simplified robust geotechnical design of soldier pile ...hsein/wp-content/uploads/2017/01/... · 1...

51
1 Simplified Robust Geotechnical Design 1 of Soldier Pile-Anchor Tieback Shoring System for Deep Excavation 2 3 Wenping Gong a,b* , Hongwei Huang a , C. Hsein Juang b,c , and Lei Wang d 4 5 Abstract 6 The shoring system that consists of soldier piles and anchor tiebacks is often used in 7 deep excavations in sandy deposits. However, uncertainties often exist in the design of such 8 shoring systems. Here, a simplified robust geotechnical design (RGD) method is proposed to 9 account for these uncertainties in the shoring system design. Specifically, for a given deep 10 excavation, uncertain soil parameters and surcharges are treated as noise factors, and the 11 parameters of soldier piles and tieback anchors are treated as design parameters. Robust 12 design is then implemented as a multi-objective optimization problem, in which the design 13 robustness is sought along with cost efficiency and safety requirements. A tradeoff between 14 design robustness and cost efficiency exists and the optimization usually leads to a Pareto 15 front. By applying the knee point concept, the most preferred design that meets the safety 16 requirements and yields the best compromise between design robustness and cost efficiency 17 can be identified on the Pareto front. Improvements made to the existing RGD method include 18 an efficient formulation of the design robustness and a new procedure for finding the most 19 preferred design in the design pool. The new simplified RGD method is illustrated with a 20 real-world excavation case study. 21 22 Keywords: Deep Excavation; Design Robustness; Knee Point; Optimization; Pareto Front; 23 Shoring System; Uncertainty. 24

Upload: dinhthien

Post on 07-May-2018

219 views

Category:

Documents


2 download

TRANSCRIPT

1

Simplified Robust Geotechnical Design 1

of Soldier Pile-Anchor Tieback Shoring System for Deep Excavation 2

3

Wenping Gong a,b*

, Hongwei Huang a, C. Hsein Juang

b,c, and Lei Wang

d 4

5

Abstract 6

The shoring system that consists of soldier piles and anchor tiebacks is often used in 7

deep excavations in sandy deposits. However, uncertainties often exist in the design of such 8

shoring systems. Here, a simplified robust geotechnical design (RGD) method is proposed to 9

account for these uncertainties in the shoring system design. Specifically, for a given deep 10

excavation, uncertain soil parameters and surcharges are treated as noise factors, and the 11

parameters of soldier piles and tieback anchors are treated as design parameters. Robust 12

design is then implemented as a multi-objective optimization problem, in which the design 13

robustness is sought along with cost efficiency and safety requirements. A tradeoff between 14

design robustness and cost efficiency exists and the optimization usually leads to a Pareto 15

front. By applying the knee point concept, the most preferred design that meets the safety 16

requirements and yields the best compromise between design robustness and cost efficiency 17

can be identified on the Pareto front. Improvements made to the existing RGD method include 18

an efficient formulation of the design robustness and a new procedure for finding the most 19

preferred design in the design pool. The new simplified RGD method is illustrated with a 20

real-world excavation case study. 21

22

Keywords: Deep Excavation; Design Robustness; Knee Point; Optimization; Pareto Front; 23

Shoring System; Uncertainty. 24

2

25

_______________________ 26 27 a Department of Geotechnical Engineering, Tongji University, Shanghai 200092, China. 28

29 b Glenn Department of Civil Engineering, Clemson University, Clemson, SC 29634, USA. 30

31 c Department of Civil Engineering, National Central University, Taoyuan 320, Taiwan. 32

33

d Department of General Engineering, Montana Tech of the University of Montana, Butte, MT 59701, 34

USA. 35

36 * Corresponding author (Email: [email protected]) 37

38

3

1. Introduction 39

In the traditional design of shoring systems using the deterministic methods, the 40

design must meet the safety or stability requirement, in terms of limiting factors of safety 41

against failures of the ground and shoring system, and the serviceability requirement, in terms 42

of limiting maximum wall and/or ground deformation for preventing damages to adjacent 43

structures (JSA 1988; PSCG 2000; TGS 2001; Kim and Lee 2005; Ou 2006; Kim et al. 2012; 44

Goh and Mair 2014). However, the deterministic methods do not allow for an explicit 45

consideration of uncertainties in the input parameters. Because of these uncertainties, however, 46

the computed system response (i.e., the response of a designed system) is subject to variation. 47

Thus, the design decision has to compensate for variation or uncertainty in the computed 48

system response, which often leads to a cost-inefficient design. 49

To explicitly account for the uncertainties in the input parameters, the probabilistic 50

design approach may be adopted (Harr 1987; Baecher and Christian 2003; Zhang et al. 2005; 51

Ang and Tang 2007; Fenton and Griffiths 2008; Juang et al. 2013b; Gong et al. 2014c). 52

However, a meaningful probabilistic design requires an accurate statistical characterization of 53

input parameters such as soil parameters, which is often difficult to attain in a typical 54

geotechnical project due to budget constraint. Furthermore, use of the probabilistic approach 55

to deal with uncertain soil parameters in the design of a complex shoring system based on 56

finite element method (FEM) is often computationally prohibitive. 57

In this paper, the robust geotechnical design (RGD) method is adopted to deal with the 58

uncertainties in the design of shoring systems for deep excavations. The robust design concept 59

is originated in the field of industry engineering for product optimization (Taguchi 1986), and 60

has attracted increasing interests from many other engineering disciplines (Phadke 1989; 61

4

Doltsinis et al. 2005; Beyer and Sendhoff 2007). The robust design concept has also been 62

applied in various geotechnical problems (Juang et al. 2012a; Juang et al. 2013a; Wang et al. 63

2013; Juang et al. 2014; Gong et al. 2014a). In the context of robust design, a design is 64

deemed robust if the system response of concern, such as the maximum wall deflection in the 65

case of deep excavations, is insensitive to the variation of uncertain input parameters (termed 66

noise factors herein). Thus, the goal of robust design is to seek a safe and cost-efficient 67

optimal design, the system response of which is robust against, or insensitive to, the 68

unforeseen variation of input parameters; such an optimal design may be achieved through a 69

proper selection of design parameters (i.e., those that can be specified by the designer). 70

This paper presents an application of RGD method to the design of a shoring system 71

for an excavation. Note that while the RGD method has been reported in various geotechnical 72

applications (Juang et al. 2012a; Juang et al. 2013a; Wang et al. 2013; Khoshnevisan et al. 73

2014; Juang et al. 2014; Gong et al. 2014a; Gong et al. 2014b), this paper has three new 74

features. First, a new gradient-based robustness measure is adopted, which, as demonstrated 75

later, eliminates the need of computing design robustness based on probabilistic analyses of 76

the system response, a step that becomes computationally prohibitive when the system 77

response can only be analyzed using numerical methods such as FEM. Second, the RGD 78

method is usually implemented as a multi-objective optimization problem (Juang et al. 2013a; 79

Juang et al. 2014), and the results are usually presented as a Pareto front (Deb 2001). When 80

the results are presented as a Pareto front, the knee point can usually be identified on this front, 81

which offers the best compromise solution or the most preferred design in the design pool. A 82

new procedure is developed in this study to find the most preferred design in the design pool; 83

this procedure involves a series of single-objective optimizations, rather than a 84

5

multi-objective optimization, and thus greatly reduces the computation efforts. In this paper, 85

the RGD implemented with these new procedures is termed the simplified RGD method. 86

Third, the simplified RGD method is applied to the design of a relatively complex shoring 87

system for a deep excavation project, in which the system response of concern is analyzed 88

with a commercially available computer program TORSA (Sino-Geotechnics 2010), and a 89

comparison with the original design that was selected by an engineering firm is made. 90

Through this real-world application, the advantages of the new simplified RGD method are 91

demonstrated. 92

This paper is organized as follows. First, the formulation of a new robustness measure 93

based on the gradient of the system response is presented. Second, the framework of the 94

simplified RGD method is presented, in which a new procedure to find the most preferred 95

design in the design pool is developed. Third, the results of a case study of a real-world deep 96

excavation project using the simplified RGD method are presented. Finally, results of a series 97

of parametric analyses to demonstrate the advantages of the simplified RGD method are 98

presented, which are followed by the concluding remarks. 99

100

2. Gradient-Based Design Robustness Measure: Sensitivity Index (SI) 101

In the construction of a deep excavation project in an urban area, the maximum wall 102

deflection is often monitored and used to assess the performance of the ground-wall-shoring 103

system. In general, an excessive wall deflection would signal poor stability and serviceability 104

performance of the shoring system, which poses a risk to the excavation project and adjacent 105

utilities (Juang et al. 2012b). Although the wall deflection is correlated well with the ground 106

settlement (Clough O’Rourke 1990; Kung et al. 2007), the former is much easier to monitor in 107

