Open access
Technical Papers
Dec 5, 2019

Wave-Induced Liquefaction and Floatation of a Pipeline in a Drum Centrifuge

Publication: Journal of Waterway, Port, Coastal, and Ocean Engineering
Volume 146, Issue 2

Abstract

This paper discusses the floatation of a buried pipe in association with wave-induced liquefaction of sand beds. Centrifuge wave tests in a drum channel were performed with viscous scaling introduced such that the time-scaling laws for fluid wave propagation and consolidation of the soil were matched. The characteristics of liquefaction under irregular waves as well as regular waves were investigated. Under severe wave conditions, the loose sand beds underwent liquefaction, and the liquefied zone propagated downward in the course of wave loading. It was found that the floatation of the buried pipe was a consequence of progressive liquefaction. With the occurrence of liquefaction at a shallow soil depth, the buried pipe started moving upward and the motion of the pipe increased markedly in association with the downward progress of liquefaction during wave loading. Finally, the pipe reached the soil surface. The effects of a gravel layer replacement over a sand bed on wave-induced liquefaction and pipe floatation were also examined. A gravel cover over the sand bed with a range of 120° above the pipe prevented significant vibratory motion of the liquefied soil and floatation of the pipe.

Introduction

Wave-induced liquefaction of seabed deposits or loosely packed backfills in coastal areas has received considerable attention in relation to the stability of submarine pipelines, cables, or other such facilities. Damgaard et al. (2006) illustrated a 1960 pipeline-floatation accident in which pipelines installed in a predredged trench floated to the soil surface owing to seabed liquefaction caused by wave action and tidal motion. Christian et al. (1974) reported that a single 3.05-m-diameter steel pipeline in Lake Ontario floated and failed several times during storms, apparently because of liquefaction. Herbich et al. (1984) reported that a 3.05-m-diameter pipeline under construction was found on the surface after a rather severe storm, confirming that liquefaction had occurred. Sumer (2014) reported an incident in which a 1.4-m-diameter plastic pipe rose to the soil surface following a severe storm about 1 month after the backfill was completed.
Some researchers have performed standard flume tests under 1g conditions and discussed the relationship between soil liquefaction and the floatation/sinking of a pipeline (Sumer et al. 1999; Teh et al. 2003; Sumer et al. 2006a). Stone protection over the backfill soil of a trench was proposed schematically as a countermeasure against pipe floatation (Sumer 2014). The effects of using cover stone over the sand bed on wave-induced liquefaction were investigated using 1g flume tests (Sumer et al. 2010). Experiments using a U-shaped oscillatory flow tunnel investigated the wave-induced instability of a submarine pipeline (Gao et al. 2003). A simple model predicting pipeline–seabed interactions was proposed by simulating the wave-induced liquefaction in the 1g wave flume test (Zhao et al. 2018). However, a fundamental problem remains, namely scaling laws for wave–soil interactions such as wave-induced liquefaction have not yet been established for 1g wave tests. As a result, the soil–pipeline responses observed in these tests cannot be properly extrapolated to field conditions.
Characteristics of the wave-induced liquefaction of sands were discussed using centrifuge wave testing with viscous scaling such that the time-scaling laws for fluid wave propagation and consolidation of the soil were matched (Sassa and Sekiguchi 1999). It was found from the centrifuge wave tests that a loose sand bed subjected to severe wave conditions underwent liquefaction and that the liquefied zone propagated in the course of wave loading. It is necessary to gain a detailed understanding of the relationship between the progressive nature of liquefaction and the stability of submarine structures. Sekiguchi et al. (2000) used a centrifuge to investigate the effects on liquefaction of soil improvement by the placement of a gravel cover. Miyamoto et al. (2018) developed an irregular-wave generation system in a drum centrifuge and applied it to the problem of pipe floatation caused by liquefaction. Recently, the application of centrifuge testing to the problems of fluid–soil-structure interactions as well as wave-induced liquefaction has become increasingly important in the fields of coastal engineering and geotechnical engineering (Sumer 2014; Sassa 2014; Sassa et al. 2016).
This paper discusses the relationship between pipe floatation and the wave-induced liquefaction of sand beds on the basis of centrifuge wave tests. It also discusses the effects of a gravel layer over a sand bed as a countermeasure against wave-induced liquefaction and pipe floatation. Emphasis was placed on describing the process of floatation of buried pipe in association with progressive liquefaction under regular and irregular wave conditions, and on investigating a sufficient range of gravel layer replacement to effectively mitigate pipe floatation and liquefaction.
This paper is organized as follows. The time-scaling law of centrifuge wave testing is described, followed by the criteria for the onset of wave-induced liquefaction, as well as the experimental conditions. Then the soil liquefaction under regular waves and the associated pipe floatation are presented. The pipe behavior in a sand bed subjected to irregular waves, similar to those in a real storm, are discussed based on the centrifuge wave tests. This is followed by a discussion of the effects of a gravel layer as a countermeasure against wave-induced liquefaction and the floatation of a pipeline buried in sand.

