Sie sind auf Seite 1von 61

CHAPTER 4

PREVIOUS EXPERIMENTAL WORK

4.1 Introduction A wide range of field and laboratory experiments has been performed by researchers attempting to provide parameters for and to validate SSPSI analytical methods. These experimental methods have been concerned with the load-deformation behavior of soil-pile systems both singly and in groups, at small to large strains, loaded statically, cyclically, dynamically, or seismically, by exciting the pile head or the soil mass, and covering a variety of pile types and soil conditions. In-situ tests have the advantage of providing correct soil and pile stress conditions, whereas laboratory tests offer the flexibility and economy of making parametric studies in a controlled environment. Taken together, field and laboratory tests of soil-pile interaction

complement each other and provide a valuable body of data where recorded SSPSI response is lacking. The following sections provide a comprehensive survey of soil-pile experimental research published in the literature; the purpose of such a review is to understand the adequacy of previous work and the dimensions of further research needs. Attention is particularly focused on experimental work applicable or directed to SSPSI in soft clays.

4.2 Full Scale Pile Test Programs Pile load test programs conducted in the field offer the distinct advantages of utilizing real soil, real piles, and realistic soil-pile stress conditions. They are limited in the sense that loading is applied in a top down fashion, concentrating the effects of

151

inertial interaction and ignoring the effects of kinematic interaction.

This section

recounts selected case studies of single pile lateral load tests, as well as a complete survey of pile group lateral load tests and pile dynamic tests reported in the english language literature.

4.2.1 Field Single Pile Lateral Load Tests Of the innumerable field load test programs performed, three are particularly notable in that they exert a disproportionate influence on engineering practice, as they have been codified in API design recommendations and are programmed as default soil properties in the COM624P and LPILE computer programs (which are universally used for the design of laterally loaded piles). These field load tests were performed by

Matlock (1962) in soft clay, Reese et al. (1975) in stiff clay, and Reese et al. (1974) in sands. A typical example of the setup for a pile lateral load test is shown in Figure 4.1; the American Society for Testing Materials publishes standardized procedures for conducting such load tests under specification ASTM D-3966 (ASTM, 1996).

Figure 4.1 - Example of Pile Load Test Set Up for Combined Lateral and Axial Load (after ASTM, 1996)

152

Matlock (1962) performed an integrated field and laboratory study that included static, cyclic, and post-cyclic lateral head loading of 12.75 in diameter steel pipe piles embedded 42 ft deep at two different soft clay sites at Lake Austin and Sabine, Texas. The undrained shear strength of the soils at these two sites ranged from 300 to 800 psf in the upper soil layers. The principal conclusions of his study were: The soil resistance-pile deflection (p-y) relationship is highly nonlinear and inelastic, with characteristic pile bending moment patterns (see Figure 4.2). Static and cyclic nonlinear soil-pile response is most severe at shallow depths, and approaches linear response at greater depths. For engineering purposes, the fundamental p-y relationship is independent of pile head fixity (although pile forces are strongly related to fixity). After a large number of cycles of loading and degradation of resistance, the soil-pile system tends to stabilize (a condition known as shakedown). An important effect of cyclic loading is gapping, with high transient pile forces developed while traversing the gap. Response during reloading after cycling is governed by soil resistance being reduced for deflections smaller than those previously attained. This seminal work has guided the field of laterally loaded pile research, and has also exerted due influence on SSPSI analyses. In their similarly conducted tests, Reese and his co-workers (1975) drove 6 and 24 in diameter steel pipe piles into a stiff, fissured, overconsolidated clay deposit near Austin, Texas; their site had been pre-excavated 3 ft, and water impounded at the surface to simulate conditions that exist at the ocean floor. Unconfined compressive strengths of

153

Figure 4.2 - Characteristic Fixed Head Laterally Loaded Pile Bending Moment Pattern (after Matlock, 1962) 2 to 4 tsf were found at this site in the upper 20 ft. Reese found a much greater degree of cyclic degradation of soil resistance in his tests than Matlock found in soft clay; the impounded water was thought to have contributed to scour of the soil in the soil-pile gap that opened during each cycle of loading, thereby further degrading resistance. It is important to consider that the period of cyclic loading in these tests was in excess of 15 seconds, and the observed scour effect could be expected to vary with loading rate. P-y curves for a series of depths in static and cyclic loading are shown in Figure 4.3, with results presented for a series of depths; they reflect the tendency for the p-y relationships to become linear at depth.

154

Figure 4.3 - P-Y Curves Developed from Static and Cyclic Lateral Load Tests on 24-in Diameter Pile in Stiff Clay (after Reese et al., 1975) Gill (1968) conducted a series of field tests on 4.5 to 16 in diameter steel pipe piles driven into San Francisco Bay Mud at Hamilton Air Force Base in Novato, California. The field tests consisted of lateral head loading tests and segmental pile tests, the latter accomplished by horizontally loading small (4 12 in tall) detachable segments of the pile which had been driven to a predetermined depth of interest. The tests were conducted in both flooded and dry areas, where the undrained shear strength measured insitu by vane shear ranged from 600 - 1200 psf in the flooded area, and up to 2600 psf in the top dessicated zone of dry sites. Figure 4.4 depicts Gills experimental results

superimposed with groundline deflections computed for this study by COM624P, the latter using pile and soil properties provided by Gill; the analysis was conducted as a blind prediction, with no iteration on input parameters to optimize the results. The analysis very well predicts deflections recorded in the field for the two smaller pile types,

155

20 Test P1 - Flooded Experimental COM624P Test P-2 Flooded

0 20 Test P-3 Flooded Test P-4 Flooded

Lateral Load (K)

0 20 Test P-5 Dry Test P-6 Dry

0 20 Test P-7 Dry Test P-8 Dry

0 0 0.5 1 1.5 2 0 0.5 1 1.5 2

Groundline Displacment (in)

Groundline Displacement (in)

Figure 4.4 - Static Lateral Load Test Results for Piles at Dry and Flooded Bay Mud Sites, Superimposed with COM624P Predicted Response (after Gill, 1968)

156

but overpredicts the stiffness of the two larger diameter piles. In fact the pile EI values published by Gill for the two larger diameter piles in particular seem 30 % high, given standard pipe sizes and steel properties, but reducing these EI values to a reasonable range has only a moderate effect on reducing lateral pile stiffness. These trends also run counter to the diameter effect observed by Stevens and Audibert (Section 3.1.2). The overall conclusion is that the lateral response of piles in Bay Mud at this site is inconsistently (and potentially unconservatively) captured by the default Matlock and Reese p-y curve criteria, and could possibly be improved with site specific p-y curve construction. After the 1985 Mexico City earthquake, Jaime et al. (1989) conducted cyclic axial load tests on friction piles driven in Mexico City clay, and concluded that the performance of these foundations during the earthquake was a function of the initially applied static load. In these tests, when the total applied cyclic load Pmax did not exceed the ultimate static capacity Pult, the piles behaved elastically, with rate-dependent stiffness and strength increasing under dynamic loading. As Pmax ranged from 1.0 to 1.2 Pult, permanent displacements ensued, and when Pmax exceeded 1.2 to 1.3 Pult, large permanent displacements and a loss of capacity occurred. The results of a parallel finite element study were in agreement with the field test results, and indicated that interaction between group piles would be greatly reduced due to nonlinearity arising from soil-pile slippage.

4.2.2 Field Pile Group Lateral Load Tests Several field load tests on pile groups are reported in the literature, and are summarized in Table 4-1. These tests have been conducted on a variety of pile types,

157

Table 4-1 Field Pile Group Loading Tests

pile type Reference (B= some battered piles) Feagin (1937) timber (B) Wagner (1954) timber Peck (1961) step-taper (B) Matsumoto et al. (1962) reinforced concrete (B) Kim and Brungraber (1976) steel HP (B) Bartolomey (1977) prestressed concrete Kim et al. (1979) steel HP (B) Stevens et al. (1979) timber Matlock et al. (1980) steel pipe Matlock et al. (1980) steel pipe Schmidt (1981) drilled shaft Holloway et al. (1982) timber Schmidt (1985) drilled shaft Meimon et al. (1986) steel HP Brown et al. (1987) steel pipe Brown et al. (1988) steel pipe Abacarius (1991) concrete fill steel pipe Abacarius (1991) concrete fill steel pipe Kobayashi et al. (1991) steel pipe Kimura et al. (1993) drilled shaft Ruesta and Townsend (1997) prestressed concrete Rollins et al. (1998) concrete fill steel pipe Weaver et al. (1998) concrete fill steel pipe pile cap above grade above grade at grade above grade partial embed above grade above grade above grade load frame load frame load frame above grade load frame load frame load frame load frame partial embed partial embed load frame above grade above grade load frame partial embed

pile diameter (in) 12 - 14 16 13 18 10 12 10 12 - 14 6 7 47 14 47 12 10.75 10.75 12 12 6 39 30x30 12.75 12.75 pile installation soil conditions driven sand driven glacial till driven sandy silt driven peat, clay driven silty clay driven clay driven silty clay jetted/driven clay, sand driven soft clay driven soft clay cast in place silt, marl jetted/driven alluvium cast in place med. dense sand driven clay driven stiff OC clay backfilled med. dense sand driven bay mud driven sandy silt driven sandy clay cast in place soft silt driven sand driven clayey silt driven clayey silt

pile length (ft) 30 34, 39 40 49 40 16, 40 40 43 - 45 45 45 52, 92 35 28 25 43 43 60 15 22 82 46 30 30

pile pile pile head group spacing fixity 6, 8, 9 3d cast 1x2 ? cast 3 4 - 8d cast 1, 4, 6 2.4 - 2.9d cast 3x3 3.6, 4.8d cast 4, 6 3d, 4d cast 3x3 3.6, 4.8d cast 8, 12 3d cast 5 3.4d fixed 10 1.8d fixed 1x2 1.3, 2.4d free 2x4 3d cast 2, 3 3d free 2x3 3d free 3x3 3d free 3x3 3d free 11, 17 3d cast 10, 12 3d cast 3x3 3.3d free 1x2 2.8d cast 4x4 ? ? 3x3 3d free 3x3 3d free/fixed

test loading static static static static cyclic load static cyclic load static cyclic displ. cyclic displ. static static to failure cyclic load cyclic load cyclic defl. cyclic defl. static to failure static to failure static cyclic static to failure static statnamic

max. # of cycles 1 1 1 1 3 1 3 1 100's 100's 1 1 40 10000 900 900 3 2 3 3 1 3 1

generally of smaller diameter and shorter length, driven in a gamut of soil conditions, and loaded both statically and cyclically. The size of the pile groups has been limited by the capacity of external loading equipment, so small pile groups at close to intermediate spacing have been the norm. The first pile group lateral load tests reported in the literature are by Feagin (1937), who performed field tests on groups of 32 ft long timber piles at Lock and Dam No. 26, Alton, Illinois. The soil conditions at this site were sandy alluvium, and the piles were installed by a combination of jetting and driving. The focus of the test program was to investigate the relative lateral resistance of vertical and battered piles, in all conceivable combinations and orientations to the direction of loading. The superior performance of battered piles under lateral load was clearly evident. Kim and Brungraber (1976) drove 2x3 groups of vertical and battered H-piles in cohesive soil, and aroused much discussion with the reporting of their test results. They compared the pile group per pile performance (fixed head condition) to single (free head) reference piles driven nearby, and computed pile group efficiencies in excess of unity, contrary to conventional notation. The group piles were joined by a pile cap, the bottom of which was cast against the ground surface, which introduced a potential for pile cap base frictional contribution to lateral resistance. In response to their critics, Kim et al. (1979) published the results of a second series of tests where 4 in of soil beneath the pile caps had been excavated to relieve any potential frictional resistance. They concluded that the pile cap base friction contribution was negligible for battered pile groups, but significant for vertical pile groups.

