of one of the
world’s largest turbine testing facilities in the world. The site con-
Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 01/29/18. Copyright ASCE. For personal use only; all rights reserved.

methodology for performing a reliability-based robust design of a


pile-raft foundation. sists of three clay layers beneath a thin sand layer at the surface. The
The procedures currently available to perform geotechnical summary of the soil profile and geotechnical parameters are pre-
design of a pile-raft foundation are broadly classified as simplified sented in Table 1. The undrained cohesion, cu, friction angle, f ,
methods, approximate computer-based methods (Sinha and Hanna and modulus of elasticity, E, tabulated and listed in Table 1 were
2017; Reul 2004), and more rigorous computer-based methods calculated using empirical correlations between these parameters
(Poulos 2001a). The simplified methods predict the behavior rea- and the cone-tip resistance obtained from the CPT7 profile of the
sonably well when the load is vertical but fail to predict behavior geotechnical report (WPC 2010). However, the moduli of elasticity
when there are lateral and moment loads. However, the computer- of the second and third layers were calculated using the empirical
based methods are widely used in practice for designing pile-raft correlation between cu and E obtained from the engineering manual
foundations subjected to combined vertical, lateral, and moment (EM 1110-1-1904) for the settlement analysis by U.S. Army Corps
loads. Such computer-based design procedures require knowledge of Engineers (USACE 1990). At this site, the groundwater table
and use of sophisticated finite- element or finite-difference com- was located at 1.52 m below the ground surface.
puter programs that may be unavailable for many practicing engi-
neers. The accurate representation of stress-strain behavior of the Design Loads
supporting soil and the interaction between the raft-soil and the
The loads for the design of the foundation for the wind turbine con-
pile-soil is the greatest challenge involved in computer-based
sist of a dead load due to the self-weight of the superstructure and
design. Consequently, practicing engineers develop simplified the lateral load and bending moment due to horizontal wind. The
models that may lead to less-than-accurate results. procedures used to determine these loads for the North Charleston,
The incorporation of uncertainties in the loading (wind speed) South Carolina, location are detailed in the following subsection.
and soil properties is also of great importance to the geotechnical
design of foundations that support tall wind turbines. Indeed, engi- Dead Load
neers have great difficulty selecting a design that is not only eco- The dead load is the total vertical load consisting of the weight of
nomical but also satisfies the performance requirements when there the tower, rotor and rotor blades, nacelle with drive train, electronic
is a significant variability in the soil properties and loading. In par- equipment, and the other wind turbine components. The weight of
ticular, the use of a pile raft introduces a significantly large number the tower is calculated based on the volume of the tower and the
of design variables (e.g., raft radius, number of piles, arrangement unit weight of the tower material. The wind turbine tower consid-
of piles, and length of piles), which greatly increase the difficulty in ered in this study was a hybrid hollow cylindrical concrete and steel
selecting the appropriate final design. In such situations, a tapering tower with the lower 93 m of concrete and the upper 37 m
reliability-based robust design optimization technique can be used of steel (Grünberg and Göhlmann 2013). The tower diameter at the
to shortlist the best candidates and select the most suitable design base and top was 12.0 and 4.0 m, respectively, and the thickness
by imposing appropriate limitations. varied between 0.04 and 1.2 m. The unit weight of the concrete and
In this study, a geotechnical design and an optimization steel used in this study was 23.6 and 78.5 kN/m3, respectively. The
procedure for a pile-raft foundation are presented for a sample appropriate weights of the nacelle and rotor for the tower height
130-m-tall hybrid wind turbine tower with a mean wind speed of were obtained from Malhotra (2011). The final dead load, P, was
201.17 km/h at a potential wind farm site in Charleston, South calculated to be 51.71 MN.
Carolina. A parametric study was also conducted to understand
the effect of uncertain parameters, known as random variables, Wind Load
such as wind speed and soil properties (undrained cohesion in The wind load was calculated considering the mean survival wind
this study) on the design variable, such as radius of raft, number speed of 201.17 km/h following the procedure described in “Standard
of piles, and length of piles, and on the material cost of the foun- ASCE/SEI 7-10” (ASCE 2010). Because most wind turbines have a
dation. In addition, a reliability-based robust design optimiza- survival wind speed of between 180.25 and 215.65 km/h (Wagner
tion procedure is presented to simplify the selection of design and Mathur 2013), a wind speed range of 143.23–259.10 km/h with a
parameters for a site condition. mean survival wind speed of 201.17 km/h was considered appropriate

Table 1. Generalized Soil Properties

Layer Soil Depth (m) Unit weight (kN/m3) cu (kPa) f 0 (degrees) E (kPa) Poisson’s ratio
1 Medium dense sand 0–1.22 17.28 — 50.1a
2.75  10 4a
0.4
2 Soft to firm clay 1.22–9.15 16.50 98.81 0 1.48  104 0.5
3 Cooper marl 9.15–25.00 19.64 106.66 0 3.20  104 0.5
a
Calculated using the CPT data but not used in the design because the bottom of the raft rests on the second layer.

© ASCE 04017143-2 Int. J. Geomech.

Int. J. Geomech., 2018, 18(2): 04017143


for this study. The wind acts on the aboveground components of the The ultimate vertical capacity calculated by adding the capaci-
wind turbine, such as the tower, blades, and the rotor-induced lateral ties of the piles and the raft was found to be lower than that calcu-
load and moment at the base of the tower. These loads were computed lated assuming the piles and raft as a single block. The final factor
by considering the wind load along the tower height and the drag force of safety for the vertical load capacity was determined to be 3.41,
acting on the nacelle. The total lateral load, V, and the bending which met the design requirement.
moment, M, were calculated to be 2.26 MN and 144.89 MN·m,
respectively. Design for the Moment Load
The ultimate moment capacity of the pile-raft foundation was esti-
Geotechnical Design of Pile-Raft Foundation mated as the lesser of (1) the ultimate moment capacity of the raft
plus the ultimate moment capacity of piles and (2) the ultimate
Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 01/29/18. Copyright ASCE. For personal use only; all rights reserved.