Experiment

Time-Scaling Law of Soil Consolidation in Centrifuge

The development and dissipation of residual pore pressure ue in the soil under wave loading may be represented by the following equation (Sassa and Sekiguchi 1999):
ue(ωt)=Kmvμωκ22ue(κz)2+1mvϵvol(ωt)
(1)
where K = intrinsic permeability coefficient; mv = coefficient of compressibility of soil skeleton; μ = dynamic viscosity of pore fluid; ω = angular frequency of waves; κ = wave number 2π/L, where L = wave length; εvol = plastic volumetric strain due to cyclic shearing; z = vertical coordinate; and t = time. To satisfy the time-scaling law of the soil consolidation, the following relation should hold for both the model and the prototype:
Kmvμωκ2=const
(2)
For the wave test of a 1/N scaled model under Ng condition, κm=Nκp and ωm=Nωp. The latter is derived from Froude’s law. The intrinsic permeability coefficient K should be constant if the same soil with the same initial state of packing is used in the model and the prototype. The parameter mv of the model is the same as that of the prototype, because it depends essentially on the effective confining pressure. Therefore, by using pore fluid with viscosity μm=Nμp, Eq. (2) is satisfied. With this technique, called viscous scaling, one can satisfy the time-scaling laws for fluid wave propagation and the consolidation of the soil (Sassa and Sekiguchi 1999).
For the wave test of a 1/N scaled model under 1g condition, the equations κm=Nκp and ωm=N0.5ωp hold. The viscosity of pore fluid usually is μm=μp. To satisfy Eq. (2), under assumptions of essentially the same compressibility mv between the model and the prototype, although the scaled 1g model cannot reproduce the effective confining pressure as in the prototype, Km=N(1.5)Kp needs to be satisfied. This indicates that the same material used in the prototype could not be used in the experiment. However, the use of different soil material brings about a difference in the compressibility mv as well as in the mechanical characteristics of the soil. Thus, the soil response observed in 1g wave tests cannot be properly extrapolated to field conditions.
Sumer (2014) compared the buildup of residual pore pressure at shallow soil depth obtained from a 1g wave test on the bed of silt (Sumer et al. 2006b) with that of centrifuge wave test on a bed of sand (Sassa and Sekiguchi 1999). Because the coefficient of consolidation [Eq. (2)] was of the same order between them, the pore-pressure response of the 1g test was in good agreement with that of the centrifuge test. However, silt was used as the soil material instead of sand in the field (Sumer et al. 2006b), as described previously.

Criteria for Onset of Wave-Induced Liquefaction

Consider a soil bed with a horizontal surface under wave loading. The time-averaged vertical and horizontal effective stresses at a generic point are
σv_ave=σv0ue
(3)
σh_ave=σh0+Δσh_aveue
(4)
where subscript_ave indicates the period-averaged value; σv0 = initial vertical effective stress; and σh0 = initial horizontal effective stress. Under initial K0-consolidation, σh0=K0σv0, where K0 is the coefficient of lateral earth pressure at rest. The value ue is the residual pore pressure due to cyclic loading, and Δσh_ave is the period-averaged increment of horizontal total stress. The increment of vertical total stress Δσv_ave may be considered to be zero, but Δσh_ave can increase under laterally confined conditions as in the seabed ground. Ishihara and Li (1972) performed triaxial torsion shear tests and showed that under the laterally confined condition, the principal stress ratio defined by K=σ3/σ1 increased in the course of cyclic loading from K0 to unity (K=K01) due to the increase in the horizontal total stress σ3 in association with buildup of residual pore pressure. This aspect was also shown in the analysis of wave-induced liquefaction, in which the mean total stress increased due to the increase in the horizontal total stress σ3, whereas the mean effective stress decreased during wave loading (Sassa and Sekiguch 2001).
From Eqs. (3) and (4), principal total stresses are
σ1_ave=σv0
(5)
σ3_ave=K0σv0+Δσh_ave
(6)
The hydrostatic pressure is excluded for representation. When residual pore pressure builds up, and the principal stress ratio K reaches unity, σ3_ave reaches σ1_ave. From this and Eqs. (5) and (6)
Δσh_ave=(1K0)σv0
(7)
Liquefaction is defined as the state in which the period-averaged mean effective stress is zero, that is, σm_ave=0. Here, σm_ave=(σv_ave+2σh_ave)/3. From Eqs. (3), (4), and (7) and σh0=K0σv0, when residual pore pressure ue reaches the level of σv0, the mean effective stress σm_ave becomes 0 (σm_ave=0). It thus follows that when liquefaction occurs, ue=σv0.
When residual pore pressure ue reaches the level of initial mean effective stress level σm0, mean effective stress does not reach zero (σm_ave0) in the seabed ground, because the principal stress ratio K increases in the laterally confined condition in association with increasing ue. The criteria for the onset of liquefaction by comparing σm0 with ue was adopted by several studies (e.g., Sumer et al. 2006b; Jeng and Seymour 2007; Jeng 2013; Sumer 2014). In this criteria, however, there is room for further buildup of residual pore pressure after the residual pore pressure ue reaches σm0 and therefore, although marginal, some effective stress remains [Sumer (2014), with reference to Sassa and Sekiguchi (2001)], and thus the soil is not in the state of liquefaction, as adopted in this study.

Wave Channel and Wave Generation System in Drum Centrifuge

A water channel installed inside the 2.2-m-diameter drum centrifuge of Toyo Construction (Nishinomiya, Japan) is shown in Fig. 1. The generation and propagation of waves in a drum centrifuge first was studied by Sekiguchi and Phillips (1991). Baba et al. (2002) developed a wave generation system based on a plunger-type device in a drum centrifuge, permitting the soil bed to be subjected to regular waves of constant and moderate amplitude. However, waves of varying amplitude or irregular waves could not be generated, and the severity of the waves generated was below the critical value at which liquefaction occurs. Miyamoto et al. (2018) introduced a piston-type wave maker in the drum channel so that the sequence of storm waves started from relatively small waves and increased to very intense ones, in which liquefaction occurred. Irregular waves, similar to real storm waves, also could be reproduced. Here, the piston-type wave maker was driven by an AC servomotor and feedback system in which the data from a laser-type displacement transducer was used to continuously vary the stroke of the wave paddle (Fig. 1).
Fig. 1. Cross section of a water channel installed inside a drum centrifuge.

