Experimental Investigation of Longitudinal Bending of ...

0 downloads 0 Views 1MB Size Report
Key words: steel pipe, buried pipe, soil, bending, ground movement ... the Pipeline and Hazardous Materials Safety Administration (PHMSA, 2010) reports that ...
1

Experimental Investigation of Longitudinal Bending of Buried

2

Steel Pipes Pulled through Dense Sand

3

Mohamed Almahakeri1; Amir Fam2; Ian D. Moore3

4

Affiliation:

5

1

Graduate Student

6

2

Professor and Canada Research Chair in Innovative and Retrofitted Structures

7

3

Professor and Canada Research Chair in Infrastructure Engineering

8

Dept. of Civil Engineering, Queen’s University, Kingston, ON K7L 3N6

9 10 11 12 13 14 15 16 17 18 19 20

Corresponding Author: Ian D. Moore Ellis Hall, Room 249 Queen’s University 58 University Ave. Kingston, On, Canada K7L 3N6 Phone: (613) 533-3160 Email: [email protected]

21 22 23

Abstract

24

North America is traversed by many high pressure oil and gas transmission pipes, and the

25

stability of that essential buried infrastructure must be maintained under a variety of earth

26

loading conditions. In this study, a series of pipe bending experiments have been conducted on

27

105 mm (4.1 inch) outside diameter and 1830 mm (6 feet) long steel pipes buried in dense sand

28

placed in a 4000x2000x2000 mm (157.5x78.7x74.7 inch) test pit. The pipe ends were pulled by 1

29

two parallel cables attached to a spreader beam outside the test region, which was pulled by a

30

hydraulic actuator. The study investigated burial depth-to-diameter (H/D) ratios of 3, 5 and 7 as

31

well as two horizontal extents for the soil behind the pipe distances of 3D and 9.5D. Special

32

consideration was made to assess the influence of friction between the pulling cables and soil. It

33

was shown that this friction is significant and may contribute about 20% of the maximum pulling

34

load for the case of H/D = 3. Consistency of results was established using four test repetitions

35

for some cases. While the horizontal extent of soil behind the pipe tested in this study had an

36

insignificant influence on the pulling forces, the burial depths significantly influenced the

37

ultimate pulling forces for the system. The failure mechanism controlling the limiting pulling

38

force in these tests was consistently governed by soil, and the pipe remained elastic.

39 40

Key words: steel pipe, buried pipe, soil, bending, ground movement

41 42

INTRODUCTION

43

Buried energy pipelines cross zones of soil instability and may need to be designed to

44

resist lateral or other forces resulting from potential soil movements. Relative pipe-soil

45

movements can result from natural phenomena such as soil creep, slope failures, landslides, and

46

earthquakes-induced faults. Soil loading may lead to pipeline failures which could negatively

47

affect the surrounding environment, local or global economy, and may even be a source of

48

catastrophic safety hazard if flammable contents ignite in populated areas. Pipeline failures due

49

to geotechnical causes are one of the main categories identified by several pipeline regulatory

50

and operation-monitoring agencies around the globe. In Canada, the National Energy Board

51

(NEB, 2008) reported that 7% of pipeline failures are due to geotechnical causes. In the USA,

52

the Pipeline and Hazardous Materials Safety Administration (PHMSA, 2010) reports that 2.5% 2

53

of significant pipeline incidents are due to earth movements. For Europe, the European Gas

54

pipeline Incident data Group (EGIG, 2008) reported that 7% of all pipeline failures were caused

55

by ground movements. Further statistical analysis of the frequency of failure data presented by

56

(EGIG, 2008) on nine pipe size categories (covering pipe diameter sizes ranges from below 102

57

mm (4 inch), up to above 1219 mm (48 inch)) yields percentages of 21% and 26% of the failures

58

being for pipe diameters of 0-102 mm (0-4 inch), and 127-254 mm (5-10 inch), respectively.

59

Such percentages make these two pipe size categories at a greater risk of failure than those in the

60

seven higher remaining size categories surveyed. Flexural behavior of buried pipes due to lateral

61

earth movement has rarely been studied, with few full-scale experimental data sets available in

62

the public domain. Some design guidelines (e.g. the American Society of Civil Engineers

63

(ASCE, 1984), and the Pipeline Research Council International (PRCI, 2004)) provide suggested

64

relations to estimate the lateral loads imposed on a buried pipeline when subjected to lateral earth

65

movement. However, these formulations are based on plane strain conditions, which assume that

66

the pipe moves as a rigid body with no flexure of the pipe. While most of the previous research

67

work has been focusing on the force-displacement characteristics, little experimental work has

68

been conducted to examine the flexural behavior of buried steel pipes (e.g. Konuk et al., 1999).

69

The objective of the current study is to measure strains and deflections along pipes at various

70

burial depths, the progressive development of deflected pipe shape, and the manner in which

71

different lateral displacements at different positions along the pipe influence the development of

72

lateral force (and peak lateral force in particular). As a start, to achieve these goals, performance

73

of test setup and consistency of results are assessed, which involves explicit evaluation of

74

friction on the pulling mechanism, the boundary effects, and reproducibility of replicate tests.

75

The experimental testing program is conducted on small diameter (102 mm (4.0 in) nominal

76

diameter, D , and 105 mm (4.1 in) for outside diameter, Do) buried steel pipes of 2.7 mm wall 3

77

thickness, subjected to lateral earth loading, to understand their flexural behavior and the

78

resistance of the overall pipe-soil system. Nine tests on steel pipes buried in dense sand were

79

conducted using the geotechnical laboratory at Queen’s University. Steel pipes with length-to-

80

diameter ratio of 18 were pulled laterally at the pipe ends. Force-displacement curves, strains,

81

and deflected profiles of the pipes were monitored and recorded during the tests. Also, the initial

82

and final positions of the pipe and the soil surface deformations were surveyed before and after

83

each test.

84

LITERATURE REVIEW

85 86

Centrifuge Modeling

87

Centrifuge modeling has proven to be a very useful tool for studying pipe-soil interaction

88

simulating large diameter pipes and deep burial depths in a relatively small and cost-effective

89

experimental setting. Several studies have been conducted examining the relative movement

90

between pipe and soil, for both lateral and oblique movement directions. Paulin et al. (1995)

91

conducted centrifuge modeling tests investigating the effects of burial depth, displacement rate,

92

and trench width for rigid pipes buried in clay. O’Rourke et al. (2005) examined axial and

93

bending strains in buried aluminum pipes with different diameter to wall thickness (D/t) ratios

94

under lateral earth fault loading, simulating behavior of larger diameter steel pipes with

95

equivalent D/t ratios. Da et al. (2008) investigated the influence of pipe-fault orientation on pipe

96

behavior under earthquake faulting. Also, the orientation of relative pipe-soil movement has been

97

studied by Daiyan et al. (2010). One of the main outcomes of their testing work is that it

98

demonstrated that coupled pipe-soil interaction due to oblique loading can considerably increase

99

the soil restraint on the pipeline. However, this interaction is minimal at very low angles.