159

Figure 4.5 - Field Pile Group Load Test Results Indicating Preferential Load Distribution to Leading Piles (after Holloway et al., 1982) The first field pile group load test program that clearly delineated group effects was performed by Holloway et al. (1982), who revisited Lock and Dam No. 26 in a study of rehabilitation schemes for the facility. They installed timber piles with the same construction techniques as originally used in the 1930s, and tested a 2x4 pile group to failure. One of the key results was experimental evidence of pile group shadowing, i.e. the preferential load carrying capacity of piles in front of the line of loading, thereby reducing load on piles at the rear of the line of loading. illustrated in Figure 4.5. Brown et al. (1987, 1988) performed cyclic lateral load tests on 3x3 pile groups in stiff clay and sand, and provided detailed evidence of pile group effects. conclusions can be summarized as follows, and are reflected in Figures 4.6a and b. Pile group deflections exceed single pile deflections under equivalent per pile load. Pile group bending moments exceed those for single piles, and are shifted deeper. Pile group maximum soil resistance is reduced relative to single piles under both static and cyclic loads, and is most pronounced at depth. Their This load distribution is

160

The greatest portion of shear on the pile group is distributed to piles in the front row, and variations in load within the group are approximately 20 %.

Pile group cyclic loading effects were found to contribute to degradation of soil resistance at the stiff clay soil site, but densification at the sand soil site prevented any loss of capacity.

(a)

(b) Figure 4.6 - Field Pile Group Load Test Results Depicting; a) Cyclic Degradation of Resistance; b) Distribution of Load by Row (after Brown et al., 1987) Abacarius (1990) statically loaded pile groups to failure that supported two bents of the then demolished Cypress Freeway that had catastrophically collapsed in the 1989 Loma Prieta earthquake. The two bents were founded on 60 ft long concrete filled steel

161

pipe piles in Bay Mud, and 15 ft long piles in sandy silt, respectively. The pile groups ranged in size from 10 to 17 piles, and the pile caps were partially embedded. The basic conclusion of the test program was that the Caltrans presumptive lateral resistance of 5 kips for this pile type at in deflection was extremely conservative, as experimental values ranged from 17.7 to 32.9 kips. This however did not take into account the passive resistance offered by the pile cap, which may be significant, and the report does not provide sufficient information to evaluate those effects. Rollins et al. (1998) investigated pile group effects with field lateral load testing of a 3x3 group (and a single pile) consisting of 12.75 in diameter concrete filled steel pipe piles spaced at 3d and driven 30 ft into lightly overconsolidated layered silts, clays, and sands. They found that pile shadowing resulted in the maximum load being

distributed to the front row of piles, and that more load was distributed to the back row than the interior piles, in contrast to Browns findings. Rollins and his co-workers also observed higher bending moments in group piles than in the single pile at the same average load level, particularly at higher load levels. As they note, this has important implications for the common engineering practice of extrapolating single pile test results to group behavior. They simulated the single pile response with the computer code LPILE, and found that using a detailed soil profile rather than an averaged one produced good results. They also proposed p-multipliers for this soft soil site ranging from 0.6 for the front row piles to 0.4 for the interior and back piles, which are lower values than those proposed by other researchers. Weaver et al. (1998) followed Rollins work by conducting statnamic lateral loading tests on the same 3x3 pile groups, in both fixed and free head conditions, and

162

with and without pile cap embedment. They found that the dynamic resistance was 30 to 50 % higher than the static resistance for a free head pile and for the fixed head group without cap embedment, and 100 to 125 % higher for the fixed head group with cap embedment. They attributed the increased resistance mainly to damping and

acknowledged that further work needs to be done in interpreting statnamic test results.

4.2.3 Field Pile Dynamic Tests In order to ascertain pile stiffness under dynamic loads, researchers have conducted three classes of tests on full-scale piles and pile groups in the field. In all three types of tests, a mass is commonly fixed to the pile head to accentuate the resonant response and damping characteristics of the pile. Ringdown tests consist of quickly releasing the pile from some imposed, initial lateral displacement, and measuring the ensuing free vibrations of the pile as it attempts to rebound to its original position. Pile stiffness and damping values can be derived from measurements of the free vibrations of the pile by the logarithmic decrement method. Impact tests are an even smaller strain version of a ringdown test, where a blow to the pile generates free vibrations in the pile to be measured. Forced vibration tests involve mounting an eccentric mass shaker to the pile head, whose motors spin eccentrically fixed masses, thereby inducing vibrations into the pile head. By adjusting the orientation, motor speed, and fixed mass, this test offers the flexibility of generating horizontal, vertical, or rocking vibrations over a range of frequencies and amplitudes. Electrodynamic oscillators are also employed in forced vibration tests, and can deliver much higher frequencies to the pile head than the mechanical type, which is limited to about 100 Hz. Soil-pile stiffness and damping can

163

be interpreted directly from the test data resonance curves with the half-power bandwidth method. Comparisons of observed and predicted behavior are good when the response remains linear and soil elastic properties are well-characterized. Conversely, when

higher load levels generate nonlinear soil-pile response, models of predicted response are less accurate. Cases of field pile dynamic tests reviewed are summarized in Table 4-2, with the following general observations followed by comments on selected test programs: Soil-pile dynamic response is highly site dependent. Soil-pile dynamic response is frequency and load level dependent. Soil-pile vertical stiffness exceeds soil-pile horizontal stiffness. Pile cap embedment increases soil-pile dynamic stiffness and damping. Soil-pile nonlinear response decreases both stiffness and damping. Pile group effects are frequency, pile spacing, and site dependent, and are more pronounced for stiffness than damping. Elastic continuum analytical models incorporating a weak zone around the pile, soil-pile gapping, and a parabolic variation of modulus with depth appear to provide a reasonably good level of accuracy for the cases studied. Petrovski and Jurokovski (1973) dynamically tested single piles and four pile groups of drilled shafts in loose sandy soil, with different conditions of pile cap embedment. The contrast between linear and nonlinear response is shown in Figures 4.7a and b, which depict stiffness and damping degradation at increasing load levels in the latter figure as a result of pile cap resistance removed relative to the first case. In a unique approach, the Soviet researcher Grib (1975) reports on field piles excited by a series of explosive charges timed to have earthquake-like characteristics. Gribs

164

results and analysis were not especially impressive, but his experimental method does hold promise as it overcomes the limitation of applying dynamic loads directly to the pile head, rather than through the free-field soil. This explosive excitation method has also been performed for soil-structure interaction tests (no piles) in the SIMQUAKE experiments in the 1970s and in centrifuge experiments by Zelikson et al. (1982).

(a)

(b)

Figure 4.7 - Dynamic Pile Response from Forced Vibration Tests: a) Linear Response; b) Nonlinear Response due to Removal of Supporting Soil Near Pile Head (after Petrovski and Jurokovski, 1973) Scott et al. (1982) conducted horizontal forced vibration and ringdown tests on an instrumented steel pipe pile driven into silty sand, and in a parallel study modeled the observed field response in centrifuge tests. The extensive field instrumentation

monitored pile bending moments, pile head displacement and acceleration, pore pressures in the surrounding soil, and ground surface velocity in the free-field (Figure 4.8).

165

Table 4-2 Field Pile Dynamic Loading Tests

pile pile diameter length Reference pile type (in) (ft) Hayashi et al. (1965) steel pipe 47 ? Hayashi et al. (1965) steel HP 12 ? Ishii and Fujita (1965) steel pipe 42, 59 109 Maxwell et al. (1969) concrete fill steel pipe 14 46 Maxwell et al. (1969) steel HP 14 52 - 88 Maxwell et al. (1969) steel HP 14 52 Alpan (1973) reinforced concrete 12 15 Hakuno (1973) steel pipe 24 197 Petrovski and Jurokovoski (1975) drilled shaft 20 49 Grib (1975) concrete, steel pipe 6 - 12 11 - 16 Gyoten et al. (1980) steel pipe 24 47, 87 Gyoten et al. (1980) concrete 24 47 Scott et al. (1982) steel pipe 24 40 Scott et al. (1982) steel pipe 24 40 Gle and Woods (1983) steel pipe 12.75, 14 50 - 160 Jennings et al. (1984) steel pipe 18 22 Mizuhata and Kusakabe (1984) steel pipe 26, 32 142 Blaney and O'Neill (1986) steel pipe 10.75 44 Blaney et al. (1987) steel pipe 10.75 43 Crouse and Cheang (1987) concrete fill steel pipe 12 38 Blaney and O'Neill (1989) steel pipe 10.75 43 Wakamatsu (1989) cast-in-place barette 28x5, 14x5 16 Hakulinen (1991) reinforced concrete 12x12 23 Hakulinen (1991) concrete fill steel pipe 10.75 23 Kobori et al. (1991) drilled shaft 23.6 24.6 Kobori et al. (1991) drilled shaft 23.6 24.6 Kobori et al. (1991) drilled shaft 23.6 24.6 Fuse et al. (1992) steel pipe 59 164 Han and Vaziri (1992) drilled shaft 12.5 25 Mizuno and Iiba (1992) steel pipe 16 20 Mizuno and Iiba (1992) prestressed concrete 14 20 Mizuno and Iiba (1992) steel pipe 4 20 Puri and Prakash (1992) reinforced concrete 18 56 Puri and Prakash (1992) reinforced concrete 18 56 Puri and Prakash (1992) reinforced concrete 18 56 Sy and Siu (1992) drilled shaft 20 27 Sy and Siu (1992) drilled shaft 20 27 Sy and Siu (1992) drilled shaft 20 27 Tuzuki et al. (1992) prestressed concrete 24 95 Kramer (1993) steel pipe 8 49 Kramer (1993) steel pipe 8 49 Kramer (1993) steel pipe 8 49 Carrubba and Maugeri (1996) drilled shaft 47 190 Han and Cathro (1996) drilled shaft 12 25 Imamura et al. (1996) precast concrete 18 49 pile group n.a. n.a. n.a. n.a. n.a. 2x2 n.a. n.a. 1, 4 2x2 1, 5 1, 6 n.a. n.a. n.a. n.a. n.a. n.a. 3x3 8, 16 3x3 n.a. 2x2 n.a. 2x2 2x2 2x2 7x8 2x3 1, 2, 4 1, 2, 4 2x2 n.a. n.a. n.a. n.a. n.a. n.a. 1, 2, 4 n.a. n.a. n.a. n.a. 2x3 1, 2x2 pile cap soil conditions test loading n.a. silty clay horizontal FV n.a. sand horizontal FV n.a. soft clay horizontal FV n.a. fat clay, sand vertical FV n.a. fat clay, sand vertical FV above grade fat clay, sand vertical FV n.a. plastic clay ringdown n.a. fine sand horizontal FV embed / above loose sand horizontal FV ? silty dense sand explosive n.a. sand FV / ringdown n.a. sand FV / ringdown n.a. silty sand horizontal FV n.a. silty sand ringdown n.a. sand, clay FV / ringdown n.a. silty sand horizontal FV n.a. sand, clay horizontal FV n.a. stiff OC clay horizontal FV above grade stiff OC clay vertical FV embedded loose sand ringdown above grade stiff OC clay horizontal FV above grade fill, loam, gravel horizontal FV above grade sand, clay horizontal FV n.a. sand, clay horizontal FV above grade sand horizontal FV at grade sand horizontal FV embedded sand horizontal FV partial embed clay, gravel, sand horizontal FV above grade silty clay horizontal FV embed / above loam, sand, clay horizontal FV embed / above loam, sand, clay horizontal FV embed / above loam, sand, clay horizontal FV n.a. clayey silt horizontal FV n.a. clayey silt vertical FV n.a. clayey silt ringdown n.a. fill, silt, sand horizontal FV n.a. fill, silt, sand vertical FV n.a. fill, silt, sand rocking FV above grade sand horizontal FV n.a. peat horizontal FV n.a. peat ringdown n.a. peat horiz. impact n.a. peaty clay horizontal FV above grade silty clay horizontal FV above grade sand horizontal FV pile spacing n.a. n.a. n.a. n.a. n.a. ? n.a. n.a. ? ? ? ? n.a. n.a. n.a. n.a. n.a. n.a. 3d 4.6d 3d n.a. 5d n.a. ? ? ? 2.6d 2.8d 3.8, 4.3d 3.8, 4.3d 3.8, 4.3d n.a. n.a. n.a. n.a. n.a. n.a. 2.8, 5.6d n.a. n.a. n.a. n.a. 2.8d 6.7d

pile head fixity ? ? ? cast cast cast fixed ? cast ? ? ? free free ? free cast free cast cast cast cast ? n.a. cast cast cast cast cast cast cast cast free free free free free free cast free free free free cast cast

freq. range (Hz.) 0.5 - 1.5 1-9 0.7 - 1.3 2 - 16 2 - 16 2 - 16 n.a. 3 - 10 2 - 40 n.a. ? ? 1.5 - 8.5 n.a. 5 - 55 2 - 14 2 - 60 1 - 15 5 - 95 n.a. 2 - 50 0.2 - 20 2 - 20 2 - 20 1 - 20 1 - 20 1 - 20 2.8 - 15 5 - 50 2 - 25 2 - 25 2 - 25 6 - 20 n.a. n.a. 6 - 30 10 - 110 20 - 55 1 - 20 1.8 - 2.9 n.a. n.a. 0 - 25 0 - 40 1 - 20

max. resonant accel. freq. damping (g) (Hz.) (%) ? 1 ? ? 5.8 ? ? 1 - 1.1 5 ? 8-9 5 ? 6-9 4-9 ? 8-9 10 ? 12.2 5.4 ? 4-5 7 0.066 13 - 18 10.5 - 39 ? 2-5 ? ? 4.2 -12.1 4 - 14 ? 5.1 - 16.2 9 - 12 0.265 1.7 - 2.9 1.5 - 7 ? 4.1 - 4.3 1.4 - 2.8 ? n.a. ? ? 8.1 7.7 ? 7.5 - 10 ? 0.1 2.1 - 2.3 10 - 11 ? 68 26 ? 3.8 - 6.3 5 - 15 ? 7.5 12 1.00 3 - 12 4 - 15 ? 13.8 - 15.8 10 - 25 ? 6 - 20 2 - 15 ? 7 ? ? 8.2 ? ? 9.8 ? ? 1.4 - 3.2 20 ? 22 ? ? 10 - 14 ? ? 10 - 15 ? ? 9 - 13 ? ? 10.3 ? ? 32.2 ? ? 11.5 ? ? 6.5 - 7.1 4 ? 46.5 - 50.4 5 ? 38.5 - 40.8 3 ? 10 - 13 ? 0.5 n.a. 15 - 20 0.15 2.3 - 5.5 25 - 33 ? 5.27 30 - 35 ? 12.5 ? ? 16 - 24 20 - 34 ? 3.2 - 6.3 4.1 - 12.8