Wave Test Cases

Two sets of centrifuge wave tests were performed on loosely packed, fresh deposits of fine sand (Table 1). Three tests (Cases 1, 2, and 5) absent from the table were the cases of a sand bed without a pipe, in which the fundamental characteristics of wave-induced liquefaction were observed, and essentially had the same behavior as in Sassa and Sekiguchi (1999). The present paper describes and discusses the cases of a sand bed with a pipe in Table 1. All the tests were carried out under a centrifuge acceleration of 70g.
Table 1. Test cases
Test setCaseWavesCountermeasure
F-casesCase 3Regular-1No measure
Case 4Irregular
C-casesCase 6Regular-2No measure
Case 7Gravel cover 60°
Case 8Gravel cover 120°
The fundamental cases (F-cases) addressed the floatation process of a pipeline, associated with wave-induced liquefaction. The characteristics of liquefaction were investigated under both regular and irregular waves. In the regular-wave experiment, a sinusoidal wave with gradually increasing amplitude was generated. The period T of the sinusoidal wave was equal to 0.1 s, corresponding to a period of 7 s in the field. In the irregular wave experiment, in which realistic storm waves were assumed, irregular wave forms were generated based on the standard frequency spectra (modified Bretschneider–Mitsuyasu equation). The forms of wave paddle motion in the regular-wave test (Case 3) and irregular-wave test (Case 4) are shown in Figs. 2(a and b). These are measured forms obtained from the laser-type displacement transducer. The wave conditions of the regular- and irregular-wave tests are listed in Table 2. The values of wave height H in the table were calculated from the wave pressures measured at the soil surface using the linear wave theory.
Fig. 2. Measured forms of wave paddle motion: (a) regular-wave form of F-cases; (b) irregular-wave form; and (c) wave form of C-cases.
Table 2. Wave conditions
WavePropertyValue
Regular wave-1 (for F-cases)Period, T0.1 s (7 s)
Wave height, HIncreasing, maximum 50 mm (maximum 3.4 m)
Irregular wavePeriod of significant wave, T(1/3)0.1 s (7 s)
Significant wave height, H(1/3)30 mm (2.0 m)
Regular wave-2 (for C-cases)Period, T0.1 s (7 s)
Wave heightIncreasing and decreasing, maximum 40 mm (maximum 2.8 m)

Note: Values in parentheses are prototype scale.

The countermeasure cases (C-cases) was designed to investigate the effects of gravel layer replacement as a countermeasure against wave-induced liquefaction and pipe floatation. In the experiments, a sinusoidal wave, whose amplitude gradually increased and then gradually decreased, was imposed on the soil surface as one severe wave group. The frequency of the sinusoidal wave was equal to 10 Hz, which corresponded to that in the regular wave tests in the F-cases. The representative form of wave paddle motion from the laser-type displacement transducer is shown in Fig. 2(c). The wave conditions of the C-cases of wave tests are listed in Table 2.

Experimental Model

The cross section of a sand deposit model with a buried pipe in the F-cases is shown in Fig. 3(a). Water was used as the exterior fluid, and Metolose (Shin-Etsu Chemical, Tokyo) with a viscosity of 7.0×105  m2/s was used as the pore fluid in order to match the time-scaling laws of soil consolidation and fluid–wave propagation (Sassa and Sekiguchi 1999). The sand used was silica sand No. 7 (specific gravity of sand grains Gs=2.66, maximum void ratio emax=1.16, minimum void ratio emin=0.70, and median grain size d50=0.15  mm). The loosely formed sand beds had relative densities Dr of 35%–38%, and the initial soil depth of 102 mm corresponded to a 7-m-thick sand bed on a prototype scale. The fluid depth was kept at 100 mm, equivalent to a 7-m water depth on a prototype scale.
Fig. 3. Representative cross sections of sand deposit model with a buried pipe: (a) F-cases; (b) countermeasure case of C-cases; and (c) schematic illustrations of countermeasures.
The pipeline model was made of aluminum pipe (diameter of 25 mm) and urethane foam. The specific gravity of the pipe model was 1.35, which was smaller than the measured specific gravity of liquefied soil, 1.83 [=γsat/γw, where γsat is saturated unit weight of soil; γsat=(Gs+e)/(1+e)γw, where e is void ratio of the sand bed; and γw is unit weight of fluid]. Here, the value of γsat is the value before wave loading, which may not exactly represent that of liquefied soil. In this respect, it is instructive to refer to Sassa et al. (2001). Namely, the wave-induced liquefied soil exhibited significant vibratory motion that centered around the original soil surface before wave loading. This shows that the volume, and thus the void state, of liquefied soil remained essentially the same as that before wave loading. The burial depth of the pipe, which represents the distance between the soil surface and top of the pipe, was 27 mm, corresponding to a 1.9-m soil depth on a prototype scale. This burial depth was nearly the same as the pipe diameter, whose experimental condition was set to be practical in view of the actual problems of buried-pipe floatation (Sumer 2014).
The sand deposit model with a gravel layer over the loose sand bed in the C-cases is shown in Fig. 3(b). The soil surface was partially improved by the gravel layer. The outline of the model was the same as that in the F-cases except for the gravel layer replacement of the soil surface. The thickness of the gravel layer was selected to be 15 mm, corresponding to about 1 m on a prototype scale. The width of the gravel layer varied in Cases 7 and 8. In Case 7, the ground surface in the range of 60° above the pipe was replaced with the gravel layer. In Case 8, the surface layer in the range of 120° above the pipe was replaced with the gravel layer. The gravel layer ranges in Cases 7 and 8 are illustrated schematically in Fig. 3(c). The gravel used had Gs=2.61 and grain size d=2.04.75  mm (d50=3.1  mm).
The setup method for the pipe was as follows. The sand bed was formed by gently pouring sand into a sea of Metolose. When the sand thickness was 50 mm, the pipe model was placed on the sand surface. Silicone grease was applied to the pipe ends to make the friction between the pipe and the wall of the water channel sufficiently small. By continuing to pour sand, a sand bed with a thickness of 102 mm was finally formed. In the case of the gravel cover, when the thickness of the sand was 85 mm, the gravel layer was carefully formed on the soil surface.
The wave pressures acting on the soil surface were measured using pressure transducers. Wave-induced pore-pressure changes in the soil bed were measured with two vertical arrays, and upper and under surfaces of the pipe. The pore-pressure transducers (PPTs) in the arrays were fixed in space using string so that they would not sink into the liquefied soil, and the PPTs on the pipe surface were fixed there. The movements of the pipe and the soil surface were observed using a high-speed charge-coupled device (CCD) camera at an image rate of 250  frames/s.