The basic idea behind the use of a pile raft is to increase the bearing moment capacity of a block containing the piles, raft, and soil. The
capacity of the foundation with the use of a raft and to decrease the ultimate moment capacity of the raft, pile group, and block of the
total and differential settlements with the use of deep foundations. pile raft were determined using the method presented in Hemsley
However, the quantification of the exact percentages of total loads (2000). The key equations are summarized in the following subsec-
carried by the raft and by the piles is the most challenging aspect in tions for the sake of completeness.
the design of a pile-raft foundation. This difficulty is mainly due to
a lack of understanding of the complex interaction among the soil, Case 1: Ultimate Moment Capacity of the Pile Raft
raft, and piles and the mobilized strengths along the interface at Considering Individual Capacity
given total and differential settlement values. Thus, a reliable design The ultimate moment capacity of the raft, Mu-R, was calculated
guideline is not yet available, especially for the foundation sub- using Eq. (2) (Hemsley 2000) as follows:
jected to combined moment, lateral, and vertical loads. "
In this study, a preliminary geotechnical design of the pile-raft  1=2 #
Mu-R 27 P P
foundation was performed following the procedure outlined by ¼ 1 (2)
Mm 4 Pu Pu
Hemsley (2000) in which the procedures proposed by Poulos and
Davis (1980) and Randolph (1994) were incorporated. The factors
considered in the preliminary design were the ultimate vertical, where Mm = maximum possible moment that the soil can support;
moment, and lateral geotechnical capacities, total elastic and differ- P = applied vertical load; Pu = ultimate centric load on the raft when
ential settlements, the rotation of the tower due to wind load, and no moment is applied. For this method, the maximum moment for
the lateral movement of the foundation. The size of the raft and the circular raft, Mm is given by
size and number of the piles required to satisfy the design require-  
ments were determined in the preliminary design stage. The qu D3 p 1
Mm ¼  (3)
capacity of the pile-raft foundation was checked for vertical load, 4 4 3
lateral load, bending moment, total and differential settlements, and
rotation. A minimum factor of safety of 2 was considered safe for where qu = ultimate bearing capacity of raft; and D = diameter of
the vertical load, lateral load, and bending moment as suggested by circular raft.
Hemsley (2000), and a vertical misalignment of 3 mm/m was con- The ultimate moment of all the piles in the foundation system,
sidered safe for the rotational stability of the tower (Grünberg and Mu-P, was estimated using Eq. (4) (Hemsley 2000) as follows:
Göhlmann 2013). A spreadsheet (not shown) was prepared to auto-
mate the iterative calculations and to perform parametric studies. X
Np
Mu-P ¼ Puui jxi j (4)
i¼1
Design for Vertical Load
The vertical capacity of the pile raft was calculated as the lesser of where Puui = ultimate uplift capacity of ith pile; jxi j = absolute dis-
the (1) sum of the ultimate capacities of the raft, Pu-R, and all the tance of ith pile from the center of group; and Np = number of piles.
piles, Pu-P, such that Pu-PR = Pu-R þ Pu-P; and (2) the ultimate The ultimate moment capacity of the pile raft, Mu-PR, system
capacity of a block, Pu-B, that consists of the piles and the raft plus considering the individual capacity is given by
that of the portion of the raft outside the periphery of the pile group
(Hemsley 2000). The ultimate capacity of raft, Pu-R, was calculated Mu-PR ¼ Mu-R þ Mu-P (5)
using the general bearing capacity equation by Vesic (1973, 1975),
and the ultimate capacity of all the piles, Pu-P, was calculated as the
sum of the ultimate downward capacity of all the piles; that is, Case 2: Ultimate Moment Capacity of the Pile Raft
Pu-P = NpPult-dn, where Np is the number of piles and Pult-dn is the Considered as a Single Block
ultimate downward capacity of a single pile, which is the sum of The ultimate moment capacity of the block, a single unit consisting
the ultimate skin resistance, Ps, and toe resistance, Pt. In this study, of the raft and the piles, MuB was estimated using Eq. (6), (Hemsley
the a and Meyerhof’s methods, provided in Das (2011), were used 2000), as follows:
to calculate the ultimate skin and toe resistance of a single pile,
respectively. Finally, the factor of safety for the vertical load was Mu-B ¼ aB 
p u BB D2B (6)
calculated using Eq. (1), as follows:
where BB and DB = width and depth of the block, respectively;  pu =
minðPu-PR ; Pu-B Þ average lateral resistance of soil along the block; and aB = factor
FSP ¼ (1)
P depending upon the distribution of the ultimate lateral pressure with
depth (0.25 for a constant distribution of 
p u and 0.2 for a linearly
where P = design vertical load. increasing 
p u with depth from zero at the surface). It is noteworthy

© ASCE 04017143-3 Int. J. Geomech.