6

the field, and is easier to analyze with better accuracy. Thus, the performance of the shoring 108

system can be monitored, and remedial actions can be taken as needed, through the wall 109

deflection monitoring during the construction phase. In this paper, the maximum wall 110

deflection is adopted as the system response of concern in the context of robust design; in fact, 111

use of the maximum wall deflection as the system response in the robust design has been 112

reported (Juang et al. 2014; Wang et al. 2014). 113

2.1. Gradient as a robustness measure 114

For a design of a shoring system for a deep excavation project with design parameters 115

of d and noise factors of as inputs (note: both are vectors), the system response of concern, 116

in terms of the maximum wall deflection (y), can be represented as a function of d and , 117

denoted as y = f(d, ). Here, the noise factors are referred to uncertain soil parameters and 118

surcharge on the ground surface behind the wall, and the design parameters are referred to the 119

dimensions, sizes, and layouts of the soldier piles and tieback anchors. Here, the system 120

response function may be constructed using an analytical method, empirical model or finite 121

element method (FEM). In this paper, a special-purpose FEM code TORSA 122

(Sino-Geotechnics 2010) is adopted. TORSA is an FEM code based on the 123

beam-on-elastic-foundation theory and has been used in hundreds of real-world excavation 124

projects. 125

Mathematically and intuitively, the sensitivity of the system response, f(d, θ), to the 126

noise factors can be captured by its gradient f, which is defined as the variation of the 127

system response caused by one unit change in the noise factors. In reference to Figure 1, two 128

different designs (referred to herein as d1 and d2) are examined to illustrate the gradient-based 129

robustness concept. Here, d2 is seen more robust than d1 against the variation of noise factors, 130

7

as the variation of the system response of d2 is much smaller than that of d1. Furthermore, the 131

gradient of the system response to the noise factors is lower in d2 than d1. That is to say, the 132

variation of the system response (due to the perceived but uncharacterized uncertainties of 133

noise factors) is proportional to the gradient of the system response to the noise factors. Thus, 134

the design robustness can be effectively represented with the gradient of the system response 135

to the noise factors. Symbolically, the gradient of the system response to the noise factors, f, 136

at a checkpoint of noise factors, denoted as ' , can be computed as follows (Gong et al. 137

2014b): 138

1 2

( , ) ( , ) ( , ), , ,

n

f f ff

'

' ' '

d d d (1) 139

where n is the number of noise factors. In a deterministic design of the shoring system for 140

deep excavation, only the nominal values of noise factors, denoted as θn, are characterized 141

and could be available to the engineer. Thus, the characterized nominal values of noise factors 142

can reasonably be assigned as the checkpoint in Eq. (1): ' = θn. 143

Since no explicit form of the system response function could be obtained in FEM 144

solutions, the closed-form solution of the partial derivatives in Eq. (1) cannot be obtained. In 145

this study, the numerical approach is adopted and implemented, within which, the partial 146

derivative of the system response to a noise factor (θi) can be approximated as: 147

( , ) ( , ) ( , )

2

i i i i

i i

f f ' d f ' d

d

' '

d d d (2) 148

where id represents a step increment of the ith noise factor, which is taken as 0.1 i' in this 149

study. It is noted that in the computation of ( , )i if ' d d and ( , )i if ' d d , the values of 150

8

all other noise factors are kept at their checkpoints ( ' ). The step increment of 0.1 i' is 151

obtained through a trial and error analysis, which yields converged results. This parameter 152

might be problem-specific and should be verified for a future problem. 153

Though formulated in a deterministic manner aiming at improving the computation 154

efficiency, a design (or system) with a lower gradient generally exhibits a lower variation in 155

the computed system response (regardless of the levels of relative variation among the noise 156

factors) and thus it yields a more robust design, and vice versa, as illustrated in Figure 1. 157

Therefore, the gradient of the system response to the noise factors is an effective measure of 158

the design robustness, and is suitable for use in a robust design optimization. 159

2.2. Formulation of sensitivity index (SI) 160

While the gradient f, defined in Eq. (1), is shown as an effective indicator of the 161

design robustness, two problems need to be resolved before the robust design optimization 162

can be implemented. First, the gradient is an n-dimensional vector; as the units of noise 163

factors are different, the mathematical operation of this vector could be a problem. Second, 164

the gradient is a vector rather than a scalar; it is not as convenient and effective as a scalar to 165

use for screening candidate designs in the design pool. 166

To solve the first problem, each partial derivative in the gradient vector,( , )

i

f

'

d, 167

is multiplied by a scaling factor of i' so that the effect of the units of noise factors on the 168

formulation of design robustness can be eliminated. Indeed, scaling the gradient term to avoid 169

the problem caused by inconsistent units is not uncommon (Gonsamo 2011). As such, the 170

gradient vector in Eq. (1) is re-written as follows, which is defined herein as the normalized 171

gradient vector (J): 172

9

1 2

1 2

( , ) ( , ) ( , ), , ,

n

n

' f ' f ' f

' ' '

d d dJ

(3) 173

To solve the second problem, the Euclidean norm of the normalized gradient vector, 174

which indicates the length of the normalized gradient vector, is adopted and defined herein as 175

the sensitivity index (SI): 176

TSI J J (4) 177

Eq. (4) yields a single value representation of the normalized gradient vector, in terms of the 178

sensitivity index (SI). As can be seen, a higher SI value signals lower design robustness, as it 179

indicates a greater variation of the system response in the face of the uncertainty in noise 180

factors. It is found that although gradient concept might be viewed as “local”, the sensitivity 181

index defined in this paper is sufficient for the proposed robust design optimization. For more 182

complex geotechnical systems that exhibit significant nonlinear behaviors, use of advanced 183

sensitivity index such as Sobol index (Marrel et al. 2009) may be worth pursuing. 184

It is noted that while the gradient-based robustness measure was introduced in a 185

previous paper (Gong et al. 2014b), three new features are introduced in this study. First, the 186

evaluation of design robustness is coupled with the deterministic analysis of the system 187

response. Second, the nominal values of noise factors are assigned as the checkpoint for 188

computing the gradient of the system response to the noise factors. Third, a full statistical 189

characterization of noise factors is not required. Whereas, in the previous study (Gong et al. 190

2014b), the system response was analyzed using probabilistic methods; the most probable 191

point (MPP) of noise factors that yields the failure of the system was taken as the checkpoint 192

for computing the gradient of the system response to the noise factors; and, the full statistical 193

information of noise factors is a prerequisite for conducting the robust design. 194

10

In general, different noise factors may have different levels of variation; and a robust 195

design should be less sensitive (or more robust) to the noise factor that has higher variability. 196

However, in the deterministic analysis and design, the uncertainty or variation of noise factors 197

is recognized but uncharacterized. Indeed, if the statistics of noise factors are available, the 198

reliability-based robust geotechnical design method (Juang et al. 2013a; Wang et al. 2013) can 199

readily be used. While the reliability-based robust geotechnical design method is theoretical 200

sounder, it requires the knowledge of the statistics of noise factors and probabilistic analyses 201

of the system response. In this paper, although the variation levels of noise factors are not 202

included in the formulation of the normalized gradient vector and sensitivity index, the new 203

robustness measure presented herein offers a practical alternative to the reliability-based 204

robustness measures, as it does not require the knowledge of the variation levels of noise 205

factors and probabilistic analyses of the system response can be avoided. The results of the 206

case study and the sensitivity analysis presented later confirm the effectiveness of this new 207

gradient-based robustness measure. 208

209

3. Simplified Robust Geotechnical Design (RGD) 210

3.1. Multi-objective optimization setting for robust design optimization 211

One unique feature of the RGD method is that it treats the geotechnical design as a 212

multi-objective optimization problem that considers explicitly and simultaneously the safety 213

requirements, cost efficiency, and design robustness (Juang et al. 2013a). Following the RGD 214

framework developed by Juang et al. (Juang et al. 2013a), the safety requirements are 215

considered as compulsory design constraints that must be satisfied, while the design 216

robustness and cost efficiency are treated as design objectives to be optimized simultaneously. 217

11

The robust design optimization algorithm for the design of a shoring system for a given deep 218

excavation can be set up as follows (modified after Juang et al. 2014): 219

1

2

factor

Find:

of safety

(design p

against push-in

arameters)