100 4

101

Model and Full Scale Testing

102

One of the first small scale test programs conducted on steel pipes (25 mm, 60 mm, and

103

114 mm in diameter) was performed by Audibert and Nyman (1977) where the soil behavior was

104

observed during lateral displacement of pipes buried in sand with different densities under a wide

105

range of embedment ratios (1.5 to 24.5). Audibert and Nyman (1977) suggested a rectangular

106

hyperbolic function for the dimensionless force-displacement relation which was in agreement

107

with the one suggested by Das and Seeley (1975) who studied vertical anchor resistance against

108

horizontal movement.

109

Trautmann and O'Rourke (1983) conducted full-scale experiments on pipes with

110

diameters of 102 mm and 324 mm subjected to lateral movements, within sand of three different

111

densities (corresponding to friction angles of 31°, 36°, and 44°). Those tests featured burial

112

depths ranging from 1.5 to 22 diameters. Their work was used as the basis for the ASCE (1984)

113

guidelines for calculating lateral soil loading on pipes. Nyman (1984) also proposed design

114

procedures to develop bilinear load-displacement relationships for soil loads on pipelines

115

subjected to horizontal and vertical movements.

116

Hsu (1993) conducted an extensive full-scale testing program (120 tests) examining

117

several preliminary variables such as burial depth, pulling rate, soil density, and pipe diameter.

118

His results showed that soil resistance and corresponding displacements exhibit a power law

119

relation with the pipe velocity. Hsu et al. (2001) presented an experimental study of the soil

120

friction loads for oblique pipeline movements in loose sand, and another on dense sand (Hsu et

121

al., 2006). Both of studies presented and discussed analytical models to predict the longitudinal

122

and transverse soil loads.

123

Paulin et al. (1998) conducted 24 tests during their large scale experimental program that

124

included five types of pipe-soil loading test conditions including upward movement, lateral 5

125

movement, downward movement, and axial movement. Since the results of that testing program

126

are proprietary, only relative pipe loading and displacement were presented. A number of other

127

studies examining the flexural behavior of buried pipelines have also been conducted at the

128

Centre for Cold Ocean Resources Engineering (C-CORE) by Konuk et al. (1999). Details of the

129

physical tests can be found in Hurley et al. (1998a, and 1998b). Konuk et al. (1999) conducted

130

two large-scale tests to assess the bending behavior of buried pipes in dense sand under lateral

131

loading. During each test, measurements were made of the deflected pipeline profile, pipeline

132

ovalization, pipeline forces, and the associated soil deformations. Plastic hinges developed in the

133

pipe for both of the tests. Most of this work conducted at C-CORE is proprietary and only partial

134

results are available (Konuk et al., 1999).

135

One of the most recent testing programs is the work conducted at the University of

136

British Columbia by Karimian (2006). The objective of that research was to investigate the

137

lateral and axial pipe-soil interaction of relatively large diameter rigid steel pipelines (324 and

138

457 mm). Special focus was given to study the effect on the interaction behavior of trenching

139

and geotextile lining of the trench.

140 141 142 143 144 145 146

6

147

EXPERIMENTAL PROGRAM

148

Nine full scale tests were conducted on small-diameter tubing of 105 mm (4.1 inch) outside

149

diameter size. These tubes (referred to as pipes throughout the remainder of the article) are

150

specified in accordance with the American Society for Testing and Materials specifications A513

151

(ASTM, 2008) for electric-resistance-welded carbon and alloy steel mechanical tubing. Each of

152

the test pipes were buried in dense Olivine sand, as shown in Table 1. Burial depths with

153

embedment ratios, defined as the ratio of soil cover height (measured from spring line of the pipe

154

to soil surface) to nominal pipe diameter (H/D), ranging from 3 to7, were examined. This range

155

covers shallow to deep pipe burial situations as examined in anchor plates by Rowe and Davis

156

(1982). Boundary limits of pipe position within the test pit were examined. This included two

157

distances to the rear wall from the back of the pipe, of 3D and 9.5D. The internal system friction

158

resulting from the interaction between the pulling cables and soil was examined by comparing

159

tests with pulling cables enclosed in hollow PVC tubes to ones with cables exposed to the soil.

160

Also, reproducibility of test data was examined by repeating some tests. As the key section of a

161

buried pipeline subjected to lateral earth movements is between points of inflection (zero-

162

moment points), this can be modelled in tests by pulling that section of the pipe at its ends, while

163

allowing them to rotate freely (Fig. 1). This analogy would not be applicable in situations where

164

non-uniform deformations develop in the soil moving past the pipe. For example, a parabolic soil

165

deformation distribution past the pipe would increase the loads at the pipe mid-section rather

166

than the pipe ends. However, in many situations, the movement of the soil relative to the buried

167

pipe is uniform along the pipe (where, for example, an intact block of soil is moving past the

168

pipe). In such situations, the testing approach is appropriate between the points of inflection

169

along the pipe.

170 7

171

Testing Facility and Setup

172

The test pit used for the current study has plan dimensions of 4000 x 2000 mm (157.5x74.7

173

inch), and is 2000 mm (74.7 inch) deep, with concrete walls (Fig.2). The pit was divided into

174

three main sections using retaining walls. The main (middle) section (section II in Fig. 2(a)),

175

with plan measurements of 2000 x 3010 mm (78.7x118.5 inch), was dedicated to the physical

176

testing of the pipe. The west end of the pit accommodated the assembly used to pull the pipe

177

laterally. It consists of a hydraulic actuator, a spreader beam, and a pair of turn buckles. The

178

actuator was situated outside the test pit, with the actuator piston extending to the spreader beam

179

through an access opening (Fig. 2(a)). On the east end of the pit a narrow, 406 mm (16 inch)

180

wide space (section III in Fig. 2(a)) running across the width of the pit was created to

181

accommodate the five string potentiometers, referred to as “string pots”, used to monitor the pipe

182

deflections during testing. The wall between Sections II and III needed to withstand passive

183

lateral soil pressure and was built using 50 x 100 mm (2x4 inch) lumber ribs bolted to the

184

concrete walls of the test pit, and then covered by 9.5 mm (3/8 inch) thick plywood sheets. On

185

the west end of section II, the retaining wall was designed to withstand the full active soil

186

loading during testing. It was built using a stack of 140 mm x 140 mm (5x5 inch) pressure

187

treated timbers. The timbers were cut to the total width of the test pit (2000 mm). A total of 14

188

pieces were used to erect a 1960 mm (77.2 inch) high wall. To transfer the reactions of this wall

189

to the concrete wall at the west end of the pit, both ends of each timber piece was supported by a

190

short timber piece of the same size, bearing directly against the concrete end-wall (Fig. 2(a)).