Int. J. Geomech., 2018, 18(2): 04017143


that Eq. (6) was proposed for designing a rectangular raft and pile stiffness of the pile raft remaining operative until the pile capacity
arrangement. However, in this study, the raft was circular in shape, was fully mobilized at load PA. With all the known values, the verti-
and therefore, the circular section was converted to an equivalent cal load versus the total elastic settlement relationships, established
rectangular section to apply Eq. (6) to calculate the ultimate by Eq. (10), as follows, were used to obtain the load versus settle-
moment capacity of the block. ment (P versus S) curve for the pile-raft foundation:

Ultimate Moment Capacity of the Pile Raft P


For P  PA ; S ¼
It was observed that the design was controlled by individual failure Kpr
because the ultimate moment capacity calculated for Case 1 was
PA P  PA
smaller than it was for Case 2. The final factor of safety for the For P > PA ; S ¼ þ (10)
moment capacity was determined to be 3.56 using Eq. (7), which Kpr Kr
Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 01/29/18. Copyright ASCE. For personal use only; all rights reserved.

met the design requirement, as follows:


The calculated vertical load versus the elastic settlement curve is
minðMu-PR ; Mu-B Þ shown in Fig. 1. It can be seen from Fig. 1 that the design vertical
FSM ¼ (7)
M load, P, was less than the load at which the pile capacity is fully mobi-
lized, PA, which means that both the piles and the raft are contributing
to the total load bearing capacity to support the applied design load.
Design for Lateral Load From the load versus total elastic settlement curve, the total elas-
tic settlement of the pile raft for the total vertical load of 51.71 MN
The lateral pile capacity of a single pile was determined using the was found to be 42.42 mm.
solutions by Broms for cohesive soil as outlined in Gudmundsdottir
(1981). The ultimate lateral load capacity and lateral deflection of a Consolidation Settlement
single pile were calculated using the horizontal coefficient of the Typically, consolidation settlement is not taken into consideration
subgrade reaction. It was assumed that all the piles would behave in for deep foundations unless the foundation design is controlled by
the same way under the application of a lateral load. Hence, the ulti- block failure. However, the proposed study deals with combined
mate lateral load capacity of a pile group was estimated as the raft and pile foundations, and therefore, the consolidation settlement
summed ultimate lateral capacities of all the piles present in the must be incorporated into the design to obtain a better design, espe-
group; that is, Vu-PR = nVu-P. The factor of safety for the lateral load cially for meeting the long-term serviceability requirement.
was calculated using Eq. (8) as follows: Because of the combined loading, the pressure distribution below
the raft is nonuniform and the calculation of the consolidation settle-
Vu-PR
FSV ¼ (8) ment due to nonuniform load is complicated. To address this issue
V accurately, the differential settlement due to the nonuniform pres-
sure distribution must be calculated. However, because of the limi-
The factor of safety was found to be 12.78 and lateral deflection tations in the procedures available in the literature, the differential
was 9.12 mm. consolidation settlement was not considered in this study. However,
the total consolidation settlement due to the average pressure below
Total Settlement the raft was calculated and incorporated in the design for this study.
Elastic Settlement For this study, the consolidation settlements of the second and
The vertical load versus the total elastic settlement response of the third clay layers due to increases in stress were calculated at the
pile-raft foundation was estimated following the approach proposed middle of each layer. The preconsolidation pressures, s c0 , of the sec-
by Poulos (2001b) in conjunction with the method of estimating ond and third layers were calculated using the correlation between
load sharing between the raft and piles presented in Randolph the cone-tip resistance and preconsolidation pressure (Mayne and
(1994). The load sharing between the raft and the piles can be esti- Kemper 1988). It was found that both layers were overconsolidated
mated on the basis of the stiffnesses of the raft, piles, and pile raft. with the final effective stress, s 0f , smaller than the preconsolidation
The stiffness of the pile raft, Kpr, was estimated using Eq. (9) as pro- pressure. The compression index, Cc, was calculated using the em-
posed by Randolph (1994) and written as pirical correlation proposed by Kulhawy and Mayne (1990).
Because he swelling index, Cs, is usually 1/5 to 1/10 of Cc (Das
Kpr ¼ XKp 2011), it was assumed to be 1/8 of Cc in this study. The initial void

1 þ ð1  2arp ÞKr =Kp


where X ¼   (9) 240
1  a2rp Kr =Kp PA = load at which pile capacity
200
Vertical load (MN)

is fully utilized

160
where Kr = stiffness of the raft; Kp = stiffness of the pile group; and
arp = raft-pile interaction factor. The raft-pile interaction factor was 120
assumed to be 0.8 because as the number of piles in the group 80
increases, the interaction factor increases and tends toward a con- Raft and piles Pile capacity is fully mobilized
40 are functioning (only raft is working)
stant value of 0.8 as reported by Randolph (1994). Among the vari-
ous methods for estimating the raft stiffness, the method outlined by 0
Randolph (1994) was used. To estimate the stiffness of the pile 0 100 200 300 400
Settlement (mm)
group, the method proposed by Poulos (2001b) was used. First, the
target stiffness of the pile raft was determined by dividing the total
Fig. 1. Calculated load versus total elastic settlement curve for a pile-
vertical load by the assumed allowable settlement. Then, Eq. (9)
raft foundation
was then solved to determine the stiffness of the pile group with the

© ASCE 04017143-4 Int. J. Geomech.

Int. J. Geomech., 2018, 18(2): 04017143


Table 2. Optimum Design Obtained from Pareto Front

Simulations Lp (m) Np Rr (m) Standard deviation of response (mm) Total cost (US$)
1,000 29.87 52 7.93 5.57 369,742
10,000 31.25 52 7.91 5.09 392,929

M
MRa Mpile
Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 01/29/18. Copyright ASCE. For personal use only; all rights reserved.