Subject to: (des

failur

ign pool)

Fs 1. e

factor of safety against basal

5 ( )

Fs 1.5 (

>

>

d

d S

)

( , ) 0.7% ( )

Objectives: min (sensitivity of

heave failure

maximum wall deflection

ma wall deflection to noise factors )

min (cost)

ximum

f y = f < H

SI

C

d

(5) 220

where Hf is the final excavation depth. The design constraint of ( , ) 0.7% fy = f < Hd is a 221

limiting condition that prohibits the maximum wall deflection (y) from exceeding 0.7% fH , as 222

per PSCG (2000). Here, the robust geotechnical design is to seek an optimal design 223

(represented by design parameters d) in the design pool (S) such that the cost (C) is 224

minimized and the design robustness is maximized (which is achieved by minimizing the 225

sensitivity index SI), while the safety requirements with respect to both stability (in terms of 226

Fs1 and Fs2) and serviceability (in terms of maximum wall deflection y) are satisfied. It is 227

noted that while there are many other potential failure models in deep excavations, such as the 228

failure of the anchor systems and the retaining structure, they are not studied in this paper for 229

simplicity. 230

In reference to the optimization setting denoted as Eq. (5), the design robustness is 231

formulated in a deterministic manner and the safety requirements are evaluated with the 232

deterministic methods. In this formulation, there is no need to perform a detailed statistical 233

characterization of noise factors and a full probabilistic analysis of the system response. Note 234

that while the uncertainties of noise factors are not explicitly considered in this optimization 235

12

setting, the design robustness of the system response against the variation of noise factors is 236

optimized. Indeed, this optimization setting can raise the design robustness by minimizing the 237

sensitivity index and thus reduces the variation of the system response. This is the essence of 238

the proposed new robustness measure, referred to herein as the simplified RGD method. 239

3.2. Pareto front and knee point 240

In general, the desire to maximize the design robustness (or to minimize the sensitivity 241

index) and the desire to minimize the cost are two conflicting design objectives in the robust 242

design optimization. Thus, a utopia design that is optimal with respect to both design 243

objectives simultaneously is not attainable; rather, a set of non-dominated designs can be 244

identified in the design pool that are superior to all others in the design pool, but within which, 245

none of them is superior or inferior to others. These non-dominated solutions collectively 246

form a Pareto front (Deb 2001), which shows a tradeoff between design robustness and cost 247

efficiency. This Pareto front may be obtained using genetic algorithms such as 248

Non-dominated Sorting Genetic Algorithm version II, NSGA-II (Deb et al. 2002). 249

As a tradeoff between design robustness and cost efficiency, the Pareto front is an 250

effective design aid. For example, either the least cost design that is above a target robustness 251

level or the most robust design that falls within a target cost level can be selected as the most 252

preferred design in a given design pool. The determination of an appropriate target level of 253

design robustness or cost, however, is problem-specific. When no such a design preference is 254

specified by the owner or client, the knee point on the Pareto front, which yields the best 255

compromise between design robustness and cost efficiency, may be taken as the most 256

preferred design in the given design pool (Branke et al. 2004; Deb and Gupta 2011; Juang et 257

al. 2014; Gong et al. 2014b). Figure 2 shows a conceptual illustration of the utopia design, 258

13

Pareto front and knee point in a bi-objective optimization problem. Note that the feasible 259

domain is defined as a domain that consists of feasible designs that meet the safety 260

requirements. 261

Oftentimes, only a single knee point can be obtained from a multi-objective 262

optimization, as shown in Figure 2(a). However, multi-knee points may be possible for some 263

complex problems, as illustrated in Figure 2(b) (Branke et al. 2004; Deb et al. 2006; Bechikh 264

et al. 2011). In such a circumstance, the global knee point, rather than the local knee point, is 265

taken herein as the most preferred design in the design pool. In this paper, the term “knee 266

point” is referred to the global knee point if no specific interpretation is given. 267

3.3. Existing approaches to identify the knee point 268

Three existing approaches to identify the knee point on the Pareto front, the reflex 269

angle approach, normal boundary intersection approach, and marginal utility function 270

approach, are briefly summarized herein to set the stage for the proposed new procedure. In 271

these methods, the objective functions are often normalized into a value ranging from 0.0 to 272

1.0, denoted as n ( )f d , through the following transformation: 273

n min

max min

( ) ( )( )

( ) ( )

i i

i

i i

f ff

f f

d dd

d d (6) 274

where max

( )if d and min

( )if d are the maximum and minimum values, respectively, of the 275

ith objective function, ( )if d . 276

Reflex angle approach 277

The reflex angle at a non-dominated design on the Pareto front indicates the bend of 278

the Pareto front from its left to right and hence can be used as a measure of the gain-sacrifice 279

in the tradeoff relationship. The knee point is identified as the non-dominated design on the 280

14

Pareto front that yields the maximum reflex angle, as shown in Figure 3(a). It is noted that the 281

reflex angle is computed using only two neighboring points, thus it can correspond to a local 282

property and may not yield a global knee point in some cases (Branke et al. 2004; Deb and 283

Gupta 2011). 284

Normal boundary intersection approach 285

To alleviate the possible local knee point using the reflex angle approach, the normal 286

boundary intersection approach can be used. As shown in Figure 3(b), the two extreme 287

designs of a Pareto front, denoted as A and B, are used to construct a straight boundary line; 288

and the non-dominated design on the Pareto front that has the maximum distance from the 289

constructed boundary line is identified the knee point (Deb and Gupta 2011; Juang et al. 290

2014). 291

Marginal utility function approach 292

As can be seen, the reflex angle approach and the normal boundary intersection 293

approach are valid and efficient only for the bi-objective optimization problem. To deal with 294

multi-objective optimization that involves more than two design objectives, the marginal 295

utility function approach (Branke et al. 2004; Gong et al. 2014b) may be used. The marginal 296

utility function, denoted as ,U' d , is formulated as: 297

, min , , , ( )i j iU' U U i j d d d (7) 298

where ,U d is a linear utility function of the design objectives, defined as: 299

n, ( )i iU f d d (8) 300

where i is a weighting parameter with a value ranging from 0.0 to 1.0, and i = 1.0; and 301

n ( )if d is the ith normalized objective function to be minimized. By means of Monte Carlo 302

15

simulations (MCS), different random values of i can be generated with an assumption of 303

uniform distribution, and then the expected marginal utility function can be computed for each 304

non-dominated design on the Pareto front. The non-dominated design with the maximum 305

expected marginal utility is taken as the knee point on the Pareto front. 306

3.4. New procedure to identify the most preferred design 307

Oftentimes, the owner or client may be only interested in the most preferred design in 308

the design pool, and not the Pareto front per se. In fact, if the most preferred design in the 309

design pool can be identified directly, use of multi-objective optimization algorithms such as 310

NSGA-II would not be needed; and thus, the RGD can be more readily adaptable for general 311

applications. In this paper, a new procedure to identify the most preferred design in the design 312

pool is developed. The new procedure is summarized in the following steps: 313

Step 1: Conduct a single-objective optimization with respect to the ith design objective 314

of concern, ( )if d , which might be set up as follows: 315

Find: (design parameters)

Subject to: (design pool)

(design pool)

Safety requirements

Objective: ( ) min ( )

i

i

i i i

f f

d *

d * S

d S

d * d

(9) 316

where id * represent the optimal design based on the ith

design objective, which meets the 317

safety requirements and yields the minimum of the ith objective function, denoted as 318

min

( )if d = ( )i if d * = min ( )if d . 319

By repeating the single-objective optimization in Eq. (9) for each and every design 320

objective, a utopia point of { 1 min( )f d , 2 min

( )f d , …, min

( )mf d } can be constructed in the 321

16

design pool, where m represents the number of design objectives to be optimized. As implied 322

by its name, the utopia design is in reality not attainable if the design objectives are 323

conflicting with each other. Note that the utopia design defined this way may also violate the 324

safety requirements or even fall outside the design pool. 325

Step 2: Determine the maximum value of each objective function among the designs 326

of {1d * , 2d * , …,

md * }. Symbolically, the maximum value of the ith objective function is 327

obtained as follows: 328

max

( ) max ( ) =1, 2, , i i jf f j m d d * (10) 329

Step 3: Normalize all the objective functions into a value ranging from 0.0 to 1.0 330

using the transformation described in Eq. (6). As a result, the coordinates of the normalized 331

utopia point are all equal to 0.0. 332

Step 4: Compute the distance from the normalized utopia point to the normalized 333

objective functions for each candidate design in the design pool. The design that meets the 334

safety requirements and yields the minimum distance is regarded as the most preferred design 335

in the design pool, as shown in Figure 3(c). 336

Note that the computed distance from the utopia point to a candidate design can be 337

interpreted as an additional “price”, in terms of a combination of the design robustness and 338

cost efficiency, one has to pay to select this candidate design, in lieu of the utopia point 339