191

Numerical calculations for the expected pulling force at the deepest burial depth indicated that

192

negligible (0.35 mm) deflections develop at the mid-span position of the 2000 mm long wooden

193

pieces, and hence the end-wall was assumed to be rigid.

194 8

195

Side Wall Friction Treatment: Tognon et al. (1999) evaluated the effect of sidewall friction for a

196

buried pipe testing facility and showed that friction treatment on the side walls of a testing

197

chamber can significantly reduce the effect of friction between the soil and the side boundaries.

198

In the literature, several configurations have been used. Trautmann and O’Rourke (1983)

199

implemented friction treatment for their testing program by using glass on one side and Formica

200

on the other side to reduce friction. Konuk et al. (1999) used plywood lining for the sidewalls of

201

a test box. Karimian (2006) used stainless steel sides and demonstrated analytically that by

202

replacing the plywood side wall lining by stainless steel, the shearing resistance of the sidewalls

203

would reduce to half of its value. In the current study, the side wall friction treatment of Tognon

204

et al. (1999) was employed. This consists of two layers of polyethylene sheets, each 0.1 mm

205

thick, with the first layer fixed to the wall at the top edge using lumber strips (Fig. 2(c)). A thin

206

film, about 1 mm thick, of silicone-based bearing grease was then spread with a paint roller to

207

act as the lubricant. The second layer of polyethylene sheet was then rolled over the grease film,

208

which had sufficient cohesion to hold that second layer in place without any other physical

209

attachment. This arrangement provides a low friction angle between the sand and the side walls

210

(Tognon et al., 1999, report that the angle of friction is as low as 5°). Both Rowe and Davis

211

(1982) and Trautmann and O’Rourke (1983) examined the effect of aspect ratio, which is in this

212

case the pipe length-to-diameter (L/D) ratio, on the measured lateral forces, accounting for the

213

effect of boundary conditions. Trautmann and O’Rourke (1983) concluded that “a length-to-

214

diameter ratio in the range of five to ten will decrease end effects to relatively small levels”. In

215

the current study, the pipes have even higher L/D ratio of 18.

216

Soil Placement and Removal: Sand was poured into the test pit in several lifts, each about 163

217

mm (6.4 inch) thick, by dumping from a height of 1.0 m using a hopper with side chute. Each

218

pour was distributed evenly using shovels and levels, and then compacted using a hand tamper 9

219

with plate dimension of 254x254 mm (10x10 inch), and weighing 7.9 kg, by dropping it from a

220

constant height of about 300 mm (11.8 inch) to ensure consistent packing effort. The soil level

221

was first brought up to a height of 500 mm (19.7 inch) before placing the pipe to minimize the

222

effect of the bottom concrete boundary. The pipe, 1830 mm (6 feet) long, was then centered

223

between the side walls of the pit leaving a side clearance of about 85 mm (about 0.85D) on each

224

side of the pipe. After each test, sand was removed by means of hand shovels and placed in sand

225

bags, which were craned out of the pit.

226

Loading Mechanism: Some of the previous full-scale testing programs (Paulin et al. 1998,

227

Konuk et al. 1999, and Karimian 2006) used two independent loading points to pull the pipe, one

228

at each end. Other investigators (Trautmann and O'Rourke 1983, and Hsu 1993) used a single

229

loading point. In the present study, a single pulling point was employed. The actuator was

230

situated outside the test pit, and positioned on the centreline of the pit. The piston was extended

231

into the testing pit and connected to the middle of a stiff steel ‘spreader beam’ fabricated from

232

HSS (hollow square section) with cross-section 102x102x6.3 mm (5x5x3/8 inch). The beam was

233

connected to 6.3 mm (1/4 inch) diameter steel cables with a rated axial load capacity of 28 kN,

234

one on each end, to pull the pipe. Each end of the spreader beam rested on a 140x200 mm

235

(5.5x7.9 inch) low-friction, lubricated Teflon strips to minimize friction as the pipe is being

236

pulled. The Teflon strips were attached to the short timber pieces used to transfer retaining wall

237

loads to the test pit wall (Fig. 2(d)). Similar Teflon strips were fixed to the side walls near the

238

ends of spreader beam to minimize friction in case the beam rotated and touched the sidewall

239

during pulling. Turnbuckles were used to connect the spreader beam to the cables through slots

240

made into the timber retaining wall (Fig. 2(e)) and sealed using rubber plugs to prevent soil

241

leakage. The turnbuckles provided the ability to adjust length and remove any cable slack. The

242

cables were attached to the pipe ends by looping each steel cable around the pipe at a distance 30 10

243

mm (1.2 inch) from the pipe end (Fig. 2a). The free end of the cable was then crimped to the

244

main cable using a number of cable clips (saddles) with the number of saddles ranged from 1 to 3

245

saddles per cable end throughout the testing program. The cables (and the accompanying

246

saddles) then extended through the soil mass, using one of two configurations (Fig. 2(a, and b)).

247

In the first configuration, the cables were left exposed to the surrounding soil. In the second

248

configuration, the cables were encased in 89 mm (3.5 inch) diameter PVC tubes fixed to the

249

retaining wall, to minimize the friction between the cables-saddles system and the soil. The PVC

250

tubes were centered around the cables at their connection points at the front wall, but were

251

gradually bent upward (using soil support to maximize clearance above the cable as the PVC

252

tubes approach the test pipe). The ends of the PVC tubes near the attachments to the steel pipe

253

were elevated upwards by almost half their diameter, to provide maximum clearance above the

254

cables close to the steel pipe, to accommodate the upward movements of the pipe during testing.

255

The effect of the friction between the pulling cables and surrounding soil has been

256

addressed and/or mitigated in previous work. Trautmann and O’Rourke (1983) enclosed threaded

257

rods, used to pull the pipe buried through the sand, in PVC tubes to eliminate soil friction.

258

Karimian (2006) employed approaches recommended by the ASCE (1984), and the PRCI (2004)

259

to calculate the axial load in the cables arising from friction and used those results to modify

260

measured pulling force. In the present study, a more quantitative assessment was made, and an

261

approach to account for system friction is presented and discussed in a subsequent subsection.

262

To achieve this goal, auxiliary tests were conducted. A pulling test was carried out on just the

263

cables (i.e. without the test pipe) buried under the soil at a depth (cover) of 304 mm (12 inch),

264

representing the case of H/D = 3. Another test without the pipe was performed after enclosing

265

the cables in the 89 mm (3.5 inch) PVC tubes.