Rotaon θRa

Equilibrium: M = MRa + MPile


Condion to be met: θRa = θPile
Rotaon θPile
(a) (b) (c)

Fig. 2. Conceptual differential settlement calculation diagram: (a) pile raft; (b) rotation of raft; (c) rotation of piles

ratio, eo, was calculated using the correlation between Cc and eo pro-
posed by Hough (1957). These consolidation parameters are listed
in Table 2. Finally, the consolidation settlement, Sc, of the second
and the third layer was calculated to be 37.62 and 16.72 mm, respec-
tively, resulting in a total consolidation settlement of 54.34 mm.
The total settlement (elastic plus differential settlements) was
calculated to be 96.76 mm, which is within the tolerable limit of
100 mm for tall structures (Raju 2015). Nevertheless, it is notewor-
thy that the uniform vertical settlement of the entire system may not
be hazardous, especially for wind turbines located away from criti-
cal infrastructure.

Differential Settlement and Rotation


There is no accurate procedure available in the literature for calculat-
ing the differential settlement of the pile-raft foundation system sub-
jected to a bending moment. In this paper, a new technique is pro-
posed to calculate the differential settlement of a pile-raft foundation.
In this method, the percentages of the bending moments carried by
the raft (MRaft) and piles (MPile = M−MRaft) were adjusted until the
differential settlements of both were equal for the applied loads,
which is considered as the differential settlement of the pile-raft foun-
dation. In practice, the piles are fixed to the bottom of the raft, and the
rotation of the piles and the raft are the same. The described idea of
adjusting the loads until the rotations are equal replicates the field
condition. The vertical shortening and extension of the piles from lat-
eral deflection were assumed as negligible in this study.

Differential Settlement of the Raft


To determine the differential settlement of the raft, the rotation,
u Raft, from the wind load was first calculated using Eq. (11),
expressed by Grünberg and Göhlmann (2013) as follows:
MRaft Es
u Raft ¼ ; cs ¼ pffiffiffiffiffiffiffiffiffi (11) Fig. 3. Designed pile-raft foundation: (a) plan; (b) 3D; (c) front view
cs IRaft f 0 ARaft

where MRaft = fixed-end moment at the soil-structure interface (the was used to determine the differential settlement of the raft, assum-
percentage of moment shared by the raft to yield an equal differen- ing that the raft rotates about its center line.
tial settlement of the piles in this study); cs = foundation modulus;
IRaft = second moment of inertia; Es = modulus of elasticity of soil; Differential Settlement of Piles
f 0 = shape factor for overturning (0.25); and ARaft = area of the foun- The differential settlement of the pile group was estimated on the basis
dation. After calculating u Raft, a simple trigonometric relationship of the individual pile-settlement profile due to the resultant vertical

© ASCE 04017143-5 Int. J. Geomech.

Int. J. Geomech., 2018, 18(2): 04017143


load induced by the moment carried by the piles, MPiles, using the Parametric Study
Fellenius method presented in Coduto (2001). First, the resultant verti-
cal loads acting on each pile were calculated as the sum or difference To account for the effect of the variations in loading and the soil
of the vertical load due to the dead load and the vertical load induced properties on the design results, a parametric study was conducted
by the bending moment. Then, the difference in the pile settlement of considering possible variations in wind speed and undrained cohe-
the outermost piles in the direction of the moment was considered as sion of the second and third layers of the soil. For each case in the
the differential settlement of the piles; this concept is graphically parametric study, only one design parameter (number of piles, Np;
shown in Fig. 2. Finally, the rotation of the pile was computed. length of a pile, Lp, or radius of the raft, Rr) was changed at a time to
meet all the design requirements. In addition, the variations in the
Differential Settlement of the Pile Raft total cost of the foundation due to variations in wind speed and
The values of MRaft and MPiles (i.e., M−MRaft) were adjusted until the
Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 01/29/18. Copyright ASCE. For personal use only; all rights reserved.

undrained cohesion were also studied. The total cost of the founda-
differential settlements of the raft and the piles were equal. The corre- tion includes material, labor, and equipment costs, and it was calcu-
sponding final values were considered as the moment carried by the lated as a sum of the total costs of the raft and piles using the unit
piles and the raft, with the final differential settlement deemed to be cost of each. The unit cost for the raft and for the prestressed con-
the differential settlement of the pile-raft foundation. This exercise crete pile used in this study was $342.19/m3 and $192.19/m, respec-
resulted in a differential settlement of 44.23 mm and a rotation of tively, as obtained from RS Means data (Phillip and Adrian 2013).
0.17°, a rotation that induced the horizontal displacement of The details of the parametric study and the results are presented the
383.7 mm at the top of the tower, which is within the acceptable limit. following sections.

Design Outcome Effect of Wind Speed on the Design Variables


The final design resulted in a raft with a radius of 7.5 m and a thick- The wind speed was varied within the range of survival wind speed,
ness of 1.2 m at a depth of 1.5 m. The raft was supported by 40 pre- which is between 143.23 and 256.10 km/h with mean, m v, of
stressed concrete piles, each with width of 0.457 m and length of 201.17 km/h and standard deviation, s v of 28.97 km/h. The designs
21.8 m, arranged equally within inner and outer circumferences of were performed for five wind speeds (143.23, 172.20, 201.17,
5.3 m and 6.7 m. The final design is shown in Fig. 3. 230.13, 256.10 km/h), which represent the m v 6 2s v range,

Fig. 4. Effect of variation in wind speed on (a) number of piles and (b) total cost of pile raft

Fig. 5. Effect of variation in wind speed on (a) length of pile and (b) total cost of pile raft

© ASCE 04017143-6 Int. J. Geomech.