Maximum pile head accelerations reached 0.265 g, which unlike many smaller amplitude tests, is representative of seismic loading. At higher loading levels, partial liquefaction was observed around the pile head, considerably reducing pile stiffness. Damping values were relatively small (1 - 7 %) and were observed to increase with the amplitude of pile motion, until the onset of liquefaction. Resonant frequencies observed in low level forced vibration and ringdown tests were considerably different, 2.3 to 2.9 versus 4.1 to 4.3 Hz, respectively. Ting (1987) computed p-y curves from the test results and

compared them to API recommended curves, which were found to overestimate the observed stiffness due to the nonlinear response, gapping, and partial liquefaction that occurred. In a recent publication, Lam and Cheang (1995) released cyclic load test data from a second test program conducted at the same site in order to compare dynamic p-y curves with cyclic p-y curves; this proprietary information had remained unpublished for a number of years. Tests were made on a pile newly installed, and on a pile previously subjected to vibratory loading; the load-deformation measurements of the two piles were nearly identical, indicating that the prior vibratory load did not result in permanent changes to the soil-pile system. Free-head resistance to lateral loading was found to be greater than fixed-head resistance, due to the mobilization of additional frictional resistance in the free-head rotational deformation mode; the authors assert this mechanism contributes to the diameter effect observed by Stevens and Audibert (1979). The soil-pile stiffness under cyclic loading compared very favorably to the low amplitude dynamic loadings, but nonlinear response under large amplitude dynamic loads reduced the apparent stiffness by 80 %. This was attributed to drained versus undrained soil behavior in the two types of tests.

167

Figure 4.8 - Field Pile Forced Vibration Test Set Up (after Scott et al., 1982) Crouse and Chang (1987) performed ringdown tests on vertical and battered concrete filled steel pipe pile-supported transformers with pile caps embedded in surficial loose, sandy, saturated soils. Observed resonant frequencies and damping values were less than those predicted by simplified numerical models by 10 30 %, and the low damping values in particular suggested suppressed pile cap-soil interaction. The authors observed that the site experienced peak ground accelerations of 0.06 - 0.1 g during the 1965 magnitude 6.5 Puget Sound earthquake, which may have induced settlement of the loose sandy soil away from contact with the pile cap. When ignoring cap embedment contact effects, predicted and observed values showed excellent agreement. Blaney et al. (1987) and Blaney and ONeill (1989) dynamically tested a 3x3 group of steel pipe piles driven into overconsolidated clay. The two publications report the results of vertical and horizontal forced vibration tests, respectively. A prime conclusion from the first study was that the average group pile frequency response was stiffer and

168

more damped than that of an equivalent single pile. This was concluded to be related to wave interference in the group, but was cautioned not to be taken as a universal result, but one highly dependent on the soil properties and pile spacing at this site. In contrast, under horizontal vibration, the average group pile frequency response was more flexible and less damped than that of an equivalent single pile. Numerical models incorporating the observed soil-pile gapping were found to more accurately capture the measured response. Kobori et al. (1991) conducted an extensive series of tests on a pile group with different pile cap contact/embedment conditions that consisted of horizontal forced vibration tests and earthquake observations, in order to evaluate both inertial and kinematic interaction effects. The pile group was composed of four drilled shafts, and is shown schematically in Figure 4.9a. The three pile cap conditions included no contact, grouted contact with the soil surface, and complete backfilled embedment. The forced vibration test results are shown in Figure 4.9b, indicating the strong influence of backfill embedment on group stiffness; damping values were not tabulated. At the completion of the forced vibration tests, the earthquake observations commenced; the maximum observed MHA at the site was 0.08 g. Transfer functions of pile cap to free-field ground surface motions are shown in Figure 4.9c, with decreasing amplitude at resonant frequency with pile cap contact/embedment. Impedance functions for the three pile cap conditions were derived, and using SHAKE (method A) and a finite element method (method B) to compute free-field input, motion at the top of the block was computed and compared favorably with the observed records, as shown in Figure 4.9d.

169

(a) Figure 4.9 - Field Pile Forced Vibration Test and Earthquake Observation: a) Test Set Up and Seismometer Arrangement; b) Forced Vibration Tests Results Illustrating Influence of Lateral Support Condition; c) Structure to Free Field Transfer Function for Three Backfill Cases; d) Observed and Computed Response Spectra for Seismic Event (after Kobori et al., 1991)

(c)

(d)

170

As part of a foundation investigation for a pile-supported bridge spanning a peatfilled slough near Seattle, Kramer (1993) performed forced vibration, ringdown, and impact tests on an 8 in diameter steel pipe pile. Unfortunately, the test results of different methods were inconsistent and in some cases ran contrary to expected trends of behavior, partially echoing Scotts findings; the lack of uniformity between test results could perhaps be attributed to shakedown effects. Radiation damping in excess of 25 % was recorded in the free vibration tests, and average horizontal stiffness was interpreted from the forced vibration test results. The latter value correlated reasonably well to static lateral load test results, and was therefore used in deriving the design dynamic stiffness of the pile groups. Finally, brief reference is made to other noteworthy experimental programs including Fuse et al. (1992) who dynamically tested a 7x8 pile group (the largest full scale group reported in the literature), and Mizuno and Iiba (1992) who reported on a welldocumented parametric study of pile cap embedment, pile spacing, number of piles, and soil nonlinearity on soil-pile dynamic response.

4.3 Model Scale Pile Test Programs Model pile tests have offered a wealth of information for SSPSI studies, but they must be carefully considered in the context of the particular scale model testing method and its inherent limitations. Scale model tests are economical, versatile, and conducive to parametric studies and repeatability tests. A technique known as modeling of models can improve the confidence of the modeling methodology. Both kinematic and inertial interaction effects may be studied, and pile groups with attached superstructures can be

171

readily constructed and tested. Principal limitations of scale model testing include the difficulty in fully satisfying all relevant scale modeling criteria, adequately replicating realistic soil-pile stress fields, and the boundary effects of test containers. This section will report on a wide variety of scale model testing programs dealing with SSPSI published in the literature; a small number of studies published in Japanese and as dissertations in England have not been reviewed here.

4.3.1 Model Pile Head Loading Tests A large number of tests on model piles are reported in the literature, and those tests consisting of static and cyclic lateral pile head loading are summarized in Table 4-3, with the inclusion of selected cases of axial head loading of model piles in clays. Wen (1955) was the first to report tests on model piles instrumented with strain gages in the literature, and focused his investigation on small groups of vertical and batter piles to ascertain distribution of axial and lateral loads. His findings regarding lateral load resistance of symmetric and asymmetric battered pile groups and the preferential distribution of load to front piles in the group were similar to Feagins (1937) test results on full scale pile groups tested in the field. As part of an integrated project investigating the lateral loading of piles in cohesive soils, Matlock and Ripperger (1957) conducted a suite of basic research tests. To appreciate the development of stress fields around a rigid cylinder laterally translating through an elastic medium, photoelastic studies of cylinders in gelatin were made and are shown in Figure 4.10. In stage (d) of the loading sequence, the rear face of the cylinder has separated from the gelatin and the resultant gap has lowered the stress in this zone;

172

Table 4-3 Model Pile Loading Tests max. # of cycles 1 1 1 ? 1 100+ 100 1 4 115 280 1 1 1 267 1 1 ? 1 1 1 ? 6 1 ? 500+ 1 1 2500 1 150 50 1 300+ 1 3 n.a free/fixed 2.5 - 7.5d free/fixed n.a. free n.a. free 48d fixed 4 - 12d fixed 3d free 3 - 10d free n.a. above grade n.a. n.a. above grade above grade above grade above grade inserted inserted inserted inserted pinned at base driven inserted inserted dry calcareous sand dry calcareous sand remolded clay soft clay soft/very soft soil dry sand silty clay marine clay static lateral static lateral static/cyclic axial static lateral static/cyclic lateral static lateral cyclic lateral static lateral 1 1 1006 1 ? 1 50 1

pile pile pile diameter wall length Reference model pile material (in) (in) (in) pile group Tschebotarioff (1953) wood rod - dolphins 1.625 - 2 n.a. 29 3, 7 Wen (1955) wood rod 1.5x1.5 n.a. 45 1x3 Matlock and Ripperger (1957) steel rod 1 n.a. 1-9 n.a. Gaul (1958) aluminum pipe 2.375 0.15 96 n.a. Shinohara et al. (1960) steel bar/steel pipe 0.7 - 11.8 ? 55.1 - 94.5 n.a. Matlock (1962) steel rod 0.375 n.a. 2 n.a. Prakash (1962) aluminum pipe 0.5 0.035 21 1, 2x2, 3x3 Kubo (1965) steel pipe/steel tube 0.7 - 4 ? 55.1 - 94.5 n.a. Davisson and Salley (1970) aluminum pipe 0.5 ? 21 1, 6, 96, 99, 245 Singh and Prakash (1971) aluminum tube 0.5x0.5 0.06 24 1, 2x2 Holmquist and Matlock (1976) aluminum pipe 1 0.03 40 n.a. Ranjan et al. (1977) aluminum pipe 0.4 - 0.8 ? 18.5 - 37.8 n.a. Allen and Reese (1980) aluminum pipe 1 composite 25 n.a. Georgiadis and Butterfield (1982) aluminum pipe 0.25 - 1 ? 24 n.a. Matlock et al. (1982) aluminum pipe 1 0.03 40 1, 6 Cox et al. (1983) stainless steel 1 0.03 2-8 1, 3x5 Franke and Muth (1985) PVC, PC, PE pipe 0.7 - 10.2 ? ? n.a. Kishida et al. (1985) acrylic resin pipe 0.8 ? 15.7 n.a. Maung (1985) PVC pipe 8.5 0.16 39.4 n.a. Meyerhof and Purkayastha (1985) steel pipe 0.5 ? 7.5 1 / 2x2 Meyerhof and Sastry (1985) steel pipe 2.9 0.28 43.3 n.a. Selby and Poulos (1985) aluminum pipe 0.6 0.05 20.5 2-9 Williams and Parry (1985) steel pipe 1.2 0.1 44.3 n.a. Smith and Slyh (1986) steel rod/steel pipe 0.7/2 n.a./0.19 3 - 25.5 n.a. Park (1987) concrete fill steel pipe 4.5 n.a. 93 - 104 1 / 1x2 Proctor and Khaffaf (1987) stainless steel pipe 1 ? 19.7 n.a. Franke (1988) PVC pipe 1.6 ? ? 1, 3, 8, 9, 16 Meyerhof et al. (1988) steel/wood/nylon pipe 0.5 ? 4 - 24 1 / 2x2 Shen et al. (1988) aluminum pipe 3.5 0.25 40 1 Shibata et al. (1989) aluminum/PVC pipe 0.8/0.9 0.06/0.09 31.5 2, 3, 4, 9, 16 Abduljauwad et al. (1990) steel pipe 1.1 0.06 30.5 n.a. Darr (1990) aluminum pipe 1 0.04 34 n.a. Adachi and Kimura (1992) aluminum tube 0.8 0.12 17.3 1, 1x2 Agaiby et al. (1992) drilled shaft 3, 6 n.a. 9 - 54 n.a. Mayne et al. (1992) drilled shaft 2 - 6.9 n.a. 10.5 - 41.3 n.a. Niiro et al. (1992) PVC pipe 1.5 0.16 11.3 1, 2, 3 Tanaka et al. (1994) steel pipe 4 0.13 78.7 1, 4, 6, 9 McManus and Chambers (1995) Poulos et al. (1996) aluminum pipe 1, 1.5, 2 0.05 - 0.08 14.8 - 26.5 n.a. Chen et al. (1996) aluminum pipe 1 0.05 26.5 2, 3, 4, 8 Bouckovalas (1996) aluminum pipe 0.75 0.12 15.7 n.a. Caliendo et al. (1996) aluminum pipe 1.315 0.331 60 n.a. Nagataki et al. (1996) reinforced concrete 4.9 n.a. 74.8 1x3 Gandhi and Selvam (1997) aluminum pipe 0.72 0.03 20 1, 2, 3, 4, 6, 9 Moss et al. (1998) aluminum pipe 1.315 0.133 60 1x5 Rao et al. (1998) aluminum/steel pipe 0.5,0.85,1 0.03 - 0.06 10 - 40 1, 2, 4, 6 pile spacing n.a. 4d n.a. n.a. n.a. n.a. 2 - 8d n.a. 3 - 4d 4d n.a. n.a. n.a. n.a. 1.8d 0.5 - 5d n.a. n.a. n.a. 3d n.a. 1.9 - 3.8d n.a. n.a. 6d n.a. 2d 3d n.a. 1.8 - 9.1d n.a. n.a. 2 - 5d n.a. n.a. 2.5, 3.5d 3d