(design), as the final design. This idea is inspired by the marginal utility concept (Branke et al. 340

2004) and the compromise programming approach (Chen et al. 1999). In reality, the utopia 341

design is not attainable. Thus, the design that meets the safety requirements and yields the 342

minimum distance (thus incurring the least additional price) can be treated as the most 343

preferred design in the design pool. As will be shown later, the most preferred design 344

17

identified with this new procedure is virtually the same as the knee point on the Pareto front 345

obtained through a multi-objective optimization. 346

Note that with this new procedure, the multi-objective optimization is solved with a 347

series of single-objective optimizations. Here, the single-objective optimization can be easily 348

performed using Matlab function fmincon (MATLAB 2010) or Microsoft Excel Solver that 349

returns the minimum of a constrained nonlinear multivariable function. Thus, use of 350

multi-objective optimization algorithms such as NSGA-II is no longer required in the RGD 351

implemented with this new procedure. It should be noted that many outstanding studies on 352

knee points have been published (Deb et al. 2006; Bechikh et al. 2011). However, the 353

development of this new procedure described herein places a high premium on simplicity, 354

aiming to enable the simplified RGD method as a practical geotechnical design tool. Further, 355

it is noted that this new procedure is applicable to the multi-objective optimization with more 356

than two design objectives, although only a bi-objective optimization problem is illustrated in 357

Figure 3(c). For convenience of description, the RGD implemented with the new robustness 358

measure and the new knee-point seeking procedure is referred to hereinafter as the simplified 359

RGD method. 360

361

4. Robust Geotechnical Design of Shoring System for Deep Excavation Case Study 362

4.1. Brief summary of the deep excavation case studied 363

To illustrate the simplified RGD method, a real-world deep excavation project in 364

Taiwan is studied herein as a demonstrative example. The selected deep excavation case is 365

located in Taipei, Taiwan, which is designed for the student dormitory and underground 366

parking of Wesley Girls High School. The plan layout of the excavation site is approximately 367

18

a rectangular shape with a length of 140 m and a width of 45 m. The excavation depths range 368

from 6.9 m to 11.9 m, as shown in Figure 4. For illustration purposes, the robust design 369

optimization is only focused on section D, although other sections can be designed using the 370

same simplified RGD method. With respect to section D, the excavation is carried out in 371

layered soils with sands and rock. The soil profiles and soil properties for each soil layer are 372

shown in Figure 5 and listed in Table 1, respectively. The effect of groundwater is negligible 373

as the groundwater table is well below the excavation depth. Based on the local experience of 374

similar projects, a soldier pile wall that consists of reinforced concrete piles and timber 375

laggings is used as the retaining structure, while the 4-strand anchor (each with diameter of 376

12.7 mm) is used as the tieback of the retaining structure. 377

4.2. Deterministic model for assessing the performance of the shoring system 378

In this study, TORSA, a commercially available computer program developed by 379

Trinity Foundation Engineering Consultants (TFEC), is adopted as the deterministic model for 380

evaluating the performance or response of the shoring system for the deep excavation. Of 381

course, other special-purpose computer programs such as DeepXcav (Deep Excavation 2014) 382

or general FEM computer programs such as PLAXIS (Likitlersuang et al. 2013; Plaxis 2014) 383

can be used. TORSA has been validated as an effective design tool in analyzing the stability 384

(i.e., push-in and basal heave) and serviceability (i.e., wall deflection) of the shoring system 385

for deep excavation through numerous projects in Taiwan (Sino-Geotechnics 2010). 386

As noted previously, TORSA is an FEM code based on the beam-on-elastic-foundation 387

theory. The retaining structure, in terms of the soldier pile wall studied in this paper, is 388

modeled as an elastic beam. The pressure acting on the back of the soldier pile wall is 389

assumed to be the active earth pressure and the resistance of soil inside the excavation is 390

19

modeled as soil springs (Ou 2006; Sino-Geotechnics 2010). The water pressure, surcharge, 391

tieback anchor (or internal strut) stiffness and preload, and the passive earth pressure caused 392

by the plastic deformation of soil inside the excavation can be considered. A limit equilibrium 393

model that considers all the forces and soil springs applied on the soldier pile wall is 394

formulated. This equilibrium model is then solved via the finite element method (FEM) 395

implemented in TORSA. With TORSA, the factor of safety against push-in failure (Fs1), the 396

factor of safety against basal heave failure (Fs2), and the maximum wall deflection (y) can 397

readily be obtained. 398

4.3. Noise factors characterization for the robust design optimization 399

For the deep excavation project illustrated in Figure 5, the main soil parameters 400

affecting the stability and serviceability behavior of the shoring system are the cohesion (c), 401

friction angle (φ) and modulus of horizontal subgrade reaction (kh). These soil parameters, 402

along with the surcharge behind the wall (q0), are the controlling input parameters. Possible 403

high variation in these parameters necessitates their treatment as noise factors in the robust 404

design. The variation of soil parameters might be attributed to the inherent variability, 405

measuring error, and transformation uncertainty (Phoon and Kulhawy 1999), while the 406

variation of the surcharge behind the wall may come from the nearby transportation and 407

heaping. 408

With limited data availability at a given site in a typical geotechnical practice, an 409

accurate statistical characterization of the aforementioned noise factors may not be possible; 410

rather, these noise factors are usually characterized with their nominal values based on the 411

available data and local experience (in a deterministic manner). Listed in Table 1 are the 412

characterized nominal values of these soil parameters in the case studied (Figures 4 and 5). 413

20

Similarly, the surcharge behind the wall (q0) is characterized with a nominal value of 1.0 414

ton/m2. Based on the characterized nominal values of noise factors, the deterministic analysis 415

and design of the shoring system for the deep excavation can readily be conducted. Note that 416

the nominal values, listed in Table 1, are the only information available to the engineer in the 417

context of deterministic design. 418

4.4. Design pool selection for the robust design optimization 419

Comparing to the hard-to-characterize (or hard-to-control) noise factors, the diameter 420

of the concrete soldier pile (D), length of the concrete soldier pile (L), interval of concrete 421

soldier piles or center-to-center spacing of soldier piles (I), vertical spacing of tieback anchors 422

(V), horizontal spacing of tieback anchors (H), and installed angle of tieback anchors with 423

respect to the horizontal direction () are the easy-to-control input parameters. These input 424

parameters can be selected by the designer; they are classified as design parameters in the 425

robust design of the shoring system for the deep excavation. 426

Apart from the aforementioned noise factors and design parameters, the stability and 427

serviceability behaviors of the shoring system can also be affected by other input parameters, 428

such as the preload of tieback anchors and the length of tieback anchors. For the limited scope 429

of this paper, the preload of tieback anchors are all set up as 20 ton/tieback, and the free 430

length and fixed length of tieback anchors are both designed as 8.0 m based on the local 431

design experience. 432

For ease of construction, a discrete design pool is often considered based on the local 433

experience. For example, the discrete design pool (S), tabulated in Table 2, may be adopted 434

for the robust design optimization of this demonstrative excavation case study. In the selected 435

design pool, a total number of 38,500 discrete candidate designs is possible (see Table 2) and 436

21

they will be analyzed and the optimal design will be derived accordingly. Given a specified 437

vertical spacing of anchor tiebacks, the excavation steps can be determined accordingly, as 438

shown in Figure 5. In Figure 5, the location of the tieback anchor is set at approximately 0.8 439

m above the excavation depth at each excavation stage, except for the last excavation stage. 440

4.5. Cost estimate of the shoring system for the deep excavation 441

As formulated in Eq. (5), cost efficiency and design robustness are the two design 442

objectives in the robust design optimization of the shoring system for the deep excavation 443

project. For a given excavation project, the site dimensions and excavation depth are fixed 444

according to either the structural or architectural requirements. The major cost of concern in 445

the robust design optimization is the cost of the shoring system (C), which is the sum of the 446

cost for the soldier pile wall (Cw) and the cost for the tieback anchor systems (Ct), expressed 447

as: 448

w tC C C (11) 449

Using the geotechnical practice in Taiwan as an example, the cost for the soldier pile 450

wall (Cw) may be estimated as (Hsii-Sheng Hsieh, personal communication 2013): 451

