obtaining ductile performance from precast, prestressed concrete piles

Upload: unix0123

Post on 05-Jul-2018

223 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/15/2019 Obtaining Ductile Performance From Precast, Prestressed Concrete Piles

    1/17Summer 2009  | PCI Journal4

    Editor’s quick points

    n  A parametric study of the inelastic seismic response of precast,

    prestressed concrete piles was conducted to determine

    whether piles with only light transverse reinforcement could act

    as ductile structural elements.

    n  A nonlinear, inelastic finite-element program written specifically

    for this project was used to validate results for both laboratory

    and in-place testing.

    n  The addition of mild-steel longitudinal reinforcement did not

    enhance ductility of the piles, though it did increase flexural

    strength.

    Obtaining

    ductile

    performancefrom precast,

    prestressed

    concrete

    piles Andrew Budekand Gianmario Benzoni

    In general, concrete piles are intended to behave as elastic

    structural elements when subjected to seismic loading. Ex-

    ceptions include single-pile columns, which are typically

    designed to support the development of a subgrade plastic

    hinge under seismic loading.1 Elastic behavior is preferred

    because of the difficulty of repairing foundations and the

    possibility of reinforcing steel being exposed to corrosive

    materials due to the extensive cracking and spalling that is

    a consequence of plastic hinging.

    Prestressed concrete piles were first used as founda-

    tion elements in the early 1950s. They offer a number

    of advantages to the designer and contractor. Inherently

    resistant to tensile stresses, prestressed concrete piles can

    be economically fabricated off-site, safely transported,

    and easily handled during the pile-driving process. Their

    ability to resist tensile stress without cracking is advanta-

    geous because inadvertent tensile stresses applied during

    construction or service will be less likely to lead to crack-

    ing of the concrete, thus reducing the risk of corrosion of

    the pile’s steel reinforcement.

    Piles carry both axial load and lateral force. Axial load is

    resisted by a combination of end bearing and skin fric-

    tion (the dominant load-resisting mechanism depends on

    the soil characteristics). Lateral forces (for instance, those

    imposed by seismic excitation of the superstructure) are

    resisted by a combination of shear and bending resistance.

    Ideally, a foundation would be designed to remain elastic

    under seismic load (as repair of damage to piles after an

    earthquake is, at best, difficult), yet this is not normally

    practical for pile-column designs, and hinging of piles in a

  • 8/15/2019 Obtaining Ductile Performance From Precast, Prestressed Concrete Piles

    2/17 6PCI Journal | Summer 2009

     

     s  =

    4 Ah

    d' s= 0.12

     f c

     f  yt 

    0.5+1.25 P 

    axial 

     f c'  A

     g 

    + 0.13l 

    0.01( ) (1)

    where

     Ah  = area of the spiral steel

    d'   = core diameter

    s  = spiral pitch

      f 

    c'   = concrete compressive strength

     f  yt   = yield strength of the transverse reinforcement

    Paxial = axial load

     Ag  = gross cross-sectional area

    ρ l  = ratio of area of distributed longitudinal reinforce-

    ment to gross concrete area perpendicular to that

    reinforcement

    However, it is not less than the minimum specified in ACI

    318 section 10.9.3, given by Eq. (2).

    pile-cap configuration may be difficult to avoid. Thus, the

    inelastic behavior of prestressed concrete piles warrants

    study. Under lateral loads imposed by an earthquake, an

    individual pile (with a fixed head condition) may develop a

    moment pattern of the shape in Fig. 1.

    Inadequate detailing of early prestressed concrete piles

    may have caused inadequate performance in several earth-

    quakes, such as the 1964 Alaska2 and 1972 Miyagi-Ken-

    Oki3 events. The relation of a pile’s detailing deficiency

    to its poor performance has been identified by several

    researchers.4–6

    Unfortunately, the response to inadequate performance has

    been either a complete ban on precast, prestressed concrete

    piles in seismic applications or the specification of ex-

    tremely conservative amounts of transverse reinforcement,which make precast, prestressed concrete piles uncompeti-

    tive in the marketplace.

    Piles in seismic regions are governed by the provisions of

    the American Concrete Institute (ACI) in Building Code

     Requirements for Structural Concrete (ACI 318-05) and

    Commentary (ACI 318R-05)7 chapter 21. The minimum

    transverse reinforcement ratio ρ s in section 21.4.4 was

    modified by Priestly et al.8 as Eq. (1).

     

    Figure 1. This drawing shows the pile moment pattern resulting from axial and lateral loads. Note: 1 m = 3.28 ft; 1 kN = 0.225 kip.

  • 8/15/2019 Obtaining Ductile Performance From Precast, Prestressed Concrete Piles

    3/17Summer 2009  | PCI Journal

       s   = 0.45

     A g 

     Ah

    1

     f c

     f  yt 

      (2)

    where

     Ah = cross-sectional area of a structural member measured

    out to out of transverse reinforcement

    In the case of a typical 610-mm-diameter (24 in.) pile,

    using twenty-four 13 mm (1 / 2 in.) special tendons and a

    76-mm-thick (3 in.) cover, Eq. (1) would be superseded by

    Eq. (2), giving a required transverse reinforcement ratio of

    3.5%.

