You are on page 1of 18

DESIGN MODEL FOR PILES IN JOINTLESS BRIDGES

By Lowell Greimann1 and Amde M. Wolde-Tinsae,2 Members, ASCE


Downloaded from ascelibrary.org by University of California, San Diego on 03/16/15. Copyright ASCE. For personal use only; all rights reserved.

ABSTRACT: Piles in integral abutment bridges (jointless bridges) are


subjected to lateral movement due to thermal expansion and contraction
of the superstructure. The reduction in vertical load-carrying capacity of
the piles due to lateral movement is an important parameter in designing
the piles and in setting length limits for the jointless bridge. A simplified
design method for analyzing piles in integral abutment bridges is
presented. The model was developed from previous analytical models
and observations of pile behavior. Two failure modes for a pile are
examined: the slip mechanism, where the pile slips through the soil; and
the lateral mechanism where the failure of the soil-pile system is
associated with lateral movement of the pile. Equations for determining
the ultimate load for the pile for each mechanism are obtained. Results
predicted by the simplified model are compared to results from a
nonlinear finite element program and shown to be conservative with
respect to thefiniteelement model.

INTRODUCTION

Conventional bridges have in the past used a system of expansion joints,


roller supports, and other structural releases to prevent damage caused by
thermal expansion and contraction of the superstructure with annual
temperature variations. Expansion joints usually increase the initial cost of
a bridge and often do not function properly after years of service unless
extensively maintained. Thus, integral abutment bridges (Fig. 1), which
have no expansion joints, provide a design alternative which potentially
offers lower initial costs and lower maintenance costs. In integral abutment
bridges founded on piles, the piles are the most flexible elements, and will
be subjected to lateral movements as the bridge expands and contracts.
The reduction in the load-carrying capacity of the piles due to lateral
movement is an important parameter in designing the piles and in setting
length limits for integral abutment bridges.
There has been little theoretical work done to determine the effects of
these movements on the vertical load capacity of the abutment piles. A
survey (Greimann 1983; Wolde-Tinsae 1983) has shown that, although
about 28 U.S. State Highway agencies design and construct integral
abutment bridges, no standard design methods or details have been
developed. In many cases, the allowable length of integral abutment
bridges has been increased solely on the basis of observed excellent

2
'Prof., Dept. of Civ. Engrg., Iowa State Univ., Ames, IA 50011.
Assoc. Prof., Dept. of Civ. Engrg., Univ. of Maryland, College Park, MD 20742.
Note. Discussion open until November 1, 1988. To extend the closing date one
month, a written request must be filed with the ASCE Manager of Journals. The
manuscript for this paper was submitted for review and possible publication on July
15, 1987. This paper is part of the Journal of Structural Engineering, Vol. 114, No.
6, June, 1988. ©ASCE, ISSN 0733-9445/88/0006-1354/$1.00 + $.15 per page. Paper
No. 22536.
1354

J. Struct. Eng. 1988.114:1354-1371.


BRIDQE DECK
REINFORCED CONCRETE APPROACH SLAB
Downloaded from ascelibrary.org by University of California, San Diego on 03/16/15. Copyright ASCE. For personal use only; all rights reserved.

FLEXIBLE PILING *-

FIG. 1. Cross-Section of Bridge with Integral Abutments

performance of existing bridges. In this paper, a design model for the piles
will be presented, and the behavior of the model evaluated.