2

10.6

w

SL DC L t

I

(12) 452

where SL is the perimeter of the excavation site (m). In this study, SL is taken as 54.08 m as 453

only section D in Figure 4 is considered. The parameter t1 is the unit price (US dollar) for a 454

0.6 m diameter pile, and in this study, t1= 66 US dollars. 455

Similarly, the cost for the tieback systems (Ct) may be estimated as (Hsii-Sheng Hsieh, 456

personal communication 2013): 457

2 2 3 t

SLC VL t l t

H

(13) 458

22

where VL is the number of vertical levels of the tieback anchor systems, which are determined 459

from Figure 5. The parameter t2 is the unit price (US dollar) for the unit length of the anchor, 460

l2 is the length of the anchor, and t3 is the unit price (US dollar) for the anchor head. In this 461

study, these unit prices are: t2 = t3 = 66 US dollars (note: it is a coincidence that t1 = t2 = t3 in 462

this design example). Note that the cost formulated above (Eqs. 11-13) is the construction cost 463

and the material cost is already included in which. 464

It should be noted that Eqs. (12) and (13) represent a rule of thumb used by Trinity 465

Foundation Engineering Consultants in its cost estimation for the design of the shoring system 466

for the excavation project that is being re-analyzed in this paper. It is adopted herein as an 467

example for the cost estimate of the robust design optimization. Any reasonable cost estimate 468

model that is applicable to the local excavation project can be used. 469

4.6. Robust design optimization results: Pareto front and knee point 470

Based on the robust optimization setting presented previously (see Eq. 5), the robust 471

design of the shoring system for the deep excavation can be conducted. For each of the 472

candidate designs in the specified design pool (see Table 2), the safety requirements (i.e., 473

factor of safety against push-in failure Fs1, factor of safety against basal heave failure Fs2, and 474

maximum wall deflection y), cost (Eq. 11), and design robustness (in terms of sensitivity 475

index SI) can be readily evaluated. Then, a Pareto front that consists of 61 non-dominated 476

designs is established using the sorting algorithm of NSGA-II (Deb et al. 2002), as shown in 477

Figure 6(a). It is noted that all the non-dominated designs on the Pareto front are superior to 478

all others in the feasible domain. Among the non-dominated designs on the Pareto front, 479

however, none of them is superior or inferior to others when all objectives are considered. 480

Thus, a tradeoff exists between design robustness and cost efficiency, as depicted in Figure 481

23

6(b). By applying the normal boundary intersection approach (Deb and Gupta 2011; Juang et 482

al. 2014), a knee point is identified as shown in Figure 6(b). Although not shown herein, the 483

same knee point is obtained using the marginal utility function approach (Branke et al. 2004; 484

Gong et al. 2014b). 485

On the left side of the knee point, a slight reduction in the cost leads to a drastic 486

increase in the sensitivity index SI, indicating a drastic reduction in design robustness. On the 487

other side of the knee point, a slight reduction in SI requires a large increase in the cost. Thus, 488

the knee point can be treated as the most preferred design in the design pool if no design 489

preference is specified by the owner or client. 490

For comparison, two other designs on the Pareto front, one representing the least-cost 491

design and the other the most robust design (which is also the most costly), are also shown in 492

Figure 6(c) along with the knee point and the original design. The original design is the one 493

designed by an experienced engineering firm (Hsii-Sheng Hsieh, personal communication 494

2013) without the knowledge of robust design. While the original design appears to be a 495

sound engineering practice, offering a compromise between the least cost design and the most 496

robust design in this case, it is inferior to the knee point on the Pareto front, as the latter is 497

more robust and cost less. Of course, the knee point is the most preferred design obtained with 498

a Pareto front that was obtained through robust design optimization. The design parameters, 499

safety performance (i.e., factor of safety against push-in failure Fs1, factor of safety against 500

basal heave failure Fs2, and maximum wall deflection y), sensitivity index (SI), and cost (C) 501

of these four designs are compared in Table 3. 502

While all four designs listed in Table 3 meet the safety requirements, such as Fs1 >1.5, 503

Fs2 > 1.5, and y < 0.7%Hf = 8.33 cm, different designs have different levels of design 504

24

robustness and cost efficiency. In this excavation problem, the least-cost design among the 505

pool of feasible designs is also the least robust design, in which a slight variation in noise 506

factors can result in a highest variation in the system response, which is not desirable. On the 507

other hand, the most robust design corresponds to the least cost efficiency, which is not 508

desirable, either. The knee point on the Pareto front, or the most preferred design in the design 509

pool, is shown as the best compromise between these two extremes on the Pareto front. 510

Compared with the original design that was selected by the local engineering firm without 511

using robust design, the knee point is cheaper and much more robust. 512

4.7. Simplified RGD with new procedure to identify the most preferred design 513

Although the RGD method shows a clear advantage of being able to find the best 514

compromise solution, or most preferred design, in the design pool, it requires use of genetic 515

algorithms such as NSGA-II (Deb et al. 2002) for the multi-objective optimization. For a 516

robust geotechnical design that involves FEM modeling in a multi-objective optimization 517

framework, the required computation effort can be overwhelming for a routine project. The 518

simplified RGD method with the proposed new procedure can find the most preferred design 519

in the design pool through a series of single-objective optimizations, without the need to use 520

genetic algorithms such as NSGA-II. 521

Using the simplified RGD method, the most preferred design in the design pool is 522

obtained without developing and sorting the Pareto front, as depicted in Figure 7. The design 523

parameters of this most preferred design, denoted as d2-2, are listed in Table 4. As a 524

comparison, the knee point obtained by the original RGD method implemented with the 525

normal boundary intersection approach (or the marginal utility function approach) is also 526

listed in Table 4. The latter two approaches within the original RGD actually yield the same 527

25

design, which is denoted as d2-1. The difference between the design parameters of d2-2 and that 528

of d2-1 is quite negligible. The results show that the most preferred design obtained with the 529

simplified RGD method is practically the same as the knee point obtained with the existing 530

RGD method that requires a two-step solution (developing a Pareto front by multi-objective 531

optimization using algorithms such as NSGA-II, and then searching for knee point using, for 532

example, the marginal utility function). This confirms the validity of the simplified RGD 533

method that is based on the new procedure for finding the most preferred design. 534

Table 5 compares the computation efforts of these three approaches. The computation 535

time is recorded using the same Windows 7® PC equipped with a 8.0 GB RAM and an Intel® 536

Core™ i7-3630QM CPU running at 2.40GHz. The simplified RGD method with the proposed 537

new procedure for knee point is seen to have drastically reduced the computation time. The 538

savings in the computation time here is mainly attributed to the use of the proposed procedure 539

to find the most preferred design in the design pool. 540

541

5. Further Discussions on the Simplified RGD Method 542

5.1. Sensitivity of the system response with respect to noise factors 543

In the design of the shoring system for the excavation project analyzed previously, the 544

system response of concern is the maximum wall deflection. A good way to examine the 545

sensitivity of the system response to each of the noise factors is to compare the normalized 546

gradient vector (Eq. 3), which might be interpreted as the variation of the system response 547

caused by one percent change in each noise factor while maintaining constants for all other 548

noise factors. This gradient could also be viewed as a measure of the prediction error in the 549

system response caused by one percent variation in the estimated noise factor. 550

26

Figure 8 shows a comparison of the variation of the system response (i.e., maximum 551

wall deflection in this case) due to the relative variation in each noise factor in each of these 552

five designs listed in Tables 3 and 4 (designated as d1, d2-1, d2-2, d3, and d0 respectively). In 553

each of these five designs, the system response is most affected by the friction angles of the 554

top three soil layers and the surcharge behind the wall. This is quite reasonable since the 555

excavation work is mainly carried out in the top three soil layers. The implication is that the 556

prediction accuracy of the maximum wall deflection is greatly dependent upon the accuracy 557

of the characterization of the top soil layers in this excavation project. 558

Also revealed in Figure 8 is that through the use of RGD method, the derived final 559

design (i.e., either d2-1 from the existing RGD method or d2-2 from the simplified RGD 560

method) yields a significant reduction of the sensitivity of the system response to the noise 561

factors, comparing to the least-cost design (d1) and the original design (d0). Further reduction 562

in the sensitivity can be achieved, for example, by adopting the most robust design (d3) in the 563

design pool; however, the cost efficiency is a major concern for this most robust design. 564