    This amount of spiral steel reinforcement would lead to

    congested construction; the spiral pitch in the plastic hinge

    regions using D9.5 (0.348 in. [8.8 mm]) wire would be

    18 mm (0.71 in.) (as opposed to 63.5 mm [25 in.] in Fig.

    2). This is not acceptable, as ACI 318 specifies Eq. (3).

      4d b ≤ s ≤ 6d b (3)

    where

    d b = main bar diameter

    Here, d b is the tendon diameter of 13.2 mm (1 / 2 in.). Use

    of D19 (0.49 in. [12 mm]) wire would give an accept-able pitch of 71 mm (2.8 in.), but forming the spiral to the

    required 229 mm (9.0 in.) radius would be difficult, slow,

    and costly.

    If the spacing requirements were relaxed to allow the

    use of a lighter-diameter wire at a smaller pitch, another

    problem would be created. The use of a small aggregate

    size would be mandated by the need to get a good aggre-

    gate connection between the core and the cover. Failure to

    provide a solid interface could result in spalled cover at the

    pile head during driving.

    It would, therefore, be advantageous to show that levels of

    displacement ductility adequate for competent foundation

    performance can be achieved using levels of transverse

    reinforcement commensurate with ACI 318 section 21.4.4,

    without the restriction that the minimums of section 10.9.3

    be met.

    The research presented in this paper gives the results of

    a parametric study using a pile modeled with transverse

    reinforcement slightly less than that required by ACI 318

    section 21.4.4. The parameters that varied are axial load,

    lateral soil stiffness, and the type of pile-cap connection.

    Figure 2. This graph compares the results of the finite-element-model prediction with the in-place pile test. Note: The pile had a 400 mm (16 in.) diameter with 50 mm

    (2 in.) of cover, transverse reinforcement ratio = 0.006, longitudinal reinforcement ratio = 0.021, and axial load ratio = 0.1 f  c '  A g . A g = gross cross-sectional area; f  c 

    '  = con-

    crete compressive strength; K  = subgrade reaction modulus. 1 m = 3.28 ft; 1 kN = 0.225 kip.

  • 8/15/2019 Obtaining Ductile Performance From Precast, Prestressed Concrete Piles

    4/17 6PCI Journal | Summer 2009

    son.13 In these tests, 0.406-m-diameter (1.33 ft) piles wereembedded in soil placed under controlled conditions into a

    purpose-built soil box and tested under combined axial and

    lateral load. Figure 2 shows the force-displacement enve-

    lope and predicted response from the first of these tests.

    The general trends of the test results—such as elastic stiff-

    ness, beginning of the softening branch, and incipient fail-

    ure—are reproduced by the prediction. Some discrepancies

    may be attributed to conservative modeling assumptions.

    First, the initial stiffness of the test pile was higher than

    that predicted. This comes from the action of the small-

    strain modulus of elasticity of the soil, which is estimatedat four times the large-strain value.14 This factor-of-four

    difference is quite close to that shown in Fig. 2. Second,

    the model underpredicted the maximum lateral force by

    about 12%. This can be directly attributed to the use of the

    28-day concrete strength to generate the prediction, rather

    than day-of-test strength. This was the prediction placed

    on the plotter at the test site.

    Use of the nondimensional system stiffness term KD6 / 

     D* EI eff  was also validated through comparison with this

    in-place testing program. One observation from previous

    analytical work was that the center of the plastic hinge

     Analytical modeling

    The analytical models of the soil-pile system used in this

    study were based on the nonlinear, inelastic finite-element

    modeling of a Winkler beam (a beam on a flexible founda-

    tion). The pile was represented using beam elements and

    the soil using lateral springs acting at the nodes (Fig. 3).

    Finite-element analysis (FEA) for inelastic soil-pile inter-

    action has been verified through field studies.9 FEA was

    chosen due to its flexibility in representing both pile and

    soil properties.

    Budek et al. used nonlinear, inelastic constitutive models

    for both the pile and soil and described them in detail.1 

    Briefly, the change in flexural stiffness of the pile as

    inelastic action took place was extracted from the moment-

    curvature data as the slope of the moment-curvature curve.

    Lateral load was applied in a series of steps. The elements’

    flexural stiffnesses were modified as necessary after each

    load step, according to the elements’ respective average

    moment. Pile yield was defined by tendon stress reaching

    85% of its ultimate value.

    A bilinear soil model was used in which the lateral stiff-

    ness of the soil (that is, individual spring stiffness) was re-

    duced to one-fourth of its original value when the displace-

    ment at a node associated with a given spring exceeded

    25.4 mm (1 in.).

    Soil stiffness as expressed by the subgrade reaction modu-

    lus K  ranged from 3200 kN/m3 to 48,000 kN/m3 

    (20 kip/ft3 to 300 kip/ft3). The nondimensional system stiff-ness KD6 /  D* EI eff  is used to describe the properties of the

    soil-pile system. It includes cracked-section flexural stiff-

    ness EI eff  of the pile shaft, and normalizes the pile diameter

     D against a reference pile diameter D* of 1.83 m (6.0 ft).