266 11

267

Test Materials

268

Backfill: Synthetic Olivine sand was used as the backfill material for all the experiments. This

269

sand can be described as medium-to-fine sand with generally angular soil particles. Synthetic

270

Olivine sand has the ability to reach a dense state with little compaction. Grain size analysis

271

indicates an average particle size (D50) of 0.67 mm (.026 inch), minimum particle size of 0.075

272

(2.9x10-3 inch) mm, and a coefficient of uniformity (Cu) of 1.84. Triaxial tests were performed to

273

assess the soil parameters, namely friction angle (), dilation angle (), and modulus of elasticity

274

(Es). The tests were performed at confining pressures of 18.3, 33, and 44 kPa. While the sand

275

was subjected to much lower initial confining stresses during pipe testing (0.8 to 3 kPa), it is

276

difficult to conduct triaxial tests at such low confining stresses, and friction angle was

277

extrapolated from the tests conducted at these higher pressures. Dry density and moisture content

278

of the sand after each lift during backfilling were checked using a nuclear densometer. Soil

279

parameters are summarized in Table 2.

280

To obtain a more accurate representation for the soil modulus of elasticity at low confining

281

stresses, the Janbu model (Janbu 1963) was used to represent the soil modulus distribution. The

282

model describes elastic soil modulus (Es) as a function of the confining stresses using the

283

following relation:

284

𝐸𝑠 𝑃𝑜

𝜎3 𝑛

=K( ) 𝑃𝑜

[1]

285

where Po is the atmospheric pressure at sea level (= 101.3 kPa); the Janbu model parameters n

286

and K are constants specific to each material; and 3 is the minimum effective principal stress.

287

Parameters n and K represent the power law parameters in the expression. They have been

288

determined for Olivine sand by plotting test data as (3/Po) versus (Es/Po) for the three triaxial

289

tests, which were then fitted with a linear relation on a log-log scale (Fig. 4). In this way, values

290

for K and n were found to be 326 and 0.86, respectively. 12

291

Steel Pipe: The steel pipe used was manufactured according to the American Society for Testing

292

and Materials Standard A513 (ASTM, 2008). Pipe dimensions and strength parameters are

293

provided in Table 3.

13

294

Instrumentation

295

Electric resistance strain gages and string pots were attached to the pipe to monitor bending

296

strains and deflections, respectively. String pot boxes were secured to the partitioning wall

297

separating sections II and III of the pit (Fig. 2). The strings were then passed through small holes

298

in the wall, into the soil, and enclosed in 50 mm (2 inch) diameter PVC tubes fixed to the rear

299

wall to avoid contact with the soil. Thin steel wire extensions were looped around and secured to

300

the test pipe and were connected to the hooks of the string pots. Displacements were measured

301

at five equidistant locations along the pipe (Fig. 3).

302

For the strain gages, an array of uniaxial strain gages was fixed to the front (passive), and

303

rear (active) spring lines of the pipe (Fig. 3). Results at the mid-span position are reported in a

304

subsequent section. A triaxial strain gage was also attached to the top (crown) of the pipe at one

305

diameter distance from the south end of the pipe. Extra slack was left in the strain gages wires by

306

embedding them with waves within the soil to allow for pipe movement.

307

To measure the pipe pulling load, a load cell (MTS System Corporation, Model 661.20A-

308

03) was used featuring an accuracy of ± 1 N. The load cell was connected between the spreader

309

beam and the actuator piston (Fig. 2). During testing, a data acquisition system (Vishay

310

Measurement Group-System 5000) was employed to collect force, strain and displacement data.

311

To check for pipe elevation and soil heave before and after testing, a laser level was used. Sand

312

density and water content were checked before each test using a nuclear densometer (by CPN

313

international Inc., model CPN MC-1DR-P) at the center of each of four quadrants of the soil for

314

each lift, as shown in the plan view, Fig 2(a).

315 316 317 14

TEST RESULTS AND DISCUSSION

318 319

Effect of Pulling System Friction

320

The pulling force on the cables buried under a soil cover of 304 mm, in the auxiliary test without

321

the pipe, was found to be 2.1 kN. This is of significant size given that it represents about 20% of

322

the total pulling force when it is attached to the pipe. In the other auxiliary test, when the cables

323

were enclosed in the 89 mm PVC tubes, a significant reduction of pulling force to 0.3 kN was

324

measured. These values are useful indicators of the expected change of force associated with

325

friction, but as will be seen in the following section, this force change varied from case to case

326

and the next section describes a unified adjustment technique developed for use with each

327

individual test.

328

Fig. 5 shows the actuator load-stroke curves for pipe specimen S3 (Table 1), which was

329

conducted with exposed cables, and the same pipe specimen for Test S4a, but with the cables

330

enclosed in PVC tubes. It can be seen that the difference in peak loads is consistent with the

331

difference between loads measured in the two auxiliary tests, 1.8 kN. It can also be seen that the

332

measured loads at any given stroke are different from the onset of loading, with the difference,

333

which represents friction between cable and soil, building up gradually to the 1.8 kN difference

334

at peak loads. In Test S3, the load measured at the instant of first sensing of load by the pipe,

335

indicated by first excitation of strain gage SG1 (Fig. 3) located at the pipe end, was about 1.145

336

kN, at an actuator displacement of 1.4 mm. On the other hand, the equivalent values in test S4a

337

were 0.23 kN and 0.2 mm, respectively. Table 4 summarizes the actuator load values recorded at

338

initial excitation of the outermost and middle strain gages SG1, SG4 and SG5 (Fig. 3). Based on

339

SG1, the point of initial load transfer to the pipe, the loads in all cases with enclosed cables

340

varied from 0.05 kN to 1.12 kN. Review of the actuator data retrieved proved that actuator stroke

341

is not relevant as it is sensitive to initial slack within the pulling cables before the start of the test. 15

342

These results have not therefore been included in this table. Data provided in Table 4 suggests

343

that variable, and unpredictable, amounts of load are needed to overcome friction. As such, an

344

approach for adjusting the actuator load-stroke curve was adopted for each test. First, the point

345

on the actuator load-stroke curve at which load was first transferred to pipe is established based

346

on first excitation of strain gage SG1. The actuator pulling force reading and stroke at this point

347

is then set as the new origin for the “Pipe Pulling Load” and “Pipe End Displacement”,

348

respectively. It is assumed that beyond this point, any further pulling would be transferred to the

349

pipe directly, except for the load used to overcome soil interaction with the short exposed portion

350

of cable. Furthermore, most of the slack in the cables would have been removed by then, except

351

for the small elongation due to the tensile loads applied during the remainder of the test. All

352

pulling loads and pipe displacements reported hereafter represent this approach of adjustment.