5.2. Effect of the uncertainty in noise factors on the system response 565

It is often difficult to characterize noise factors with certainty in geotechnical practices 566

due to limited data availability at a given site. The uncertainty in the estimated noise factors 567

can lead to the variation in the predicted system response. To investigate the effect of the 568

robust design on the system response, all the non-dominated designs on the Pareto front 569

obtained with the existing RGD method (see Figure 6) are analyzed herein for their 570

feasibilities assuming that there is a 10% coefficient of variation (COV) in each and every 571

noise factor. For this series of analysis, the feasibility is defined as the probability that the 572

design remains meeting the serviceability requirement in the face of uncertainties, which is 573

27

denoted as Pr 0.7% fy H . For each non-dominated design on the Pareto front, 10,000 574

Monte Carlo simulations (MCS) are performed herein to sample the noise factors from the 575

joint distribution of noise factors; as such, the feasibility (or the probability of meeting the 576

design constraint of 0.7% fy H ) can be computed. 577

Figure 9 shows the feasibilities computed for all the non-dominated designs on the 578

Pareto front versus their corresponding sensitivity indexes. The feasibility of the design is 579

seen to decrease with the increase in the sensitivity index when there is a 10% variation in 580

noise factors. The implication is that a design with a lower sensitivity index (or higher design 581

robustness) has a higher probability to remain feasible in the face of the uncertainty in noise 582

factors. That is to say, a system might remain feasible even if the input parameters undergo 583

some variation, if the system is designed with high robustness against the unforeseen variation 584

of noise factors. Thus, the robustness measure in terms of the gradient-based sensitivity index, 585

proposed in this paper, is consistent with the well-accepted feasibility robustness measure 586

(Parkinson et al. 1993; Park et al. 2006; Juang et al. 2013a; Wang et al. 2013), which further 587

confirms the validity of Eqs. (1) through (4). 588

589

6. Concluding Remarks 590

In this paper, the authors present a simplified robust geotechnical design (RGD) 591

method. In the context of robust design, the safety requirements, cost efficiency, and design 592

robustness are simultaneously considered and implemented in a multi-objective optimization 593

framework. The results of such optimization are usually expressed as a Pareto font, as cost 594

efficiency and design robustness are conflicting design objectives. The knee point is then 595

identified that yields the best compromise design among all the non-dominated designs on the 596

28

Pareto front. This existing RGD method is improved in this paper in two aspects: 1) a more 597

efficient measure of design robustness based on gradient is developed, and 2) a new procedure 598

to identify the most preferred design in the design pool is introduced. The simplified RGD 599

method essentially turns the multi-objective optimization into a series of single-objective 600

optimizations. The following conclusions are reached based on the results presented: 601

1) The simplified RGD method is demonstrated as an effective and practical design tool 602

that considers the safety requirements, cost efficiency and design robustness 603

simultaneously. This simplified RGD method is formulated so that no reliability or 604

probabilistic analysis is required, and as such, is more readily adaptable for general 605

applications. The advantages of the simplified RGD method over the existing RGD 606

method are demonstrated in the example design of a shoring system for deep 607

excavation. 608

2) The new gradient-based robustness measure, in terms of sensitivity index, is shown to 609

be intuitive and effective, as it corresponds to the variation of the system response due 610

to a relative variation in noise factors. The design with a lower sensitivity index 611

always yields a lower variation of the system response regardless of the variation 612

levels of noise factors, and thus higher design robustness and a lower potential risk 613

(indicated by a higher feasibility) in the face of the uncertainty in noise factors. This 614

definition is easier to implement in a RGD and computationally more efficient than the 615

existing robustness measures. 616

3) With the proposed new procedure, the most preferred design in the design pool can be 617

identified directly through a series of single-objective optimizations, which does not 618

involve genetic algorithms such as NSGA-II, as such the simplified RGD method can 619

29

be easily implemented with Matlab function fmincon or Microsoft Excel Solver. The 620

proposed new procedure for finding the most preferred design in the design pool is 621

shown to be effective and computationally more efficient than the existing methods 622

that are based on the Pareto front obtained through a multi-objective optimization, 623

such as normal boundary intersection approach and marginal utility function approach. 624

4) Compared to the traditional deterministic design approach, the simplified RGD 625

method shows an advantage for being able to consider the effect of the uncertainty in 626

noise factors on the system response (through minimizing the sensitivity index of the 627

system response to the noise factors). This advantage allows for a more robust design 628

to be achieved with a balancing consideration of cost. Compared to the probabilistic 629

design approach, the simplified RGD method shows an advantage for being able to 630

achieve similar design results (i.e., to achieve a desired level of feasibility in the face 631

of uncertainties) without having to perform a probabilistic analysis of the system 632

response with a full statistical characterization of noise factors. The superiorities of the 633

simplified RGD method are demonstrated. 634

635

Acknowledgments 636

The study on which this paper is based was supported in part by the National Science 637

Foundation through Grant CMMI-1200117 (“Transforming Robust Design Concept into a 638

Novel Geotechnical Design Tool”) and the Glenn Department of Civil Engineering, Clemson 639

University. The results and opinions expressed in this paper do not necessarily reflect the 640

views and policies of the National Science Foundation. The TORSA code was provided by Dr. 641

Hsii-Sheng Hsieh of Trinity Foundation Engineering Consultants (TFEC), Taipei, Taiwan. Dr. 642

30

Hsieh is also thanked for his untiring effort in answering our countless questions regarding 643

TORSA and the engineering practice of braced excavations during the course of this study. 644

645

31

References 646

Ang, A.H.S., Tang, W.H., 2007. Probability Concepts in Engineering: Emphasis on 647

Applications to Civil and Environmental Engineering, 2nd edition, Wiley, New York, 648

USA. 649

Baecher, G.B., Christian, J.T., 2003. Reliability and Statistics in Geotechnical Engineering, 650

Wiley, New York, USA. 651

Bechikh, S., Said, L.B., Ghédira, K., 2011. Searching for knee regions of the Pareto front 652

using mobile reference points. Soft Computing 15(9), 1807-1823. 653

Beyer, H.G., Sendhoff, B., 2007. Robust optimization – A comprehensive survey. Computer 654

Methods in Applied Mechanics and Engineering 196(33), 3190-3218. 655

Branke, J., Deb, K., Dierolf, H., Osswald, M., 2004. Finding knees in multi-objective 656

optimization. Parallel Problem Solving from Nature-PPSN VIII, 722-731. 657

Chen, W., Wiecek, M.M., and Zhang, J., 1999. Quality utility—A compromise programming 658

approach to robust design. Journal of Mechanical Design 121(2), 179-187. 659

Clough, G.W., O’Rourke, T.D., 1990. Construction induced movements of in situ walls. In: 660

Proceedings of Design and Performance of Earth Retaining Structure, GSP 25, ASCE, 661

New York, 439-470. 662

Deb, K., 2001. Multi-objective Optimization Using Evolutionary Algorithms. John Wiley and 663

Sons, New York, USA. 664

Deb, K., Gupta, S., 2011. Understanding knee points in bicriteria problems and their 665

implications as preferred solution principles. Engineering Optimization, 43(11), 666

1175-1204. 667

Deb, K., Sundar, J., Udaya Bhaskara Rao, N., Chaudhuri, S.. 2006. Reference point based 668

32

multi-objective optimization using evolutionary algorithms. International Journal of 669

Computational Intelligence Research 2(3), 273-286. 670

Deb, K., Pratap, A., Agarwal, S., Meyarivan, T., 2002. A fast and elitist multi-objective 671

genetic algorithm: NSGA-II. IEEE Transactions on Evolutionary Computation 6(2), 672

182-197. 673

Deep Excavation LLC, 2014. DeepXcav User’s Manual, 30-62 Steinway St. #188, Astoria, 674

New York. 675

Doltsinis, I., Kang, Z., Cheng, G., 2005. Robust design of non-linear structures using 676

optimization methods. Computer Methods in Applied Mechanics and Engineering 677

194(12), 1779-1795. 678

Fenton, G.A., Griffiths, D.V., 2008. Risk Assessment in Geotechnical Engineering. John 679

Wiley and Sons, New York, USA. 680

Goh, K.H., Mair, R.J., 2014. Response of framed buildings to excavation-induced movements. 681

Soils and Foundations 54(3), 250-268. 682

Gonsamo, A., 2011. Normalized sensitivity measures for leaf area index estimation using 683

three-band spectral vegetation indices. International Journal of Remote Sensing 32(7), 684

