Group2022TechnicalManual Compressed
Group2022TechnicalManual Compressed
GROUP
Version 2022
Lymon C. Reese
Shin-Tower Wang
Luis Vasquez
for
ENSOFT, INC
3003 W. Howard Lane
Austin, Texas 78728
February - 2022
ii
The user is not entitled to copy this software unless copying for
backup purposes. The user may not loan, rent, lease, or transfer
this software to any other person, company, or location. This
software and documentation are copyrighted materials and
should be treated like any other copyrighted material (e.g. a
book or musical recording). This software is protected by the
United States Copyright Law and International Copyright
Treaty.
FOREWORD
The analysis of a group of piles that support a cap or a structure where the
system is subjected to axial and lateral loading is an important problem in soil-
structure interaction. The pile may be of different sizes and may be installed
vertically or on a batter. The piles are assumed to be far enough apart so that there
is no significant pile-soil-pile interaction or, if the piles are spaced closely together,
such interaction is reflected by an appropriate modification of soil resistance
curves.
Computer program GROUP allows the user to analyze the behavior of pile
groups either using a 2-dimensional model or using a 3-dimensional model.
Table of Contents
4.6.5 Modified p-y Formulation for Stiff Clay with no Free Water ........................................... 4-35
4.7 Response of Sand Above and Below the Water Table ............................................................. 4-36
4.7.1 Field Experiments............................................................................................................. 4-36
4.7.2 Recommendations for Computing p-y Curves ................................................................. 4-37
4.7.3 Recommended Soil Tests ................................................................................................. 4-43
4.7.4 Example Curves ................................................................................................................ 4-43
4.7.5 Response of Sand Above the Water Table ....................................................................... 4-44
4.8 Response of Sand Above and Below the Water Table (API RP2A Recommendation) ............. 4-45
4.8.1 Background ...................................................................................................................... 4-45
4.8.2 Recommendations for Computing p-y Curves ................................................................. 4-45
4.8.3 Example Curves ................................................................................................................ 4-49
4.8.4 Other Recommendations for p-y Curves ......................................................................... 4-51
4.9 Response of Soil with Both Cohesion and Internal Friction–Simplified Procedure ................. 4-52
4.9.1 Background ...................................................................................................................... 4-52
4.9.2 Recommendations for Computing p-y Curves ................................................................. 4-53
4.9.3 Discussion......................................................................................................................... 4-59
4.10 Response of Soil in Liquefied Sand .......................................................................................... 4-61
4.10.1 Background ...................................................................................................................... 4-61
4.10.2 Recommended Equations by Rollins et al. (2005) ........................................................... 4-63
4.10.3 Simplified Hybrid p-y Model ............................................................................................ 4-65
4.10.3.1 Equations for Simplified Hybrid p-y Curve ............................................................... 4-66
4.10.3.2 Correlation for 33% residual strength as function of SPT (N1)60-cs -value ................ 4-67
4.10.4 Consideration of Lateral Spreading.................................................................................. 4-72
4.11 Response for Loess Soils .......................................................................................................... 4-73
4.11.1 Background ...................................................................................................................... 4-73
4.11.1.1 Description of Load Test Program............................................................................ 4-73
4.11.1.2 Soil Profile from Cone Penetration Testing .............................................................. 4-74
4.11.2 Procedure for Computing p-y Curves in Loess ................................................................. 4-75
4.11.2.1 General Description of p-y Curves in Loess.............................................................. 4-75
4.11.2.2 Equations of p-y Model for Loess ............................................................................ 4-76
4.11.2.3 Step-by-Step Procedure for Computing p-y Curves ................................................. 4-81
4.11.2.4 Limitations on Conditions for Validity of Model ...................................................... 4-82
List of Figures
Figure 4.8 Characteristic shape of p-y curve for static loading in stiff clay in the presence of free
water ............................................................................................................................ 4-21
Figure 4.9 Values of Constants As and Ac...................................................................................... 4-23
Figure 4.10 Characteristic shape of p-y curve for cyclic loading in stiff clay in the presence of free
water ............................................................................................................................ 4-26
Figure 4.11 Soil profile used for example p-y curves for stiff clay .................................................. 4-28
Figure 4.12 Plotted p-y curves ........................................................................................................ 4-29
Figure 4.13 Characteristic shape of p-y curve for static loading in stiff clay with no free water ... 4-31
Figure 4.14 Characteristic shape of p-y curve for cyclic loading in stiff clay with no free water ... 4-33
Figure 4.15 Example curves for stiff clay with no free water, cyclic loading .................................. 4-34
Figure 4.16 Characteristic shape of a family of p-y curves for static and cyclic loading in sand .... 4-36
Figure 4.17 Values of coefficients Ac and As .................................................................................. 4-40
Figure 4.18 Non-dimensional coefficient B for soil resistance versus depth.................................. 4-41
Figure 4.19 Example p-y curves for sand below the water table, cyclic loading ............................ 4-44
Figure 4.20 Coefficients as Function of ′ ....................................................................................... 4-46
Figure 4.21 Initial modulus of subgrade reaction, k, used for API sand criteria ............................. 4-48
Figure 4.22 Example of p-y curves computed based on API sand criteria ...................................... 4-49
Figure 4.23 Proposed p-y curves for c− soil ................................................................................. 4-53
Figure 4.24 Representative Values of k for c- soil ........................................................................ 4-58
Figure 4.25 p-y curves computed based on the simplified procedure ........................................... 4-60
Figure 4.26 Example p-y Curve for Liquefied Sand ......................................................................... 4-63
Figure 4.27 Method for Computing Residual Shear Strength of Liquefied Soil for Use in Hybrid p-y
Model ........................................................................................................................... 4-68
Figure 4.28 Factor 50 as Function of SPT Blowcount ..................................................................... 4-69
Figure 4.29 Possible Intersection Patterns of Residual and Dilative p-y Curves in Hybrid p-y Model 4-
70
Figure 4.30 Example of Non-intersecting Curves............................................................................ 4-70
Figure 4.31 Example of Curves with One Intersection of Dilative and Residual Curves ................. 4-71
Figure 4.32 Example of Curve with One Intersection of Dilative Curve and Residual Plateau ....... 4-71
Figure 4.33 Example of Curve with Two Intersection Points .......................................................... 4-72
Figure 4.34 Idealized Tip Resistance Profile from CPT Testing Used for Analyses. ........................ 4-75
Figure 4.35 Generic p-y curve for Drilled Shafts in Loess Soils ....................................................... 4-76
Figure 4.36 Variation of Modulus Ratio with Normalized Lateral Displacement ........................... 4-79
Figure 4.37 p-y Curves for the 30-inch Diameter Shafts ................................................................. 4-80
Figure 4.38 p-y Curves and Secant Modulus for the 42-inch Diameter Shafts ............................... 4-80
Figure 4.39 Cyclic Degradation of p-y Curves for 30-inch Shafts .................................................... 4-81
Figure 4.40 Recommended p-y curve for design of drilled shafts in vuggy limestone ................... 4-84
Figure 4.41 Initial Moduli of Rock from Pressuremeter, San Francisco Test (After Reese,1997) ... 4-88
Figure 4.42 Engineering properties for intact rock, after (Horvath & Kenney, 1979); (Peck, 1976);
and (Deere, 1968) ........................................................................................................ 4-89
Figure 4.43 Sketch of p-y curve for weak rock (after Reese, 1997) ................................................ 4-91
Figure 4.44 Comparison of experimental and computed values of pile head deflection, Islamorada