353 354

Effect of Distance to Rear Test Cell Boundary

355

Tests S1 and S2, both using exposed pulling cables, were compared to examine the effect of

356

front and rear boundary limits on the load-pipe end displacement curve (Fig. 6). Configurations

357

for both tests were the same, except for the position of the pipe relative to the rear and front end

358

walls. For test S1, the length of soil behind the pipe, measured from the centerline of the pipe to

359

the rear wall, was 305 mm (3D), and the length of soil in front of the pipe 2705 mm (26.6D). In

360

test S2, there was 965 mm (9.5D) of soil behind the pipe, and 2045 mm (20.1D) in front. Fig. 6

361

shows a difference of just 4% in the peak load. Also, the extent of soil surface deformation seen

362

behind the pipe after the test was concluded did not exceed distances of 105 mm and 94 mm

363

(equivalent to 0.92D and 1.02D) for tests S1 and S2, respectively. This indicates that in

364

situations where tight space is of a concern, a three pipe diameters to the rear boundary limit

365

should be sufficient. For the side boundary limits, a distance of 85 mm (0.85D) was left between 16

366

each end of the pipe and the side wall. While no physical testing was performed to assess the

367

effect of those side boundary limits, numerical calculations not included here show that side

368

clearance as low as 0.35D has a minimal effect on the load-deflection curvature of the pipe.

369 370

Test Repeatability

371

To assess the reproducibility of test results under the same burial conditions, tests S4a, S4b, S4c,

372

and S4d with embedment ratio of H/D = 3 were replicates. The only minor difference between

373

the tests is the number of clips used to grip the cable when looped around the pipe end, or when

374

threaded through the turnbuckle at the other end. Tests S4a and S4b used one cable clip for each

375

cable end, while tests S4c, and S4d used two clips instead after minor slippage (2 mm) of the

376

cable was noticed after test S4b. Fig. 7(a) shows the load-displacement curves, which depict

377

almost identical behavior in the elastic range, but with some variability in the peak load, which

378

ranged from 9.3 kN for test S4c to 10.4 kN for test S4a (mean of 9.92 kN and standard deviation

379

of 0.11 kN). Fig. 7(b) shows the pipe bending responses (i.e. mid-span deflection of the pipe

380

relative to a reference line connecting both ends). Some variability can be seen at the initial parts

381

of the curves. The maximum deflection varied from 2.5 mm in test S4b to 3.5 mm in test S4a.

382 383

Effect of Burial Depth

384

In the following section, the effect of embedment ratios of H/D = 3, 5 and 7 on several aspects of

385

the pipe-soil interaction are examined. Test S4a (Table 1) is selected to represent the burial depth

386

for H/D = 3.

17

387

Load-Displacement Responses: Fig. 8(a) and (b) show the load-displacement curves for the

388

three burial depths, for both pipe end, and pipe mid-span displacements, respectively. Clearly,

389

the peak pulling force increases with embedment ratio (10.4, 18.5, and 30.4 kN, for embedment

390

ratios 3, 5, and 7, respectively). It is important to note that failure in all cases was governed by

391

the soil and not the pipe as will be explained from the load-strain responses of the pipes at mid-

392

span. Fig. 8(c) shows the load-pipe bending deflection responses at mid-span, . The responses

393

are generally linear elastic and having similar slopes. Only the peak load and deflection increase

394

with increasing burial depth. Also, the three curves show some nonlinearity (concave down), i.e.,

395

there is less incremental load needed for the same incremental deflection of the pipe, even

396

though the pipe material is believed to still be behaving elastically (based on the measured

397

strains). This nonlinearity is not usually observed when conventional ‘in-air’ beam bending tests

398

are used (beam-bending experiments on the pipe featuring simple supports and traditional

399

characterization methods such as 3-point and 4-point bending, or uniformly distributed loads).

400

This difference between buried pipe response and ‘in-air’ response is consistent with earlier

401

studies that found that pipe bending response when buried in soil is different to simple beam-

402

bending behavior (e.g. Konuk et al., 1999; and Mahdavi, 2008).

403

Load-Strain Responses: Fig. 8(d) compares the load-bending strain responses at mid-span of the

404

pipes tested for embedment ratios 3 and 5 (difficulties were experienced with the strain gages for

405

the test at embedment ratio of 7, so the comparison is not available for that case). Strain

406

development with pipe loading for both tests are somewhat similar, except for the small shift in

407

initial load for the test with H/D = 5. It can be seen that beyond the peak load as the load drops,

408

the strain reduces suggesting that the curvature of the tube and hence its net deflection reduces.

409

The maximum axial strain reached at the mid-span of the pipe was for Test S5 to be 1100 micro

410

strain, which is far less than the yield strain of the pipe. It is important to note that these pipes 18

411

were not tested with internal fluid pressure, which would induce circumferential tensile strains

412

placing the pipe in a state of bi-axial stresses. Also, a number of geometrical factors are expected

413

to influence the longitudinal bending behavior of the pipe, such as its length to diameter ratio,

414

and its diameter to thickness ratio (the experimental results presented here are for one particular

415

choice of pipe diameter, length, and wall thickness). Numerical calculations can be employed to

416

investigate the effect of these parameters once the test data are used to establish the effectiveness

417

of the computer modeling (Fatemi et al., 2008, provide an example of this use of computer

418

analysis to establish the influence of pipe parameters).

419 420

Fig. 9 shows the development of pipe displacement (by measuring string pot extensions) at

421

different sections along the pipe versus pipe loading for embedment ratio H/D = 3 (Test S4a).

422

This specific test illustrates a situation where the pipe experiences slight rotation during pull. In

423

the ideal loading case, the extension of string pot pairs at pipe ends (curves for 1, and 5) should

424

be identical, as should displacements at the quarter positions (curves for 2, and 4). In this test,

425

the North end of the pipe advanced ahead of the South end (ahead by 7.4 mm at peak pipe

426

loading). Considering this pipe tilting or pipe rotation together with the bending deflection of the

427

pipe itself of about 3.5 mm, yields a total oblique orientation angle between the pipe span

428

relative to the pipe pulling direction of about 0.3 (from 90o). At post-peak loading, this value

429

increased to a maximum of 1.2. Phillips et al. (2004), and Daiyan et al. (2010) examined the

430

effect of combined axial and lateral pipe-soil interaction and developed interaction graphs for

431

combined loading with the calculated and measured horizontal and axial bearing capacity factors

432

(Nqh, and Nt, respectively) for the soil. One of the main outcomes from both studies is that

433

coupled pipe-soil interaction due to oblique loading is of minimal effect on the lateral soil

434

restraint on the pipeline at such low rotation angles. 19

435

Soil Surface Deformation: profile limits of the soil deformed surface were recorded after each

436

test. Typical features of deformed soil surface after lateral pipe pull reported in previous

437

experimental work (e.g. Rowe and Davis 1982, Trautmann and O’Rourke 1983, and Karimian

438

2006) were observed in the current study. The main characteristics are formation of a passive,

439

heaving, wedge in front of the pipe undergoing general shear failure, and an active, sinking,

440

wedge behind the pipe. Degree of surface deformation varied as pipe burial depth changed. Fig.