Int. J. Geomech., 2018, 18(2): 04017143


following the procedure presented in the previous section and The parametric studies presented in these section took into
keeping the undrained cohesion constant at mean value. The account the effect of a single variable at one time. A three-
required Np, Lp, and Rr values to fulfill the design requirements dimensional (3D) graph can be developed to visualize the effect of
for five wind speeds and corresponding total cost of the founda- two variables. However, in reality, more than two variables may
tion are presented in Figs. 4, 5, and 6, respectively. The results affect the system simultaneously. In such a situation, a reliability-
show that Np, Lp, and Rr increased with increasing wind speed as based robust optimization procedure can be used to produce an
did the total cost. The radius of the raft at the lower three wind easy-to-use graph for selecting a suitable design.
speeds was the same, as shown in Fig. 6, because it was the mini-
mum radius requirement according to the bottom diameter of the
tower. Likewise, the total cost also showed a similar trend at the Design Optimization
lower wind speeds. An investigation of the total cost of the founda-
Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 01/29/18. Copyright ASCE. For personal use only; all rights reserved.

tion for different numbers and lengths of piles revealed that for the Although the parametric study results presented in the preceding
higher wind speeds, the design requirement is met more economi- section shows the effect of variations in soil properties and wind
cally by increasing either Np or Lp rather than by increasing Rr. speed on the design outcomes, they do not consider the change in
more than one variable at the same time. Moreover, they do not give
Effect of Undrained Cohesion on the Design Variables a clear indication on how to select the most cost-efficient and robust
design for a given performance criterion. In such a situation, a
Considering the medium site variability, a coefficient of variation reliability-based robust design optimization can be used to develop
(COV) of 25% was assumed to determine the variation in undrained
a criterion to select the most suitable design for the given perform-
cohesion, cu, for the second and third layers of soil. For the second
ance and cost criteria. One such framework is presented in this
layer, the standard deviation of the undrained cohesion (cu), s cu,
paper.
was calculated as 24.70 kPa using 25% COV and the mean of the
In reliability methods, risk or reliability is calculated for a given
cu, m cu, was 98.81 kPa. Similarly, the standard deviation of the third
performance criterion or a performance function. The computational
layer was calculated as 26.66 kPa using the same COV as for the
second layer and a mean cu of 106.66 kPa. The parametric study
was conducted by varying cu by 62s cu above and below the mean
value for both layers. The variation of cu used in this parametric
study is also shown in Fig. 7.
As shown in Fig. 7, for each case in the parametric study, the cu
values of the second and third layer were changed simultaneously
while the wind speed was kept constant. Although only the varia-
tions up to 25 m in depth are shown in Fig. 7, the cu for a depth
greater than that was assumed as the same for the parametric study.
The Np, Lp, and Rr required to meet all the design requirements at
five levels undrained cohesion and the corresponding total costs of
the foundation are presented in Figs. 8, 9, and 10, respectively. The
results indicate that Np, Lp, and Rr decreased with an increase in cu
along with the total cost of the foundation. For the lowest cu, in
Fig. 8(a), piles were arranged in three circumferences to meet all the
design requirements without facing a group effect. In Fig. 10(a), it
can be seen that Rr remains the same even with an increase in cu
because it is the minimum radius requirement according to the bot-
tom diameter of the tower. Similar to the results for the variations in
wind speed, adjusting Np or Lp, but not Rr, is the most economical Fig. 7. Soil profile showing variations in undrained cohesion
method for meeting all of the design requirements.

Fig. 6. Effect of variation in wind speed on (a) radius of raft and (b) total cost of pile raft

© ASCE 04017143-7 Int. J. Geomech.

Int. J. Geomech., 2018, 18(2): 04017143


Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 01/29/18. Copyright ASCE. For personal use only; all rights reserved.

Fig. 8. Effect of variation in undrained cohesion on (a) number of piles and (b) total cost of pile raft

Fig. 9. Effect of variation in undrained cohesion on (a) length of pile and (b) total cost of pile raft

Fig. 10. Effect of variation in undrained cohesion on (a) radius of raft and (b) total cost of pile raft

Table 3. Consolidation Parameters and Consolidation Settlement

Middle of layer Thickness (m) s 0c s 0o s 0f Cc Cs eo Sc (mm)


2 7.93 424.05 50.53 229.92 0.0907 0.0113 0.57 37.62
3 15.85 504.89 154.92 216.97 0.0907 0.0113 0.57 16.72

© ASCE 04017143-8 Int. J. Geomech.

Int. J. Geomech., 2018, 18(2): 04017143


approaches available for conducting a reliability analysis of geotech- is performed. Various methods are described in the literature for per-
nical or structural engineering systems with implicit performance forming reliability-based robust design optimization. Some of the
functions can be grouped into three types: (1) the Monte Carlo simu- latest procedures used for optimizing geotechnical systems along
lation, (2) response-surface approach, and (3) sensitivity-based anal- with their robustness measure, safety constraint, random variables,
ysis. The Monte Carlo and response-surface approaches are widely and mathematical models are presented in Table 3.
used in geotechnical engineering. The Monte Carlo approach is Although the advantages and disadvantages of these methods
mostly used when a closed-form solution is achievable with reasona- vary with the problem, the reliability-based robust geotechnical
ble computational effort because the function is developed on the ba- design (RBRGD), with the nondominated sorting genetic algo-
sis of thousands of simulation results. This method may not be effec- rithm, Version II (NSGA II), developed by Deb et al. (2002), was
tive if obtaining a deterministic solution is time consuming, as in the used in this study. This method considers the soil properties and
wind speed as random variables, the standard deviation of the dif-
Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 01/29/18. Copyright ASCE. For personal use only; all rights reserved.

case of the FEM for complex problems. In such a situation, the


response-surface method is used; the function is developed to ap- ferential settlement and the total cost of the foundation as robustness
proximate the performance through a few selected simulations. The measures, and an allowable differential settlement and the reliabil-
inputs for these simulations are selected in the neighborhood of the ity index as safety constraints. A framework of the design optimiza-
most likely failure point. Then, a regression analysis of these results tion procedure for the pile-raft foundation is presented in Fig. 11,
and the procedure discussed in the following section.