pile head fixity pile cap pile installation model soil test loading fixed above grade driven sand/silty clay static lateral fixed above grade driven dry fine sand static lateral fixed n.a. inserted remolded clay static lateral free n.a. pinned at base bentonite clay static/cyclic lateral free n.a. embedded saturated sand static lateral fixed n.a. inserted remolded clay static/cyclic lateral cast above grade embedded dry sand static/cyclic lateral free n.a. embedded saturated sand static lateral free/fixed at / above grade embedded dry sand static/cyclic lateral fixed ? ? medium sand static/cyclic lateral fixed n.a. inserted remolded clay static/cyclic axial free n.a. inserted remolded clay static lateral free n.a. inserted stiff, soft clay static lateral free n.a. inserted kaolinite clay static lateral fixed above grade inserted remolded clay cyclic axial fixed above grade inserted kaolinite/bentonite clay static lateral free n.a. embedded dry sand static lateral free n.a. inserted dry sand / remolded clay cyclic lateral free n.a. inserted soil-cement slope face static lateral free/fixed above grade inserted clay, sand static inclined fixed n.a. inserted clay/sand static inclined fixed above grade embedded/inserted dry sand static/cyclic lateral free n.a. inserted dense sand cyclic lateral fixed n.a. inserted clay / loose/dense sand static torsional/lateral free/fixed above grade embedded loose / medium sand static/cyclic lateral fixed n.a. inserted remolded clay cyclic axial free/fixed above grade embedded dry sand static lateral free/fixed above grade inserted loose sand / soft clay static lateral free n.a. embedded silty clay / sand static/cyclic lateral free above grade embedded dry sand static lateral free n.a. embedded saturated sand static/cyclic lateral free n.a. inserted fire clay static/cyclic inclined free none embedded dry sand static lateral free n.a. drilled/cast-in place dry sand static/cyclic lateral free n.a. drilled/cast-in place kaolinite/supersil static/cyclic lateral free/fixed above grade ? silty sand static lateral cast above grade driven dry sand, gravel cyclic lateral

the authors correctly deduced that this behavior would only apply to the near surface region of cohesive soil-pile systems. Tests were also made of fully buried model pile segments laterally pulled through a remolded clay soil, and rigid vertical pile segments laterally loaded in the remolded clay. These two test methods were intended to evaluate the deep soil-pile and near surface soil-pile response to lateral loading. Rate of loading, length to diameter ratio, and embedment depth were varied, and the nonlinear loaddeformation response was compared with results of field tests. In his project summary report, Matlock (1962) extended the laboratory test technique to cyclic lateral loading, and performed pot tests of rigid cylindrical pile segments vertically embedded in remolded clay, shown schematically in Figure 4.11 along with a representation of a typical loading cycle. These tests, in conjunction with field studies, significantly

contributed to the current API recommendations for the construction of p-y curves in soft clay under static and cyclic lateral loads.

Figure 4.10 - Stress Fringe Patterns of Rigid Cylinder Laterally Translating in Elastic Medium (after Matlock and Ripperger, 1957)

174

Figure 4.11 - a) Schematic of Pot Test; b) Typical Loading Cycle with Slack Zone while Traversing Gap (after Matlock, 1962) Gaul (1958) provided a detailed analysis of scale model similitude in his reporting of static and cyclic lateral loading tests of model piles in bentonite clay. Figure 4.12a depicts data from 1 Hz cyclic loading tests with the interesting result of the bending moment envelope progressively decreasing with the application of an overburden pressure of 50 psf, and subsequent removal. A comparison of static and cyclic loading bending moment envelopes is provided in Figure 4.12b, indicating only minor cyclic degradation of lateral soil resistance in these tests; this result is no doubt a function of the very high plasticity index of the foundation medium, in this case 550 %. Prakash (1962) in his Ph.D. dissertation performed static and cyclic tests on groups of model piles embedded in sands. Pile spacing was varied from 2d to 8d in his tests, and he concluded that group effects were negligible for spacings greater than 8d in

175

the direction of loading, and greater than 3d normal to loading. He also observed that the effect of cyclic loading was to increase, at a decreasing rate, the deflections and moments under a constant load level. These findings have been quite influential, as they have been commonly cited for the design of pile groups.

Figure 4.12 - Model Pile Head Loading Test Bending Moment Diagram: a) Variation with Overburden Pressure; b) Dynamic and Static Loading (after Gaul, 1958) . Davisson and Salley (1970) performed lateral load tests on single model piles and

very large groups in sand, including battered piles, to develop design criteria for foundations for locks and dams for the U.S. Army Corps of Engineers. Their objective was to develop a better understanding of group behavior, so that model tests could be correlated to single pile field load tests, which could then be extrapolated to group design. They found that cyclic loading of a single pile caused deflections roughly twice that of static loading, and that bending moment distribution increased moderately at depth, but the maximum moment was unaffected. Tests on 6 pile groups (vertical/battered)

176

illustrated the effects of pile head fixity and cap contact with the soil surface. The largescale group tests again demonstrated increased deflections under cyclic load (150 % of static loading deflections) and analysis of group effects with Hrennikoffs (1950) method proved reasonably accurate. Allen and Reese (1980) conducted lateral load tests on model piles in soft clay (Test 1) and soft clay covered with a progressively stiffening upper layer (Tests 2 - 5). They compared their test data to results predicted by the COM623 computer program using default p-y curve criteria for stiff and soft clays. The match between predicted and observed results was good, and is illustrated in Figure 4.13. As observed by the authors, some inaccuracy may have resulted from the dependence of deflection on pile diameter, as suggested by Stevens and Audibert (1979), and borne out by lateral load tests on large diameter piles.

Figure 4.13 - Comparison of Experimental and Analytical Model Pile p-y Curves (after Allen and Reese, 1980)

177

Matlock et al. (1982) performed an important study of model pile groups in remolded clay instrumented to measure pore pressures and frictional resistance at the pile wall during installation, subsequent consolidation, and cyclic axial loading. The pore pressure measurements were consistent with consolidation theory, but showed inconsistent and minor changes during pile loading, not correlating well with effective stress concepts of degradation of frictional resistance. The authors therefore postulated that the mechanism of cyclic degradation was concentrated in a thin shear zone at some finite distance from the pile wall, as shown in Figure 4.14. Figure 4.15 depicts the reduction in shear transfer of a single pile over 100 cycles of constant deflection; Figure 4.16 illustrates the reduction of shear transfer over progressively increasing displacements. Group efficiency under vertical loading greater than 1.0 was measured, thought to be the result of consolidation between the closely spaced piles; per pile group displacements exceeded those of equivalent single piles.

Figure 4.14 - Shear Zone Behavior in Axially Loaded Model Pile in Remolded Clay (after Matlock et al., 1982)

178

Figure 4.15 - Shear Transfer Behavior During Cyclic Axial Loading of Model Pile in Remolded Clay (after Matlock et al., 1982)

Figure 4.16 - Shear Transfer Under Progressively Increasing Displacements During Cyclic Axial Loading of Model Pile in Remolded Clay (after Matlock et al., 1982)

Figure 4.17 - Group Efficiency As a Function of Pile Spacing As Determined by Model Pile Tests (after Cox et al., 1983)

179

Cox et al. (1983) described a parametric study of pile spacing and orientation to loading on lateral group efficiency using pile segments inserted into very soft clay. The pile segments were chosen for testing efficiency, and with the belief that the lateral loading effects are well modeled by considering only the near surface portion of the soilpile system. This approach is similar to Matlocks pot tests, although the rigid body pile deflection mode seems better suited to studying single pile response, as the interaction of a rigid pile group may be quite different than a flexible pile group. A summary of their computed group efficiencies is presented in Figure 4.17. Kishida et al. (1985) made x-ray pictures of cyclic lateral loading tests of model piles in sand and clay, illustrating gap infill and compaction behavior in sand, and a gap standing open in clay (see Figure 4.18). In a similar vein, Hughes and Goldsmith (1977) made a study of pile-soil displacement fields under lateral loading with a stereoscopic photogrammetric technique.

Figure 4.18 - Diagram of Laterally Loaded Model Soil-Pile Displacement Vectors Obtained by X-Ray Technique Illustrating Gap Infill in Sand and Open Gap in Clay (after Kishida et al., 1985)

180

Park (1987) published a comprehensive study of the seismic performance of steelencased concrete piles, primarily focused on the structural behavior of these composite members under lateral loading1. Both free-head and fixed-head intermediate-scale piles were tested in loose and medium dense sands. Observed trends included higher pile forces developed in medium than in loose sands, sand densification in a gap around the pile head, expansion and migration of the depth of peak curvature during cyclic loading, lead piles load-shadowing trailing piles, and the formation of distributed zones of yielding in the piles. Interestingly the increased soil stiffness due to densification in the gap was partially offset by the lowered soil surface and hence increased unsupported length of the pile; this effect is also a function of pile installation technique. Despite the formation of local buckling at low ductility levels, strength, ductility, and energydissipating characteristics were judged to be equal or superior to equivalent reinforced concrete members. Under contract to the Electric Power Research Institute, researchers at Cornell University undertook an extensive series of static and cyclic lateral loading tests on large model drilled shafts in sand and clay, as reported by Agaiby et al. (1992) and Mayne et al. (1992), respectively. In the case of the tests in clay soil, large test chambers were filled with clay slurry, which were then consolidated over a period of weeks and even months to obtain the desired overconsolidated strength profiles. The model drilled shafts were constructed with similar care to replicate soil/concrete interface roughness. The static test results showed nonlinear load deformation response, with no apparent yielding,

Other researchers who have studied the structural integrity of full-scale precast piles under seismic loading include Banerjee et al. (1987), Priestly and Park (1990), Kokusho et al. (1984), and Sheppard (1983); Meyersohn (1994) researched the structural performance of piles subject to liquefaction induced loading.

181

except in dense sands where post-peak softening occurred. Cyclic loading led to an increase in accumulated displacements, which increased at a decreasing rate, except in dense sand, which was considered a metastable condition. Subsequent post-cyclic static loading to failure showed no loss of lateral capacity, save a minor reduction in initial stiffness.