    Budek et al. describes its derivation.10

    The FEA program used was written specifically for this

    research program and was validated through comparison

    with laboratory testing10–12 and against results from in-place

    testing.9,13

    The laboratory testing program consisted of a series of 16tests,10 which included a number of precast, prestressed

    concrete pile shafts of a size and configuration similar to

    those examined in this study. The purpose of the investiga-

    tion was to characterize the effect of external confinement

    (as may be provided by competent soil) on the flexural

    ductility available in the pile shaft at the subgrade hinge

    (Fig. 1). The finite-element model in this study successful-

    ly predicted pile-shaft response under the imposed loading.

    The analytical model was further validated by comparison

    with a series of four in-place tests of free-head reinforced

    concrete pile-columns performed by Chai and Hutchin-

    Figure 3. The prestressed pile analytical model is shown with indicated forces.

    The notch in the cap of the prestressed model is for illustrative purposes, to allow

    the top spring to be shown. Note: K 1 = subgrade reaction modulus associated with

    node 1; K 2 = subgrade reaction modulus associated with node 2; K 3 = subgrade

    reaction modulus associated with node 3; K 4  = subgrade reaction modulus as-

    sociated with node 4; K 5  = subgrade reaction modulus associated with node 5; K n  

    = subgrade reaction modulus associated with node n; K n -1 = subgrade reaction

    modulus associated with node n – 1.

  • 8/15/2019 Obtaining Ductile Performance From Precast, Prestressed Concrete Piles

    5/17Summer 2009  | PCI Journal

    center of the plastic region after testing, fall almost directly

    on the curves.

    In this study, the structure was modeled as a 0.61-m-

    diameter (2 ft) continuous pile. Depth to the pile base was

    set at 24.4 m (80 ft), which is equal to 40 D and exceedsthe depth corresponding to long-pile response. The pile

    head was assumed to connect to a cap at ground level.

    Axial load varied from zero to 0.4  f 

    c'  Ag to represent forces

    from global overturning moments transferred from the

    superstructure.

    The pile section used in the analysis was a typical 0.61-m-

    diameter (2 ft) round pile as used in California. Figure

    2 shows the details. A transverse reinforcement ratio of

    1% was chosen for this study, slightly less than the 1.2%

    that would result from the use of Eq. (1). Nominal effec-

    tive section prestress after losses was 9.3 MPa (1.35 ksi).Figure 5 shows the section geometry.

    Three different types of pile–pile cap connections were

    considered: pile head embedded in cap, tendons embed-

    ded in cap, and tendons stressed through cap. Each was

    examined with and without the presence of mild-steel

    longitudinal bars providing dowel reinforcement through

    the pile-cap connection and down through the area of the

    subgrade plastic hinge (Fig. 4). The piles’ flexural respons-

    es were analyzed using the Mander15 model for confined

    concrete modified for prestressed sections. Modifications

    allowed for appropriate positioning of tendons with the

    in the pile shaft (that is, the subgrade hinge) would move

    toward the surface as plasticity progressed from the origi-

    nal location of the point of maximum subgrade moment.

    Its terminal location when the ultimate inelastic capacity

    of a free-head pile is known (in which the subgrade hinge

    controls response) can therefore be predicted. It is consis-tently at a depth of 70% of the point of maximum subgrade

    elastic moment, a value that is insensitive to either soil

    stiffness or above-grade height of the pile head. The actual

    depth is, of course, a function of both of these variables.

    Figure 4 shows curves predicting subgrade hinge depths

    for the above-grade heights tested in the relevant range of

    nondimensional system stiffness for the four tests con-

    ducted by Chai and Hutchinson. The depths of the centers

    of the plastic hinges, determined by digging down to the

    Figure 4. This graph compares the predicted and observed plastic-hinge depths from in-place pile tests. Note:D  = pile diameter; H  = depth of plastic hinge.

    Figure 5. This drawing illustrates the prestressed concrete pile section.

    Note: f  c '  = concrete compressive strength; D9.5 = 8.8 mm. 1 mm = 0.394 in.; 1 m

    = 3.28 ft; 1 MPa = 0.145 ksi.

  • 8/15/2019 Obtaining Ductile Performance From Precast, Prestressed Concrete Piles

    6/17 6PCI Journal | Summer 2009

    was handled similarly. Active prestress went from zero

    to its full value over the transfer length, and the effect of

    stressing the tendons was modeled by applying an extra

    increment of axial load to each section (in addition to the

    design load), such that the combination of this additional

    axial and prestress load would remain constant (that is,

    40% of prestress at 40% of d b was supplemented by an

    axial load equivalent to 60% of the prestressing force).

    Strain penetration of the longitudinal reinforcement into

    the cap is an important part of modeling. In the response of

    actual structures it permits a larger rotation at the pile-cap

    connection. It was modeled in this study by decreasing the

    stiffness of the pile’s top element as yielding of the tensile

    reinforcement occurred.

    For plastic hinges forming against supporting members,

    such as footings or cap beams, theoretical and experi-

    mental studies have led to the development of Eq. (4) for

    plastic hinge length l p.8

      l p = 0.08 L + 0.022 f  yd bl  (4)

    correct level of prestressing force applied through the sec-

    tion analysis, which produced the moment-curvature data

    used as input for the FEA.