Test ( after Reese, 1997) .............................................................................................. 4-95
Figure 4.45 Plot of Computed Curves of Deflection and Bending Moment versus Depth, Islamorada
Test, Lateral Load of334kN (after Reese, 1997)........................................................... 4-96
Figure 4.46 Comparison of Experimental and Computed Values of Pile-Head Deflection for
Different Values of EI, San Francisco Test .................................................................... 4-97
Figure 4.47 Values of EI for three methods, San Francisco Test ..................................................... 4-98
Figure 4.48 Comparison of Experimental and Computed Values of Maximum Bending Moments for
Different Values of EI ................................................................................................... 4-99
Figure 4.49 p-y Curve in Massive Rock. ........................................................................................ 4-101
Figure 4.50 Equation for Estimating Modulus Reduction Ratio from Geological Strength Index.4-104
Figure 4.51 Poisson’s Ratio as Function of Stress Wave Velocity Ratio ........................................ 4-106
Figure 4.52 Model of Passive Wedge for Drilled Shafts in Rock ................................................... 4-107
Figure 4.53 Degradation Plot for Es .............................................................................................. 4-111
Figure 4.54 p-y Curve for Piedmont Residual Soil......................................................................... 4-111
Figure 4.55 Illustration of Equivalent Depths in a Multi-layer Soil Profile.................................... 4-113
Figure 4.56 Example problem of soil response for layered soil .................................................... 4-115
Figure 4.57 Example p-y curves for layered soil ........................................................................... 4-116
Figure 4.58 Equivalent depths of overlying soil for use in computing p-y curves for a layered
soil .............................................................................................................................. 4-117
Figure 4.59 Modification of p-y curves for battered piles, after (Kubo, 1964), and (Awoshika &
Reese, 1971)............................................................................................................... 4-118
Figure 4.60 p-y curves for single pile and pile group without shadowing effect .......................... 4-121
Figure 4.61 p-y curves for single pile and pile group with group effect ....................................... 4-122
Figure 5.1 Typical load-settlement curve........................................................................................ 5-3
Figure 5.2 Curve showing typical distribution of load along the length of an axially-loaded pile .. 5-4
Figure 5.3 Idealized load-distribution and load-settlement curves for an axially-loaded deep
foundation...................................................................................................................... 5-5
Figure 5.4 Mechanical model of axially-loaded deep foundation .................................................. 5-6
Figure 5.5 Normalized curves showing load transfer in side resistance versus settlement for drilled
shafts in clay (after Reese and O’Neill, 1987) ................................................................ 5-9
Figure 5.6 Normalized curves showing load transfer in end bearing versus settlement for drilled
shafts in clay ( after Reese and O’Neill,1987) .............................................................. 5-13
Figure 5.7 Load transfer-movement curves for piles in sand from field tests(after Coyle and
Sulaiman,1967) ............................................................................................................ 5-15
Figure 5.8 Normalized curves showing load transfer in side resistance versus settlement for drilled
shafts in cohesionless soil (after Reese and O'Neill, 1987) .......................................... 5-16
Figure 5.9 Tip resistance-displacement function .......................................................................... 5-19
Figure 5.10 Normalized curves showing load transfer in end bearing versus settlement for drilled
shafts in cohesionless soil (after Reese and O'Neill, 1987) .......................................... 5-21
Figure 6.1 Definition of fm ............................................................................................................... 6-2
Figure 6.2 p-y curves for single pile and pile group without shadowing effect .............................. 6-3
Figure 6.3 p-y curves for single pile and pile group with shadowing effect ................................... 6-4
Figure 6.4 Curve giving reduction factors βa for piles in a row ....................................................... 6-7
Figure 6.5 Curve giving reduction factors βbl for leading piles in a line ......................................... 6-8
Figure 6.6 Curve giving reduction factors βbt for trailing piles in a line .......................................... 6-9
Figure 6.7 System for computing reduction factor for skewed piles ............................................ 6-11
Figure 6.8 Pile group for example computations.......................................................................... 6-12
Figure 7.1 Results of model tests on groups of instrumented driven piles in granular soils
(fromVesic,1969( ............................................................................................................ 7-3
Figure 7.2 Vertical efficiencies for model groups ins and(After O’Neill,1983) ............................... 7-5
Figure 7.3 Full-scale groups in sand ................................................................................................ 7-6
Figure 7.4 Results of model tests on pile groups in clay under compression (After Mello,1969) .. 7-9
Figure 7.5 Vertical efficiencies for model groups in clay .............................................................. 7-10
Figure 7.6 Non-end bearing full-scale groups in clay .................................................................... 7-11
Figure 8.1 p-y curves for pile cap and individual pile...................................................................... 8-3
Figure 8.2 Distribution of soil resistance on the front side of pile cap ........................................... 8-4
Figure 9.1 Element of beam subjected to pure bending ................................................................ 9-3
Figure 9.2 Stress-Strain relationship for concrete. ......................................................................... 9-6
Figure 9.3 Stress-Strain relationship for confined concrete (Mander, Priestley, & Park, 1988). ... 9-9
Figure 9.4 Stress-Strain relationship for steel............................................................................... 9-11
Figure 9.5 Stress-Strain relationship for steel with hardening (Thompson & Park, 1978). .......... 9-12
Figure 9.6 Stress-Strain relationship for prestressing strands recommended by PCI Design
Handbook, 5th Edition .................................................................................................. 9-14
Figure 9.7 Load and elongation diagram for Carbon Fiber Composite Cables (CFCC) by Tokyo Rope
...................................................................................................................................... 9-15
Figure 9.8 Equivalent Elastoplastic Moment Curvature by CALTRANS. ........................................ 9-17
List of Tables
The computer program GROUP has been well accepted as a useful tool for analyzing the
behavior of grouped piles subjected to both axial and lateral loadings. The program was developed to
compute the distribution of loads (vertical, lateral, and overturning moment) from the pile cap to piles
in a pile-group. The piles may be installed vertically, or on a batter and their heads may be fixed, pinned,
or elastically restrained by the pile cap. The cap is assumed to act as a rigid body, and may settle,
translate, and rotate. GROUP allows the user to analyze the behavior of pile groups using either 2-or 3-
dimensional models. The new version of the program GROUP includes an option to model the pile-cap
as flexible for three-dimensional models.
The program internally generates the nonlinear response of the soil, in the form of t−z curves for axial
loading, torque-rotation curves for torsional loading, and p-y curves for lateral loading. A solution
requires iteration to accommodate the nonlinear response of each of the piles. In the solution, the
equations of equilibrium are satisfied, and compatibility is achieved between pile movement and soil
response, and between the movement of the cap and the pile-head movements.
Foundation engineering at its best can predict the manner and amount a foundation will deform
and deflect in response to future loadings. This information meshes with that from analysis of the
superstructure, providing a global analysis. Thus, designs ensuring that all elements of a structure are in
equilibrium and compatible with each other for all times can be made. This concept is termed
compatible design.
Modeling a group of piles is important, since in practice piles are usually used in groups. The
behavior of piles in a closely spaced group is influenced by two forms of interaction. The first is pile-soil-
pile interaction, which causes the piles in a group to be less efficient than in an isolated pile. The second
occurs through the cap, or superstructure, and addresses the distribution of the axial load, lateral load,
and moment of each individual pile as a result of the displacement and rotation of the cap. The latter
problem is one of mechanics, and an exact solution can be found if the models for the individual piles
are valid. New developments in computer technology allow a complete solution to be readily developed,
including automatic generation of the nonlinear response of soil, and iteration to achieve force
equilibrium and compatibility. This program provides a complete solution for analysis of pile groups.