START Design Variables, Random Variables, and


Objective Functions
Select design and random variables The design optimization of the pile-raft foundation was performed
considering Lp, Np, and Rr as design variables and V, cu(2), and cu(3)
Set range for random variables as random variables. The ranges of design variables used in optimi-
zation that would satisfy design requirements for all the possible
design sets, considering variations in random variables, were Lp
Set range for design variables between 28.4 and 31.4 m, Np between 44 and 52, and Rr between
7.5 and 10.5 m. The maximum and minimum values and the stand-
Identify design cases and calculate response (Sdiff) ard deviation of random variables used in the optimization are the
same as those used in the parametric study. Because the excessive
differential settlement, Sdiff, can cause a collapse of the entire tur-
No
Sdiff ≤ Sdiff-all bine, it is considered the response of concern in this study. A bi-
objective optimization was performed using NSGA-II to reduce the
Yes effect of uncertainties on the response and to capture a set of designs
Develop a response function and calculate its in terms of cost efficiency and insensitivity to uncertain parameters.
standard deviation using Monte Carlo simulation To achieve this, two objective functions of the total cost of the pile-
raft foundation and the standard deviation of the predicted differen-
tial settlements were computed and minimized through optimiza-
Calculate total cost of the foundation
tion. The total cost of the foundation was considered as one of the
objectives because in the wind-turbine tower construction, the foun-
Perform bi-objective optimization considering cost dation is the only component that is dependent on the site subsur-
and standard deviation of response as objectives face condition. All the other components of the wind turbine are
prefabricated and assembled at the site. This situation implies that
Develop Pareto front the cost of the wind-turbine tower is mostly controlled by the cost
of the foundation. Hence, using the total cost as an objective helps
the user compare costs of different foundations within a range of the
SELECT DESIGN performance requirement (variation in differential settlement).
Thus, the result from the design optimization can be used by clients
Fig. 11. Design optimization procedure to select the site-specific optimal design within their allocated
budgets.

480,000 480,000
Total cost ($)

Total cost ($)

440,000 440,000

400,000 400,000

360,000 360,000

320,000 320,000
3 4 5 6 7 8 3 4 5 6 7 8
Std. dev. of differential settlement (mm) Std. dev. of differential settlement (mm)
(a) (b)

Fig. 12. Pareto front optimized to both total cost and standard deviation: (a) 1,000 simulations; (b) 10,000 simulations

© ASCE 04017143-9 Int. J. Geomech.

Int. J. Geomech., 2018, 18(2): 04017143


480,000 Development of Response Function
A A = Robust but the most expensive design
B = Cheapest but the least robust design To develop a response function for optimization, the responses
440,000 C = Optimum design (knee point) (Sdiff in this study) of all the possible design sets, considering varia-
Total cost ($)

tions in design variables and random variables, were first calculated.


C
400,000 A regression analysis was then performed using the results of the
differential settlement analysis to establish a response function in
360,000 terms of the design variables (Lp, Np, and Rr) and random variables
B
[V, cu(2), and cu(3)]. The resulting response function is shown in Eq.
320,000
(12) as
3.5 4.5 5.5 6.5 7.5
Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 01/29/18. Copyright ASCE. For personal use only; all rights reserved.

Std. dev. of differential settlement (mm) Sdiff ¼ exp f11:53 þ 3:55lnðV Þ  0:13ln½cuð2Þ   1:05ln½cuð3Þ 
*Pareto front used is for 10,000 simulations.
 2:07lnðLp Þ  3:02lnðNp Þ  1:10lnðRr Þg (12)
Fig. 13. Application of Pareto front for design selection

Pareto Front and Design Selection


Pareto Front
The bi-objective optimization considering the total cost and stand-
ard deviation of the response as objectives for the preferred number
of design sets, coupled with Monte Carlo method, was used to de-
velop Pareto front. The first objective (i.e., the total cost of the foun-
dation) was calculated on the basis of the unit cost of the raft and
prestressed concrete pile, as detailed in the parametric study section
of this paper. The second objective (i.e., the standard deviation of
response for each design set) was calculated on the basis of the
response function developed as shown in Eq. (12) for the desired
number of Monte Carlo simulations. In this case, the number of sim-
ulations is the total number of a set of random variables, selected
randomly within the provided limits. In this study, 1,000 simula-
tions and 10,000 simulations were used to compute the standard
deviation of the response for each design set. The preferred designs
resulting from the optimization procedure are presented graphically
Fig. 14. NBI approach to determine knee point using a Pareto front in Figs. 12(a and b) for the 1,000 and 10,000
simulations. All the designs in the Pareto front are considered as

Table 4. Applications of Reliability-Based Robust Design Optimization to Various Geotechnical Systems