4.3.2 Model Pile Dynamic Tests A limited number of dynamic tests have been conducted on model scale piles and pile groups in field and laboratory conditions, and are summarized in Table 4-4. The experimental results have generally reinforced those obtained in field pile dynamic tests. Several researchers have prepared artificial clay soils as a testing medium, including Kana et al. (1986), who carefully considered scale model similitude in the development of their model soil and pile, and successfully compared the experimental test results with prototype test data, providing some validation for the use of 1-g dynamic model tests of piles in cohesive soils (using a scaling factor of 10.75). Novak has concentrated his experimental efforts on the dynamic testing of model pile groups, and most notably has conducted a test on a 102 pile group, which represents the largest full scale or model pile group whose dynamic response has been recorded (El Sharnouby and Novak, 1992). The layout of this closely-spaced group is shown in Figure 4.19. The group was placed in a hole excavated in the field, and a fly ash/sand mixture designed to have similar dynamic properties to the free-field was backfilled around the group. In this manner, boundary effects typically imposed by a laboratory test container were eliminated, but the experiment can be faulted for not properly replicating the soil

182

Table 4-4 Model Pile Dynamic Loading Tests

Reference Moore and Crossley (1972) Agarwal (1973) Novak and Grigg (1976) Novak and Grigg (1976) Novak and Grigg (1976) Novak and Grigg (1976) Novak and Grigg (1976) Richart and Chon (1977) El Sharnouby and Novak (1984) El Sharnouby and Novak (1984) El Sharnouby and Novak (1984) Butterfield and Khan (1985) Butterfield and Khan (1985) Kana et al. (1986) Kim et al. (1987) Gao et al. (1988) Stanton et al. (1988) Hassini and Woods (1989) El-Marsafawi et al. (1992) El-Marsafawi et al. (1992) Burr et al. (1997) 31 12 92.5 / 88.5 92.5 / 88.5 92.5 / 88.5 82 82 36 - 60 41.7 41.7 41.7 62.5 18.8, 31.3 59.3 60, 90, 120 16.5 45 78 114.2 114.2 ?

model pile material steel bar brass pipe steel pipe steel pipe steel pipe steel pipe steel pipe steel pipe / steel tube steel pipe steel pipe steel pipe steel pipe steel pipe aluminum pipe drilled shaft plexiglass rod stainless steel pipe steel pipe steel pipe steel pipe steel pipe

pile diameter (in) 0.94x0.06 0.4 2.4 / 3.5 2.4 / 3.5 2.4 / 3.5 2.4 2.4 3.5 / 4.5 1.1 1.1 1.1 1.3 1.3 1 6 1.2 1.2 2.4 4 4 1, 1.5, 2

pile length (in)

pile freq. resonant pile pile head pile range freq. group spacing fixity pile cap installation model soil test loading (Hz.) (Hz.) n.a. n.a. free n.a. inserted clayey silt vertical FV 15 - 30 ? n.a. n.a. free n.a. inserted plastellina, clay horizontal FV 15 - 30 27 n.a. n.a. fixed n.a. driven silty sand vertical FV 7 - 60 40 - 48 n.a. n.a. fixed n.a. driven silty sand horizontal FV 7 - 60 7-8 n.a. n.a. fixed n.a. driven silty sand rocking FV 7 - 60 7-8 2x2 7.5d fixed at / above grade driven silty sand horizontal FV 7 - 60 11 - 13 2x2 7.5d fixed at / above grade driven silty sand rocking FV 7 - 60 38 - 46 n.a. n.a. free n.a. inserted drained, saturated sand ringdown n.a. 7.1 - 20 102 3d cast above grade backfilled artificial sand vertical FV 6 - 60 32 102 3d cast above grade backfilled artificial sand horizontal FV 6 - 60 18 - 27 102 3d cast above grade backfilled artificial sand torsional FV 6 - 60 28 - 30 n.a. ? ? ? driven soft remolded clay horizontal FV 6 - 20 10 - 13 2x2 ? ? ? driven soft remolded clay horizontal FV 10 - 70 40 - 48 n.a. n.a. fixed above grade driven bentonite/aerosil/veegum horizontal FV 10 - 158 17 - 19 n.a. n.a. ? n.a. cast-in-place sand horizontal FV 10 - 1500 60 - 120 1, 2, 3 2 - 5d free/fixed above grade embedded fine to medium sand horizontal FV 10 - 70 35 - 60 n.a. n.a. free n.a. embedded dry sand horizontal FV 10 - 110 ? 2, 4 2d - 10d fixed ? backfilled sand vertical FV 5 - 60 22 - 25 1, 6 3, 4d cast above grade driven silty sand horizontal FV 10 - 70 16 - 18 1, 6 3, 4d cast above grade driven silty sand vertical FV 10 - 70 45 - 70+ 2x2 2.3 - 15d fixed above grade driven soft, stiff clay horizontal FV 5 - 21 12 - 20

damping (%) ? 12.4 ? ? ? ? ? 4 - 53.1 ? ? ? ? ? 0.9 - 5.5 ? 14 - 22 ? 5.4 - 6.8 ? 7.8 - 9.2 ?

displacement and densification that occurs during pile installation (though this is compensated by the fact that elastic continuum theories do not consider this effect). Forced vibration, impact, and static lateral load tests were conducted on the pile group, and a seismic cross-hole survey was made to verify the shear wave velocities of the freefield and backfill soil. The forced vibration and impact test results showed good

agreement, and even under higher force amplitudes the response was seen to remain linear elastic.

Figure 4.19 - Layout of 102 Model Pile Group Subjected To Dynamic Testing (after Novak and El Sharnouby, 1992)

Novak and El Sharnouby (1992) evaluated group effects from these tests with static interaction factors (Poulos), dynamic interaction factors (Kaynia and Kausel), complete dynamic analysis (Waas and Hartman), and an equivalent pier concept (Novak and El Sharnouby). Stiffness and damping of a single pile were computed by Novaks method with PILAY2, ignoring the top 10 cm of soil resistance, incorporating a weak zone around the pile, and proscribing a parabolic variation of soil modulus with depth.

184

The pile group impedance was then assembled with interaction factors for the first two methods, and computed directly for the latter two. For vertical response, the Waas and Hartman method, which transforms the pile group into an axisymmetric one for solution efficiency, provided the best match to the test results without modification. Good

performance of the other three models required consideration of the total or apparent mass, which included the soil between the piles, no modification to stiffness, and adjustment of damping as follows: x 10 % for static interaction, x 50 % for equivalent pier, and x 200 % for dynamic interaction methods, respectively. For horizontal

excitation, the Waas and Hartman and Kaynia and Kausel methods provided good results without modification. The equivalent pier model required consideration of apparent mass and reduced damping (x 40 %) to replicate the test data; the static interaction factors provided poor correlation to the test results. A summary of the observed horizontal response and model predicted behavior is shown in Figure 4.20.

Figure 4.20 - Experimental Model Pile Group Horizontal Response Curve Compared With Theoretical Models: P, Equivalent Pier; K, Kaynia and Kausel Interaction Factors; and W, Waas and Hartmann Direct Analysis (after Novak and El Sharnouby, 1992)

185

The conclusion that can be drawn from this work is that dynamic response of large closely spaced pile groups is a very complex matter, and currently available theories provide a fair estimation of response, with some modifications required. Considering static interaction only may provide an approximate estimate of dynamic group stiffness for small groups at low frequencies, but may otherwise underestimate stiffness. It

appears that the theories systematically overpredict damping, as they do not account for formation of a soil-pile gap or soil nonlinearity; inclusion of a weak zone surrounding the pile may alleviate this problem. For the particular soil conditions and pile spacing at this test site, the total mass of piles and intervening soil appeared to vibrate as a rigid body.

4.3.3 Model Pile Centrifuge Tests Centrifuge studies of scale model piles have provided a valuable means for understanding and validating aspects of SSPSI that are not readily accomplished by other experimental methods. A centrifuge apparatus consists of a rotating arm with an

experiment package fixed to a swivel at one end; the centrifugal acceleration of the rotating arm induces an elevated gravitational field onto the model, which swivels to a position normal to the arm (see Figure 4.21). The principal advantage of centrifuge testing is that the gravitational stress field in the model can replicate the prototype. This consideration is crucial when testing materials such as cohesionless sands whose stressstrain behavior is a function of confining pressure. In clay soils, where overburden stresses are not as significant, the centrifuge does offer the important capability of consolidating the deposit during spin-up, thereby achieving a more realistic soil strength profile. Centrifuge tests of model piles have been conducted with both head loading

186

schemes, and beginning in the 1980s, base excitation by shaker devices. Another recent innovation is the installation of piles in-flight while the centrifuge is spinning at the test g-level, so as to properly reproduce the displacement and stress fields around the model pile. Craig (1985) reviewed a range of model pile installation procedures and concluded that for model pile axial loadings in sand it is imperative that installation be carried out at appropriate acceleration levels.... He noted that for lateral, cyclic, and dynamic loadings, the effect may be less critical, and considered the undrained behavior of model piles in clay to be relatively insensitive to the installation method. He also stated that the method and rate of in-flight installation can be important variables. Although not

commonly observed, it would seem to follow that in-situ soil characterization should also be accomplished at the proper in-flight stress level. General limitations of centrifuge model studies include boundary effects from the test container, spatial and temporal variation in the induced gravity field, and only 1-D shakers are presently available. Nonetheless, centrifuge model pile tests have been employed successfully, particularly to study liquefaction phenomena, for which they are well-suited.

Figure 4.21 - Representation of Centrifuge Testing Scheme (after Scott, 1994)

187

A summary of centrifuge model pile test programs reviewed is given in Table 4-5, with a comment as to what degree scale model similitude has been addressed in the reference. It can be seen in this table that several references make use of a laminar or hinged box in an attempt to reduce container boundary effects. Fiegel et al. (1994) provide a discussion of this design consideration with the conclusion that all types of model containers, whether rigid or flexible, have unique dynamic characteristics which must be considered in the evaluation of test results. A popular technique is to design a box that has a stiffness profile equivalent to the expected soil stiffness profile during shaking; the challenge in this approach is to choose an appropriate soil stiffness profile representative of perhaps a range of shaking intensities. Unfortunately no researchers, including these ones, compare measured model site response with theoretical free-field response; this comparison would provide an excellent index of the containers effectiveness in suppressing boundary effects. Another feature apparent from Table 4-5 is that only 11 test programs with earthquake base excitation are reported; of these, only three have been in cohesive soils, two of which were conducted in hinged containers, and both only of moderate shaking intensity. A third point of interest is that several

researchers used model piles of rectangular cross section (noted as a bar in Table 4-5) which do not present the correct profile to the soil, resulting in altered soil-pile interaction from the prototype conditions. The conclusion to be emphasized is that to date, only one centrifuge testing program (at U.C. Davis) of model piles in soft clay has been reported with high level base excitation, an effective flexible model container, and rigorous consideration of scale model. Comments on particular experimental programs follow.