DESIGN MODEL

The model used to describe the soil-pile system is shown in Fig. 2a. This
model consists of only one pile. Lateral loading group effects can be
ignored if the spacing of the piles perpendicular to the direction of loading
is greater than 2.5 to 3 times the pile diameter or width (Davisson 1970;
Parkash 1962). Vertical loading group effects, however, must still be
considered. The pile is idealized as a beam column with an elastic-perfectly
plastic, moment-curvature relationship, as shown in Fig. 2b. Boundary
conditions at the top of the pile are assumed to provide lateral restraint,
and either zero (pinned condition) or complete (fixed condition) rotational
restraint. The tip of the pile is assumed to be free rotationally.
The soil is idealized by three sets of springs: lateral springs, vertical
springs, and a point spring. Soil resistance-displacement relationships for
the springs are shown in Fig. 2c. Parameters needed to describe these
relationships are the ultimate soil resistance and the initial stiffness. These
parameters are shown in Table 1. Shear strength reduction factors are
shown in Fig. 3 (Tomlinson 1957). The ultimate lateral soil resistance p„
and the initial stiffness kh are assumed to be either constant with depth or
linearly increasing with depth. The parameters for the vertical springs are
the maximum skin friction developed between the pile and soil fmax and the
initial stiffness k„ . The point spring is described by the maximum bearing
stress of the pile tip qmax and the initial point stiffness kq .
The behavior of this model will first be described analytically and then
be compared with results from an Integral Abutment Bridge Two-Dimen-
sional (IAB2D) Finite Element Program (Greimann 1986). This program
was developed with materially and geometrically nonlinear, two-dimen-
sional beam elements, and with a nonlinear, Winkler soil model with
lateral, vertical, and pile tip springs. These elements are used to simulate
the pile in Fig. 2a (A three-dimensional version, IAB3D, has also been
developed). In IAB2D, the soil resistance-displacement relationships are
approximated by the modified Ramberg-Osgood equation, as opposed to

1355

J. Struct. Eng. 1988.114:1354-1371.


V, A v
Downloaded from ascelibrary.org by University of California, San Diego on 03/16/15. Copyright ASCE. For personal use only; all rights reserved.

H.V

x
nfn
LATERAL
/ SPRINGS

VERTICAL CURVATURE
SPRINGS '
rtjTt
<b)

nrn

1
nrn POINT
^a^SPRING

fTtff P. f . q PU''max'"max

(a)

rt\> "•V "Q

y. z
(C)

FIG. 2. Design Model: (a) Model of Soil-Pile System; (to) Elastic, Perfectly Plastic
Moment-Curvature Relationship for Pile; (c) Bilinear Soil Resistance-Displacement
Relationships for Soil Springs

the design model which assumes the relationships are elastic-perfectly


plastic (Fig. 2b and c). Several experimental and analytical examples
(Greimann 1986) have been used to verify the program.
Considering several IAB2D analyses and experimental results, it was
observed that a pile can fail by basically two mechanisms: the slip
mechanism and the lateral mechanism. The slip mechanism occurs when
the soil/pile interface fails, and the pile slips through the soil while
remaining essentially undeformed. The lateral mechanism occurs when the
pile deflects laterally because of the interaction of geometric instability and
plasticity effects. The ultimate axial load V„ is the load associated with the
mechanism that forms first.

Slip Mechanism
For the slip mechanism, the load capacity of the pile is equal to the sum
of the maximum load carried by skin friction along the length of the pile,
1356

J. Struct. Eng. 1988.114:1354-1371.


TABLE 1. Parameters for Soil Springs
Case
Parameter Clay Sand
Downloaded from ascelibrary.org by University of California, San Diego on 03/16/15. Copyright ASCE. For personal use only; all rights reserved.

(1) (2) (3)


(a) Lateral springs (Davisson 1970; Ha and O'Neil 1981;
Poulos and Davies 1980)
P„ 9c„B 3 7 Bk p x
k„ 67c„ n ; ,x
(b) Vertical springs (Greimann 1986; Highway Structures Design
Handbook 1965; Poulos and Davies 1980; Wolde-Tinsae 1982)
tmax Least of: 0.02N[2(d + 2b,)]
H-piles Z(d + \>f)cu
2(d + 2b/)c„
2(dc„ + bfcj
f Lesser of: 0.04Nlg(klf)
Others

10fraax 10fmax
K Zc Zc

(c) Point spring (Greimann 1986; Ha and O'Neill 1981;


Wolde-Tinsae 1982)

4max 8N co „(ksf)
k„ 10o

and the maximum load carried by end bearing at the pile tip, as shown in
Fig. 4. This load can be calculated from
»« tmax-L' "r" Imax-^-e (1)
where L is the embedded length of the pile and Ae is the effective pile tip
area. For an H-pile, Ae is assumed to equal the rectangular area formed by
the section depth and the flange width.
Lateral Mechanism
Failure of the soil-pile system can also be associated with lateral
movement of the pile, which activates the lateral soil springs. As an
example, consider the pile in Fig. 5a. (Note that the slip mechanism is
eliminated here by the bottom support.) This pile has a lateral restraint,
representing the abutment, at the pile head. The pile is given a horizontal
displacement AA to simulate the movement of the bridge superstructure
due to a temperature change. If this movement is sufficiently large, a
plastic hinge may form near the top of the pile. An axial load V,
representing the live load on the bridge, is then applied to the pile. If
geometric instability were the only collapse consideration (i.e., no material
yielding), the ultimate load would equal the elastic buckling load V c r , i.e.,
the perfectly elastic case illustrated on the left in Fig. 5b. If collapse were
due to plasticity effects only (i.e., no geometric instability), the ultimate
load V p would occur when a plastic hinge forms and produces a plastic
mechanism. The rigid-perfectly plastic case on the right of Fig. 5b