System Random variables Robustness measure Safety constraint Optimization method Reference
Drilled shafts Soil properties, construction variations, Weighted sensitivity Target reliability Simplified RBRGD in Khoshnevisan
in clay loading index of response index spreadsheet et al. (2016)
Monopile Undrained shear strength, friction angle, lat- Total cost Failure probability RBDO by coupling SS Overgård et al.
foundation eral load method with SA stochastic (2016)
optimization algorithm.
Drilled shaft Friction angle, coefficient of lateral earth Variation in failure Failure probability RGD with NSGA-II Juang et al.
pressure probability, feasibility (2013)
robustness
Shallow Undrained shear strength, loads (moment, Volume of concrete Target probability d-RBD with MCS Ben-Hassine
foundation vertical, horizontal) of failure and Griffiths
(2012)
Shallow Undrained shear strength, effective friction Variation in failure Failure probability RGD with NSGA-II Juang and Wang
foundation angle, COV compressibility, vertical central probability, feasibility (2013)
load robustness
Shallow Geotechnical parameters (unit weight, fric- SI based on gradient Safety margin (dif- RGD with NSGA-II Gong et al.
foundation tion angle, Young’s modulus, Poisson’s ra- of system response to ference between (2014)
tio), loading parameters (dead load, live noise factors resistance and
load), construction tolerance (width, length, load)
depth), model error (ULS and SLS solution)
Spread Unit weight, effective friction angle, opera- Construction cost ULS and SLS RBD Wang (2009)
foundation tive horizontal stress coefficient requirement
Note: d-RBD = direct reliability-based design; MCS = Monte Carlo simulation; RBDO = reliability-based design optimization; RGD = robust geotechnical
design; SA = simulated annealing; SLS = serviceability limit state; SS = subset simulation; ULS = ultimate limit state.

© ASCE 04017143-10 Int. J. Geomech.

Int. J. Geomech., 2018, 18(2): 04017143


equally optimum. A clear trade-off relationship between the total selecting the preferred design for the given condition using the knee
cost of the foundation and the standard deviation of the response point concept.
can be inferred from the resulting Pareto front shown in Fig. 12. In
other words, decreasing the cost of the foundation may result in
designs with a higher vulnerability and response variability against References
uncertainties. From the 1,000 simulation results, it can be seen that
ASCE. (2010). “Minimum design loads for buildings and other structures.”
a decrease in the standard deviation of differential settlement, from Standard ASCE/SEI 7-10, ASCE, Reston, VA.
7.6 to 3.5 mm, increased the total foundation cost from $340,000 to Ben-Hassine, J., and Griffiths, D. V. (2012). “Reliability based design of
$460,000. Similarly, the 10,000 simulation results showed that a foundations subjected to combined loading with applications to wind
decrease in the standard deviation of the differential settlement, turbine foundations.” Proc., 11th Int. Congress on Numerical Methods
Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 01/29/18. Copyright ASCE. For personal use only; all rights reserved.