441

10 illustrates the soil surface deformation for the different burial depths tested. It can be seen that

442

for embedment ratio of H/D=3, soil dip behind the soil and the soil heave in front of the pipe is

443

very distinctive and of order of magnitude comparable to pipe diameter size, D, (1.1D below,

444

and 0.4D above initial soil surface, respectively). For embedment ratio of H/D=5, the soil

445

dipping behind the pipe is 0.9D, whereas soil heave in front of the pipe reduces considerably to

446

just 0.1D. For the deeper burial depth of H/D=7, soil surface movement was barely detectable

447

(0.1D, and 0.1D for soil dip and soil heaving, respectively). Such minimal soil surface

448

disturbance for embedment ratio of H/D=7 suggests that the failure mechanism here is close to

449

the deep, punching, failure mechanism observed by Audibert and Nyman (1977) with H/D

450

values ranging from 12-24, Rowe and Davis (1982) for deeply buried laterally loaded anchor

451

plates (with H/D values >3), and Trautmann and O’Rourke (1983) for deeply buried pipes with

452

H/D values ranging from 8-11.5. Ideally in this type of failure, only local failure of soil around

453

pipe takes place with no change occurring at the soil surface.

454

As for the pipe translation within the soil mass, Fig. 11 shows pipe initial and final positions at

455

pipe mid-span for the same tests under examination, which depicts a similar trend of change as

456

the one observed with soil surface heave in Fig. 10. At burial depth ratio of H/D=3, the pipe

457

elevated by 37 mm after a horizontal displacement of 102 mm, while for a similar pipe

458

horizontal displacement of 98 mm at burial depth ratio of H/D=5, the pipe elevated 17 mm only. 20

459

When the pipe was pulled for burial depth ratio of H/D=7 (which required longer horizontal

460

travel of 154 mm to reach peak loading), pipe also moved up by 17mm. Such upward movement

461

is expected since the pipe travels in the direction of least resistance. Since soil strength and

462

stiffness increases with depth (it is a function of the earth pressures), as the pipe moves forward,

463

it is easier to form a path which has an upward component, taking it into the weaker, less stiff

464

soil.

465

Other Pipe results: progress of the deflected shape of the pipe for all the tests was monitored and

466

recorded by plotting the instantaneous string pot readings along the pipe. Fig. 12 shows deflected

467

shape of the pipe at 50%, 75%, and 100% of peak pulling load, Pu, for pipe burial depth H/D=3

468

(Test S4a). The figure also illustrates the amount of tilt (exaggerated due to scale effect)

469

experienced by the pipe as discussed earlier.

470 471

Conclusion

472

A series of full-scale tests on high strength, steel pipes buried in dense sand under three burial

473

depth-to-diameter (H/D) ratios of 3, 5 and 7 has been conducted to study the pipe-soil

474

interaction, with focus on the flexural behaviour of the pipe under lateral earth loading with

475

idealizing earth lateral loading conditions on the pipe by pulling the pipe at its end instead. The

476

testing facility and experimental setup were discussed and demonstrated to meet the intended

477

purposes of the current study program with adequate accuracy. Other conclusions can also be

478

drawn from analyzing the test results:

479

a) Consistency of test setup configurations (geometrical setup, and material preparation)

480

was assessed with replicate tests, and four tests produced maximum pulling forces that

481

varied by less than 6%.

482

b) Friction within the pipe pulling system could lead to unpredictable, and potentially 21

483

significant (up to 20% of total pulling force), amount of load consumed in overcoming

484

cable-soil interaction. Mitigation of such factor is presented and discussed.

485

c) In similar testing configurations where tight space is a concern, a distance of three pipe

486

diameters to the rear boundary of the test box should be sufficient. Reduction to just three

487

diameters changed the measurement of pulling force by about 4%, which is less than the

488

amount seen in the replicate tests.

489

d) Bending deflection of the pipe is not linear with pulling force, even within the elastic

490

range of the steel material tested, which gives more significance to the study of bending

491

behavior of buried pipes in situ conditions than studying it at “in-air” loading

492

configurations.

493

e) By characterizing the load-bending deflection of flexible buried pipelines and

494

establishing bending tolerance estimations, better decisions about excavating,

495

straightening, and reburying pipes could be then conducted. The current two-dimensional

496

approach does not provide direct evidence of pipeline tolerance to these deflections).

497 498

Notation

499

The following symbols and abbreviations are used in this paper:

500

Cu

= coefficient of conformity of soil

501

D

= nominal pipe diameter

502

Do

= outside pipe diameter

503

D50

= average particle size of soil grains

504

Es

= soil modulus of elasticity

505

H

= burial depth to pipe spring line

506

K

= Janbu modulus multiplier at stress of 1 atmosphere (dimensionless)

507

L

= pipe length

508

n

= Janbu model index controlling level of stress dependence (dimensionless) 22

509

Nqh

= horizontal bearing capacity factor for sand

510

Pu

= peak soil resistance load to lateral pipe movement

511

Po

=atmospheric pressure at sea level

512

Pa

=actuator load

513

t

= pipe wall thickness

514



= bending deflection at mid span of the pipe

515



= sectional displacement of pipe

516

a

= actuator stroke

517



= friction angle of sand

518



= bulk unit weight of soil

519



= Poisson ratio

520



= minimum effective principal stress

521



= dilation angle of sand

522 523

References

524 525

ASTM A513. (2008).” Standard Specification for Electric-Resistance-Welded Carbon and Alloy Steel Mechanical Tubing.” American Society for Testing and Materials specification.

526 527 528 529

American Society of Civil Engineers. (1984). “Guidelines for the seismic design of oil and gas pipeline systems.” Committee on Gas and Liquid Fuel Lifelines, Technical Council on Lifeline Earthquake Engineering, ASCE, New York.

530 531 532 533

Audibert, J.M.E. and Nyman, K.J. (1977). “Soil Restraint against Horizontal Motion of Pipes”, Journal of the Geotechnical Division, Proceedings of the American Society of Civil Engineers, Vol. 103, No. GT10, pp. 1119-1142

534 535 536 537

Da, H., Abdoun, T. H., O'Rourke, M. J., Symans, M. D., O'Rourke, T. D., Palmer, M. C., & Stewart, H. E. (2008). “Centrifuge Modeling of Earthquake Effects on Buried High-Density Polyethylene (HDPE) Pipelines Crossing Fault Zones.” Journal of Geotechnical and Geoenvironmental Engineering, 134(10), 1501-1515.

538 539

Daiyan, N., Kenny, S., Phillips, R., and Popescu, R. (2010). “Numerical Investigation of Oblique Pipeline/soil Interaction in Sand.” Proceedings, International Pipeline Conference, Calgary.