The development of computational methods has been limited because of a lack of knowledge
on single-pile behavior. In order to meet the practical needs of designing structures with grouped piles,
various computational methods were developed by making rational assumptions that would allow for
analysis of the problem.
The simplest way to treat a grouped-pile foundation is to assume that both the structure and
the piles are rigid, and to only consider the axial resistance of the piles, while excluding the lateral
resistance of the piles from computation. Using these assumptions, Culmann presented a graphical
solution in 1866 for the equilibrium state of the resultant external load and axial reaction of each group
of similar piles by drawing a force polygon (Terzaghi, 1956). However, the application of Culmann's
method is limited to the case of a foundation with only three groups of similar piles. A supplemental
method to this graphical solution was proposedin1930 by Brennecke and Lohmeyer (Terzaghi, 1956).
The vertical component of the resultant load is distributed in a trapezoidal shape that makes the total
area equal to the magnitude of the vertical component, and has the center of gravity lie on the line of
the vertical component of the resultant load. The vertical load is distributed to each pile assuming that
aside from the end piles, the trapezoidal load is separated into independent blocks at the top of the
piles. Unlike Culmann's method, the latter method can handle more than three groups of similar piles,
but is restricted to the case where all of the pile tops are at the same level.
The elastic displacement of pile tops was first taken into consideration by Westergaard in 1917
(Karol, 1960). Westergaard assumed linear-elastic displacement of pile tops under a compressive load,
but the lateral resistance of the piles was not considered. He developed a method that finds the center
of rotation of a pile cap, which allows for computation of the displacements and forces in each pile.
Later, in 1953, Vandepitte applied the concept of the elastic center in developing the ultimate-
design method, which was further formulated by (Hansen B. , 1959). The transitional stage where some
of the piles in the foundation reach their ultimate bearing capacity, while the remainder are in an elastic
range, can be computed by a purely elastic method if the reactions of the piles in their ultimate stages
are regarded as constant forces on the cap. The failure of the cap is reached after successive failures of
all but the last two piles. Then the cap can rotate around the intersection of the axis of the two elastic
piles. Vandepitte resorted to a graphical solution to directly compute the ultimate load of a two-
dimensional cap. Hansen extended the method to the three-dimensional case. Although the plastic-
design method is unique and rational, its assumptions simplifying the real soil-structure system need
examination. It was assumed that a pile had only axial resistance, but no lateral resistance, and
rotational restraint of the pile tops on the cap was not considered. The axial load versus displacement of
each individual pile was represented by a bilinear relationship.
A comprehensive, modern structural treatment was presented by Hrennikoff for the two-
dimensional case (Hrennikoff, 1950). He considered the axial, transverse, and rotational resistance of
piles on the cap. The load-displacement relationship of the pile top was assumed to be linearly elastic.
One restrictive assumption was that all piles must have the same load-displacement relationship.
Hrennikoff substituted a free-standing elastic column for an axially loaded pile. A laterally loaded pile
was regarded as an elastic beam on an elastic foundation with uniform stiffness. Even with these crude
approximations of pile behavior, the method is significant since it potentially allows for an analytical
treatment of the soil-pile-interaction system. Hrennikoff’s method consisted of obtaining influence
coefficients for cap displacements by summing the influence coefficients of individual piles in terms of
their spring constants, letting the spring constants represent the pile-head reactions onto the pile cap.
Almost all subsequent work follows his approach.
Radosavljevic (1957) also regarded a laterally loaded pile as an elastic beam in an elastic
medium with a uniform stiffness. He advocated for the use of the results of tests of single piles under
axial loading. In this way a designer can choose the most practical spring constant for the axially-loaded-
pile head, and nonlinear behavior can also be considered. Radsoavljevic showed a slightly different
formulation than Hrennikoff in deriving the coefficients for the equations of force equilibrium. Instead
of using unit displacement of a cap, he used an arbitrary set of displacements. Still, his structural
approach is essentially analogous to Hrennikoffs method. However, Radosavljevics method is restricted
to the case of identical piles in identical soil conditions.
Turzynski (1960) presented a formulation using a matrix method for the two-dimensional case.
Neglecting the lateral resistance of pile sand soil, he considered only the axial resistance of piles.
Further, he assumed piles were elastic columns pinned at the top and at the tip. He derived a stiffness
matrix and inverted it to obtain the flexibility matrix. Except for the matrix method, Turzynskis method is
not practical because it oversimplifies the soil-pile interaction system.
(Asplund, 195) formulated the matrix method for both two-and three-dimensional cases. His
method also starts by calculating a stiffness matrix to obtain a flexibility matrix by inversion. In an
attempt to simplify the final flexibility matrix, Asplund defined a pile-group center by which the
flexibility matrix is diagonalized. He stressed the importance of the pile arrangement for an economical
grouped-pile foundation and contended that the pile-group-center method allowed for better
visualization of the effect of the geometric factors. He employed the elastic-center method for the
treatment of laterally loaded piles. Any transverse load through the elastic center causes only transverse
displacement of the pile head, and rotational load around the elastic center gives only the rotation of
the pile head. In spite of the elaborate structural formulation, there is no particular correlation with the
soil-pile system. Laterally loaded piles are merely regarded as elastic beams on an elastic bed with a
uniform spring constant.
(Francis, 1964) computed the two-dimensional case using the influence-coefficient method. The
lateral resistance of soil was considered either uniform throughout or increasing in proportion to depth.
Assuming a fictitious point of fixity at a certain depth, elastic columns fixed at both ends are substituted
for laterally loaded piles. The axial loads on individual piles are assumed to have an effect only on elastic
stability, and no effect on settlement or uplift at the pile tips.
(Saul, 1968) gave the most general formulation of the matrix method for a three-dimensional
foundation with rigidly connected piles. He employed the cantilever method to describe the behavior of
laterally loaded piles. He left it to the designer to set the soil criteria for determining the settlement of
axially loaded piles and the resistance of laterally loaded piles. Saul indicated a possible application of
his method to dynamically loaded foundations.
(Reese & Matlock, Numerical analysis of laterally loaded piles, 1960), (Reese & Matlock, 1966)
presented a method for coupling the analysis of the grouped-pile foundation with the analysis of
laterally loaded piles by the finite-difference method. The method of Reese and Matlock presumes the
use of computers. The finite-difference method of analyzing a laterally loaded pile can handle a pile of
varying size and flexural rigidity in any complex profile of highly nonlinear soils. The method can account
for the behavior of any soil system providing that the soil behavior can be described analytically or
numerically. Any type of boundary conditions of the pile head can be treated; namely, the fixed, pinned,
or elastically-restrained pile head. A nonlinear curve showing axial load versus pile-head deflection is
employed in the analysis. The curve may either be derived by computations based on proper
assumptions, or it may be obtained from field-loading tests. The formulation of equations giving the
movement of the pile cap is done by the influence coefficient method, similar to Hrenikoff’s method.
Reese and Matlock devised a convenient way to represent both the pile-head moment and lateral
reaction using only spring forces in terms of the lateral displacement of a pile top. The effect of pile-
head rotation on the pile head reactions is included implicitly in the force-displacement relationship. The
significance of the Reese and Matlock method lies in the fact that it can predict the bent cap behavior
continuously for incremental loading until the bent cap fails by excessive movement, while incorporating
the nonlinear relationships between pile-head loads and displacement in to the analysis.
Using the method of Reese and Matlock, example problems were worked out by (Robertson,
1961) and (Parker & Cox, 1969). Robertson compared the method with Vetter’s method and Hrenikoff’s
method and Parker and Cox integrated typical soil criteria for laterally loaded piles into Reese and
Matlock’s method.