2069-2080. 685

Gong, W., Wang, L., Juang, C.H., Zhang, J., Huang, H., 2014a, Robust geotechnical design of 686

shield-driven tunnels. Computers and Geotechnics 56, 191-201. 687

Gong, W., Khoshnevisan, S., Juang, C.H., 2014b, Gradient-based design robustness measure 688

for robust geotechnical design. Canadian Geotechnical Journal 51(11), 1331-1342. 689

33

Gong, W., Luo, Z., Juang, C. H., Huang, H., Zhang, J., Wang, L. 2014c. Optimization of site 690

exploration program for improved prediction of tunneling-induced ground settlement in 691

clays. Computers and Geotechnics 56, 69-79. 692

Harr, M.E., 1987. Reliability-Based Design in Civil Engineering, McGraw-Hill Book, New 693

York, USA. 694

Japanese Society of Architecture (JSA), 1988. Guidelines of Design and Construction of Deep 695

Excavations. Japanese Society of Architecture, Tokyo, Japan. 696

Juang, C.H., Wang, L., Atamturktur, S., Luo, Z., 2012a. Reliability-based robust and optimal 697

design of shallow foundations in cohesionless soil in the face of uncertainty. Journal of 698

GeoEngineering 7(3), 75-87. 699

Juang, C.H., Luo, Z., Atamturktur, S., Huang, H., 2012b. Bayesian updating of soil 700

parameters for braced excavations using field observations. Journal of Geotechnical and 701

Geoenvironmental Engineering 139(3), 395-406. 702

Juang, C.H., Wang, L., Liu, Z., Ravichandran, N., Huang, H., Zhang, J., 2013a. Robust 703

Geotechnical Design of drilled shafts in sand – New design perspective. Journal of 704

Geotechnical and Geoenvironmental Engineering 139(12), 2007-2019. 705

Juang, C.H., Ching, J., Luo, Z., 2013b. Assessing SPT-based probabilistic models for 706

liquefaction potential evaluation: a 10-year update. Georisk: Assessment and management 707

of risk for engineered systems and geohazards 7(3), 137-150. 708

Juang, C.H., Wang, L., Hsieh, H.S., Atamturktur, S., 2014. Robust geotechnical design of 709

braced excavations in clays. Structural Safety 49, 37-44. 710

Kim, C., Kwon, J., Im, J.C., Hwang, S., 2012. A method for analyzing the self-supported 711

earth-retaining structure using stabilizing piles. Marine Georesources and Geotechnology 712

34

30(4), 313-332. 713

Kim, D.S., Lee, B.C., 2005. Instrumentation and numerical analysis of cylindrical diaphragm 714

wall movement during deep excavation at coastal area. Marine Georesources and 715

Geotechnology 23(1-2), 117-136. 716

Khoshnevisan, S., Gong, W., Wang, L., Juang, C.H., 2014. Robust design in geotechnical 717

engineering–an update. Georisk: Assessment and Management of Risk for Engineered 718

Systems and Geohazards 8(4), 217-234. 719

Kung, G.T.C., Juang, C.H., Hsiao, E.C.L., Hashash, Y.M.A., 2007. A simplified model for 720

wall deflection and ground surface settlement caused by braced excavation in clays. 721

Journal of Geotechnical and Geoenvironmental Engineering 133(6), 731-747. 722

Likitlersuang, S., Surarak, C., Wanatowski, D., Oh, E., Balasubramaniam, A., 2013. Finite 723

element analysis of a deep excavation: A case study from the Bangkok MRT. Soils and 724

Foundations 53(5), 756-773. 725

Marrel, A., Iooss, B., Laurent, B., Roustant, O., 2009. Calculations of Sobol indices for the 726

gaussian process metamodel. Reliability Engineering and System Safety 94(3), 742-751. 727

MATLAB version 7.10.0. Natick, Massachusetts: The MathWorks Inc; 2010. 728

Oka, Y., Wu, T.H., 1990. System reliability of slope stability. Journal of Geotechnical 729

Engineering 116(8), 1185-1189. 730

Ou, C.Y., 2006. Deep Excavation-Theory and Practice. Taylor and Francis, England. 731

Park, G.J., Lee, T.H., Lee, K.H., Hwang, K.H., 2006. Robust design: an overview. AIAA 732

journal 44(1), 181-191. 733

Parkinson, A., Sorensen, C., Pourhassan, N., 1993. A general approach for robust optimal 734

design. Journal of Mechanical Design 115(1), 74-80. 735

35

Phadke, M.S., 1989. Quality Engineering Using Robust Design. Prentice Hall, New Jersey, 736

USA. 737

Phoon, K.K., Kulhawy, F.H., 1999. Characterization of geotechnical variability. Canadian 738

Geotechnical Journal 36(4), 612-24. 739

Plaxis bv. PLAXIS-2D, 2014. User’s Manual, 2600 AN Delft, The Netherlands. 740

Professional Standards Compilation Group (PSCG), 2000. Specification for Excavation in 741

Shanghai Metro Construction. Professional Standards Compilation Group, Shanghai, 742

China. 743

Sino-Geotechnics, 2010. User Manual of Taiwan Originated Retaining Structure Analysis for 744

Deep Excavation. Sino-Geotechnics Research and Development Foundation, Taipei, 745

Taiwan. 746

Taguchi, G., 1986. Introduction to Quality Engineering: Designing Quality into Products and 747

Processes. Quality Resources, New York, USA. 748

Taiwan Geotechnical Society (TGS), 2001. Design Specifications for the Foundation of 749

Buildings. Taiwan Geotechnical Society, Taipei, Taiwan. 750

Wang, L., Hwang, J.H., Juang, C.H., Atamturktur, S., 2013. Reliability-based design of rock 751

slopes – A new perspective on design robustness. Engineering Geology 154, 56-63. 752

Wang, L., Juang, C.H., Atamturktur, S., Gong, W., Khoshnevisan, S., Hsieh, H.S., 2014. 753

Optimization of design of supported excavations in multi-layer strata. Journal of 754

GeoEngineering 9(1), 1-12. 755

Zhang, L.M., Li, D.Q., Tang, W.H., 2005. Reliability of bored pile foundations considering 756

bias in failure criteria. Canadian Geotechnical Journal 42(4),1086-1093. 757

758

36

Lists of Tables

Table 1. Basic soil properties adopted in the case study of deep excavation

Table 2. Design pool for the robust design optimization

Table 3. Analysis of the four designs of the shoring system for the deep excavation project

Table 4. Most preferred design (i.e., knee point) obtained with different approaches

Table 5. Comparison of the computational efficiency among the three approaches

37

Lists of Figures

Figure 1. Illustration of the robustness concept

Figure 2. Conceptual illustration of Pareto front and knee point in a bi-objective optimization

(a) Optimization problem with a single knee point

(b) Optimization problem with multi-knee points

Figure 3. Different approaches to identify the most preferred design (i.e., knee point)

(a) Reflex angle approach

(b) Normal boundary intersection approach

(c) New procedure proposed in this paper

Figure 4. Plan layout of the excavation site in the case study

Figure 5. Four possible tieback anchor layouts of the shoring system

(a) Vertical spacing, V = 2.0 m

(b) Vertical spacing, V = 2.5 m

(c) Vertical spacing, V = 3.0 m

(d) Vertical spacing, V = 3.5 m

Figure 6. Results of the robust design of the shoring system for deep excavation

(a) Pareto front obtained from robust design optimization

(b) Normal boundary intersection approach

(c) Preferred designs on the Pareto front and the original design

Figure 7. New simplified RGD procedure to find the most preferred design

Figure 8. Sensitivity of the system response of concern (i.e., maximum wall deflection) with

respect to noise factors (Note: 1 = φ of 1st soil layer; 2 = kh of 1

st soil layer; 3 = φ

of 2nd

soil layer; 4 = kh of 2nd

soil layer;5 = φ of 3rd

soil layer; 6 = kh of 3rd

soil

layer;7 = φ of 4th

soil layer; 8 = kh of 4th soil layer; 9 = su of 4

th soil layer; 10 =

q0 behind the wall; the designs d1, d2-1, d2-2, d3, and d0 are listed in Tables 3 and 4)

Figure 9. Design feasibility versus sensitivity index (Note: the noise factors are subject to a

10% variation)

38

Table 1. Basic soil properties adopted in the case study of deep excavation

Layer

No.