188

Table 4-5 Model Pile Centrifuge Tests

Reference Scott et al. (1977) Scott (1979) Scott (1981) Prevost and Abdel-Ghaffer (1982) Prevost et al. (1982) Scott et al. (1982) Zelikson et al. (1982) Barton (1984) Ko et al. (1984) Ko et al. (1984) Ting and Scott (1984) Luong (1984) Nunez and Randolph (1984) Oldham (1984) Finn and Gohl (1987) Steedman and Maheetharan (1989) Terashi et al. (1989) Chang et al. (1990) Bouafia and Garnier (1991) Cafe (1991) Gohl (1991) Hamilton et al. (1991) Lenke et al. (1991) Miura et al. (1991) Terashi et al. (1991) Kotthaus and Jessberger (1993) Levacher and Schoefs (1994) McVay et al. (1994) Stewart et al. (1994) Chacko (1995) Liu and Dobry (1995) Wilson et al. (1995) Abdoun et al. (1996) Abdoun et al. (1996) Fukuoka et al. (1996) Horikoshi and Randolph (1996) Pinto et al. (1997) Ohtsuki et al. (1998) Wang et al. (1998) ? ? ? ? ? ? ? rigid cylinder rigid cylinder rigid cylinder ? ? rigid cylinder ? ? ? ? ? ? ? ? ? ? laminar box rigid box ? ? ? ? hinged box ? laminar box laminar box laminar box laminar box rigid box ? laminar box hinged box

model pile material steel bar aluminum rod aluminum bar aluminum pipe aluminum pipe stainless steel pipe aluminum rod aluminum pipe aluminum pipe tapered wood rod stainless steel pipe steel pipe alum. pipe, polypropelene rod stainless steel pipe stainless steel pipe dural tube steel, aluminum, acrylic bar steel pipe aluminum pipe annealed steel rod/rubber stainless steel pipe aluminum rod aluminum pipe stainless steel pipe steel, aluminum, acrylic bar aluminum pipe aluminum pipe aluminum pipe brass rod aluminum pipe brass pipe stainless steel pipe brass pipe PE rod steel pipe brass pipe aluminum pipe, some battered aluminum pipe aluminum pipe

pile pile pile pile diameter wall length pile pile head structural pile (in) (in) (in) group spacing fixity pile cap mass installation model soil 0.17x0.24 n.a. 7.9 n.a. n.a. free n.a. yes inserted dry sand 0.16x0.16 n.a. 8 n.a. n.a. free n.a. yes inserted saturated sandy silt 0.24x0.14 n.a. 8 n.a. n.a. free n.a. yes inserted saturated fine sand 0.22 0.02 4.8 - 10.6 1, 2x2 4.5 - 18.3d free/fixed above grade no ? loose/dense/dry/sat. sands 0.22 0.02 9.9 n.a. n.a. free n.a. yes ? loose/dense/dry/sat. sands 0.5 0.01 8 n.a. n.a. free n.a. no inserted dry, saturated silty sand 0.4 n.a. 4.7 10 ? fixed above grade yes ? saturated sand 0.375 - 0.63 ? ? 1, 2, 3, 6 2 - 8d fixed n.a. no ? saturated fine sand 0.11 - 0.22 ? 5.2 - 10.3 3x3 3d fixed above grade no driven in flight kaolinite clay 0.14 - 0.28 n.a. 4.8 - 9.6 2x4 3d fixed above grade no driven in flight dry sand 0.53 0.01 10 1, 2, 4 2 - 7d free/fixed above grade no inserted saturated sand 0.53 0.1 13.8 2x4, 4x2 ? free/fixed above grade ? ? ? 0.42 - 0.54 0.01 9.3 - 15.6 n.a. n.a. free above grade yes installed in flight consolidated clay 0.75 0.05 13.5 n.a. n.a. free n.a. no driven in flight dry sand 0.375 0.01 8.25 1x2 2 - 6d free/fixed above grade yes inserted loose/dense dry sand 0.375 ? 8 n.a. n.a. free above grade yes embed/inserted dry sand 0.4 - 1.3x0.1 - 0.3 n.a. 3.9 - 16.1 n.a. n.a. free n.a. no fixed at base uniform dense sand 0.375 0.03 4 2x2 10.7d fixed above grade yes inserted medium dense dry sand 1.1 - 1.8 0.04 - 0.08 11.8 - 13.2 n.a. n.a. free n.a. no embed/inserted medium dense, dense sand 0.25 n.a. 8 2x4 8d fixed above grade yes inserted peat 0.375 0.01 8.25 1x2, 2x2 2 - 6d free/fixed above grade yes inserted loose/dense dry sand 0.52 - 1.1 n.a. 4.5 - 9 n.a. n.a. pinned/fixed n.a. no inserted kaolinite clay w/free water 0.25 0.04 5 n.a. n.a. free n.a. no ? dry uniform silica sand 0.38 0.01 8.4 2x2 2.5d fixed above grade yes inserted dry/saturated sand 0.4 - 1.3x0.1 - 0.3 n.a. 3.9 - 16.1 n.a. n.a. free n.a. no fixed at base sloping sand 1.2 0.08 23.6 1x3 3, 4d free above grade no embedded dry fine sand 1.1 ? 15.4 n.a. n.a. free n.a. no embedded dry sand 0.375 0.04 11.5 1 / 3x3 3, 5d free / fixed above grade no driven in flight loose, medium dense dry sand 0.125x0.125 n.a. 8.9 2x7 4.9, 6.5d fixed above grade no ? soft clay, dense sand 0.25 0.03 12 n.a. n.a. free above grade yes inserted dense sand, remolded Bay Mud 0.375 0.014 6.625 n.a. n.a. free / fixed n.a. no embedded medium dense saturated sand 0.875 0.05 22 1, 2x2 4d fixed embed / above yes driven loose, dense saturated sand 0.375 0.01 7.1 n.a. n.a. fixed n.a. no fixed at base layered saturated sand 0.375 n.a. 8.7 n.a. n.a. fixed n.a. no ? layered saturated sand 0.6 0.01 18.1 2x2 5.1d fixed above grade yes pinned at base saturated sand 0.125 0.028 6 5 - 69 5d, 8d fixed embed / above yes inserted kaolin clay 0.375 0.035 11 1 - 21 3, 5d free / fixed above grade yes driven in flight loose, medium dense dry sand 0.4 0.04 10 2x2 15d fixed embedded yes pinned at base oil saturated sand, gravel 0.25 0.03 8.5 n.a. n.a. free n.a. yes inserted dense sand, remolded Bay Mud test container

max. centrifugal accel. model accel. (g) input excitation (g) similitude 100 free vibration head loading n.a. yes 100 cyclic lateral head loading n.a. yes 100 cyclic lateral head loading n.a. no ? FV head load -> 1200 Hz. n.a. no 100 free/forced vibration head loading n.a. yes+ 50 FV head load -> 500 Hz. n.a. yes 100 explosive excitation 0.8 no 30 - 120 cyclic lateral head loading n.a. yes 50, 70, 100 static axial loading n.a. yes+ 50, 70, 100 static axial loading n.a. yes+ 48 free/forced vibration head loading n.a. no 100 EQ head loading 0.5 yes+ 100 static/cyclic axial tension loading n.a. yes 52.5 static/cyclic lateral head loading n.a. no 60 EQ base excitation 0.15 yes+ ? base excitation ? no 20 - 75 cyclic lateral head loading n.a. no 50 base excitation/cyclic head loading 0.2 no 18 - 20 static lateral head loading n.a. no 60 Loma Prieta base excitation 0.48 yes 60 sine waves/EQ base excitation 0.14 yes+ 46 - 93 static lateral head loading n.a. no 60 FV vertical head loading ? no 50 El Centro base excitation 0.25 yes 25 - 75 static lateral head loading n.a. no 50 static lateral head loading n.a. no 18 cyclic lateral head loading n.a. yes 45 static lateral head loading n.a. no 110 in-flight embankment loading n.a. no 50 Santa Cruz, Landers base excitation 0.32 no 40 base excitation + lateral head loading 0.4 no 30 EQ base excitation 0.55 no 40 cyclic lateral head loading n.a. no 50 base excitation 0.2 no 45 El Centro base excitation 0.11 no 100 static lateral head loading n.a. yes 45 static lateral head loading n.a. no 25 Tokachi-oki base excitation 0.2 yes 50 Kobe base excitation 0.25 no

Scott et al. (1977) were the first to test model piles in the centrifuge, and conducted a pilot study of free vibration tests on steel bars instrumented with strain gages inserted in dry sand. Basic scaling relations and trends of damping and subgrade reaction were observed. In following work, Scott (1979) improved the loading system to deliver cyclic lateral loading to the pile head; degradation of soil resistance was clearly discerned. In Scott (1981), the centrifuge experiment was designed to model Reese et al.s (1974) Mustang Island full scale pile tests, with moderate success (see Figure 4.22). Scott et al. (1982) carried out a parallel field/laboratory program of dynamic forced vibration pile tests. The field test was discussed in section 4.2.3, and a typical

comparison of prototype and model results is shown in Figure 4.23. The discrepancy can possibly be attributed to pile installation effects, container boundary reflections, and variable pore pressure response, illustrating some of the difficulties in centrifuge modeling. Ting and Scott (1984) investigated small pile group lateral interaction factors which they judged to compare favorably to Poulos elastic interaction factors; they found that dynamic pile group interaction was less pronounced than cyclic interaction. They did note pore pressure dissipation was more rapid in the model test than the prototype, and different gapping and gap infill behavior than observed in their field tests. Prevost et al. (1982) reported on centrifuge experiments of model piles excited by forced and free vibration tests, including small pile groups. Static lateral loading of the model piles again designed to replicate the Mustang Island tests resulted in p-y curves roughly the stiffness of the prototype. Under dynamic loads, nonlinear soil-pile

response, frequency dependent stiffness, and frequency independent damping were detected. Theoretical stiffness predictions were not well met, and the authors suspected

190

Figure 4.22 - Laterally Loaded Model Pile Centrifuge Test Data Compared with Prototype Results of Mustang Island (MI) Test (after Scott, 1981)

Figure 4.23 - Centrifuge Test Model Pile Forced Vibration Displacement and Bending Moment Response Compared with Prototype (P9) Test Results (after Scott et al., 1982)

191

that waves reflected from the rigid walls of the test container may have corrupted their results. As researchers developed ways to deliver earthquake shaking to centrifuge models, Zelikson et al. (1982) advanced a novel approach which consisted of generating a programmed series of explosions inside an echo box equipped with filters situated adjacent to the model container on the centrifuge. The resultant earthquake-like

excitations could be scaled in magnitude and frequency content. Barton (1984) investigated Poulos static group lateral interaction factors in her centrifuge tests, and found nonlinear response that did not correlate well with Poulos elasticity-based theory. The author suggested that nonlinear soil-pile response initiates with the onset of yielding in tension behind the displacing pile, which occurs at much smaller strains than compressive yielding in front of the pile. In addition, Barton

performed tests with three differently scaled models at three g-levels in a technique known as modeling of models to successfully validate the scale modeling technique. Oldham (1984) was the first to laterally load piles installed in flight, and lateral load results from piles installed at 1 g and at 52.5 g illustrate a measurable difference in stiffness (see Figure 4.24). Static and cyclic p-y curves were constructed, and stiffer response under cyclic loads was observed, postulated to be the result of sand densification in the soil-pile gap. Finn and Gohl (1987) implemented a base shaker device that imparted earthquake and sine wave motions to their centrifuge models; they also positioned bender elements in their models to measure in-situ in-flight soil moduli. Excitations were moderate so that elastic response would be achieved. Results from single pile tests showed that pile

192

head and free-field motions were magnified relative to the base input. The predominant period of the pile head response exceeded the free-field, indicating that the pile filtered out high frequency components of the ground motion and was dominated by the inertial response. Tests of small pile groups revealed no interaction effects for offline shaking, and increasing interaction and influence on the bending moment diagram as pile spacing was varied from 6d to 2d. In his Ph.D. dissertation, Gohl (1991) provided a detailed examination of the complete series of centrifuge model pile group tests, with the conclusion that elastic interaction factors underestimated group deflections at close spacings (d/b < 4) and overestimated interaction for larger spacings. He also compared cyclic p-y curves to those constructed according to API recommendations, with the latter implying substantially stiffer response, particularly at depth. Simulation of the single pile experimental results with the computer code SPASM 8 was very successful.

Figure 4.24 - Influence of In-Flight Pile Installation on Subsequent Load Deformation Response of Model Pile in Centrifuge Test (after Craig, 1985)
193

Terashi et al. (1989) performed static lateral head loading of model piles in sand in a modeling of models approach at six different scale levels, obtaining remarkably consistent results. Significantly, the researchers recognized the influence of pile diameter on the results of their calculated p-y curves, and they concurred with Stevens and Audibert (1979) that soil-pile resistance is proportional to the square root of pile diameter. Hamilton et al. (1991) were the first to report on centrifuge tests of laterally loaded piles in clay. Much of their analysis focused on the computation of ultimate soil resistance, incorporating mechanisms of soil-pile suction and adhesion in their model. Normalized experimental p-y curves were compared with curves constructed by Matlocks soft clay criteria (1970). In response to observed pile performance and site response during the 1989 Loma Prieta earthquake, a joint research project by CUREe and Kajima Corporation (1991) performed two series of centrifuge tests of model pile groups in liquefiable sands and soft peat. The first project endeavored to replicate excess pore pressure generation and

liquefaction of the sandy test soil, but was challenged by the model scaling laws which imply different rates of pore pressure dissipation in the model and prototype. This is typically overcome by using a finer sized soil or more viscous fluid in the model, but these can also influence the soil constitutive relationship and the pore pressure response. The experimental approach taken was to use a uniform fine sand. Special features of these tests were that they were conducted in a laminar box, a set of stacked rings that is free to translate horizontally, thereby reducing boundary effects, (see Figure 4.25), and that they used a base shaker to impart earthquake motions to the model. The piles were

194

inserted into the soil at 1 g, which raises the question as to whether correct stress fields and realistic pore pressure response were truly modeled. Although purely horizontal input motion was delivered to the model, a substantial component of vertical acceleration was measured in the tests. General trends of site amplification, nonlinear pile response, and pore pressure response were observed, and a 2-D effective stress model (TARA-3) was applied to the results. Increasing bending moments toward deeper positions on the pile suggested that liquefaction-induced kinematic loads dominated the pile response.