(Reese & O’Neill, 1967) developed the theory for the general analysis of a three-dimensional
group of piles using matrix formulations. Their theory is an extension of the theory of (Hrennikoff, 1950),
in which springs are used to represent the piles. Representation of piles by springs imposes the
superposition of two independent modes of deflection of a laterally loaded pile. The spring constants for
the lateral reaction and the moment at the pile top must be obtained for a mode of deflection where a
pile head is given only transitional displacement without rotation, and also for a mode of deflection
where a pile head is given only rotation without translation. While the soil-pile-interaction system has
highly nonlinear relationships, the pile material also exhibits nonlinear characteristics when it is loaded
near its ultimate strength.
The method of analysis used herein is based on the concept presented by (Reese & Matlock,
1966), but the problem of superimposing nonlinear systems was resolved by means of a successive-
displacement-correction method for foundations with a rigid cap. The successive-displacement-
correction method does not require the superposition of independent modes of pile deflection, which is
sometimes inefficient for nonlinear systems.
The influence of spacing between piles can be illustrated by referring to Figure 1.1. We make the
assumption that all of the piles are fastened to a cap or superstructure, and that the lateral deflection of
all the piles is equivalent, or nearly so. Figure 1.1a shows three closely spaced piles that are in line. It is
evident that the resistance of the soil against Pile 2 is less than that for an isolated pile because of the
presence of Piles 1 and 3. Pile 2 may be considered to be in the shadow of Pile 3; the shadow-effect on
soil resistance is obviously related to pile spacing.
Similarly, the soil resistance against Pile 2 in Figure 1.1b is influenced by the presence of Piles 1
and 3. The edge-effect on soil resistance is again influenced by pile spacing.
Some authors have used the theory of elasticity, or a modified version of that theory, to develop
influence coefficients that are related to the geometry shown in Figure 1.1c (Poulos & Davis, Pile
Foundation Analysis and Design, 1980); (Focht & Koch, 1973). Those coefficients show that the influence
of Pile 1 on Pile 2 can be related to the pile spacing nb and to the angle β.
Research continues on developing quantitative expressions from full-scale experiments that will
give the influence of pile spacing on pile-soil-pile interaction (Brown & Reese, Behavior of a large-scale
pile group subjected to cyclic lateral loading, 1985); (Morrison & Reese, 1988); (O’Neill, Group action in
offshore piles, 1983). No definite conclusions have been reached, but a procedure that appears to be
useful is illustrated in Figure 1.2. For piles such as Pile 2 in Figure 1.1, the p-y curve is reduced by
multiplying the soil resistance p for an isolated pile by a reduction factor △1 and/or by multiplying the
deflection y by an expansion factor △2. The p-y curves that are generated internally in the computer can
be weakened according to recommendations that are being developed by research studies. Thus, the
portion of the total load that is taken by Pile 2, for example, will be computed to be less than that taken
by an isolated pile.
The analysis employed in this program for closely spaced pile groups is generally called the
hybrid method. A more detailed discussion of this method is presented in Chapter 4.
Most of the analyses can be simplified to the two-dimensional condition, where the loading is
along a plane of symmetry. The analysis using a two-dimensional model is even more convenient when a
pile group consists of more than 20 individual piles. However, if piles in a pile group are not
symmetrically arranged, or piles are battered in various directions, a three-dimensional model is
recommended.
The analysis of three-dimensional batter piles under a rigid cap becomes complex, not because
the theory is complex, but because the additional degrees of freedom requires more bookkeeping. The
basic theory for the nonlinear model of a three-dimensional pile was presented by Reese et al., in 1970.
The transformation matrix used to convert between local and global coordinates for three-dimensional
batter piles requires more attention than those for two-dimensional cases. The details of the methods of
analysis are presented in the next chapter.
As a foundation problem, the analysis of a pile group under axial and lateral loading is
complicated by the fact that soil reaction (resistance) is dependent on pile movement, which is in turn
dependent on soil reaction. Thus, the problem is one of soil-structure interaction and is nonlinear
because soil response is a nonlinear function of pile deflection.
The use of models for the analysis of bridge behavior is shown in Figure 2.1a. A simple, two-span
bridge is shown with spans in the order of 30m and with piles supporting the abutments and central
span. The girders and columns are modeled by lumped masses, and the foundations are modeled by
nonlinear springs, as shown in Figure 2.1b. If the loading is three-dimensional, the cap at the central
span will undergo three translations and three rotations. A simple matrix-formulation for the group of
piles is shown in Figure 2.1c. It assumes two-dimensional loading and a set of mechanisms for the
modeling of the foundation. Three springs are shown as symbols for the response of the pile cap to
loading: one for axial load, one for lateral load, and one for moment. The assumption is made that the
nonlinear curve for axial loading is not influenced by lateral loading (shear) and moment. The
assumption is not strictly true because lateral loading can cause a molding away of over-consolidated
clay around the top of the pile, causing a loss of load transfer in skin friction along the upper portion of
the pile. However, in that case the soil near the ground surface can be ignored above the first point of
zero lateral deflection. Typically, the curve of axial load versus settlement, and the term K11, are
negligibly affected.
The curves representing the response to shear and moment at the top of the cap are certainly
multi-dimensional and unavoidably so. Figure 2.1c shows a curve and identifies one of the stiffness
terms K32. A single-valued curve is shown only because a given ratio of moment M1 and shear V1 was
selected in computing the curve. Therefore, because such a ratio would be unknown in the general case,
iteration is required between the solutions of the superstructure and the foundation. The normal
procedure is to select values for shear and moment at the pile cap and then to compute the initial
stiffness terms so that the solution of the superstructure can proceed for the most critical cases of
loading. With revised values of shear and moment at the pile cap, the model for the group can be
resolved, and then the revised terms for stiffness can be used to obtain a new solution for the model of
the superstructure. Such a procedure can be performed automatically by a global computer program,
but most designers prefer to use independent models in order to exercise engineering judgment in
achieving compatibility and equilibrium in the entire system for a given case of loading.
An interesting presentation of the forces that resist the displacement of an isolated pile is
shown in Figure 2.2 (Bryant & Matlock, 1977).Figure 2.2a shows a single pile beneath a cap, along with
its three-dimensional displacements and rotations. The assumption is made that the top of the pile is
fixed, or partially fixed, to the cap, so that bending moments and torque will develop as a result of the
three-dimensional rotation of the cap. The various reactions of the soil along the pile are shown in
Figure 2.2b, and the soil-resistance curves are shown in Figure 2.2c. The argument given earlier about
the curve for axial displacement being single-value also pertains to the curve for torque. However, the
curve for lateral deflection is a function of the shears and moments that cause such deflection. In regard
to the lateral deflection, a complication can arise because the deflection may not be in a two-
dimensional plane. The recommendations that have been made for correlating the lateral resistance
with pile geometry and soil properties all depend on the results of loading in a two-dimensional plane.
As discussed in the User Manual, GROUP provides the necessary information for evaluating the
stiffness of a grouped-pile foundation. Stiffness matrices are often used to model foundations in
structural analyses. An option in GROUP allows the user to create curves, similar to those shown in
Figure 2.3, of pile-cap movements and rotations as functions of incremental loadings. The program
divides the loads specified at the pile cap into 10 unequal increments and then computes the pile cap
response for each individual loading. The deflection and rotation of the pile cap is computed for each
lateral load and bending moment increment. The user can thus define the flexibility matrix directly
based on the relationship between computed deformation and applied load. For instance, the flexibility
coefficient S23, shown in Figure 2.3, can be obtained by dividing the computed deflection Yi by the
applied moment Mi at the pile cap.