Soil

type

Depth

(m)

Weight, γt

(ton/m3)

cohesion, c (ton/m2)

Friction

angle, φ (°)

Modulus of horizontal

subgrade reaction, kh (ton/m3)

1 SM 3.35 1.80 0.0 28.0 500

2 GP 11.15 1.90 0.0 30.0 4000

3 GP 15.50 2.30 0.0 36.0 5000

4 Rock 20.0 2.40 2.0 34.0 5000

39

Table 2. Design pool for the robust design optimization

Design parameter Design pool (defined by ranges of design parameters)

Diameter of the solider pile, D

(m) {0.3 m, 0.4m, 0.5m, 0.6 m, 0.7 m}

Length of the solider pile, L

(m) {14 m, 15 m, 16 m, 17 m, 18 m, 19 m, 20 m}

Horizontal interval of the

solider pile, I (m) {D, D + 0.1 m, D + 0.2 m, …, D + 1.0 m}

Vertical spacing of tieback

anchors, V (m) {2.0 m, 2.5 m, 3.0 m, 3.5 m}

Horizontal spacing of tieback

anchors, H (m) {1.5 m, 2.0 m, 2.5 m, 3.0 m, 3.5 m}

Installed angle of the tieback

anchor, () {10, 15, 20, 25, 30}

40

Table 3. Analysis of the four designs of the shoring system for the deep excavation project

Resulting design Design parameters Safety performance Cost, C

(10,000 USD)

Sensitivity

index, SI D (m) L (m) I (m) V (m) H (m) () Fs1 Fs2 y (cm)

Least cost design, d1 0.5 17 1.3 3.0 3.5 10 4.46 2.75 8.20 8.66 19.06

Most preferred design, d2-1 0.6 18 1.6 3.0 2.0 10 5.67 2.96 3.48 13.12 4.28

Most robust design, d3 0.7 20 0.7 3.0 2.0 15 7.82 3.34 1.49 23.10 1.67

Original design, d0 0.5 17 0.6 3.0 2.5 20 4.46 2.75 4.13 14.42 11.22

Note: d1: least cost design on Pareto front; d2-1: most preferred design on Pareto front (identified by the marginal utility function

approach); d3: most robust design on Pareto front; and d0: original design in an actual project.

41

Table 4. Most preferred design (i.e., knee point) obtained with different approaches

Adopted approach Design parameters Safety performance Cost, C

(10,000 USD)

Sensitivity

index, SI D (m) L (m) I (m) V (m) H (m) () Fs1 Fs2 y (cm)

Normal boundary intersection

approach, d2-1 0.6 18 1.6 3.0 2.0 10 5.67 2.96 3.48 13.12 4.28

Marginal utility function

approach, d2-1 0.6 18 1.6 3.0 2.0 10 5.67 2.96 3.48 13.12 4.28

Proposed new procedure, d2-2 0.5 18 1.4 3.0 2.0 10 5.67 2.96 4.94 12.31 5.58

Note: d2-1: most preferred design on Pareto front (identified by the marginal utility function approach or the normal boundary

intersection approach); d2-2: most preferred design by the simplified RGD with the proposed new procedure.

42

Table 5. Comparison of the computational efficiency among the three approaches

Adopted approach

Computation time

Sorting Pareto front

using NSGA (sec)

Identifying knee

point (sec)

Total time

(sec)

Normal boundary intersection

approach, d2-1 110.6 0.5 111.1

Marginal utility function approach,

d2-1 110.6 0.8 111.4

Proposed new procedure,

d2-2 5.7 5.7

Note: d2-1: most preferred design on Pareto front (identified by the marginal utility function

approach or the normal boundary intersection approach); d2-2: most preferred design by the

simplified RGD with the proposed new procedure.

43

f (d1, )

Sensitive

Design response, f (d, )

Noise factors,

Robust

System with high gradient

f (d2, )

System with low gradient

Variation level 2

Variation level 1

Figure 1. Illustration of the robustness concept

44

Pareto front

Ob

ject

ive

2,

min

[f2(d

)]

Objective 1, min [f1(d)]

Feasible domain

Knee point

Utopia point

(a) Optimization problem with a single knee point

Local knee point

Pareto front

Ob

ject

ive

2,

min

[f2(d

)]

Objective 1, min [f1(d)]

Knee point

Utopia point

Global knee point

Local knee point

(b) Optimization problem with multi-knee points

Figure 2. Conceptual illustration of Pareto front and knee point in a bi-objective optimization

45

Pareto front

Ob

ject

ive

2,

min

[f2(d

)]

Objective 1, min [f1(d)]

B

AKnee point

Maximum reflex angle

(a) Reflex angle approach

Pareto front

Obje

ctiv

e 2,

min

[f2(d

)]

Objective 1, min [f1(d)]

Maximum distance

A

B

Knee point

Boundary line

(b) Normal boundary intersection approach

Feasible design

Ob

ject

ive

2,

min

[f2(d

)]

Objective 1, min [f1(d)]

Knee point

Minimum distance

Utopia point

(c) New procedure proposed in this paper

Figure 3. Different approaches to identify the most preferred design (i.e., knee point)

46

54.08 m

48. 90 m

28.74 m

AA

BB

DD

CC

41.79 m 27.68 m

Section E 8.5 m

Section D 11.9 m

Section C 11.4 m

Section B 9.15 m

Section A 6.9 m

Section Excavation depth

EE

Figure 4. Plan layout of the excavation site in the case study

47

Surcharge, q0

3.35 m

11.15 m

15.50 m

20.00 m

SM

GP

GP

Rock

10.30 m

8.30 m

6.30 m

4.30 m

2.30 m

11.90 m

4 . 8 0 m

S u r c h a r g e , q0

3 . 3 5 m

1 1 . 1 5 m

1 5 . 5 0 m

2 0 . 0 0 m

SM

GP

GP

Rock

2.30 m

11.90 m

7.30 m

9.80 m

(a) Vertical spacing, V = 2.0 m (b) Vertical spacing, V = 2.5 m

Surcharge, q0

3.35 m

11.15 m

15.50 m

20.00 m

SM

GP

GP

Rock

2.30 m

11.90 m

5.30 m

8.30 m

9 . 3 0 m

5 . 8 0 m

S u r c h a r g e , q0

3 . 3 5 m

1 1 . 1 5 m

1 5 . 5 0 m

2 0 . 0 0 m

SM

GP

GP

Rock

2.30 m

11.90 m

(c) Vertical spacing, V = 3.0 m (d) Vertical spacing, V = 3.5 m

Figure 5. Four possible tieback anchor layouts of the shoring system

48

5 10 15 20 250

4

8

12

16

20

24

Sen

siti

vit

y i

nd

ex,

SI

Cost, C (10,000 USD)

Pareto front

Feasible design

(a) Pareto front obtained from robust design optimization

5 10 15 20 250

4

8

12

16

20

24

Sen

siti

vit

y i

ndex

, SI

Cost, C (10,000 USD)

Pareto front

Maximum distance

Knee point

(b) Normal boundary intersection approach

5 10 15 20 250

4

8

12

16

20

24

Sen

siti

vit

y i

nd

ex,

SI

Cost, C (10,000 USD)

Pareto front

Preferred designs

Original design

Knee point on

the Pareto front Most robust design

Least cost design

(c) Preferred designs on the Pareto front and the original design

Figure 6. Results of the robust design of the shoring system for deep excavation

49

5 10 15 20 250

4

8

12

16

20

24

Sen

siti

vit

y i

nd

ex,

SI

Cost, C (10,000 USD)

Utopia point

Feasible design

Minimum distance

Most preferred design

Figure 7. New simplified RGD procedure to find the most preferred design

50

-16

-12

-8

-4

0

4

10

9876543

2

Sen

siti

vit

y v

ecto

r, J

i

d1

d2-1

d2-2

d3

d0

1

Figure 8. Sensitivity of the system response of concern (i.e., maximum wall deflection) with

respect to noise factors (Note: 1 = φ of 1st soil layer; 2 = kh of 1

st soil layer; 3 = φ of 2

nd soil

layer; 4 = kh of 2nd

soil layer;5 = φ of 3rd

soil layer; 6 = kh of 3rd

soil layer;7 = φ of 4th soil

layer; 8 = kh of 4th soil layer; 9 = su of 4

th soil layer; 10 = q0 behind the wall; the designs d1,

d2-1, d2-2, d3, and d0 are listed in Tables 3 and 4)

51

0 4 8 12 16 200.0

0.2

0.4

0.6

0.8

1.0

1.2

Non-dominated design

Fea

sib

ilit

y,

Pr[

y <

0.7

%H

f ]

Sensitivity index, SI

Most preferred design (d2-1

)

Most preferred design (d2-2

)

Figure 9. Design feasibility versus sensitivity index (Note: the noise factors

are subject to a 10% variation)