1357

J. Struct. Eng. 1988.114:1354-1371.


1.25

^ ^ AVERAGE CURVE FOR


1.00
Downloaded from ascelibrary.org by University of California, San Diego on 03/16/15. Copyright ASCE. For personal use only; all rights reserved.

^ ^ CONCRETE AND
^ ^ TIMBER PILES
o
I-
o 0.75 -

0.50 —
-

AVERAGE CURVE / ^ ^ - ~ ^ _
8 °- 25 FOR STEEL PILES

0.0 I I I I I
0.0 0.5 1.0 1.5 2.0 2.5 3.0

cu (ksf)

FIG. 3. Reduction Factor a

illustrates this situation. Load-displacement curves for each extreme case


are illustrated in Fig. 5c.
In general, both geometric and material effects interact such that the
actual load-displacement behavior as observed from IAB2D and experi-
mental results is similar to that illustrated in Fig. 5c. The actual curve is
bounded by the curves for Vp and V c ,. The resulting lateral mechanism
load V„ , sometimes called the inelastic buckling load, is lower than either
the elastic buckling load V„. or the plastic mechanism load Vp . The design
method described in this paper is not intended to predict this complete
curve. However, a reasonable and conservative estimate to the ultimate
load V„ can be obtained using the Rankine equation (Chajes 1974; Poulos
and Davies 1980).

= 1.0 (2)
v„ V,
This equation combines both geometric and material instabilities.
Elastic Buckling Load
The elastic buckling load for initially bent columns approaches the
elastic buckling load for straight columns, providing the initial imperfec-
tions are relatively small (Chajes 1974). Following this rationale, the elastic
buckling load for a pile with a lateral head displacement will be calculated
using expressions developed for straight piles. If there are no vertical
springs along the pile, the pile axial load is constant, and the lateral
stiffness kh is constant, the buckling load is given by
U'EI
V,.,-= (3)

1358

J. Struct. Eng. 1988.114:1354-1371.


Downloaded from ascelibrary.org by University of California, San Diego on 03/16/15. Copyright ASCE. For personal use only; all rights reserved.

t
t SKIN
FRICTION

t
W-. END BEARING

FIG. 4. Vertical Load on Pile is Carried by Skin Friction and End Bearing

where I is the moment of inertia of the pile, E is the modulus of elasticity


of the pile material, U' is a nondimensional buckling coefficient, and

L
'max
R'
where L = pile length. For similar conditions, excepting k,,, which
increases linearly with depth, the buckling load is given by
_ V'EI
* cr ~ nn2 (4)

where V is a nondimensional buckling coefficient and

T = J—

^max np

1359

J. Struct. Eng. 1988.114:1354-1371.


Downloaded from ascelibrary.org by University of California, San Diego on 03/16/15. Copyright ASCE. For personal use only; all rights reserved.

(b)

PERFECTLY ELASTIC

RIGID, PERFECTLY PLASTIC


ACTUAL

(c)

FIG. 5. Example Illustrating Lateral Mechanism: (a) Schematic Drawing of Pile and
Soil; (b) Failure Modes; (c) Load-Displacement Curves for Pile

Nondimensional graphs for U' and V have been constructed for a variety
of pile head boundary conditions: free (no restraints); pinned (only lateral
restraints); fixed, no translation (rotational and lateral restraints); and
fixed, translating (only rotational restraints). The pile tip condition will be
viewed as pinned. Considering that there is uncertainty in the buckling
1360

J. Struct. Eng. 1988.114:1354-1371.


TABLE 2. Pile Head Boundary Conditions
Variable Pinned Fixed
d) (2) (3)
U' 2.0 2.5
Downloaded from ascelibrary.org by University of California, San Diego on 03/16/15. Copyright ASCE. For personal use only; all rights reserved.