Most structural analyses can employ either the stiffness matrix or the flexibility matrix to define
the support condition at the pile cap. If the user prefers to use the stiffness matrix in the structural
model, Figure 2.4 illustrates a basic procedure to convert the flexibility matrix to the stiffness matrix.
The initial coefficients for the flexibility matrix may be defined based on the magnitude of the service
load. The user may need to repeat several iterations before achieving comfortable agreement between
flexibility coefficients.
Figure 2.2 Three-dimensional soil-pile interaction (After Bryant and Matlock, 1977)
The correct modeling of the problem of the single pile in response to axial and lateral loading is
challenging and complex, and the modeling of a group of piles even more so. However, in spite of the
fact that much more still needs to be learned, the following chapters will demonstrate that useable
solutions are at hand.
3.1 Introduction
A structural theory is formulated herein for computing the behavior of a pile foundation with
arbitrarily arranged piles that possess nonlinear force-displacement characteristics. Coupled with the
structural theory of a pile cap are the theories for a laterally loaded pile and an axially loaded pile. In this
chapter each theory is developed separately. Solutions for all of the theories depend on the use of
computers for the actual computations.
The general system of a two-dimensional pile foundation is illustrated in Figure 3.1a. Three piles
with arbitrary spacing and inclination are connected to an arbitrarily shaped pile cap. Sectional
properties of a pile such as the width, cross-sectional area, and moment of inertia can vary not only
from pile to pile, but also along the axis of a pile. The pile material may be different from pile to pile but
we assume that each individual pile is made of the same material.
There are three conceivable cases of pile connection to the pile cap. Pile 1 in Figure 3.1a and
Figure 3.1c illustrates a pin connection, Pile 2 shows a fixed-head pile with its head clamped by the pile
cap, and Pile 3 represents an elastically restrained pile, which is the typical case for an offshore bent
(Figure 3.2). The piles can be grouted into the superstructure, but in the case shown the piles rest
against knife-edge supports and can deflect freely between those supports. Elastic restraint is provided
by the flexural rigidity of the pile itself. This treatment of a laterally loaded pile with an elastically
restrained top provides a useful tool for handling the real foundation.
Piles are frequently embedded into a monolithic, concrete-reinforced pile cap with the
assumption that complete fixity of the pile to the pile cap is obtained (Figure 3.3). However, the
elasticity of the reinforced concrete and local failure due to stress concentrations allow for the rotation
of a pile head within the pile cap. The magnitude of the restraint on the pile from the pile cap is
indeterminate, but perhaps a range of values can be estimated to bracket the behavior of the pile group.
The pile cap is subject to two-dimensional external loads. The line of action of the resultant
external load may be inclined and may assume any arbitrary position with respect to the structure. The
external loads cause displacement of the pile cap, resulting in axial, lateral, and rotational displacements
of each individual pile (Figure 3.1c). The displacements of individual piles in turn result in loads on the
pile cap, as illustrated in Figure 3.1b. These pile reactions are highly nonlinear in nature. They are
functions of pile properties, soil properties, and the boundary conditions at the pile top.
The structural theory of pile groups uses a numerical method to seek compatible displacement
of the pile cap, which should satisfy the equilibrium of the applied external loads and the nonlinear pile
reactions.
3.1.2 Assumptions
Some of the basic assumptions employed for the treatment of pile groups are presented below:
We first assume a two-dimensional arrangement of the bent cap and the piles. The usual design
practice is to arrange piles symmetrically with a plane or planes, with loads acting in this plane of
symmetry. The assumption of a two-dimensional case considerably reduces the number of variables to
be handled. However, in theory there is no essential difference between the two-dimensional case and
the three-dimensional case. GROUP allows the user to analyze the behavior of pile groups using either
the 2-or 3-dimensional model. If the piles are not symmetrical to the loading plane, the 3-dimensional
model should be considered.
The second major assumption is the non-deformability of the pile cap. A pile head encased in a
monolithic pile cap (Figure 3.3) or supported by a pair of knife-edge supports (Figure 3.2) can rotate or
deflect within the pile cap, but the shape of the pile cap itself is assumed to always be the same. This
means that the relative positions of the pile tops remain the same for any pile-cap displacement. If the
pile cap is deformable, the structural theory of the pile group must include the compatibility condition
of the pile cap itself. No treatment for a foundation with a deformable pile cap is included in this study,
but the theory could be extended to the case where the pile cap consists of a structural member for
which the analytical computation of the deformation of the pile cap is possible.
As was explained in Chapter 1, the program is directed principally at the case where the
individual piles are so widely spaced that the piles do not influence one another. However, in many pile
designs, the piles are close enough for pile-soil-pile interaction to occur. While some recommendations
are included in this manual to reflect the effect of pile-soil-pile interaction, options are also provided in
the program to allow the user to analyze a group of closely spaced piles. The user can refer to technical
literature cited in Chapter 6 for further information.
3.1.2.4 Behavior Under Lateral Load and Under Axial Load are Independent
The assumption is made that there is no interaction between axial-pile and lateral-pile behavior. In
other words, the relationship between axial load and displacement is not affected by the presence of
lateral deflection, and vice versa. The validity of this assumption is discussed by (Parker & Reese, 1971).
The generally accepted argument is that the soil near the ground surface determines the lateral
response, and the soil depth determines the axial response. If over-consolidated clay exists at the
ground surface and pile deflection is sizable, the recommendation is made that the soil above the first
point where lateral deflection is zero be discounted in computing axial capacity.
The equilibrium on a pile cap of applied loads and pile reactions are found using the successive
correction of pile-cap displacements. After each correction of the displacement, the difference between
the load and pile reaction is calculated. The next correction is obtained through the calculation of a new
stiffness matrix at the previous pile-cap position. The elements of a stiffness matrix are obtained by
giving a small virtual increment to each component of displacement, one at a time. The proper
magnitude of the virtual increment may be set at 110−5 times a unit displacement to attain acceptable
accuracy.
Figure 3.4a shows the coordinate systems and sign conventions. The superstructure and pile cap
use the global structural coordinate system (X, Y) where the X and Y axes are vertical and horizontal,
respectively. The resultant external forces act at the origin 0 of this global structural coordinate system.
The positive directions of the components of the resultant load P0, Q0, and M0 are shown by the arrows.
The positive curl of the moment was determined by the usual right-hand rule. The pile head of each
individual pile in a group is referred to the local structural coordinate system (x'i, y'i), whose origin is the
pile head and has axes running parallel to those of the global structural coordinate system. The member
coordinate system (xi, yi) is further assigned to each pile. The origin of the member coordinate system is
the pile head. Its xi axis coincides with the pile axis and its yi axis is perpendicular to the xi axis. The xi axis
makes an angle λi with the vertical. λi is positive when it is measured counterclockwise.
Figure 3.4b shows the positive directions of the forces, Pi, Qi, and Mi that are exerted from the
pile cap onto the top of an individual pile in the ith individual pile group. The forces Pi and Qi are acting
along the xi and yi axes of the member coordinate system, respectively.
3.2.2.1 Displacement
Figure 3.5 illustrates the pile-head displacement in the structural, local structural, and member
coordinate systems. Due to the pile-cap displacement from point 0 to 0' with rotation α, the ith pile
moves from the original position P to the new position P'. The rotation of the pile head depends on the
way it is fastened to the cap. The components of pile-cap displacement are expressed by (U, V, α) with
regard to the structural coordinate system. The pile-head displacement is denoted by (u'i, v'i, α) in the
local coordinate system, and by (ui, vi, α) in the member coordinate system.