Figure 4.25 - Laminar Box for Centrifuge Testing (after Hushamand et al., 1988) The second portion of the research project is reported by Cafe (1991) in her Masters thesis. In these tests, a model of the Struve Slough Bridge, which suffered major damage in the Loma Prieta earthquake, and which was supported on peaty soil, was tested. Tests of a remolded peat deposit without a structural model were first made to calibrate the dynamic properties of the peat with iterative SHAKE analyses. The bridge tests consisted of 8 model piles supporting a single span bridge deck, in braced and unbraced configurations, to represent conditions of lateral restraint near the

195

abutments and no restraint near the center of the bridge. In the braced tests, the measured deck motions were comparable to the input motions, and large pile moments were observed both at the pile head and beneath the ground surface, indicating large kinematic loading from the soil. In the unbraced configuration, deck accelerations were of lower magnitude and longer period, similar to the surface free-field, and pile moments were much larger than the braced tests. A simplified finite element model of the soil-pile response was applied with fair agreement to the observed response. McVay et al. (1994) performed centrifuge tests of laterally loaded 3x3 pile groups driven in-flight in loose and medium dense sand. Figures 4.26a - d provides direct experimental evidence of the effects of a) soil relative density, b) load distribution among lead, middle, and trailing piles, c) pile spacing, and d) in flight installation. They found their test results compared very favorably with those of Brown et al.s (1988) tests on a full scale pile group in sand (see section 1.2.2) and that the p-multiplier method for group effects matched their results closely, with minor adjustment of the multiplier factors. They also investigated Reeses (1984) pile group equivalent pier concept, which was found to substantially overpredict observed group deflections. Pinto et al. (1997) used the same apparatus to conduct tests with varied pile head fixity and batter. Pile group efficiencies were verified to be independent of soil density but a function of spacing. Batter pile performance was found to be related to vertical dead load. Hoit et al. (1997) described the FLPIER computer program for analysis and design of pile supported bridge piers, which incorporated the results of the concurrent centrifuge testing program for use as pile design and group interaction parameters.

196

(a)

(b)

(c)

(d)

Figure 4.26 - Centrifuge Modeling of Laterally Loaded Pile Groups in Sand: a) Effect of Relative Density on Group Capacity; b) Load Distribution By Rows; c) Effect of Pile Spacing on Total Lateral Resistance; d) Influence of Acceleration Level During Driving on Total Lateral Resistance (after McVay et al., 1994) The centrifuge testing facilities at U.C. Davis have been the site of a series of research projects dealing with SSPSI. Chacko (1993) described model tests of single piles embedded in remolded Bay Mud in a hinged container on the small centrifuge, and analyzed the results with the free-field and pile response computer codes SRANG and NONSPS from Kagawa (1980, 1983). The analyses showed only fair agreement with the

197

test results (related to the limitations of the codes), and emphasized the dependence of the soil-pile interaction analysis on the accuracy of the computation of the free-field motions. Wang et al. (1998) applied several numerical codes to Chackos centrifuge test results, including PAR, NONSPS, and a p-y formulation in the general finite element code DRAIN-2D. They demonstrated that the SHAKE free-field analysis was superior to the SRANG results, and that a Novak-Nogami representation of radiation damping in the DRAIN-2D model provided the most accurate simulation of the centrifuge data (see Figure 4.27). In fact the results were found to be fairly sensitive to the implementation of radiation damping; placing the linear viscous dashpots in series with the hysteretic p-y springs was found to be superior to a parallel arrangement. Wilson et al. (1995) performed model tests of pile supported structures in liquefiable sands on the large centrifuge facility at U.C. Davis. This machine is equipped with a 1-D base shaker, an equivalent shear beam laminar box to suppress boundary effects, and special container end conditions to provide complementary shear stresses and reduce unintended vertical accelerations. There is however no capability for driving piles in flight. An extensive series of tests was made of single piles and small groups in saturated sands with a range of shaking intensities and development of very high excess pore pressure ratios, and in some cases consequent liquefaction. Pile motions were dominated by inertial forces from the superstructure, pile cap embedment had a significant effect on response, and peak pile bending moments during liquefaction events occurred close to the soil surface. Further work is reported by Wilson (1998) including the successful derivation of p-y curves from the experimental data and the favorable comparison of the tests results to the Caltrans pseudo-static analysis method.

198

Figure 4.27 - Comparison of Centrifuge Test Experimental and DRAIN-2D Computed Acceleration Response Spectra at Pile Head and Superstructure (after Wang et al., 1998)

199

4.3.4 Model Pile Shaking Table Tests Shaking table tests of model piles have provided another important means for understanding and validating SSPSI effects. The principal feature of shaking table tests are that they are conducted in a 1-g environment, and therefore cannot achieve the elevated stress field suitable for tests of cohesionless soils, as in centrifuge tests. The 1-g test environment is most appropriate for tests involving cohesive soils, whose undrained stress-strain behavior is not dependent on confining pressure. This has, however, not prevented researchers from conducting the great majority of shaking table tests on problems of pile response to liquefaction. Like centrifuge tests, shaking table model tests are sensitive to boundary effects imposed by the test container, and hinged, laminar, and shear boxes have been employed to overcome these effects. Researchers have rarely published comparisons of recorded shake table and theoretical free-field site response, which would be indicative of the containers effectiveness in suppressing boundary effects. The greatest advantage of shaking table model tests is that a number of

experimental facilities have the capability for two- and three-dimensional shaking, a distinctly more realistic condition than the 1-D shaking capability presently offered by centrifuges. Shaking table model pile tests do not require special in-flight procedures for pile installation or verification of in-situ soil properties. Scale model similitude is more complex in shaking table testing than in centrifuge testing, as will be discussed in Chapter 5. A summary of shaking table model pile test programs reviewed is given in Table 4-6, with a comment as to how scale model similitude has been addressed in the reference.

200

Table 4-6 Model Pile Shaking Table Tests

Reference Prakash and Aggarwal (1965) Tajime et al.(1965) Kubo (1969) Yamashita and Inatomi (1970) Hakuno et al. (1977) Tatsuoka et al. (1978) Sugimura (1980) Yao (1980) Kagawa and Kraft (1981) Mizuno and Iiba (1982) Ranjan et al. (1982) De Alba (1983) Heng-Li (1985) Korgi (1986) Gohl and Finn (1987) Mizuno et al. (1988) Stanton et al. (1988) Tamori and Kitagawa (1988) Yoshikawa and Arano (1988) Gohl (1991) Liu and Chen (1991) Nomura et al. (1991) Yan et al. (1991) Finn and Gohl (1992) Mori et al. (1992) Taga et al. (1992) Tazoh and Shimizu (1992) Tokida et al. (1992) Yamamoto et al. (1992) Yao and Kobayashi (1992) Sreerama (1993) Kagawa et al. (1994) Ohtomo and Hamada (1994) Sakajo et al. (1995) Dou and Byrne (1996) Hideto et al. (1996) Ohtomo (1996) Makris et al. (1997) Tao et al. (1998) n.a. 2x2 3x? 2x2 2x2, 2x3, 3x3 2x2, 2x3, 3x3 1x2 1, 3x3 n.a. 1x2 2x2, 3x3, 4x4 2x2 10 2x? n.a. 1x2 n.a. 2x? 2x2 1, 1x2, 2x2 3x9, 5x9, 9x9 2x? n.a. 1x2, 2x2 3x3 3x3 3x3 1 - 16 8 2x2 1, 1x2, 2x2 1x2 n.a. 6x6 n.a. n.a. 1x2 n.a. n.a.

pile pile pile diameter wall length model pile material (in) (in) (in) aluminum pipe 0.6 0.1 13 steel pipe 1 - 1.2 0.04 - 0.06 7.9 steel pipe 3.9 0.1 118 steel bar 3.9x0.4 n.a. 45.3 aluminum rod 0.8 n.a. 27.6 aluminum rod 0.8 n.a. 27.6 steel bar 2x0.2 n.a. 28.2 aluminum pipe 0.4 0.04 21.3 steel bar 2.4x0.2-0.4 n.a. 31.5 steel bar 2x0.2 n.a. 28.2 ? 0.3 ? 13 brass pipe 1 ? 2.5 acrylic pipe 0.3 0.1 12.6 steel bar 2x0.2 n.a. 28.1 aluminum pipe 0.25 0.05 24 steel bar 2x0.2 n.a. 28.1 stainless steel pipe 1.2 0.19 45 steel bar 2x0.2 n.a. 28.1 aluminum pipe 4 ? 21 aluminum pipe 0.25 0.05 16 aluminum pipe 0.4 ? 11.8 ? 1.9 - 3 ? 78.7 aluminum pipe 0.25 - 0.5 ? 13.8 aluminum pipe 0.25 0.05 24 aluminum rod 2 n.a. 39.4 aluminum bar 0.2x0.4 n.a. 15.7 aluminum pipe 1.2 0.04 35.4 PVC pipe 0.9 ? 31.1 brass pipe 0.8 0.12 26.3 aluminum tube 1x2 0.08 68.9 aluminum pipe 0.44 0.03 17.5 steel rod 2x0.24 n.a. 36.1 polycarbonate, PE pipe 0.7 - 1 ? 24.4 - 26 plastic pipe 1 0.1 35.4 aluminum pipe 0.25 0.03 15 acrylic resin 1.2 0.08 23.6 aluminum, stainless pipe 1 ? 27.6 aluminum pipe 0.8 0.04 38.2 steel pipe 12 0.25 240 pile group

pile max. pile head structural accel. model spacing fixity pile cap mass pile installation model soil test container input excitation (g) similitude n.a. free none ? embedded sand rigid box 2 -10 Hz. ? no ? fixed embedded yes fixed at base sand/plaster/water rigid box 1.4 - 8 Hz. ? no ? fixed above grade yes fixed at base cinder sand/oil rigid box 1 - 4Hz. 0.1 yes 50d fixed above grade yes inserted dry sand ? sine waves 0.15 no 3d fixed above grade yes fixed at base saturated loose sand hinged box 10 - 20 Hz. 0.13 yes 3d fixed above grade yes fixed/free at base saturated loose sand hinged box 5 - 45 Hz. 0.1 no ? ? embed / above yes ? polyacrylamide/bentonite ? 3 - 15 Hz. 0.1 no ? free/fixed above grade yes inserted clayey silt hinged box 4 - 45 Hz. 0.1 no n.a. free ? yes ? saturated sand hinged box 10 Hz. 0.1 no ? fixed embedded yes pinned at base polyacrylamide/bentonite rigid cylinder sine waves, Japan EQ 0.1 yes 2 - 5d ? ? ? inserted saturated sand rigid box 4 Hz. 0.3 no 2.5d fixed above grade n.a. inserted saturated medium sand laminar cylinder cyclic ? no 2 - 2.7d fixed above grade yes ? silicon gum ? 2 - 24 Hz., Japan EQ ? yes ? fixed embedded yes pinned at base polyacryl./bentonite/fly ash rigid cylinder 1 - 30 Hz., Japan EQ 0.12 no n.a. free above grade yes inserted dense dry sand rigid box 5 - 50 Hz. 0.6 no ? fixed embed / above yes ? polyacrylamide/bentonite rigid cylinder 1 - 30 Hz., EQ 0.12 yes+ n.a. free ? ? embedded dry sand flexible cylinder 1 - 40 Hz. ? no ? hinged embedded yes pinned at base plasticine/oil, polyacr./ben. rigid cylinder 0 - 30 Hz., EQ 0.8 yes 4.4d ? embedded yes ? loose saturated sand shear box 10 - 25 Hz., w/ vertical 0.1 yes 2 - 8d fixed above grade yes inserted dry sand rigid box 5 - 70 Hz., Taft 0.69 yes+ 3d ? ? ? inserted saturated sand ? ? 0.6 no ? ? above grade yes embedded saturated clean silica sand shear box Taft, El Cento EQ 0.3 yes n.a. free none yes ? saturated uniform fine sand hydraulic gradient 10, 20 Hz. 0.51 yes 3 - 8d fixed above grade yes inserted dry sand rigid box 7.5, 10 Hz. 0.45 no ? fixed above grade yes fixed at base saturated sand shear box EQ 0.14 yes 22d pin partial embed yes ? dry sand rigid box 1 - 10 Hz. 0.2 yes 2.5d fixed above grade yes fixed at base dry sand shear box 1 - 30 Hz. 0.25 no ? free none no fixed at base sloping saturated sand rigid box 2 Hz. 0.25 no ? free/fixed above grade yes fixed at base N-methylol propene, sand shear box El Centro EQ ? no 15d ? above grade yes embedded saturated sand shear box 1 - 7 Hz. 0.3 no 3 - 8d free/fixed above grade yes inserted low plasticity remolded clay rigid box 1 - 12 Hz. ? no 16d fixed above grade yes pinned at base saturated sand shear box 0 - 25 Hz. 0.16 no n.a. free none no fixed at base sloping saturated sand rigid box 10 Hz. 0.68 no 2.5d fixed above grade no fixed at base saturated sand/gravel shear box 10 Hz. 0.1 no n.a. free above grade yes inserted saturated uniform fine sand hydraulic gradient 17.5, 30 Hz 0.49 no n.a. free n.a. yes embedded dry sand shear box 10 - 60 Hz., EQ 0.3 no 10d fixed embedded no ? saturated sand rigid box 10 Hz. 0.4 no n.a. fixed embedded yes fixed at base dry sand shear box frequency sweep ? no n.a. free n.a. no pinned at base loose moist sand laminar shear box sweep 1-20 Hz., KPI EQ 0.45 no