    Previous experimental work 6,16–20 has shown that similar

    pile-cap connections can support ductile response up to a

    displacement ductility level of six.

    Modeling of the different connections was addressed

    through variation of either section prestress or axial load

    in calculating moment-curvature data. For the pile head

    embedded in the cap, full prestress was assumed at the

    bottom of the cap (that is, full transfer was assumed at the

    interface) (Fig. 6). To model embedment of the tendons,

    effective prestress was assumed to go from zero at the top

    of the pile to its full value at the end of the transfer length

    of 115d b.21,22 This part of the pile was, therefore, modeled

    in 10 sections, adding 10% of the effective prestress each

    time. Therefore, the input data for this case consisted of 11

    sets of moment-curvature data (output from the Mander

    model analysis) applied to the relevant section of the pile.

    The practice of stressing the tendons through the pile cap

    Figure 6. Three prestressed concrete pile–pile cap connections were considered in this study and are shown without reinforcing steel dowels. Note: The nominal embed-ment length of a pile head embedded in cap is two pile diameters. d b  = main bar diameter.

  • 8/15/2019 Obtaining Ductile Performance From Precast, Prestressed Concrete Piles

    7/17Summer 2009  | PCI Journal

    Figure 7. This graph compares Eq. (1) results with nominal design values for a prestressed pile with head embedded in cap with no nonprestressed, longitudinal reinforce-

    ment. Note: A g  = gross section area; D  = pile diameter; D * = reference pile diameter; EI eff  = cracked-section bending stiffness of the pile; f  c '   = concrete compressive

    strength; K  = subgrade reaction modulus; P axial  = axial load.

    Figure 8. This graph plots the moment versus height for a pile head embedded in cap with no reinforcing dowels. Note: f  c ' 

     = concrete compressive strength; K  = subgradereaction modulus; P axial  = axial load. 1 m = 3.28 ft; 1 kN = 0.225 kip.

  • 8/15/2019 Obtaining Ductile Performance From Precast, Prestressed Concrete Piles

    8/17 7PCI Journal | Summer 2009

    where

     L  = distance from the critical section to the point of con-

    traflexure (point A in Fig. 1)

     f  y  = yield strength

    d bl  = longitudinal bar diameter

    The first term in Eq. (4) represents the spread of plasticity

    resulting from variation in curvature with distance from the

    critical section and assumes a linear variation in moment

    with distance. The second term represents the increase in

    effective plastic hinge length associated with strain pen-

    etration into the supporting member.

    Figure 7 shows that Eq. (4) overestimates the plastic hinge

    length for the hinge occurring at the pile–pile cap connec-

    tion. Using typical values from an elastic analysis for depthto point of contraflexure at yield for a pile with the pile head

    embedded into the cap (both with and without reinforcing

    dowels), the plastic hinge length is substantially greater than

    that resulting from the present inelastic analysis.

    Used in this context, Eq. (4) demonstrates one of the

    inaccuracies associated with elastic analyses of piles:

    when yielding is reached in the pile’s critical section, the

    structure softens and a lesser depth of soil needs to be mo-

    bilized, thus moving the point of contraflexure toward the

    surface. If an appropriate correction is made, Eq. (4) gives

    a reasonable prediction of l p.

    The second term in Eq. (4), representing strain penetration,

    is straightforward in cases in which dowels are present

    because the dowels’ yielding is concurrent with the yield

    point of the structure. Where prestressing tendons alone

    form the longitudinal steel, however, interpretation of

    strain penetration is complicated by the fact that tendons

    in general require a development length about twice that ofdeformed, nonprestressed reinforcing steel. Also, ultimate

    tensile strains in prestressing steel are about half those of

    ordinary reinforcing steel, and even this is compromised

    by proximity to anchorages, with their attendant stress ris-

    ers. Thus, the strain penetration length should be consider-

    ably longer in the case of prestressing steel. Accordingly,

    in this study the strain penetration length for prestressing

    tendons was assumed to be roughly equivalent to that

    resulting from the use of ordinary reinforcing bar of twice

    the diameter of the prestressing steel used, in this case

    455 MPa (Grade 60) 29M (no. 9) deformed bar.

    Results

    Flexural responseand moment patterns

    The response of piles to lateral loading is best introduced

    through examination of moment patterns. In the case of

    a fixed-head pile, the maximum moment, which controls

    overall response, is generated at the pile-cap connection,

    and a secondary moment maximum forms below grade.

    Modeling inelastic pile response resulted in the formation

    of a plastic hinge at the pile-cap connection, after which

    Figure 9. This graph plots the moment versus height for a pile head embedded in cap with no reinforcing dowels. Note: A g  = gross section area; f  c '   = concrete compres-

    sive strength; K  = subgrade reaction modulus; P axial  = axial load. 1 m = 3.28 ft; 1 kN = 0.225 kip.