The coordinate transformation between the structural and the local structural coordinate
system is derived from the following simple geometric consideration:
𝑢′ 𝑖 = 𝑈 − 𝑌𝑖 𝛼 (3.1)
𝑣 ′ 𝑖 = 𝑉 + 𝑋𝑖 𝛼 (3.2)
where (Xi, Yi) is the location of the ith pile head in the structural coordinate system.
The transformation of pile-head displacement from the local structural coordinate system to the
member coordinate system is obtained from another geometric relationship (Figure 3.5):
Substitution of equations 3.1 and 3.2 into equations 3.3 and 3.4 yields the transformation relationship
between the pile-cap displacement in the structural coordinate system and the corresponding pile-top
displacement of the ith individual pile group in the member coordinate system.
Equation 3.7 presents in matrix notation the case where the pile head is fixed to the cap, causing
the cap and the pile head to rotate the same amount.
The expressions for Ui and Vi will remain the same for cases where the pile head is free to rotate
or is partially restrained, but the expression for αi must be modified. The matrix expression above is
written concisely:
𝑢𝑖 = T𝐷,𝑖 𝑈 (3.8)
where:
ui = displacement vector of the head of the pile in the ith individual pile group,
3.2.2.2 Force
Figure 3.5 illustrates the action of the load on the pile cap and the pile reactions. The load is
expressed in three components (P0, Q0, M0) with regard to the structural coordinate system. The
reactions in the ith individual pile group are expressed in terms of the member coordinate system (Pi, Qi,
Mi). Decomposition of the reactions of the ith pile with respect to the structural coordinate system gives
the transformation of the pile reaction from the member coordinate system to the structural coordinate
system.
or more concisely:
where:
P'i = reaction vector of the pile of the ith individual pile group in the structural coordinate
system,
It is observed that the force transformation matrix TF,i is obtained by transposing the
displacement transformation matrix TD,i. Thus,
After successive correction, the pile cap moves from the initial position 0 to the last position 0'
with new displacement components (U, V, α). If the three components of the displacement of a pile cap
are given, the displacement of each pile head may be computed by Equation 3.8 or one that is modified
to reflect the way the pile head is fastened to the cap. Then, the theories for laterally loaded and axially
loaded piles, presented in the following sections, may be used to solve for the reaction vector fi of each
pile numerically. If curves are available for axial load versus the displacement of the pile top, the axial-
pile reactions may be directly obtained by reading the curves. The summation of the reaction vector
with respect to the structural coordinate system is given by:
𝑛 𝑛
′
𝑅 = ∑ 𝐽𝑖 𝑝 𝑖 = ∑ 𝐽𝑖 𝑇𝐹,𝑖 𝑝′ 𝑖
𝑖=1 𝑖=1
(3.15)
where
The difference of the applied load and the pile reactions, or the force correction, is
𝑑𝐹 = 𝐿 − 𝑅 (3.16)
where
Each element of the reaction vector is a highly nonlinear function of the pile-cap displacement
(U, V, α). So
𝑃𝑅 = 𝑃𝑅 (𝑈, 𝑉, 𝛼) (3.17)
𝑄𝑅 = 𝑄𝑅 (𝑈, 𝑉, 𝛼) (3.18)
𝑀𝑅 = 𝑀𝑅 (𝑈, 𝑉, 𝛼) (3.19)
in matrix notation
or in concise form:
𝑑𝑅 = 𝐾 𝑑𝑉 (3.24)
where:
The elements of the stiffness matrix K, or the partial derivatives of the total reaction forces with
respect to each element of the pile-cap displacement, are obtained by giving small displacement dX, dY,
and dα one at a time to the pile cap (Figure 3.6).
Giving a small displacement dU in the x direction, three elements in the first column of the
stiffness matrix K are determined.
Equating the reaction variation vector dR with the force correction vector dF and substituting
the displacement variation vector dU for the displacement correction vector dV, the necessary
correction of the displacement is obtained from Equation 3.24.
𝑑𝑈 = 𝐾 −1 𝑑𝐹 (3.28)
The new displacement of the pile cap is given by adding the displacement correction vector dU
to the displacement vector U.
The size of the stiffness matrix K is only three by three. Therefore, the inversion of the matrix K
is most conveniently done by the Cramer rule. That is, an element of the flexibility matrix or the inverted
matrix K-1 is expressed by:
𝐴𝑖𝑗
𝐾𝑖𝑗 −1 =
det 𝐾
(3.29)
where
The solution is obtained through an iterative numerical procedure and GROUP makes the
necessary computations. The following describes the logic of the computer program:
5. Compute the difference between the applied load and the pile reactions to obtain the force
correction vector.
8. Multiply the flexibility matrix (Step 7) with the force correction vector (Step 5) to get the
displacement correction vector.
Repeat Steps 2 through 9 until the displacement-correction vector becomes sufficiently small.
The analytical prediction of the axial and lateral behavior of a pile may be made by finite
difference methods. The remaining portion of this chapter is devoted to formulation of the finite
difference methods for laterally loaded and axially loaded piles.
GROUP is internally equipped with the finite difference method for a laterally loaded pile, but
not for an axially loaded pile. The main reason for this decision is that the relationship between axial
load and displacement is expressed by a single curve, while the lateral and rotational pile reactions
consist of families of curves in terms of multi-parameters such as the lateral and rotational
displacements, and the type of pile connections to the pile cap.
The analysis of three-dimensional batter piles under a rigid cap is complex because additional
degrees of freedom require more “bookkeeping.” The basic model for a three-dimensional nonlinear
pile was presented by Reese et al. (1970). The transformation matrix used to convert between local and
global coordinates for three-dimensional batter piles requires more attention than those for two-
dimensional cases.
We consider a structure supported by piles. The pile-cap and the individual piles use the
coordinate systems shown in Fig. 3.7. We describe them:
The superstructure and the pile-cap use the right-handed Cartesian coordinate system (0; X, Y,
Z). The origin 0 of this system is a point in the pile-cap. The ZY-plane is the plane of the pile cap. The X-
axis is perpendicular to the ZY -plane and is positive downwards.
With respect to this coordinate system, the coordinates of the point Pi, which is at the
intersection of the longitudinal axis of pile i and the pile-cap are (Xi, Yi, Zi).
The point Pi on the longitudinal axis of the pile at the pile-head is taken as the origin of a local
structural coordinate system (Pi ; X'i, Y'i, Z'i), whose axes are parallel to those of the global structural
coordinate system.
To each pile i one assigns the pile coordinate system (Pi ; Xi, Yi, Zi) with origin Pi. The Xi -axis
coincides with the longitudinal axis of the pile and is positive downwards. The Yi -axis is perpendicular to
the Xi -axis and is fixed in the Xi Yi-plane, such that the angle between the positive Yi -direction and the
positive Y'i -direction is less than 90◦. The Xi -axis is perpendicular to the XiYi -plane and oriented such
that (Pi ; Xi , Yi , Zi ) is a right-handed coordinate system.
The orientation of each pile i with respect to its local structural coordinate (Pi ; Xi , Yi , Zi ) is
described by the direction angles i, i, i. More precisely:
i is the angle between the positive directions of the X'i and the Xi -axis,
i, is the angle between the positive directions of the Y'i and the Xi -axis, and
i is the angle between the positive directions of the Z'i and the Xi -axis.
The angle between two oriented axes is always less than or equal to 180◦.