V .2.3 4.2

analysis, and that nondimensional graphs are not available for the fixed,
no-translation case (rigid girders), the following approximations in Table 2
can be used. The values for the fixed head are taken from Fig. 6 (Reddy
and Valsangkar 1970; Toakley 1965) for the constant axial load case (^ =
0) and apply only for piles where z max or lmax are greater than four.
The assumption that the axial load is constant along the pile is only true
for short piles or for stiff end-bearing piles (Poulos and Davies 1980). For
other piles, vertical load transfer occurs along the pile and therefore, the
axial load varies with depth. This nonlinear variation of axial load is
approximated by the linear variation

V, = v ( l - * ( £ ) ) (5)
where x is the depth below the ground surface, Vv is the axial load at any
depth x, V is the axial load at the pile head, and \\i is a coefficient that
represents the rate of decrease in axial load between the pile head and pile
tip. The axial load at the pile tip is equal to V(l — i|»). The axial load at the

(a) (b)

FIG. 6, Effect of Skin Friction on Buckling Load for (a) Constant Lateral Soil
Stiffness; (b) Linearly Varying Lateral Soil Stiffness

1361

J. Struct. Eng. 1988.114:1354-1371.


Downloaded from ascelibrary.org by University of California, San Diego on 03/16/15. Copyright ASCE. For personal use only; all rights reserved.

CONSTANT

FIG. 7. Development of Collapse Mechanism, Assuming Rigid, Perfectly Plastic


Behavior

pile tip is also equal to the force in the point spring, which is obtained by
solving the differential equation for an axially loaded pile. Equating the two
expressions for the axial load at the pile tip gives
kgAf
*=1- (6)
k 9 A e cosh p + -\/k,AE sinh p
The value A is the cross-sectional area of the pile, kq is the initial stiffness
of the soil at the pile tip, and p = Vkt,/AE L. Fig. 6 is used to obtain a set
of nondimensional buckling coefficients U' and V , which take into
account the axial load transfer along the pile. Fig. 6 can only be used for
piles where z max or lmax are greater than or equal to four.

Plastic Mechanism Load


To calculate the ultimate load using the Rankine equation (Eq. 2), the
plastic mechanism load Vp must also be determined. This is the load at
which the pile collapses because a sufficient number of plastic hinges form
to create a complete mechanism assuming rigid-perfectly plastic behavior.
The value Vp will be derived using the pile shown in Fig. 7. The pile head
in this figure is first displaced from point a to point b because of the
expansion/contraction of the bridge superstructure. This movement causes
1362

J. Struct. Eng. 1988.114:1354-1371.


a plastic hinge to form at depth L2 below the surface. The movement of the
pile head from a to b causes a change in external work and internal energy
as expressed by the following equation for a soil having p„ constant with
depth:
Downloaded from ascelibrary.org by University of California, San Diego on 03/16/15. Copyright ASCE. For personal use only; all rights reserved.

1
HA,, = iV
i/i = M„V L " + 2 (P»L2) (7)

By minimizing the lateral load with respect to L 2 , one finds

U = (8)

When the vertical load V is applied, the pile head moves to point c. This
results in a second hinge forming at a distance L^ below the first hinge. The
first hinge is assumed to remain at L 2 . It is important to note that two
plastic hinges must form when rigid-perfectly plastic behavior is assumed.
This does not mean that two hinges form in the real pile at the ultimate
load. In general, the second hinge forms only at a very large displacement.
The plastic moment capacity of the pile at the plastic hinges is reduced
to Mp because of the presence of the compressive axial load. For a
rectangle, M^ is given by (Neal 1963):

M„ = M„ 1 (9)

where V is the compressive axial load in the pile, Vy is the yield load of the
pile, and Mp is the plastic moment capacity with no axial load. For
H-shaped sections, the following approximate expressions apply ("Com-
mentary of plastic design in steel" 1961):
Strong axis bending:
M; M„ V < 0 . 1 5 Vv (10)

M; = I . I 8 M J I-^- 0.15 V y < V < V y (ID

Weak axis bending:


M; = MP V<0.4V, (12)

M; = I.19M P
-<r 0.4 Vy =£ V < Yy

The change in external work and internal energy that occurs as the pile
(13)

head moves from b to c can be expressed by the following equation for a


soil with a constant p„ :

VA„ = M;e, + M;I e, + e2 - ^ j + P„L 2 (|J + P 0 L , ( | (14)

Simplifying and solving for V gives


1363

J. Struct. Eng. 1988.114:1354-1371.


l
V= V { , (15)
_y_ JA,
Downloaded from ascelibrary.org by University of California, San Diego on 03/16/15. Copyright ASCE. For personal use only; all rights reserved.