Floatation of Pipe Caused by Regular-Wave-Induced Liquefaction

A typical set of experimental results from Case 3 is discussed here.

Liquefaction Process

The measured time histories of the excess pore pressures at three different soil depths are shown in Fig. 4, together with the wave pressure at the soil surface. These measurements were from Array A near the pipe [Fig. 3(a)]. Fig. 4(b) shows the response at a shallow soil depth (z=22  mm). No significant buildup of residual pore pressure occurred during the moderate level of wave loading (t=3.03.6  s). The increase in wave severity brought about a buildup of residual pore pressure. When the wave pressure at the soil surface reached 8.1 kPa, the residual pore pressure reached the level of the initial vertical effective stress σvo, indicating the occurrence of liquefaction at this depth. Because the amplitudes of the wave pressure at the soil surface u0+ (trough) and u0 (crest) were slightly asymmetric due to the nonlinearity of the wave, their average pressure was adopted as the wave pressure on the occurrence of liquefaction.
Fig. 4. Time histories of (a) wave pressure acting on the soil surface; and (b–d) excess pore pressures at three different depths (Case 3).
In this study, liquefaction occurred at the wave pressure 8.1 kPa as a result of the increase in amplitude of wave pressure. Under constant wave loading, liquefaction might occur at a lower level of wave pressure, because the occurrence of liquefaction depends not only on the maximum shear stress arising from wave pressure but also on the number of effective waves pertaining to liquefaction. The effective number of waves, together with the conventional maximum shear stress, is capable of representing the severity of given cyclic loading on liquefaction due to the buildup of residual pore pressure (Sassa and Yamazaki 2016). The effective number of waves of constant regular waves with a certain wave pressure becomes larger than that of regular waves whose amplitude increases.
After liquefaction occurred at the shallow soil depth, it occurred at the middle soil depth and then at the bottom, indicating a downward advance of the liquefied zone [Figs. 4(c and d)]. The process of the downward progress of liquefied zone is more clearly illustrated in Fig. 5, in which the advance of liquefaction is based on the time records of the occurrence of liquefaction at three different soil depths.
Fig. 5. Measured propagation of liquefaction in the course of regular wave loading.
The pore-pressure responses in Array B far from the pipe [Fig. 3(a)] also indicated the downward progress of liquefaction. The rate of progressive liquefaction was essentially the same as that observed in Array A.

Observation of Floatation of Pipe

The floatation process of the buried pipe under severe wave loading was observed using a high-speed camera. Photos of the floatation at four different times are shown in Fig. 6. After the occurrence of liquefaction in the sand bed, the pipe gradually moved upward. In the course of wave loading, the pipe finally reached the soil surface. A closer observation of the movement of the pipe is shown in Fig. 6(e). The trajectory of the pipe movement was in the shape of an ellipse, showing that the pipe advanced upward in the significant fluidized motion of liquefied sand under severe waves. Therefore, the behavior of the pipe essentially was caused by the dynamic interaction between the waves and liquefied soil.
Fig. 6. Observed pipe floatation under regular wave loading: (a–d) successive pictures of pipe floatation; and (e) trajectory of the pipe movement.

Relationship between Liquefaction and Pipe Floatation

Floatation of the buried pipe to the soil surface was a consequence of liquefaction. This is illustrated in Fig. 7, which shows the vertical movement of the top of the pipe during wave loading based on the observations using the high-speed camera. No significant movement of the pipe occurred before the soil bed underwent liquefaction. Upon the occurrence of liquefaction at the shallow soil depth the buried pipe started to move upward. The motion of the pipe increased markedly with the downward progress of the liquefied zone in the course of continued wave loading and the pipe eventually reached the soil surface. Because the pipe moved upward with an elliptical motion [Fig. 6(e)], the vertical vibratory motion of the pipe in Fig. 7 represents the vertical component of the elliptical motion of the pipe. The period of the elliptical motion of the pipe corresponded to that of the input waves, indicating that the vibratory motion of the pipe was induced by the significant vibratory motion of the liquefied sand caused by wave loading.
Fig. 7. Measured time history of vertical movement of the pipe in association with propagation of liquefaction during regular wave loading.
The observed pipe behavior had a certain similarity to the previous observation in a 1g wave test. Namely, the video of the 1g test (Sumer 2014) showed that in the course of wave loading, the pipe moved upward and reached the soil surface in association with the significant vibratory motion of liquefied soil. However, the detailed process of the pipe motion could not be observed from the video, because the pipe was in the liquefied soil. In addition, there was a significant difference in the velocity of pipe floatation: 20  mm/s in the present study (Case3) and 0.5  mm/s in the 1g test (Sumer et al. 1999, Fig. 10). As stated previously, scaling laws of wave–soil interactions involving the time-scaling law of the soil consolidation as well as the mechanical similarity have not yet been established for 1g wave tests, so no further quantitative discussion should be made here.