540 541 542

Das, B. M., Seeley, G. R. (1975). “Load displacement relationship for vertical anchor plates.” Journal of Geotechnical Engineering Division, ASCE, 101(GT7), 711-715. 23

543 544

EGIG. (2008). “7th Report of the European Gas Pipeline Incident Data Group.” (http://www.egig.nl/downloads/7th_report_EGIG.pdf).

545 546 547 548

Fatemi, A., Kenny, S., Taheri, F., Duan, D., and Zhou, J. (2010). “End Boundary Effects on Local Buckling Response of High Strength Linepipe.” Proceedings, International Pipeline Conference, Calgary.

549 550 551

Hsu, T. W. (1993). “Rate effect on lateral soil restraint of pipeline.” Soils and Foundations, 33(4), 159-169.

552 553 554 555

Hsu, T.W., Chen, Y.J., and Hung, W.C. (2006). “Soil Friction Restraint of Oblique Pipelines in Dense Sand”, J. Transp. Eng. 132(2), pp. 175-181.

556 557

Hsu, T.W., Chen, Y.J., and Wu, C.Y. (2001). “Soil Friction Restraint of Oblique Pipelines in Loose Sand.”, J. Transp. Eng. 127(1), pp. 82-87.

558 559 560 561

Hurley, S., Zhu, F., Phillips, R. and Paulin, M. J. (1998a). “Large scale modeling of soil/pipe interaction under moment loading,” Test GSC 01 data report. Contract Report for Terrain Sciences Division. Geological Survey of Canada, C-CORE Publication 98 C-20.

562 563 564 565

Hurley, S., Zhu, F., Phillips, R. and Paulin, M. J. (1998b). “Large scale modeling of soil/pipe interaction under moment loading,” Test GSC 02 data report. Contract Report for Terrain Sciences Division. Geological Survey of Canada, C-CORE Publication 98 C-21.

566 567 568 569

Janbu, N. (1963). “Soil Compressibility as Determined by Odometer and Triaxial Test.” Proceedings of the European Conference on Soil Mechanics and Foundation Engineering, Vol. 1, pp 19-25, Weisbaden.

570 571 572 573

Karimian, S.A. (2006). “Response of Buried Steel Pipelines Subjected to Longitudinal and Transverse Ground Movement.”, PhD Thesis, University of British Columbia, Vancouver, Canada.

574 575 576

Konuk, I., Phillips, R., Hurley, S., and Paulin, M. J. (1999). “Preliminary Ovalisation of Buried Pipeline Subjected to Lateral Loading.” OMAE99/PIPE,-5023.

577 578 579 580

Mahdavi, H, Kenny, S., Phillips, R., and Popescu, R. (2008). “Influence of Geotechnical Loads on Local Buckling Behavior of Buried Pipelines.” Proceedings, International Pipeline Conference, Calgary. 24

581 582

National Energy Board. (2008, July). Focus on safety: A comparative analysis of pipeline safety performance 2000-2006.

583 584 585

Nyman, K.J. (1984), “Soil response against oblique motion of pipes”, Journal of Transportation Engineering, ASCE, Vol. 110, No. 2, pp190-202.

586 587 588 589

O’Rourke, M., Gadicherla, V., and Abdoun, T., (2005). “Centrifuge Modeling of PGD Response of Buried Pipe.” Earthquake Engineering and Engineering Vibration, Vol. 4, No. 1, pp. 6973

590 591 592 593

Paulin, M.J., Phillips, R., and Boivin, R. (1995). “Centrifuge modeling of lateral pipeline/soil interaction-Phase II.” OMAE 95, Proceedings of the 14th International Conference on Offshore Mechanics and Arctic Engineering, Copenhagen, Denmark, Volume 5, pp.107-123

594 595 596 597

Paulin, M.J., Phillips, R., Clark, J.I., Trigg, A., and Konuk, I. (1998). “A Full-scale Investigation into Pipeline/Soil Interaction.” Proceedings, International Pipeline Conference, Calgary, AB, Canada

598 599 600

Phillips, R., Nobahar, A., and Zhou, 1. (2004) "Combined Axial and Lateral Pipe-Soil Interaction Relationships." Proceedings, International Pipeline Conference, Calgary, AB, Canada.

601 602 603 604

PHMSA (2010). “Significant Pipeline Incidents by Cause.” (http://primis.phmsa.dot.gov/comm/reports/safety/SigPSIDet_1991_2010_US.html?nocache =1111).

605 606 607 608

Pipeline Research Council International, PRCI (2004) "Guidelines for the Seismic Design and Assessment of Natural Gas and Liquid Hydrocarbon Pipelines." Pipeline Research Council International, Inc. (PRCI), Catalogue No. L51927.

609 610 611

Rowe, R. K., and Davis, E.H. (1982). “The behaviour of anchor plates in sand.” Geotechnique, 32(1), 25-41.

612 613 614

Tognon, A.R., Rowe, R.K., and Brachman, R.W.I. (1999). “Evaluation of Sidewall Friction for a Buried Pipe Testing Facility.” Geotextiles and Geomembranes, 17: 193–212.

615 616

Trautmann, C.H., and O’Rourke, T.D. (1983). “Behaviour of pipe in dry sand under lateral and uplift loading”. Geotechnical Engineering Report, Cornell University, Ithaca, N.Y., 83–7.

617

25

618

619 620 621 622 623 624 625 626

627 628 629

631 632

Table 1. Test Matrix Distance to rear Test I.D. H/D wall Boundary** S1 3.0 D 3 S2 9.5 D 3 S3 9.5 D 3 S4a 9.5 D 3 S4b 9.5 D 3 S4c 9.5 D 3 S4d 9.5 D 3 S5 9.5 D 5 S6 9.5 D 7 * BL SF Rep BD

Pulling Cables Confinement Exposed* Exposed* Exposed Enclosed Enclosed Enclosed Enclosed. Enclosed Enclosed

Parameters Examined BL BL BD

SF SF

Rep Rep Rep Rep

BD BD

Uninstrumented Pipe ** Distance is in multiples of pipe diameter (D) Boundary Limit “System Friction” due to pulling cables confinement arrangement, by being either Exp. or Enc. Repeatability check test Burial Depth

Table 2. Soil Parameters Parameter Modulus of Elasticity (E), MPa Poisson Ratio () Bulk unit Weight (), kN/m3 Friction Angle () (Peak/Residual) Angle of Dilation ()

Value 16 0.3 15.1 53°/45° 16°

Table 3. Steel Pipe Parameters

Parameter Nominal Diameter (D), mm Outside Diameter (Do), mm Length (L), mm D/t Ratio L/D Ratio Wall Thickness (t), mm