L, l
where 1 = Lj + L 2 . The mechanism begins at y, which equals zero and
corresponds to the point Vp in Fig. 5c.

v
» = ^ ( ' + £) + :£; L *. + L>> <16)
The locations of the plastic hinges, that is, Lr and L 2 , are, in a typical
problem of this type, selected to minimize the mechanism load. This leads
to a negative value of L 2 . In lieu of this approach, each term on the right
of Eq. 16 will be bounded by a conservative estimate. In the first term, Lj
is taken as much larger than L 2 . This is a valid approximation, since L 2 is
usually small. In the second term L1 will be assumed to be small with
respect to L 2 . This is not true, but is certainly conservative; that is, it gives
a lower bound to the second term. The value L2 will be assumed to be
V2M;/P„, as given by Eq. 9. Eq. 16 thereby reduces to

2M:
v (17)
* = ^f
Similar derivations can be made for piles with linearly varying p„.
Because of the simplifying approximations, p„ drops out of these equations
for Vp. Therefore, V p for pinned-headed piles with either constant or
linearly varying p„ is given in Eq. 17. Similarly, for all fixed-headed piles
the following equation can be derived:

v (18)
* =^

The effect of vertical springs (skin friction) on Vp can be accounted for


by using a reduced value of V in the expressions for M^ in Eqs. 9-13. For
example, a reduced value of V at the hinge location L 2 could be found from
Eq. 5. This will result in larger values of M^ at the plastic hinges, and
reduce some of the conservatism in the design method.

VERIFICATION OF THE DESIGN MODEL

Numerical examples will be presented to compare the results obtained


using the design method with those obtained using the finite element
program IAB2D. For all the examples, an HP10 x 42 pile was used. The
pile was bent about the weak axis and had a modulus of elasticity of 29,000
ksi and a yield stress of 50 ksi.
To illustrate the slip mechanism, a 40-ft long, axially loaded pile was
used. Properties of the soil are listed in Fig. 8 and are typical values for a
stiff clay. The load-displacement curve for this pile is also shown in Fig. 8.
The ultimate vertical load calculated using the design method explained in
Eq. 1 agrees with the finite element results.
1364

J. Struct. Eng. 1988.114:1354-1371.


Downloaded from ascelibrary.org by University of California, San Diego on 03/16/15. Copyright ASCE. For personal use only; all rights reserved.

A v (INCHES)

FIG. 8. Load-Displacement Curve for 40-ft Long HP10 x 42 Pile in Stiff Clay
Illustrating Slip Mechanism

The lateral mechanism, as predicted by the Rankine equation, was


checked using the 40-ft long pile shown in Fig. 5a. Figure 5A shows the pile
with a pinned support at the pile tip and with lateral springs only (no
vertical springs). This configuration was used with values of A/, of one inch
and two inches. One example was run with the pile head fixed against
rotation and horizontal translation. In another example, vertical springs
were added to a pile with a pinned head, and the vertical support at the pile
tip was removed.
The soil properties that were used are given in Table 3. The values n are
the shape factors used with the Ramberg-Osgood curves for the finite
element program (Wolde-Tinsae 1988). Even though soil types with
parameters 1/5 the values of soft clay and loose sand are somewhat
unrealistic, these were used in order to check the Rankine equation for a
greater range of values. Note that clay is approximated as having a uniform
lateral soil stiffness, whereas sand is approximated as having a linearly
varying modulus. For the case with the vertical springs, a large value for
fmax was used for all clay soils to insure that the slip mechanism would not
occur.