  • 8/15/2019 Obtaining Ductile Performance From Precast, Prestressed Concrete Piles

    9/17Summer 2009  | PCI Journal

    Analyses of other end conditions and considering the

    presence of dowels show similar trends. Ultimate moment

    versus height for three levels of axial load Paxial (0,

    0.2  f 

    c'  Ag, and 0.4 

     f c'  Ag) and the yield moment for the pile

    shafts is represented. Increasing soil stiffness increased the

    the moment was redistributed into the shaft. This resulted

    in the formation of a secondary hinge in the shaft. Figures

    8 through 10 demonstrate this, and the effect of varying

    soil stiffness and axial load on pile response is represented

    for the case of the pile head embedded into the cap.

    Figure 10. This graph plots the moment versus height for a pile head embedded in cap with no reinforcing dowels. Note: A g  = gross section area; f  c '   = concrete compres-

    sive strength; K  = subgrade reaction modulus; P axial  = axial load. 1 m = 3.28 ft; 1 kN = 0.225 kip.

    Figure 11. This graph plots the moment versus height in comparing plastic and elastic analyses for a pile head embedded in cap with reinforcing dowels. Note: The plastic

    analysis implies the development of a subgrade hinge. A g  = gross section area; f  c ' 

     = concrete compressive strength; K  = subgrade reaction modulus; P axial  = axial load.1 m = 3.28 ft; 1 kN = 0.225 kip.

  • 8/15/2019 Obtaining Ductile Performance From Precast, Prestressed Concrete Piles

    10/17 7PCI Journal | Summer 2009

    Figure 12. This graph plots the shear versus height in comparison of elastic and inelastic analyses for pile head embedded in cap with reinforcing dowels. Note: A g  = gross

    section area; f  c '   = concrete compressive strength; K  = subgrade reaction modulus; P axial  = axial load. 1 m = 3.28 ft; 1 kN = 0.225 kip.

    Figure 13. This graph shows the depth of maximum subgrade moment of the three pile–pile cap connections tested with no dowel reinforcement. Note: A g  = gross sec-

    tion area; D  = pile diameter; D*  = reference pile diameter; EI eff  = cracked-section bending stiffness of the pile; f  c ' 

     = concrete compressive strength; K  = subgrade reactionmodulus; P axial  = axial load.

  • 8/15/2019 Obtaining Ductile Performance From Precast, Prestressed Concrete Piles

    11/17Summer 2009  | PCI Journal4

    Figures 13 and 14 show the depth of the point of maxi-

    mum subgrade moment for piles without and with longitu-

    dinal mild-steel reinforcement, respectively. This is an im-

    portant parameter, as it shows the length of pile for which

    detailing for inelastic flexural action should be provided in

    the form of increased levels of transverse reinforcement.

    Depth of the maximum subgrade moment or hinge, if oneforms, is strongly affected by system stiffness but weakly by

    structural details such as the presence of mild-steel reinforce-

    ment or the specific type of pile-cap connection evaluated.

    The maximum depth seemed to approach a limiting value as

    system stiffness (functionally, in the case of the piles exam-

    ined in this study, soil stiffness) increased. The overall range

    was from 9 diameters (for soft soils) to 4 diameters. The anal-

    ysis was not extended for softer soils on the assumption that

    pole behavior (that is, rotation of the pile as a whole) would

    begin to take place as stiffness was dropped to low levels.

    Comparison of pile-cap connections

    Figure 15 compares ultimate moment patterns for the

    three pile–pile cap configurations examined in which

    reinforcing dowels were not included: pile head embedded

    in cap, prestressing tendons embedded in cap, and ten-

    dons stressed through cap. Figure 16 examines the effect

    of the presence of longitudinal mild-steel reinforcement

    maximum magnitude of the subgrade moment. Stiffening

    soil reduced the shear span between the two moment maxi-

    mums, which required greater mobilization of the pile-

    shaft flexural capacity at the subgrade moment maximum.

    Also, lower axial load with a commensurate reduction of

    section flexural strength placed more demand on the pile

    shaft to assist in resisting the applied moment.

    The effect of moment redistribution can only be assessed

    through an inelastic analysis. Figures 11 and 12 show an

    important aspect of this. Figure 11 compares an inelastic

    analysis and an elastic analysis of the system with the

    maximum flexural strength. The ultimate moments at the

    pile head are similar, but the shaft maximum predicted

    by elastic analysis was much lower. The consequence of

    this is seen in Fig. 12, which shows the difference in shear

    predictions from inelastic and elastic analyses.

    The redistribution of moment down the pile shaft after for-

    mation of the hinge at the pile-cap connection created much

    higher levels of shear (approaching twice as much, in the

    worst cases) in the pile shaft, which was missed by a purely

    elastic analysis. The increase in shear was also caused by

    the point of maximum moment in the shaft moving upward,

    toward ground level, as the subgrade hinge formed. This

    effect reduced the shear span and has been previously ob-

    served in both analytical10 and experimental13 studies.

    Figure 14. This graph shows the depth of maximum subgrade moment of the three pile–pile cap connections tested with dowel reinforcement. Note: A g  = gross section

    area; D  = pile diameter; D*  = reference pile diameter; EI eff  = cracked-section bending stiffness of the pile; f  c ' 

     = concrete compressive strength; K  = subgrade reactionmodulus; P axial  = axial load.