𝐿𝑒 = (𝐹 𝑒 𝑋 , 𝐹 𝑒 𝑌 , 𝐹 𝑒 𝑍 , 𝑀𝑒 𝑋 , 𝑀𝑒 𝑌 , 𝑀𝑒 𝑍 ) (3.30)
where 𝐹 𝑒 𝑋 , 𝐹 𝑒 𝑌 , 𝐹 𝑒 𝑍 , 𝑀𝑒 𝑋 , 𝑀𝑒 𝑌 , 𝑀𝑒 𝑍 are the external forces and moments with respect to the
global structural coordinate system (0; X, Y, Z).
The goal is to obtain the equilibrium state on the pile cap of the applied external load and the
pile reactions. This can be achieved by using the successive-displacement-correction method. It is an
iterative numerical procedure which comprises of the following sequence of steps.
2. For each individual pile i, compute the induced displacement of the pile-top.
3. For each individual pile i, compute the pile reaction vector for the given pile-top displacement.
4. Compute the total pile reaction vector on the pile-cap by summing the individual pile reaction
vectors.
5. Compute the force correction vector by taking the difference of the applied load and the total
pile reaction vector.
6. Compute the stiffness matrix (to do that, one has to take small increments for each component
of the pile-cap displacement vector, one at a time, and go through steps 1 to 4).
7. Compute the flexibility matrix by taking the inverse of the stiffness matrix.
8. Get the displacement correction vector by applying the flexibility matrix to the force correction
vector.
9. Correct the pile-cap displacement by adding the displacement correction vector computed in
(8).
Repeat steps (2) through (9) until the displacement correction vector becomes very small.
𝑈 = (𝑈, 𝑉, 𝑊, 𝑥 , 𝑦 , 𝑧 ) (3.31)
The vectors 𝑈(𝑈, 𝑉, 𝑊) and (𝑥 , 𝑦 , 𝑧 ) represent the translation and rotation vectors of the
pile-cap with respect to the global structural coordinate system (0; X, Y, Z) (all displacement
parameters are assumed small).
The displacement of the pile-cap induces a displacement of each pile i. The pile-head Pi moves
to the new position P'i . We describe the displacement of the pile-head with respect to the pile
coordinate system (Pi ; xi , yi , zi) by a six-component displacement vector:
The first three components 𝑢𝑖 , 𝑣𝑖 , 𝑤𝑖 represent translations along the x, y, and z-axis, respectively, and
the remaining ones, 𝑥𝑖 , 𝑦𝑖 , 𝑧𝑖 , represent rotations around their corresponding axes. The
displacement vector ui of the pile-top Pi is related to the displacement vector U of the pile-cap by the
deformation-transformation matrix Ti:
𝑢𝑖 = 𝑇𝑖 𝑈 (3.33)
𝑖 𝑖 𝑖 𝑌𝑖 𝑖 − 𝑍𝑖 𝑖 𝑍𝑖 𝑖 − 𝑋𝑖 𝑖 𝑋𝑖 𝑖 − 𝑌𝑖 𝑖
𝑖 𝑖 𝑖 𝑖 𝑌𝑖 𝑖 + 𝑋𝑖 𝑖
− 0 −𝑍𝑖 −𝑍𝑖
𝑖 𝑖 𝑖 𝑖 𝑖
𝑖 𝑖 𝑖
− − 𝑖 𝑖 𝑖 𝑌𝑖 𝑖 + 𝑍𝑖 𝑖 𝑖 −𝑋𝑖 𝑖 − 𝑍𝑖 𝑖 𝑖 (𝑌 − 𝑋𝑖 𝑖 )
𝑖 𝑖 𝑖 𝑖 𝑖 𝑖 𝑖
𝑇𝑖 =
𝑖 𝑖 𝑖
0 0 0 𝑖
− 𝑖 0
0 0 0 𝑖 𝑖
0 0 𝑖 𝑖
− − 𝑖 𝑖 𝑖
[ 𝑖 𝑖 ]
(3.34)
We let
denote the reaction vector at the pile-top (forces f and moments m) with respect to the pile
coordinate system (Pi ; xi , yi , zi) and let
denote the reaction vector at the pile-top with respect to the global structural coordinate
system (0; X, Y, Z) . This gives the following matrix relationship:
𝑅𝑖 = 𝑇𝑖 𝑡 𝑟𝑖 (3.37)
The pile reaction vector 𝑟𝑖 is calculated under the assumption of independence between the
behavior of the pile and its response to axial, lateral and torsional deformations.
3.4.1 Introductions
In the framework presented in Section 3.3, the pile-cap is modeled as a rigid body. Under this
framework, the pile-cap deformation can be represented only by one point (the origin of
coordinates) since the deformation of all other points can be estimated using rigid body
equations. For a three-dimensional analysis the rigid pile-cap can be modeled using a 6 degree of
freedom system, that correspond to 3 translations and 3 rotations.
For flexible pile-cap, the pile-cap is discretized using a finite element approach. Plate elements
and/or beam elements can be used in order to model the flexible pile cap
Each element considers the contribution of in-plane stresses (membrane effects) and out of plane
stresses (bending effects). Bending, also called plate, effects consider three degrees of freedom
per node, out of plane displacement (ux) and two bending rotations (y and z). Membrane effects
consider two degrees of freedom per node, in-plane displacements (uy and uz). By consider an
element with drilling degrees of freedom, the membrane effects can be associated to three
degrees of freedom per node, two in-plane displacements (uy and uz) and one drilling rotation
(x). Therefore, each node that belongs to the finite element mesh that defines the flexible pile-
cap can be represented for 6 degrees of freedom, three displacements and three rotations. In the
program, the finite element used is called plate that includes both effects: in-plane (membrane)
and out of plane (bending)
- Stresses
Considering an infinitesimal cubic element isolated from within a deformable body such that its
faces normal to the coordinate axes. Then the traction acting along each face can be decomposed in
terms of individual components along each axis referred to as the stress components.