630 Value 102 (4 in.) 105 (4.1 in.) 1829 (6 ft.) 37 18 2.7 (0.106 in)*

Modulus of Elasticity (GPa)

200

Yield Strength (MPa)

379

Ultimate Strength (MPa)

427

0.28 Poisson Ratio () Material Specifications ASTM A513 * measured (different from the 2.1mm (0.083 in.) value from ASTM A513 and the supplier) 26

633 634

635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672

Table 4. Load Cell Reading (kN) at 1st Excitation of Strain Gages 3 H/D Test I.D # S3 S4a S4b S4c S4d

5 S5

7 S6

SG1

1.145

0.232

1.074

0.345

0.052

0.446

1.117

SG5

1.471

0.232

1.373

0.534

0.052

0.476

1.520

SG4 (Mid-span) 1.264 0.485 2.191 0.736 0.723 Note: Shaded areas identify tests with enclosed pulling cables

0.446

2.130

27

673

Soil Loading

674

Inflection Points (Zero Moment)

Tested Portion 675 676 677 678 679 680 681 682 683 684 685

Fig. 1. Plan View of Lateral Earth Movement past a Buried Pipeline and the Experimental Test Section between Points of Inflection Being Modeled in this Study

686

Quadrant Quadrant Quadrant Quadrant

N

3 2

6

7

A

A

4

(b)

A

ua dra nt

(e)

4000 mm

3

H/D=7

4

694

696 697 698 699

5

584 mm

3010 mm

4

500 mm

Soil level @ H/D = 3

H

H/D=5

695

(d)

5

1

688 689 690 691 692 693

(c) 2000 mm

( a A)

Section III

Section II

(a) Section I

2000 mm

687

1: Actuator piston 2: Spreader beam 3: Retaining wall (front end wall) 4: Pulling cable (in exposed configuration) 5: Pulling cable with encasing PVC tube (in enclosed configuration) 6: Rear end wall 7: String potentiometer

406 mm

Fig. 2. Experimental Test Setup: a) Plan View, b) Elevation view (section A-A), c) View of Test Pit, d) Friction Treatment for Spreader Beam, and e) Spreader Beam-Turnbuckle Connection

700 701 28

702 703 704

SP5

SP3

SP4

String Potentiometer (SP1)

SP2

705

455 mm

706 707 708

SP Wires

Protecting Tubes

709 710 711 712

SG4R

SG5R

SG-Triaxial SG4F

SG5F

713

456 mm

714 715 716

SG2R Strain Gauge 1-Rear

SG3R

SG2F

SG3F

C.L .

Strain Gauge 1-Front (SG1F) 270 mm 102 mm

Fig. 3. Pipe Instrumentation: Top View Schematic Showing Strain Gauges and Stringpot Locations

717 718 719 720 721 722

Soil Modulus (Es)/ Pa

1000

y = 326.42x0.8603

100

10

Experimental 1

723 724 725

0.1

Confining Stress/ Pa

1

Fig. 4. Determination of Janbu Model Parameters 29

726

Actuator Pulling Load (kN)

14 Exposed Cables Setup (Test S3)

12 10 8

Enclosed Cables Setup (Test S4a)

6 4 Indicates point of 1st sensing of load

2 0 0

727

20 40 60 80 Actuator Stroke, a (mm)

728 729 730

Fig. 5. Effect of System Friction on Load-Displacement Curve Based on Actuator Stroke

731 732 733 734 10 9 8 7 6 5 4 3 2 1 0

Pipe Pulling Load (kN)

H/D =3 d= 9.5D (Test S1) d= 3D (Test S1)

d d

29.5 D 0

735 736 737

20

40

60

80

Pipe End Displacement (mm)

Fig. 6. Effect of Rear Soil Boundary on Load-Displacement Curve

738 739 740 30

741 12

H/D=3

(a)

Pipe Pulling Load (kN)

10 8 6 Test S4a Test S4d Test S4b Test S4c

4 2 0 0

742

12

20 40 60 Pipe End Displacement (mm)

H/D=3

80

(b)

Pipe Pulling Load (kN)

10 8 6 Test Test8S4c Test Test7S4b Test Test9S4d Test6S4a Test

4 2 0 0

743

1 2 3 Pipe Mid-span Deflection,  (mm)

4

744 745 746 747

Fig. 7. Effect of Test Repeatability on (a) Load-Displacement Curves, and (b) Pipe Bending Deflection Curves at Mid-Span

748 749 750 751 752 753 754 755 31

756 757 758 40

30

missing data

H/D =7

25 20

H/D =5

15 10

H/D =3

5 0 0

759

20 40 60 80 100 Pipe End Displacement (mm)

120

35

Pipe Pulling Load (kN)

30

(b) missing data H/D = 7

20

10

H/D = 5 H/D = 3

5 0

760

(c)

30 25 missing data

20 15

H/D=7 (Test S6) Load

10

H/D=5 Load (Test S5)

(kN) (kN)

H/D=3 Load (Test S4a)

5

(kN)

0 -1

1

0 20 40 60 80 Mid-span Stringpot Extension () (mm)

100

3 5 7 9 Deflection,  (mm)

11

20 Tension Compression 18 16 14 H/D = 5 12 10 8 6 H/D = 3 4 2 0 -1200 -800 -400 0 400 800 1200 Micro Strain (at Pipe Mid-span)

(d)

25

15

35

Pipe Pulling Load (kN)

(a)

Pipe Pulling Load (kN)

Pipe Pulling Load (kN)

35

761 762 763 764 765

Fig. 8. Effect of Burial Depth on (a) Load-Displacement Curves at Pipe Ends, (b) LoadDisplacement Curves at Mid-Span String Pot, (c) Load-Bending Deflection Curves at MidSpan, and (d) Load-Pipe Strain at Mid-Span Curves

766 767 768 769 770 771 32

772 773 774 12



Pipe Loading (kN)

10







(South End)

8 6 4 2 0 0

775 776 777

20 40 Stringpot Extension, , (mm)

60

Fig. 9. Sectional Displacements along the Pipe for H/D = 3

778 779 780

Soil Heave (mm)

60 40 20 0 -200 -20 0 -40 -60 -80 -100 -120 -140

781 782 783 784

H/D=3 (Test S4a) H/D=7 200

400

600

800

1000

H/D=5

Horizontal Distance from Initial Pipe Position (mm)

Fig. 10. Soil Surface Heave at Different Burial Depths (at Pipe Centerline) (Line markers indicate measured locations)

785 786 787 788 33

789 790 791

Burial Depth (mm)

-50

100 0 -100 0 -200 -300 -400 -500 -600 -700 -800

50

100

150

200

H/D=3 H/D=5 H/D=7

Horizontal Distance from Initial Pipe Position (mm)

Fig. 11. Initial and Final Pipe Position of Pipe at Different Burial Depths

792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 34

813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836

35