from 7.0 to 3.7 mm, increased the total foundation cost from in Engineering and Scientific Applications, E. Dávila, G. Uzcátegui,
$360,000 to $460,000, which is a $10,000 difference between the and M. Cerrolaza, eds., Venezuelan Society of Numerical Methods in
two amounts. Engineering, Caracas, Venezuela, 17–23.
Coduto, D. P. (2001). Foundation design: Principles and practices, 2nd
Design Selection Ed., Prentice Hall, Upper Saddle River, NJ.
Das, B. M. (2011). Principles of foundation engineering, 7th Ed., Cengage
The resulting Pareto front can be judged by designers, and the final Learning, Stamford, CT.
design can be selected on the basis of the performance requirements Deb, K., Pratap, A., Agarwal, S., and Meyarivan, T. (2002). “A fast and elit-
and available funds for the construction of the foundation. The dif- ist multi objective genetic algorithm: NSGA-II.” IEEE Trans. Evol.
ferent possible optimum design outcomes that can be extracted Comput., 6(2), 182–197.
from the Pareto front are presented in Fig. 13. In this section, the Gong, W., Khoshnevisan, S., and Juang, C. H. (2014). “Gradient-based
Pareto front for 10,000 simulations is used for demonstration. design robustness measure for robust geotechnical design.” Can.
Generally, to compare the concept of a Pareto front with a conven- Geotech. J., 51(11), 1331–1342.
Grünberg, J., and Göhlmann, J. (2013). Concrete structures for wind tur-
tional design, it should be noted that the least costly design, which
bines, K. Bergmeister, F. Fingerloos, and J.-D. Wörner, eds., Ernst &
is the most sensitive design on a Pareto front (marked as B in Sohn, Berlin.
Fig. 13), is usually considered as the final design in conventional Gudmundsdottir, B. (1981). “Laterally loaded piles.” M.S. thesis, Univ. of
practices where uncertainties are not involved. Similarly, the design Alberta, Edmonton, AB, Canada.
that is the least sensitive but the most expensive of all on the Pareto GWEC (Global Wind Energy Council) (2015). “Global wind report:
front (marked as A in Fig. 13) can also be obtained if the client Annual market update 2015.” Brussels.
desires to have the most robust design. Nevertheless, the most opti- Hemsley, J. A. (2000). Design applications of raft foundations, Thomas
mal design that meets the given performance and cost requirements Telford Ltd., London,.
Hough, B. K. (1957). Basic soils engineering, Ronald Press, New York.
can be obtained from the Pareto front using the knee point concept.
Juang, C. H., and Wang, L. (2013). “Reliability-based robust geotechnical
In this study, the normal boundary intersection (NBI) approach, design of spread foundations using multi-objective genetic algorithm.”
shown in Fig. 14, was used to determine the knee point in the result- Comput. Geotech., 48(Mar), 96–106.
ing Pareto front. In this method, for each point on the Pareto front, Juang, C. H., Wang, L., Liu, Z., Ravichandran, N., Huang, H., and Zhang, J.
the distance from the boundary line, which connects two extreme (2013). “Robust geotechnical design of drilled shafts in sand: New
upper and lower points of the Pareto front, was computed in the nor- design perspective.” J. Geotech. Geoenviron. Eng., 10.1061/(ASCE)GT
malized space of the Pareto front. The knee point on the Pareto .1943-5606.0000956, 2007–2019.
front, which has the greatest distance from the boundary line, was Khoshnevisan, S., Wang, L., and Juang, C. H. (2016). “Simplified proce-
dure for reliability-based robust geotechnical design of drilled shafts in
then determined. The knee point determined using the NBI
clay using spreadsheet.” Georisk Assess. Manage. Risk Eng. Syst.
approach for the Pareto front with 10,000 simulations is also shown Geohazards, 10(2), 121–134.
in Fig. 13 as point C. Kulhawy, F. H., and Mayne, P. W. (1990). “Manual on estimating soil prop-
The optimum designs for 1,000 and 10,000 simulations obtained erties for foundation design.” Rep. EL-6800, Project 1493-6, Cornell
via the NBI approach are given in Table 4. It is observed that the op- Univ., Ithaca, NY.
timum design for both numbers of simulations has a similar value Lewin, T. J. (2010). “An investigation of design alternatives for 328-ft
for both the cost and the standard deviation of the response. (100-m) tall wind turbine towers.” M.S. thesis, Iowa State Univ.,
Ames, IA.
Malhotra, S. (2011). “Chapter 10: Selection, design and construction of off-
Conclusion shore wind turbine foundations.” Wind turbines, I. Al-Bahadly, ed.,
InTech, London.
A reliability-based robust design optimization of the pile-raft foun- Mayne, P. W., and Kemper, J. B. (1988). “Profiling OCR in stiff clays by
CPT and SPT.” Geotech. Test. J., 11(2), 139–147.
dation for a tall wind turbine on clayey soil was presented in this pa-
Overgård, I. E. V., Depina, I., and Eiksund, G. (2016). “Reliability-
per. According to the deterministic geotechnical design outcomes, based design of a monopile foundation for offshore wind turbines
it was found that the final design was controlled by a differential set- based on CPT data.” Proc., 17th Nordic Geotechnical Meeting:
tlement and rotation. The use of the pile raft, which takes advantage Challenges in Nordic Geotechnics, Icelandic Geotechnical Society,
of both the raft and piles to control the bearing capacity and settle- Reykjavik, Iceland, 495–502.
ment, respectively, was found to be the best option for meeting the Phillip, W., and Adrian, C. (2013). RSMeans building construction cost
design requirements. The results of the parametric study showed data, 71 Ed., The Gordian Group, Rockland, MA.
that the design requirements can be met by increasing either the Poulos, H. G., and Davis, E. H. (1980). Pile foundation analysis and design,
T. W. Lambe and R. V. Whitman, eds., Wiley, New York.
number of piles, length of piles, or radius of the raft at high wind
Poulos, H. G. (2001a). “Methods of analysis of piled raft foundations.”
speeds. The results of the Pareto front created from the design opti- Technical Committee TC-18 on Piled Foundations, International
mization results showed a clear trade-off relationship between the Society of Soil Mechanics and Geotechnical Engineering.
total cost of the foundation and the standard deviation of the Poulos, H. G. (2001b). “Piled raft foundation: Design and applications.”
response (differential settlement). Such a relationship is useful for Geotechnique, 51(2), 95–113.

© ASCE 04017143-11 Int. J. Geomech.

Int. J. Geomech., 2018, 18(2): 04017143


Raju, V. S. (2015). “Piled raft.” V. S. Raju Consultants, Geotechnical and Vesic, A. S. (1973). “Analysis of ultimate loads of shallow foundations.” J.
Structural Engineers, Chennai, India. Soil Mech. Found. Div., 99(1), 45–73.
Randolph, M. F. (1994). “Design methods for pile groups and piled rafts.” Vesic, A. S. (1975). “Bearing capacity of shallow foundations.” Foundation
Proc., 13th Int. Conf., Soil Mechanics and Foundation Engineering, engineering handbook, 1st Ed., H. F. Winterkorn and H.-Y. Fang, eds.,
CRC Press, Boca Raton, FL, 61–82. Van Nostrand Reinhold Company, New York.
Reul, O. (2004). “Numerical study of the bearing behavior of piled Wagner, H.-J., and Mathur, J. (2013). Introduction to wind energy systems:
rafts.” Int. J. Geomech., 10.1061/(ASCE)1532-3641(2004)4:2(59), Basics, technology and operation, 2nd Ed., Springer, Berlin.
59–68. Wang, Y. (2009). “Reliability-based economic design optimization of
Sinha, A., and Hanna, A. M. (2017). “3D numerical model for piled raft spread foundations.” J. Geotech. Geoenviron. Eng., 10.1061/(ASCE)GT
foundation.” Int. J.Geomech., 10.1061/(ASCE)GM.1943-5622.0000674, .1943-5606.0000013, 954–959.
04016055. WPC. (2010). “Clemson Wind Turbine Testing Facility North Charleston,
USACE. (1990). “Engineering and design—Settlement analysis.” Engineer South Carolina.” Geotechnical Engineering Rep., WPC Project No.
Downloaded from ascelibrary.org by National Taiwan University of Sci and Tech on 01/29/18. Copyright ASCE. For personal use only; all rights reserved.

Manual 1100-1-1904. EN105060, Charleston, SC.

© ASCE 04017143-12 Int. J. Geomech.

Int. J. Geomech., 2018, 18(2): 04017143

S-ar putea să vă placă și