Liquefaction and Pipe Floatation under Irregular Wave Loading

A typical set of experimental results from Case 4 is discussed here. The measured time histories of the excess pore pressures at three different soil depths under irregular wave loading are shown in Fig. 8, together with the wave pressure at the soil surface. These measurements were from Array A near the pipe [Fig 3(a)]. In the case of regular wave loading (Case 3), liquefaction took place at a wave pressure of 8 kPa. When the severe waves with magnitudes exceeding this pressure acted on the soil surface, the residual pore pressures in the sand bed built up significantly. In fact, liquefaction took place at a shallow soil depth at t=2.5  s, and then the liquefied zone propagated downward, leading to liquefaction of the entire sand bed. The process of downward propagation of the liquefied zone during irregular wave loading is shown more clearly in Fig. 9. The rate of progression of liquefaction was far less constant than that in the regular-wave case (Fig. 5). Although the wave pressure for the onset of liquefaction was essentially the same as that in the regular-wave case, the effective number of waves which affected the build-up of residual pore pressure (Sassa and Yamazaki 2016) was fewer than that in the regular-wave case. As a result, the progress of liquefaction under irregular waves was slower than that under regular waves.
Fig. 8. Time histories of (a) wave pressure acting on the soil surface; and (b–d) excess pore pressures at three different depths (Case 4).
Fig. 9. Measured propagation of liquefaction in the course of irregular wave loading.
The pore-pressure responses in Array B, which was far from the pipe [Fig. 3(a)], indicated the same pattern of downward progress of liquefaction. No significant difference was observed in the responses of Arrays A and B.
The vertical movement of the pipe top during irregular wave loading is shown in Fig. 10, together with the measured form of the wave pressure acting on the soil surface. No significant movement occurred before liquefaction took place. At the onset of liquefaction at the shallow soil depth, the pipe started to move upward. In the course of the downward advance of the liquefied zone, the vibratory motion of the pipe gradually became more significant. The vibration of the pipe was an elliptical motion whose period corresponded to that of the waves. The increase in such elliptical motion of the pipe was brought about by the increase in vibratory soil motion associated with the downward propagation of the liquefied zone. Sassa et al. (2001) distinctly observed the relationship between such vertical movement of liquefied sand and the advance of the liquefied zone. Specifically, the soil surface started vibrating at the onset of liquefaction and the amplitude of the vibratory soil motion increased markedly during the downward progress of the liquefaction. Therefore, it is clear that the behavior of the pipe was closely related to the progress of the liquefaction.
Fig. 10. Measured time history of vertical movement of the pipe in association with propagation of liquefaction during irregular wave loading.

Effects of Gravel Layer Replacement on Wave-Induced Liquefaction and Pipe Floatation

A typical set of experimental results from Cases 6–8 is discussed here.

Improvement of Liquefaction Resistance Owing to Gravel Layer

The measured time histories of the excess pore pressures at the shallow soil depth from Cases 6–8 are shown in Fig. 11 together with the wave pressures at the soil surface, u0. These measurements were from Array A near the pipe [Fig. 3(b)]. Fig. 11(a) shows the waveform from Case 6 for the no-countermeasures case. The residual pore pressure developed continuously with the increase in wave severity, and liquefaction occurred at a wave pressure of 7.3 kPa. A closer observation of the soil response shows that the pore pressure decreased after the onset of liquefaction, indicating the dilatancy of the soil induced by the movement of the pipe. In other words, the decrease in the pore pressure was a consequence of dynamic interaction between the pipe movement and the liquefied soil. Such a decrease in residual pore pressure after the occurrence of liquefaction was not measured by Array B, which was far from the pipe. Array A in the C-cases [Fig. 3(b)] was closer to the pipe than in F-cases [Fig 3(a)]. In F-cases, the effects of pipe movement were not detected by Array A.
Fig. 11. Buildup of residual pore pressure at shallow soil depth: (a) no-countermeasures case (Case 6); (b) case of a 60° gravel layer (Case 7); and (c) case of a 120° gravel layer (Case 8).
Fig. 11(b) shows the soil response for Case 7, in which the gravel layer was laid in a 60° range. As in Case 6, the residual pore pressure developed gradually and liquefaction occurred at a wave pressure of 7.8 kPa. This suggests that the effect of the gravel layer in Case 7 on the onset of liquefaction was marginal. Finally, Fig. 11(c) shows the soil response for Case 8, in which the gravel layer was laid in a 120° range. The rate of development of residual pore pressure was much lower than that in Case 6, and liquefaction occurred at a wave pressure of 10.2 kPa. The wave pressure level at which liquefaction occurred in Case 8 was greater than that in Cases 6 and 7. This comparison means that the liquefaction resistance became much greater owing to the gravel layer replacement in a sufficient range over the soil surface.
The onset of liquefaction was accompanied by wave pressure attenuation. The waveform on the soil surface in Fig. 11(a) shows that the amplitude of the wave pressure increased gradually before the soil bed underwent liquefaction. However, upon the occurrence of liquefaction at the shallow soil depth, the rate of increase of the wave pressure amplitude at the soil surface decreased. This indicates a wave damping effect of the liquefied sand bed, which is one form of wave–soil interaction. Wave pressure damping was also seen after the occurrence of liquefaction in Cases 7 and 8.
The increase in liquefaction resistance caused by the gravel layer now is discussed. The severity of the action of travelling sinusoidal waves on a given soil deposit may be expressed in terms of the cyclic stress ratio χ0 (Sassa and Sekiguchi 1999), which is expressed in terms of poroelasticity theory (Madsen 1987; Yamamoto et al. 1978) as
χ0=(τσv0)z=0=κu0γ
(8)
τ=u0·κ·z·exp(κz)for  z0
(9)
σv0=γz
(10)
where τ = wave-induced maximum shear stress in soil; u0 = amplitude of fluid pressure fluctuation at soil surface (z=0); κ = wavenumber; γ = submerged unit weight of soil; and z = vertical coordinate taken downward from soil surface.
The relationship between the cyclic stress ratio χ0 and the residual pore pressure ratio ue/σv0 at the shallow soil depth is depicted in Fig. 12 on the basis of the results from three cases: the no-countermeasures case (Case 6), the case of gravel layer in the 60° range (Case7), and the 120° case (Case 8). Looking first at the soil response in the no-countermeasures case (Case 6) (Fig. 12, hollow circles), the residual pore-pressure ratio ue/σv0 increased with increasing χ0 and reached unity at a critical value χcr, 0.106; ue/σv0=1.0 indicates the occurrence of liquefaction, and the critical value χcr, beyond which liquefaction occurs, indicates the liquefaction resistance of a given sand bed. Looking at the soil response in the case of the 60° gravel layer (Case 7) (Fig. 12, hollow diamonds), the critical value χcr was 0.112, which was somewhat higher than the χcr value in the no-countermeasures case (Case 6). This means that the liquefaction resistance became slightly higher when the gravel layer was applied in the 60° range. Finally, looking at the soil response in the case of 120° gravel layer case (Case 8) (Fig. 12, solid circles), the critical value χcr was considerably higher than that in the no-countermeasures case. The critical value in Case 8 was 0.146, whereas the value in the no-countermeasures case (Case 6) was 0.106. This indicates that the liquefaction resistance of the sand bed increased by 40% owing to the gravel layer replacement in a 120° range over the soil surface.
Fig. 12. Increase in liquefaction resistance owing to a sufficient range of gravel layer.