TABLE 3. Soil Properties Used to Check the Lateral Mechanism


K P„ K, n n
Soil types (ksi) (k/in) (ksi) (P-y) (f-z)
(1) (2) (3) (4) (5) (6)
Very stiff clay 15.6 3.75 20.06 2.0 1.0
Soft clay 0.5 0.24 4.10 1.0 1.0
1/5 soft clay 0.1 0.05 0.82 1.0 1.0
Dense sand 0.0840X 0.0104X — 3.0 1.0
Loose sand 0.0095X 0.0058X — 3.0 1.0
1/5 loose sand 0.0019X 0.0012X — 3.0 1.0

1365

J. Struct. Eng. 1988.114:1354-1371.


500
Downloaded from ascelibrary.org by University of California, San Diego on 03/16/15. Copyright ASCE. For personal use only; all rights reserved.

400

N
300 S

v (k)

200

100 VERY STIFF CLAY


SOFT CLAY
____ 1 / 6 SOFT CLAY

n V I I I I I I
0.0 0.2 0.4.0.6 0.8 1.0 1.2 1.4

FIG. 9. Typical Load-Displacement Curves for Lateral Mechanism Example

Typical load displacement curves for the example with vertical springs
are shown in Fig. 9. These curves, which were obtained by specifying
vertical displacements at the pile head, exhibit a typical beam-column-type
behavior as illustrated schematically in Fig. 5c. Notice that the descending
branch of the curves falls most rapidly for the softest soil. This instability
is even more pronounced for the cases with vertical support at the pile tip
and no vertical springs. The ultimate load V„ for the finite element results
is considered the maximum load from curves like Fig. 9. The results are
plotted in Fig. 10 by normalizing with respect to V „ (Eqs. 12 and 4) and V p
(Eqs. 17 and 18).
The Rankine equation is also shown in Fig. 10. It gives conservative
results, even for the clay with parameters 1/5 those of soft clay. Uncon-
servative results, however, were obtained when a soil type with parame-
ters 1/50 those of soft clay was used (not shown).
1366

J. Struct. Eng. 1988.114:1354-1371.


FINITE ELEMENT RESULTS
VERY STIFF CLAY
1.4 -
Downloaded from ascelibrary.org by University of California, San Diego on 03/16/15. Copyright ASCE. For personal use only; all rights reserved.

SOFT CLAY
1/§ SOFT CLAY
DENSE SAND
1.2 - LOOSE SAND
1/5 LOOSE SAND

1.0

A h = 2 " , FIXED HEAD


a
> 0.8

0.6 -

0.4 RANKING EQUATION

0.2

0.0
0.0 0.2 0.4 0.8 0.8 1.0

FIG. 10. Comparison of Rankine Equation and Finite Element Results

As described earlier, the plastic mechanism load V p is conservative


because of the conservative estimates used in bounding Eq. 16 to obtain
the simplified Eq. 17. The method is particularly conservative for the case
with vertical springs; that is, when the axial load varies along the pile
length. As noted previously, a reduced axial load at the hinge locations
increases the plastic moment capacity and, thereby, V p . This correction
was not used in Fig. 10.
Fig. 10 shows that plasticity effects tend to dominate the behavior of
piles in realistic soil types and that elastic buckling is unlikely to occur;
that is, the points tend to be in the upper left corner of the figure.

APPLICATIONS

The effect of lateral pile head movement A/( on the ultimate vertical load
carrying capacity typical of piles was examined using both the design
method and the finite element program. Friction piles and end-bearing piles
1367

J. Struct. Eng. 1988.114:1354-1371.


Downloaded from ascelibrary.org by University of California, San Diego on 03/16/15. Copyright ASCE. For personal use only; all rights reserved.

o °-6 - 0 SOFT CLAY


> £t STIFF CLAY
-3 0 4 • VERY STIFF CLAY
> ' FINITE ELEMENT
DESIGN METHOD
0.2 -

0.0 1.0 2.0 3.0 4.0


A h , IN.