A feature immediately apparent from Table 4-6 is that the great majority of shaking table test programs have studied the seismic response of piles in cohesionless soils, with very few studies conducted in cohesive soils. Single piles and small groups have been tested, with only two tests reported of larger groups. A number of tests have used flat bars as model piles, which as noted earlier, present an incorrect pile crosssection to the soil. A recent trend toward using shear boxes for the model container can be distinguished. Input motions have primarily consisted of sine waves, with a limited number of earthquake record base excitations, and only a handful of cases could be considered to be at strong levels of shaking. Reporting of adherence to scale model similitude has been inconsistent and often incomplete. To date, no shaking table testing program of model piles in soft clay has been reported with a high level earthquake base excitation, an effective flexible model container, and rigorous consideration of scale model similitude. Comments on particular experimental programs follow: Kubo (1969) was the first to report on shaking table model pile tests conducted with attention to scale model similitude. The results of his large model tests emphasized the kinematic component of soil-pile interaction, and he detected bending moment profiles and deflections consistent with computed prototype behavior. Yao (1980) drove groups of aluminum pipes into silty clay and performed static lateral load tests in addition to shaking table tests. The static test results showed evidence of pile group behavior, but they did not directly correlate to the observed dynamic pile group response. Strong resonant dynamic response was observed when the natural

frequencies of the superstructure and soil deposit coincided. Kagawa and Kraft (1981) proposed a nonlinear p-y type analysis method that

202

incorporated pore pressure response, and compared their analytical technique against shaking table test results of model piles in sand. Through the shaking events, the

measured fundamental resonance frequency of the soil-pile-structure system decreased from 34 to 4 Hz, passing through the resonant condition of the 10 Hz input motion. This behavior illustrates the dramatic softening behavior of the system during liquefaction. The numerical model compared favorably to the experimental results in the early and late stages of the loading sequence, but underestimated the response during the onset of liquefaction. Mizuno and Iiba (1982) were the first to subject their models to earthquake time history base excitations, and published a comprehensive discussion of scale model similitude. They attempted to fabricate an elastic soil medium with a mixture of

polyacrylamide and bentonite, and used model piles of rectangular cross-section supporting single mode model structures. They simulated the response of structures with mat, fixed pile head, and free pile head foundations, and discovered important variability in response (Figure 4.28a). Greater forces were detected in the fixed head model piles, and higher earth pressures on the embedded mat foundation. Parametric studies of three building models confirmed the effects of building frequency on dynamic interaction, with pile response dominated by kinematic interaction effects at the relatively low levels of shaking in these tests. They also conducted tests on a model that corresponded to an eleven story prototype structure, and compared the experimental results to the actual recorded motions at the site. Figure 4.28b depicts recorded and experimental

building/pile tip transfer functions that compare favorably, given the simplifications made in the model construction. Korgi (1986) extended these tests to investigate the

203

effects of different soil deposits and pile cap embedment on the system response. A trench around the footing removed the lateral resistance afforded by the pile cap, and this condition was found to have the strongest system response and induced the greatest pile bending moments. The structures fixed base period was also found to lengthen due to SSPSI effects.

(a)

(b) Figure 4.28 - SSPSI Shaking Table Model: a) Influence of Three Foundation Conditions on Superstructure Response; b) Comparison of Experimental and Recorded Seismic Response (after Mizuno and Iiba, 1982) Stanton et al. (1988) fabricated a large cylindrical shear bag 48 in diameter and 48 in tall to conduct shaking table tests of model piles in dry sand. Shortcomings in the container design and loading system performance contributed to experimental difficulties, but general trends of static and dynamic response were obtained and found to be in agreement with established methods of analysis. The authors correctly observed that pile

204

head-loading response is most sensitive to the near-surface soil properties, and pile seismic response is sensitive to the properties of the entire soil deposit. In addition to the centrifuge tests previously described, Gohl (1991), in his Ph.D. dissertation, performed shaking table tests of model piles. Ringdown, impact, frequency sweep, and free vibration tests were first made of single piles embedded in sand to ascertain whether the mode of loading, kinematic or inertial, affects the computation of the piles natural frequency and damping. The different methods computed a range of values that could be equally ascribed to the influence of amplitude, number of cycles of vibration, and the effect of previous strain history. Base excitation tests of small groups verified trends of group interaction for inline and offline shaking. Comparing shaking table test results to centrifuge test results, a deeper distribution of bending moments was observed in the former case, underscoring the effects of lower confining pressure in the shaking table tests. Cyclic p-y curves constructed by the API method were found to be considerably softer than the experimental results, which displayed nonlinear and strong hysteretic characteristics. Again, the computer code SPASM 8 was successfully

employed to simulate the test results. Liu and Chen (1991) tested large groups of model piles in liquefiable sands, and investigated the effects of pile installation on densification of the foundation soils. Excess pore pressure ratios were measured to be lower in the piled zone than in the freefield, as would be expected, but quantitative conclusions could not be drawn from the test results due to erroneous stress fields in this 1-g model. The authors did acknowledge that although driven piles may densify the immediate surrounding soil, global liquefaction mechanisms can still render such foundations susceptible to lateral or bearing capacity

205

failures. Of the numerous researchers studying pile response during liquefaction, one of the best efforts has been made by Nomura et al. (1991). A laminar box model container was used to allow shear deformations of the soil deposit, and different pile rigidities were tested. An effective stress free-field response analysis was coupled to a horizontal

subgrade reaction pile model to simulate the experimental results, and excellent agreement between the observed and computed response was obtained (Figure 4.29). The two basic conclusions these researchers derived were: Regardless of pile rigidity, the dynamic response of the pile-structure system before liquefaction is significantly influenced by the response of the soil. and Unless the pile is rigid enough, the dynamic response of the pile-structure system after liquefaction is also influenced by the response of the soil. Other researchers have come to a variety of conclusions from soil-pile liquefaction shaking table tests, as soil relative density and layering, pile spacing and installation procedure, and input motions can all contribute to unique test results.

Figure 4.29 - Comparison of Shaking Table Model Pile Liquefaction Response to Analytically Computed Fourier Amplitude Spectra (after Nomura et al., 1991)

206

Sasaki et al. (1991) and later Tokida et al. (1992) and Ohtomo and Hamada (1994) have studied the effects of lateral spreading on model pile groups in shaking table tests. They have considered variations in the slope of the ground surface, the slope length, the thickness of the liquefiable layer, and pile group configurations on the development of drag forces acting on the resisting piles. A clever variation of standard shaking table tests is presented by Yan et al. (1991), who applied a hydraulic gradient to a cohesionless model soil deposit in order to increase the stress level in the model with depth in a similar fashion as to centrifuge tests. Special scaling laws were derived, which Dou and Byrne (1996) validated in a modeling of models approach, while conducting free and forced vibration tests on model piles in the hydraulic gradient apparatus. Yamamoto et al. (1992) utilized shaking table model tests to investigate the performance of a pile foundation viscous damping device. Tests in both soft ground and liquefied sand verified the performance of the viscous dampers in attenuating ground motions (Figure 4.30), but such equipment must be designed to be multidirectional to be effective.

Figure 4.30 - Fourier Spectra Illustrating Effect of Viscous Damping Device in Shaking Table Model Test of Pile Foundation (after Yamamoto et al., 1992)

207

Sreerama (1993) tested small pile groups embedded in soft clay at different spacings subjected to small amplitude base excitations in order to investigate pile group dynamic interaction. Pile stiffness and damping were computed as a function of soil shear strain, to account for nonlinearity of response. He proposed a dynamic group interaction factor as a function of pile spacing and number of piles in the group, but independent of frequency, and compared his results to methods proposed by other researchers (Figure 4.31). Pile group stiffness was found to increase with pile spacing, and decrease with increasing strain (though the experimental strain range was small). Total pile group damping was found to decrease with increasing pile spacing, but radiation damping was computed with elastic interaction factors ignoring wave interference effects. Makris et al. (1997) compared the results of shaking table tests on single piles in sands with an advanced dynamic Winkler foundation model that incorporated frequency dependent stiffness and damping terms. The authors also applied an alternate method for calculating bending moments and axial strains that was not computed from the second derivatives of the pile deflected shape, but rather was based on inertial forces acting on the pile-superstructure system (the Clough method), and was found to give superior results. Tao et al. (1998) conducted large-scale shaking table tests at the new NIED facility in Tsukuba, Japan. This table is 15 m x 15 m, with a payload capacity of 500 tons at 0.5 g. A large-scale laminar shear box and full-size pile were tested, and the results compared to the Kagawa suite of analyses (SHAKE21, SRANG, NONSPS).

208

Figure 4.31 - Shaking Table Model Pile Group Interaction Factor Versus Pile Spacing, Experimental Data, and as Computed by a Variety of Methods (after Sreerama, 1993)

From full-scale tests to the other extreme, Konagai et al. (1998) performed virtual soil-pile interaction shaking table tests by mounting a beam to a shaking table and using analog circuits to simulate the soil-structure interaction effects.

209

4.4 Summary of Experimental Findings Together, field and laboratory pile load test programs have made a vital contribution to our understanding of SSPSI. Field pile tests have delivered inertial pile head loading, and have shed light on single pile, pile group, and dynamic response. Matlocks original 1962 work emphasizing the nonlinearity of soil pile response, along with cyclic degradation and gapping, are the basis for this entire field of study. Analysis of Gills results (1968) does suggest that universal p-y curves may be inappropriate for some cases, including Bay Mud. Tests of full-scale pile groups have clearly illustrated shadowing effects on load sharing, cyclic degradation, batter pile performance and pile cap contributions to resistance. Dynamic testing of single piles and pile groups has shown that response is site, frequency, load level, pile cap, and pile group spacing dependent. The variability in dynamic test results by various experimental methods has been demonstrated by several researchers. Scale model tests have successfully been calibrated against prototype test results, importantly demonstrating the viability of the scale modeling technique. Matlocks laboratory soil-pile experimental work from 1957 to 1962 greatly complemented his field studies, and stands as a complete investigation of its own. At the same time, scale models have been demonstrated to be sensitive to container boundary effects, scale modeling techniques, and adherence to similitude laws. The results of many scale model experiments are compromised by not paying proper heed to these issues. Nonetheless, advanced scale models can deliver both kinematic and inertial loads to soil-pile-superstructure systems. Novaks (1992) series of dynamic

experiments with model pile groups demonstrated the limitations in our present ability to accurately predict complex pile group behavior. Gohl (1991) conducted both centrifuge

210

and shaking table tests of small model pile groups, successfully reproducing single pile and group response, but with important differences in the two methods. Wilson (1998) has performed the highest quality centrifuge experiments dealing with SSPSI to date, with high level shaking and strong inertial response. Sreerama (1993) produced shaking table results of small pile groups in clay, and showed the variability in currently available dynamic pile group analysis methods. With this background, the next chapter will

describe the development of scale modeling criteria and test components for this experimental program.

211

Das könnte Ihnen auch gefallen