Effects of Gravel Layer on Progressive Liquefaction

The process of wave-induced liquefaction is progressive in nature. This section discusses the progressive nature of liquefaction in a soil bed with gravel layer replacement. Measured time histories of the downward progress of the liquefied zone in the course of wave loading are shown in Fig. 13, in which the advance of liquefaction is based on the time records of the occurrence of liquefaction at three different soil depths. In these tests, the differences in the time of occurrence of liquefaction at the shallow soil depth corresponded to those in the liquefaction resistance χcr of each case, which was discussed in the previous section. The form of progressive liquefaction in the case of a 60° gravel layer (Case 7) was the same as that of the no-countermeasures case (Case 6). By contrast, in the case of a 120° gravel cover, liquefaction at a shallow soil depth occurred significantly later than in the no-countermeasures case. This is because of the increase in the liquefaction resistance χcr caused by gravel layer replacement. However, the rate of downward progress of the liquefaction was more rapid because the sand at the greater depth was not improved by the gravel layer at the soil surface. This indicates that it is important for practical designs to prevent the onset of liquefaction against design waves by increasing the liquefaction resistance χcr in a sufficient manner by effective countermeasures, such as gravel layer replacement.
Fig. 13. Measured propagation of liquefaction in the course of wave loading from Cases 6–8.

Effects of Gravel Layer on Pipe Floatation

The instability of the buried pipe under severe wave loading was observed using the high-speed camera. Figs. 14(a–d) show the pipe–soil–gravel behavior during wave loading at four different times in the case of a gravel layer in the 120° range (Case 8). A gravel cover with a sufficient range brought about less development of floatation of the pipe and prevented it from reaching the soil surface. Closer observation of the sand–gravel–pipe behavior shows that after the occurrence of liquefaction in the sand bed, the soil surface started vibrating. Both ends of the gravel layer began to sink in association with the vibratory soil motion [Fig. 14(b)]. The gravel layer directly above the pipe prevented the liquefied sand from enhancing the vibratory motion, leading to less movement of the pipe [Fig. 14(c)]. Thus, the pipe did not float to the soil surface [Fig. 14(d)].
Fig. 14. Observed pipe floatation during a wave group: (a–d) successive pictures in the case of a 120° gravel layer (Case 8); and (e–h) successive pictures in the case of a 60° gravel layer (Case 7).
By contrast, a gravel cover with an insufficient range could not prevent the pipe from floating to the soil surface. This aspect can be clearly seen in Figs. 14(e–h), which show the floatation of the pipe in Case 7. The insufficient gravel layer could not prevent the liquefied sand from enhancing vibratory motion, leading to significant pipe floatation.
The elliptical trajectory of the pipe movement of Case 7 is shown in Fig. 15. As a result of the dynamic interaction between the waves and liquefied soil near the soil surface, the pipe moved laterally in the same direction as the progressive wave. The pipe passed the end of the gravel, reached the soil surface, and was exposed. Thus, a gravel layer with an insufficient range could not prevent pipe exposure.
Fig. 15. Relationship between the range of gravel layer and the trajectory of the pipe movement (Case 7).