FIG. 11. Nondimensional Forms of Ultimate Vertical Load V„ versus Specified


Lateral Displacement A,,, in Clay Soils (End-Bearing Pile). Value V„„ is Ultimate
Vertical Load When A;, = 0

with both free and fixed rotation at the pile head were studied. Typical soil
types ranging from loose to dense sand and from soft to very stiff clay have
been used. A 40-foot long HP10 x 42 steel pile bent about the weak axis
with the pile head fixed against rotation was used. Both the design method
and the program predict no reduction in the ultimate vertical load for
values of A;, up to four inches for the friction pile case for all soils. For the
end-bearing case, the finite element program and the design method both
predict some reduction in load-carrying capacity (Figs. 11 and 12).
Agreement between the two methods is good, with the design method
results being relatively conservative. Besides predicting a greater relative
reduction in capacity due to A,,, the design method also gives conservative
magnitudes of the capacity (Fig. 10), as discussed previously. Although not
shown here, similar behavior was observed for both the friction and
end-bearing cases for HP10 x 42 piles bent about the weak axis with
rotationally free heads and for HP10 x 42 piles bent about the strong axis.
Also, there was no reduction in the load-carrying capacity for 1-ft diameter
timber and concrete friction piles for values of A;, up to two inches.
The purpose for developing this simplified method was to predict the
change in the ultimate load capacity of a pile due to a lateral pile head
displacement, and thereby determine the maximum allowable length for
bridges with integral abutments. Two failure mechanisms are possible: the
slip mechanism and the lateral mechanism. The ultimate load for the slip
mechanism, as demonstrated in Eq. 1, does not depend on the lateral
displacement of the pile, while the ultimate load for the lateral mechanism
does, since Vp (Eqs. 17 and 18) decreases with increasing A;,. The ultimate
load for the pile is the smaller of the two mechanism loads. The slip
mechanism will tend to control for friction piles with relatively small A,,
values, while the lateral mechanism will tend to control for end-bearing
piles and for friction piles with large A,, values. As long as the slip
mechanism controls the ultimate load, the piles will be unaffected by A,,.
Methods for determining A,, are not discussed in this paper, but will depend
1368

J. Struct. Eng. 1988.114:1354-1371.


Downloaded from ascelibrary.org by University of California, San Diego on 03/16/15. Copyright ASCE. For personal use only; all rights reserved.

LOOSE SAND
0.4 MEDIUM SAND
DENSE SAND
FINITE ELEMENT
0.2 _____ DESIGN METHOD

0.0 J_
0.0 1.0 2.0 3.0 4.0

A h . IN.

FIG. 12. Nondimensional Forms of Ultimate Vertical Load V„ versus Specified


Lateral Displacement A,,, in Sand Soils (End-Bearing Piles)

upon the cyclic temperature changes; the bridge girder stiffness; the pile
stiffness; and the passive soil pressure on the abutment backwall (Grei-
mann 1986; Emerson 1981; Integral, no joint structures and required
provisions for movement 1980; Jorgenson 1981). The IAB2D program
permits these features to be incorporated into one model.

CONCLUSipNS AND RECOMMENDATIONS


A method for analyzing the piles in an integral abutment bridge has been
presented in this paper. The maximum allowable length of such a bridge
depends on whether the piles are friction or end-bearing, as well as on the
properties of the soil and piles. The design method gives conservative
results for the ultimate load capacity of the piles.
It is important to note that the design method considers only the
structural integrity of the piles. Other factors, notably the effects of the
abutment movement on the approach slab and fill, and the effects of the
induced axial stresses in the superstructure, need to be studied. While these
factors have a relatively small effect on shorter bridges, as longer bridges with
integral abutments are built, these problems will grow in importance. In
addition, the expressions for the plastic mechanism load could be refined by
including the effects of axial load transfer and differing soil types.

ACKNOWLEDGMENTS
The work reported in this paper was carried out by the writers through
the Department of Civil Engineering and the Engineering Research Insti-
tute at Iowa State University with funds provided by the Iowa Department
of Transportation.
1369

J. Struct. Eng. 1988.114:1354-1371.


APPENDIX I. REFERENCES

Chajes, A. (1974). Principles of structural stability theory. Prentice-Hall, Inc.,


Englewood Cliffs, N.J.
Downloaded from ascelibrary.org by University of California, San Diego on 03/16/15. Copyright ASCE. For personal use only; all rights reserved.