  • 8/15/2019 Obtaining Ductile Performance From Precast, Prestressed Concrete Piles

    12/17 7PCI Journal | Summer 2009

    Figure 15. This graph compares ultimate inelastic moment patterns for different pile–pile cap connections with no reinforcing dowels. Note: A g  = gross section area; f  c '   =

    concrete compressive strength; K  = subgrade reaction modulus; P axial  = axial load. 1 m = 3.28 ft; 1 kN = 0.225 kip.

    Figure 16. This graph compares ultimate inelastic moment patterns for different pile–pile cap connections with reinforcing dowels. Note: A g  = gross section area; f  c ' 

     =concrete compressive strength; K  = subgrade reaction modulus; P axial  = axial load. 1 m = 3.28 ft; 1 kN = 0.225 kip.

  • 8/15/2019 Obtaining Ductile Performance From Precast, Prestressed Concrete Piles

    13/17Summer 2009  | PCI Journal

    Figure 17. This graph shows the displacement ductility capacity versus the nondimensional system stiffness for different pile–pile cap connections with no reinforcing

    dowels. Note: 3200 kN/m3 < K  < 48,000 kN/m3 (20 kip/ft3 < K  < 300 kip/ft3). A g  = gross section area; D  = pile diameter; D * = reference pile diameter; EI eff  = cracked-

    section bending stiffness of the pile; f  c '   = concrete compressive strength; K  = subgrade reaction modulus; P axial  = axial load.

    Figure 18. This graph shows the displacement ductility capacity versus the nondimensional system stiffness for different pile–pile cap connections with reinforcing dowels.

    Note: A g  = gross section area; D  = pile diameter; D * = reference pile diameter; EI eff  = cracked-section bending stiffness of the pile; f  c ' 

     = concrete compressive strength; K  =subgrade reaction modulus; P axial  = axial load.

  • 8/15/2019 Obtaining Ductile Performance From Precast, Prestressed Concrete Piles

    14/17 7PCI Journal | Summer 2009

    Prestressed concrete piles can be considered to be ductile

    structural elements. This has been shown experimentally

    through previous research and has been verified through

    the analytical work described in this study. With appropri-

    ate detailing, prestressed concrete piles may be used as

    energy-dissipating elements in structures in which ductile

    behavior may be demanded. Detailing that would allow

    sufficient ductility capacity can be specified in both rein-forcement specification (through relevant design codes)

    and location (through knowledge of the location of the

    point of maximum subgrade moment).

    The ductility capacity of prestressed concrete piles is

    limited by the performance of the soil-pile system. Con-

    sideration of ductile response of individual parts (pile-cap

    connection or pile shaft) may be misleading. Ductility is

    affected by both the connection type chosen and axial load.

    The connection type is intrinsic to the structure of which

    the piles form a part, but axial load may be subject to

    variation in a given pile.

    Equation (4) is inappropriate for plastic hinges forming in

    pile shafts because of the following:

    Inelastic curvature can be expected to spread both•

    above and below the critical section.

    The slope of the moment profile at the section of•

    maximum moment is zero, invalidating the assump-

    tion of a linear decrease in moment with distance from

    the critical section.

    There should be no strain-penetration effect. This is•because there should be no significant slip of tension

    reinforcement past the critical section (which results

    in the strain-penetration effect for a fixed-base plastic

    hinge) due to the approximate symmetry of the mo-

    ment profile about the critical section.

    Analytical modeling, validated by in-place testing, has

    shown that the plastic hinge forming in a pile shaft will

    be located at a depth 70% of that predicted by an elastic

    analysis to maximum pile capacity. Results of previous

    experimental work have shown that a precast, prestressed

    concrete pile-shaft plastic hinge can give displacementductility in excess of six for a pile of the configuration and

    reinforcement levels examined in this study.

    The transverse reinforcement requirements given by ACI

    318 section 21.4.1 applied to piles of the configuration

    examined in this study will allow the piles to perform as

    structural elements of limited ductility. It is therefore rec-

    ommended that transverse reinforcement that would sup-

    port formation of a plastic hinge in the pile shaft (ideally

    equal to that provided at the pile-cap connection) should

    on moment patterns for these configurations. Curves are

    shown for both soft and stiff soil at a moderate axial load.

    The greatest demand on the pile shaft was from embedding

    the tendons into the pile cap (Fig. 15) due to the reduced

    flexural strength of the embedded tendon condition. In

    this case, prestress was assumed to be developed over the

    transfer length of 115d b, beginning at the bottom of the cap

    and extending upward into the cap. The effective prestressforce at the connection is, therefore, zero, which gives a

    lower strength and flexural stiffness over the plastic hinge

    length of 0.5 D at the pile-cap connection, requiring greater

    mobilization of the pile shaft. The transfer length was

    about 2.2 pile diameters, so the active prestress over the

    hinge length of 0.5 D did not exceed 30% of total prestress.

    The presence of reinforcing dowels in the three categories

    of pile–pile cap connection reduced the differences between

    subgrade moment maxima (Fig. 16). In both cases, the dif-

    ferences between subgrade moment maximums among the

    different end conditions were larger for softer soils.