Processes of Pipe Floatation with Gravel Layer

The effects of a gravel layer on floatation of the pipe are clearly shown in Fig. 16, which shows the time histories of vertical movement of the pipe centre during wave loading for Cases 6, 7, and 8 and the measured wave pressure at the soil surface for Case 8. The rate of development of pipe displacement in Case 8, which had a gravel layer in the 120° range, was much lower than that in the no-countermeasures case (Case 6). The vertical displacement of the pipe in Case 8 was 18 mm after passage of the storm wave group, and it remained at 15 mm below the soil surface, corresponding to a 1-m depth of burial on a prototype scale. Closer observation of its development indicates that the amplitude of pipe vibration of this case was much smaller than that in the no-countermeasures case, indicating that the sufficient range of the gravel layer prevented liquefied soil motion caused by severe wave loading. By contrast, in Case 7, which had a gravel layer in the range of 60°, the pipe floated upward at the same rate as in the no-countermeasures case. This floatation was closely associated with the development of the elliptical motion of the pipe, which was brought about by the significant vibratory motion of the liquefied sand, and it finally reached the soil surface. It was found that a gravel cover with a range of 120° proved sufficient in preventing significant vibratory motion of the liquefied soil and floatation of the pipe and could thus be an effective countermeasure against wave-induced liquefaction and pipe floatation.
Fig. 16. Effects of a gravel layer: time histories of development of pipe floatation during waves in Cases 6–8 and wave pressure at the soil surface in Case 8.

Conclusions

The relationships between wave-induced liquefaction, the floatation of a buried pipe in a sand bed, and the countermeasures against pipe floatation associated with liquefaction were investigated by performing a range of centrifuge wave tests in a drum channel. This was done in such a way that the time-scaling laws for fluid wave propagation and consolidation of the soil were matched. The principal findings obtained from the present study may be summarized as follows:
1.
Under severe wave loading, the loosely packed fresh deposit of sand underwent liquefaction. The liquefied zone propagated downward in the course of wave loading. The floatation of the buried pipe was a consequence of progressive liquefaction. In the regular-wave test, with the occurrence of liquefaction at a shallow soil depth, the buried pipe started moving upward and the motion of the pipe increased markedly in association with the progress of liquefaction during severe wave loading. Finally, the pipe reached the soil surface.
2.
The onset and spread of liquefaction in the irregular-wave condition were essentially the same as those observed in the regular-wave condition. When more-severe waves than those beyond which liquefaction occurred (in the case of the regular-wave condition) acted on the soil, residual pore pressures built up significantly, leading to the onset and downward spread of liquefaction. The pipe floated to the soil surface with a significant elliptical motion, which took place owing to the vibratory soil motion associated with the progress of liquefaction.
3.
The effects of a gravel layer over the sand bed on wave-induced liquefaction and pipe floatation were also examined. A sufficient range of gravel layer replacement increased the liquefaction resistance of the sand bed, lessened the development of floatation of the pipe and prevented it from reaching the soil surface. A gravel layer in the range of 120° above the pipe increased the liquefaction resistance by 40% compared with the no-countermeasures case. By contrast, the increase in liquefaction resistance caused by a gravel layer in the range of 60° was marginal. A gravel layer in the range of 120° prevented the liquefied sand from enhancing vibratory motion, leading to less movement of the pipe. Thus, the pipe did not float to the soil surface. In the case of a gravel layer in the range of 60°, the pipe floated upward at the same rate as in the no-countermeasures case. This floatation was associated with the vibratory motion of the pipe caused by the significant motion of the liquefied sand, and the pipe finally reached the soil surface. Gravel cover with a range of 120° proved sufficient in preventing significant vibratory motion of the liquefied soil and floatation of the pipe and could thus be an effective countermeasure in a realistic setting against wave-induced liquefaction and pipe floatation.

Data Availability Statement

Some data used during the study are available from the corresponding author by request (test movies).