"Commentary of plastic design in steel." (1961). Manuals of engineering practice,


No. 41. ASCE, New York, N.Y.
Davisson, M. T. (1963). "Estimating buckling loads for piles." Proc. Second Pan
American Conference on Soil Mechanics and Foundation Engrg., 1, 351-71.
Davisson, M. T. (1970). "Lateral load capacity of piles." Transp. Res. Record No.
333, Transp. Res. Board, Nat. Res. Council, Nat. Academy of Sciences,
Washington, D.C.
Emerson, M. (1981). "Thermal movements of concrete bridges; Field measurement
and methods of prediction." Joint sealing and bearing systems for concrete
structures. Special Publication 70, American Concrete Inst., Detroit, p. 77-102.
Greimann, L. F., Wolde-Tinsae, A. M., and Yang, P. S. (1983). "Skewed bridges
with integral abutments." Bridges and culverts, Transp. Res. Record, No. 903,
Transp. Res. Board, Nat. Res. Council, Nat. Academy of Sciences, Washington,
D.C.
Greimann, L. F., Yang, P. S., and Wolde-Tinsae, A. M. (1986). "Nonlinear
analysis of integral abutment bridges." / . Struct. Engrg., ASCE, 112(10),
2263-2280.
Ha, N. B., and O'Neill, M. W. (1981). Field study of pile group action, Appendix
A, Federal Highway Administration, D.C.
Highway structures design handbook, Vol. 1. (1965). United States Steel Corpo-
ration, Pittsburgh, Pa.
Integral, no-joint structures and required provisions for movement. (1980). Federal
Highway Administration, D.C.
Jorgenson, J. L. (1981). Behavior of abutment piles in an integral abutment bridge.
Engineering Experiment Station, North Dakota State University, Fargo, N.D.
Neal, B. G. (1963). The plastic methods of structural analysis, 2nd Ed., John Wiley
and Sons, Inc., N.Y.
Poulos, H. G., and Davies, E. H. (1980). Pile foundation analysis and design. John
Wiley and Sons, Inc., N.Y.
Prakash, S. (1962). "Behavior of pile groups subjected to lateral loads," thesis
presented to the University of Illinois, at Urbana, 111. in partial fulfillment of the
requirements for the degree of Doctor of Philosophy.
Reddy, A. S., and Valsangkar, A. J. (1970). "Buckling of fully and partially
embedded piles." J. Soil Mech. and Foundation Div., ASCE, 96(SM6),
1951-1965.
Toakley, A. R. (1965). "Buckling loads for elastically supported struts." J. Engrg.
Mech. Div., ASCE, 91(EM3), 205-231.
Tomlison, M. J. (1957). "The adhesion of piles driven in clay soils." Proc. 4th Int.
Conference on Soil Mechanics and Foundation Engrg., ASCE, 2, 66.
Wolde-Tinsae, A. M., Greimann, L. F., and Yang, P. S. (1982). "Nonlinear pile
behavior in integral abutment bridges." Report 1501, Engrg. Res. Inst., Iowa
State Univ., Ames, IA.
Wolde-Tinsae, A. M., Greimann, L. F., and Johnson, B. (1983). "Performance of
integral abutment bridges." / . Int. Assoc, for Bridge and Struct. Engrg., 1,
(P-S8-83), 17-34.
Wolde-Tinsae, A. M., Greimann, L. F., and Yang, P. S. (1988). "End-bearingpiles
in jointless bridges." J. Struct. Engrg., ASCE (in press).

APPENDIX il. NOTATION

The following symbols are used in this paper:

B = pile width;
by = flange width of H-piie (fi);

1370

J. Struct. Eng. 1988.114:1354-1371.


c„ = adhesion between soil and pile; ac„ (psf);
c„ = undrained cohesion of the clay soil; 97.ON + 114.0 (psf);
d = section depth of H-pile or diameter of pipe pile (ft);
J = 200 for loose sand; 600 for medium sand; 1,500 for dense sand;
Downloaded from ascelibrary.org by University of California, San Diego on 03/16/15. Copyright ASCE. For personal use only; all rights reserved.

kp = tan2 (45° + <f>/2);


\g = gross perimeter of the pile (ft);
N = average standard penetration blow count;
N t . 0)r = corrected standard penetration test (SPT) blow count at depth
of pile tip; N (uncorrected) if N < 15; 15 + 0.5(N-15) if N >
15;
nh = constant of subgrade reaction; J/1.35;
x = depth from soil surface;
zc = relative displacement required to develop fmax or qmax ; 0.4 in.
(0.033 ft.) for sand; 0.2 in. (0.021 ft.) for clay;
a = shear strength reduction factor (see Fig. 3);
7 = effective unit soil weight; and
4> = angle of internal friction.

1371

J. Struct. Eng. 1988.114:1354-1371.

You might also like