    System ductilities

    Figures 17 and 18 show displacement ductility as a func-

    tion of nondimensional system stiffness, axial load, and

    pile-cap connection type. Figure 17 shows displacement

    ductility capacity for piles without longitudinal mild-steel

    reinforcement, and Fig. 18 shows displacement ductility

    capacity for piles with mild-steel reinforcement. Accord-

    ing to these two figures, displacement ductility capacity is

    relatively insensitive to the parameters studied. The plots

    lie on a fairly narrow range.

    The most prominent general trends were an increase in

    ductility capacity with both axial load and soil stiffness

    (expressed here as nondimensional system stiffness). Inclu-

    sion of mild-steel reinforcement had a significant effect.

    Comparing Fig. 17 and 18 shows that such reinforcement

    both increased ductility at the lower end of its range and

    decreased it at the top end of the range.

    In both cases, embedding the pile head into the cap gave

    the largest displacement ductility capacity. For piles with-

    out mild-steel reinforcement, the maximum range of ductil-

    ity for this case was from 2.5 to 4. With reinforcement, therange was about 2.7 to 3.4.

    Embedding the tendons (and dowels, where present) gave

    the next-highest ductility, followed by the rarely used prac-

    tice of stressing the tendons through the footing.

    Conclusion

    The results from this research program allow the formula-

    tion of several conclusions.

  • 8/15/2019 Obtaining Ductile Performance From Precast, Prestressed Concrete Piles

    15/17Summer 2009  | PCI Journal

    11. Budek, A. M., M. J. N. Priestley, and G. Benzoni.

    2004. The Effect of External Confinement on

    Flexural Hinging in Drilled Pile Shafts. Earthquake

    Spectra, V. 20, No. 1 (February): pp. 1–20.

    12. Budek, A. M., and M. J. N. Priestley. 2005. Ex-

    perimental Analysis of Flexural Hinging in Hollow

    Marine Prestressed Pile Shafts.Coastal Engineering

     Journal, V. 47, No. 1 (March): pp. 1–20.

    13. Chai, Y. H., and T. C. Hutchinson. 2002. Flexural

    Strength and Ductility of Extended Pile-Shafts—

    Experimental Study. Journal of Structural Engineer-

    ing, V. 128, No. 3 (March): pp. 595–602.

    14. Bowles, J. 1995. Foundation Analysis and Design.

    5th ed. New York, NY: McGraw-Hill.

    15. Mander, J. B., M. J. N. Priestley, and R. Park. 1988.

    Observed Stress-Strain Behavior of Confined Con-

    crete. Journal of Structural Engineering, V. 114, No.

    6 (June): pp. 1827–1849.

    16. Ikeda, S., T. Tsubaki, and T. Yamaguchi. 1982. Duc-

    tility Improvement of Prestressed Concrete Piles. In

    Transactions of the Japan Concrete Institute, V. 4,

    p. 538. Tokyo, Japan: Japan Concrete Institute.

    17. Falconer, T. J., and R. Park. 1982. Ductility of

    Prestressed Concrete Piles under Seismic Loading.

    Research report no. 82-6. Department of Civil Engi-

    neering, University of Canterbury, New Zealand.

    18. Muguruma, H., F. Watanabe, and M. Nishiyama.

    1987. Improving the Flexural Ductility of Preten-

    sioned High Strength Spun Concrete Piles by Lateral

    Confining of Concrete. In Proceedings of the Pacific

    Conference on Earthquake Engineering, V. 1, pp.

    385–396. Wairakei, New Zealand.

    19. Pam, H. J., R. Park, and M. J. N. Priestley. 1988.

    Seismic Performance of Precast Concrete Piles and

    Pile-Cap Connections. Research report 88-3, Depart-

    ment of Civil Engineering, University of Canterbury,

    New Zealand.

    20. Silva, P., F. Seible, and M. J. N. Priestley. 1997. Seis-

    mic Response of Standard Caltrans Pile-to-Pile Cap

    Connections under Simulated Seismic Load. Report

    no. SSRP-97/05, Department of Structural Engineer-

    ing, University of California at San Diego, La Jolla,

    CA.

    21. Zia, P., and T. Mostafa. 1977. Development Length

    of Prestressing Strands. PCI Journal, V. 22, No. 5

    (September–October): pp. 54–65.

    be provided to a depth of at least one pile diameter below

    the calculated point of maximum subgrade moment taken

    from an elastic analysis.

    References

    1. Budek, A. M., M. J. N. Priestley, and G. Benzoni.

    2000. Inelastic Seismic Response of Bridge Drilled-Shaft RC Pile/Columns. Journal of Structural Engi-

    neering, V. 126, No. 4 (April): pp. 510–517.

    2. Kachedoorian, R. 1968. Effects of March 27, 1964

    Earthquake on the Alaska Highway System. U.S.

    Department of the Interior, U.S. Geological Survey

    professional paper 545-C, Washington, DC.

    3. Kishida, H., T. Hanazato, and S. Nakai. 1980. Dam-

    age of Reinforced Precast Piles during the Miyagi-

    Ken-Oki Earthquake of June 12, 1972. In Proceed-

    ings of the Seventh World Conference on Earthquake

     Engineering. Istanbul, Turkey.

    4. Falconer, T. J., and R. Park. 1982. Ductility of

    Prestressed Concrete Piles under Seismic Loading.