References

Baba, S., M. Miyake, K. Tsurugasaki, and H. Kim. 2002. “Development of wave generation system in a drum centrifuge.” In Proc., Int. Conf. on Physical Modelling in Geotechnics (ICPMG ‘02), 265–270. Rotterdam, Netherlands: A.A. Balkema.
Christian, J. T., P. K. Taylor, J. K. C. Yen, and D. R. Erali. 1974. “Large diameter underwater pipeline for nuclear power plant designed against soil liquefaction.” In Proc., Offshore Technology Conf., 597–606. Houstan: Offshore Technology Conference.
Damgaard, J. S., B. M. Sumer, T. C. Teh, A. C. Palmer, P. Foray, and D. Osorio. 2006. “Guidelines for pipeline on-bottom stability on liquefied noncohesive seabeds.” J. Waterway, Port, Coastal, Ocean Eng. 132 (4): 300–309. https://doi.org/10.1061/(ASCE)0733-950X(2006)132:4(300).
Gao, F. P., X. Y. Gu, and D. S. Jeng. 2003. “Physical modeling of untrenched submarine pipeline instability.” Ocean Eng. 30 (10): 1283–1304. https://doi.org/10.1016/S0029-8018(02)00108-7.
Herbich, J. B., R. E. Schiller, W. A. Dunlap, and R. K. Watanabe. 1984. Seafloor scour: Design guidelines for ocean-founded structures. New York: Marcel Dekker.
Ishihara, K., and S. Li. 1972. “Liquefaction of saturated sand in triaxial torsion shear test.” Soils Found. 12 (2): 19–39. https://doi.org/10.3208/sandf1972.12.19.
Jeng, D. S. 2013. Porous models for wave-seabed interactions. Berlin: Springer.
Jeng, D. S., and B. R. Seymour. 2007. “Simplified analytical approximation for pore-water pressure buildup in marine sediments.” J. Waterway, Port, Coastal, Ocean Eng. 133 (4): 309–312. https://doi.org/10.1061/(ASCE)0733-950X(2007)133:4(309).
Madsen, O. S. 1987. “Wave-induced pore pressure and effective stress in a porous bed.” Géotechnique 28 (4): 377–393. https://doi.org/10.1680/geot.1978.28.4.377.
Miyamoto, J., S. Sassa, and K. Tsurugasaki. 2018. “Wave-induced liquefaction and floatation of pipeline buried in sand beds.” In Proc., 9th Int. Conf. on Physical Modelling in Geotechnics 2018 (ICPMG 2018), 571–576. London: CRC Press.
Sassa, S. 2014. “Tsunami-seabed-structure interaction from geotechnical and hydrodynamic perspectives.” Geotech. Eng. J. 45 (4): 102–107.
Sassa, S., and H. Sekiguchi. 1999. “Wave-induced liquefaction of beds of sand in a centrifuge.” Géotechnique 49 (5): 621–638. https://doi.org/10.1680/geot.1999.49.5.621.
Sassa, S., and H. Sekiguchi. 2001. “Analysis of wave-induced liquefaction of sand beds.” Géotechnique 51 (2): 115–126. https://doi.org/10.1680/geot.2001.51.2.115.
Sassa, S., H. Sekiguchi, and J. Miyamoto. 2001. “Analysis of progressive liquefaction as a moving-boundary problem.” Géotechnique 51 (10): 847–857. https://doi.org/10.1680/geot.2001.51.10.847.
Sassa, S., H. Takahashi, Y. Morikawa, and D. Takano. 2016. “Effect of overflow and seepage coupling on tsunami-induced instability of caisson breakwaters.” Coastal Eng. 117 (Nov): 157–165. https://doi.org/10.1016/j.coastaleng.2016.08.004.
Sassa, S., and H. Yamazaki. 2016. “Simplified liquefaction prediction and assessment method considering waveforms and durations of earthquakes.” J. Geotech. Geoenviron. Eng. 143 (2): 04016091. https://doi.org/10.1061/(ASCE)GT.1943-5606.0001597.
Sekiguchi, H., and P. Phillips. 1991. “Generation of water waves in a drum centrifuge.” In Proc., Int. Conf. Centrifuge 1991 (CENTRIFUGE 91), 345–350. Rotterdam, Netherlands: A.A. Balkema.
Sekiguchi, H., S. Sassa, K. Sugioka, and J. Miyamoto. 2000. “Wave-induecd liquefaction, flow deformation and particle transport in sand beds.” In Proc., Int. Conf. GeoEng2000. Lisbon, Portugal: International Society for Rock Mechanics. CD-ROM.
Sumer, B. M. 2014. “Liquefaction around marine structures.” In Vol. 39 of Advanced series on ocean engineering. Singapore: World Scientific.
Sumer, B. M., F. H. Dixen, and J. Fredsoe. 2010. “Cover stones on liquefiable soil bed under waves.” Coastal Eng. 57 (9): 864–873. https://doi.org/10.1016/j.coastaleng.2010.05.004.
Sumer, B. M., J. Fredsoe, S. Christensen, and M. T. Lind. 1999. “Sinking/floatation of pipelines and other objects in liquefied soil under waves.” Coastal Eng. 38 (2): 53–90. https://doi.org/10.1016/S0378-3839(99)00024-1.
Sumer, B. M., F. Hatipoglu, J. Fredsoe, and N.-E. O. Hansen. 2006a. “Critical floatation density of pipelines in soils liquefied by waves and density of liquefied soils.” J. Waterway, Port, Coastal, Ocean Eng. 132 (4): 252–265. https://doi.org/10.1061/(ASCE)0733-950X(2006)132:4(252).
Sumer, B. M., F. Hatipoglu, J. Fredsoe, and S. K. Sumer. 2006b. “The sequence of soil behaviour during wave-induced liquefaction.” Sedimentology 53 (3): 611–629. https://doi.org/10.1111/j.1365-3091.2006.00763.x.
Teh, T. C., A. Palmer, and J. Damgaard. 2003. “Experimental study of marine pipelines on unstable and liquefied seabed.” Coastal Eng. 50 (1–2): 1–17. https://doi.org/10.1016/S0378-3839(03)00066-8.
Yamamoto, T., H. L. Koning, H. Sellmeijer, and E. Hijum. 1978. “On the response of a poro-elastic bed to water waves.” J. Fluid Mech. 87 (1): 193–206. https://doi.org/10.1017/S0022112078003006.
Zhao, K., H. Xiong, G. Chen, D. Zhao, W. Chen, and X. Du. 2018. “Wave-induced dynamics of marine pipelines in liquefiable seabed.” Coastal Eng. 140 (Oct): 100–113. https://doi.org/10.1016/j.coastaleng.2018.06.007.

Information & Authors

Information

Published In

Go to Journal of Waterway, Port, Coastal, and Ocean Engineering
Journal of Waterway, Port, Coastal, and Ocean Engineering
Volume 146Issue 2March 2020

History

Received: Jan 30, 2019
Accepted: Jun 25, 2019
Published online: Dec 5, 2019
Published in print: Mar 1, 2020
Discussion open until: May 5, 2020

Authors

Affiliations

Junji Miyamoto, Dr.Eng. [email protected]
Senior Research Engineer, Technical Research Institute of Naruo, Toyo Construction, 1-25-1 Naruohama, Nishinomiya 663-8142, Japan (corresponding author). Email: [email protected]
Shinji Sassa, Dr.Eng. [email protected]
Head of Soil Dynamics Group and Research Director, Port and Airport Research Institute, National Institute of Maritime, Port and Aviation Technology, National Research and Development Agency, 3-1-1 Nagase, Yokosuka 239-0826, Japan. Email: [email protected]
Kazuhiro Tsurugasaki, Dr.Eng.
Head of Geo-Disaster Prevention Laboratory, Technical Research Institute of Naruo, Toyo Construction, 1-25-1 Naruohama, Nishinomiya 663-8142, Japan.
Hiroko Sumida
Research Engineer, Technical Research Institute of Naruo, Toyo Construction, 1-25-1 Naruohama, Nishinomiya 663-8142, Japan.

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share