    Research report no. 82-6, Department of Civil Engi-

    neering, University of Canterbury, New Zealand.

    5. Sheppard, D. A. 1983. Seismic Design of Prestressed

    Concrete Piling. PCI Journal, V. 28, No. 2 (April–

    May): pp. 20–49.

    6. Banerjee, S., J. F. Stanton, and N. M. Hawkins. 1987.

    Seismic Performance of Precast Concrete BridgePiles. Journal of Structural Engineering, V. 113, No.

    2 (February): pp. 381–396.

    7. American Concrete Institute (ACI) Committee 318.

    2005. Building Code Requirements for Structural

    Concrete (ACI 318-05) and Commentary (ACI 318R-

    05). Farmington Hills, MI: ACI.

    8. Priestley, M. J. N., F. Seible, and G. Calvi. 1996.

    Seismic Design and Retrofit of Bridges. New York,

    NY: John Wiley and Sons.

    9. Priestley, M. J. N. 1974. Mangere Bridge Foundation

    Cylinder Load Tests. Ministry of Works and Develop-

    ment Central Laboratories report no. 488. Wellington,

    New Zealand: Ministry of Works and Development.

    10. Budek, A. M., G. Benzoni, and M. J. N. Priestley.

    1997. Experimental Investigation of Ductility of In-

    Ground Hinges in Solid and Hollow Prestressed Piles.

    Report no. SSRP 97/17. Department of Structural

    Engineering, University of California at San Diego,

    La Jolla, CA.

  • 8/15/2019 Obtaining Ductile Performance From Precast, Prestressed Concrete Piles

    16/17 7PCI Journal | Summer 2009

    22. Ghosh, S. K., and M. Fintel. 1986. Development

    Length of Prestressing Strands, Including Debonded

    Strands and Allowable Concrete Stresses in Preten-

    sioned Members. PCI Journal, V. 31, No. 5 (Septem-

    ber–October): pp. 38–57.

    Notation

     Ah  = cross-sectional area of a structural member mea-

    sured out to out of transverse reinforcement

     Ag  = gross cross-sectional area

     Ah  = area of the spiral steel

    d'   = core diameter

    d b  = main bar diameter

    d bl  = longitudinal bar diameter

     D  = pile diameter

     D*  = reference pile diameter

     EI eff   = cracked-section bending stiffness of the pile

      f 

    c'    = concrete compressive strength

     f  y  = yield strength

     f  yt   = yield strength of the transverse reinforcement

     H   = depth of plastic hinge

    K   = subgrade reaction modulus, expression of soil

    stiffness

    l p  = plastic hinge length

     L  = distance from the critical section to the point of

    contraflexure

    Paxial  = axial load

    s  = spiral pitch

    ρ l  = ratio of area of distributed longitudinal reinforce-

    ment to gross concrete area perpendicular to that

    reinforcement

  • 8/15/2019 Obtaining Ductile Performance From Precast, Prestressed Concrete Piles

    17/17

     About the authors

    Andrew Budek, P.E., PhD, is an

    assistant professor for theDepartment of Civil and Environ-

    mental Engineering at New

    Mexico Institute of Mining and

    Technology in Socorro, N.Mex.

    Gianmario Benzoni, PhD, is a

    research scientist for the Depart-

    ment of Structural Engineering at

    the University of California at

    San Diego in La Jolla, Calif.

    Synopsis

    A parametric study of the inelastic seismic response

    of precast, prestressed concrete piles was conducted

    to determine whether piles with only light transverse

    reinforcement could act as ductile structural elements.

    A nonlinear, inelastic finite-element program written

    specifically for this project was used to validate results

    for both laboratory and in-place testing. The study

    examined single piles using several types of pile-cap

    connections, the addition of mild-steel reinforcement,

    varying levels of axial load, and a range of soil stiff-ness.

    The piles were modeled with 1% transverse reinforce-

    ment, which is less than 1 / 3 of that required by ACI

    318. The results indicated that modest levels of trans-

    verse reinforcement will allow for ductile response.

    Assuming that the pile-cap connection is detailed to

    allow and support the formation of a plastic hinge,

    with subsequent redistribution of moment down the

    shaft to form a secondary subgrade hinge in the pile

    shaft, the pile configurations analyzed provided aminimum displacement ductility of 2.

    The addition of mild-steel longitudinal reinforce-

    ment did not enhance ductility, though it did increase

    flexural strength. The optimum pile-cap connection to

    maximize ductility is embedment of the pile head into

    the cap. Rotation capacity is maximized by embed-

    ment of the prestressing tendons and any mild-steel

    longitudinal reinforcement present into the pile cap.

    Keywords

    Ductility, footing, pile, seismic, soil structure, trans-

    verse reinforcement.

    Review policy 

    This paper was reviewed in accordance with the

    Precast/Prestressed Concrete Institute’s peer-review

    process.

    Reader comments

    Please address any reader comments to PCI Journal 

    editor-in-chief Emily Lorenz at [email protected] orPrecast/Prestressed Concrete Institute, c/o PCI Journal,

    209 W. Jackson Blvd., Suite 500, Chicago, IL 60606.  J