These components can be referred as the stress tensor [𝝈] which is represented in Cartesian
coordinates as:
The stress tensor is considered to be symmetric, then 𝜏𝑥𝑦 = 𝜏𝑦𝑥 , 𝜏𝑥𝑧 = 𝜏𝑧𝑥 and 𝜏𝑦𝑧 = 𝜏𝑧𝑦 . Due
to its symmetry, it may be convenient to express the stress tensor in a vector form
- Compatibility equations
These equations provide the relationship between strains and displacements (ux, uy, uz). For small
deformations the relations can be written as:
𝜕𝑢𝑥
𝜀𝑥𝑥 =
𝜕𝑥
(3.40)
𝜕𝑢𝑦
𝜀𝑦𝑦 =
𝜕𝑦
(3.41)
𝜕𝑢𝑧
𝜀𝑧𝑧 =
𝜕𝑧
(3.42)
𝜕𝑢𝑥 𝜕𝑢𝑦
𝛾𝑥𝑦 = +
𝜕𝑦 𝜕𝑥
(3.43)
𝜕𝑢𝑥 𝜕𝑢𝑧
𝛾𝑥𝑧 = +
𝜕𝑧 𝜕𝑥
(3.44)
𝜕𝑢𝑦 𝜕𝑢𝑧
𝛾𝑦𝑧 = +
𝜕𝑧 𝜕𝑦
(3.45)
Similarly, to the stress tensor, the strain tensor can be written in a vector form as:
𝜕
0 0
𝜕𝑥
𝜕
0 0
𝜀𝑥𝑥 𝜕𝑦
𝜀𝑦𝑦 𝜕
0 0 𝑢𝑥
𝜀 𝜕𝑧 𝑢
{𝜀} = 𝛾𝑧𝑧 = { 𝑦 } = [𝐿] {𝑈}
𝑥𝑦 𝜕 𝜕
𝛾𝑥𝑧 0 𝑢𝑧
𝜕𝑦 𝜕𝑥
{ 𝛾𝑦𝑧 } 𝜕 𝜕
0
𝜕𝑧 𝜕𝑥
𝜕 𝜕
0
[ 𝜕𝑧 𝜕𝑦]
(3.47)
- Constitutive equations
For an isotropic linear elastic model, the constitutive equations can be expressed as:
1−𝜈 𝜈 𝜈 0 0 0
𝜈 1−𝜈 𝜈 0 0 0
𝜎𝑥𝑥 𝜀𝑥𝑥
𝜈 𝜈 1−𝜈 0 0 0
𝜎𝑦𝑦 1 − 2𝜈 𝜀𝑦𝑦
𝜎 𝐸 0 0 0 0 0 𝜀𝑧𝑧
{𝜎} = 𝜏 𝑧𝑧 = 2 𝛾𝑥𝑦 = [𝐷] {𝜀}
𝑥𝑦 (1 + 𝜈)(1 − 2𝜈) 1 − 2𝜈
𝜏𝑥𝑧 0 0 𝛾𝑥𝑧
0 0 0 2
{ 𝜏𝑦𝑧 } { 𝛾𝑦𝑧 }
1 − 2𝜈
[ 0 0 0 0 0
2 ]
(3.48)
For a membrane problem, the thickness is considered to a small, then the out of plane stresses
are negligible (𝜎𝑥𝑥 = 𝜏𝑥𝑦 = 𝜏𝑥𝑧 = 0). Then the constitutive relations can be written as:
𝜎𝑦𝑦 1 𝜈 0 𝜀𝑦𝑦
𝐸 𝜈 1 0
{𝜎} = { 𝜎𝑧𝑧 } = [ 𝜀
1 − 𝜈 ] { 𝑧𝑧 } = [𝐷] {𝜀}
𝜏𝑦𝑧 (1 − 𝜈 2 ) 0 0 𝛾𝑦𝑧
2
(3.49)
To find the contribution of membrane effects a finite element model is considered. The 4-node
rectangular element with drilling nodes is used. Please refer to (Ibrahimbegovic, et. al., 1990) for more
details about the element.
Figure 3.9 Finite element model for membrane effects. (Ibrahimbegovic, et. al., 1990)
- Introduction
A typical plate which has the dimension of 𝑎 and 𝑏 on each side and thickness of ℎ under
uniform surcharge load 𝑝𝑥 (𝑦, 𝑧) is shown in Figure 3.10. Detailed derivation of the governing equations
for a plate under various conditions can be reviewed in (Szilard, 1973) and (Blaauwendraad, 2010).
Equations for a simple rectangular plate under vertical loads are presented here for reference.
Development of this software is based on the first order Mindlin thick plate theory. In this theory, there
exist three assumptions:
1. Deflection is much smaller than the plate thickness and the middle plane of the plate keeps
being free of stress and strain.
2. A straight line normal to the middle plane remains straight but may have a rotation angle
against the normal of the middle plane of plate due to warping.
3. Stress and strain perpendicular to the plate plane are small to be neglected.
𝑢𝑥 : deflection in 𝑥 direction
Figure 3.11 shows the definition of 𝜑𝑦 and 𝜑𝑧 , where 𝑢𝑦 , 𝑢𝑧 are displacement in y and z direction
respectively.
- Compatibility equations
These equations provide the relationship between strains and displacements (rotations). For small
deformations the relations can be written as:
𝜕𝑢𝑦 𝜕𝜑𝑦
𝜀𝑦𝑦 = = 𝑥
𝜕𝑦 𝜕𝑦
(3.50)
𝜕𝑢𝑧 𝜕𝜑𝑧
𝜀𝑧𝑧 = = 𝑥
𝜕𝑧 𝜕𝑧
(3.51)
Similarly, to the stress tensor, the strain tensor can be written in a vector form as:
0 0 0
𝜕
0 𝑥 0
𝜕𝑦
𝜀𝑥𝑥 𝜕
𝜀𝑦𝑦 0 0 𝑥
𝜕𝑧 𝑢𝑥
𝜀𝑧𝑧
{𝜀} = 𝛾
𝑥𝑦
= 𝜕 1 0 {𝜑𝑦 } = [𝐿] {𝑈}
𝛾𝑥𝑧 𝜕𝑦 𝜑𝑧
{ 𝛾𝑦𝑧 } 𝜕
0 1
𝜕𝑧
𝜕 𝜕
0 𝑥 𝑥
[ 𝜕𝑧 𝜕𝑦]
(3.56)
- Constitutive equations
For an isotropic linear elastic model, the constitutive equations can be expressed as:
𝐸 𝐸 𝜕𝜑𝑦 𝜕𝜑𝑧
𝜎𝑦𝑦 = 2 (𝜀𝑦𝑦 + 𝜈𝜀𝑧𝑧 ) = 𝑥 2 ( +𝜈 )
1−𝜈 1−𝜈 𝜕𝑦 𝜕𝑧
(3.57)
𝐸 𝐸 𝜕𝜑𝑧 𝜕𝜑𝑦
𝜎𝑧𝑧 = (𝜀𝑧𝑧 + 𝜈𝜀𝑦𝑦 ) = 𝑥 ( + 𝜈 )
1 − 𝜈2 1 − 𝜈 2 𝜕𝑧 𝜕𝑦
(3.58)
𝐸 𝐸 𝜕𝜑𝑦 𝜕𝜑𝑧
𝜏𝑦𝑧 = 𝛾𝑦𝑧 = 𝑥 ( + )
2(1 + 𝜈) 2(1 + 𝜈) 𝜕𝑧 𝜕𝑦
(3.59)
𝐸 𝐸 𝜕𝑢𝑥
𝜏𝑥𝑦 = 𝛾𝑥𝑦 = (𝜑𝑦 + )
2(1 + 𝜈) 2(1 + 𝜈) 𝜕𝑦
(3.60)
𝐸 𝐸 𝜕𝑢𝑥
𝜏𝑥𝑧 = 𝛾𝑥𝑧 = (𝜑𝑧 + )
2(1 + 𝜈) 2(1 + 𝜈) 𝜕𝑧
(3.61)
- Stress Resultant
Stresses on a plate element can be seen in Figure 3.12.
ℎ/2
𝑄𝑦𝑧 = ∫ 𝜏𝑦𝑧 𝑑𝑥 = 0
−ℎ/2
(3.65)
ℎ/2 ℎ/2
𝐸 𝜕𝑢𝑥
𝑄𝑦 = ∫ 𝜏𝑦𝑥 𝑑𝑥 = ∫ 𝑥 2 (𝜑𝑦 + ) 𝑑𝑥
−ℎ/2 2(1 + 𝜈) −ℎ/2 𝜕𝑦
(3.66)
ℎ/2 ℎ/2
𝐸 𝜕𝑢𝑥
𝑄𝑧 = ∫ 𝜏𝑧𝑥 𝑑𝑥 = ∫ 𝑥 2 (𝜑𝑧 + ) 𝑑𝑥
−ℎ/2 2(1 + 𝜈) −ℎ/2 𝜕𝑧
(3.67)
𝜕𝜑𝑦
𝑀𝑦 1 𝜈 0 𝜕𝑦
𝐸 ℎ3 𝜈 1 0 𝜕𝜑𝑧
𝑀
{ 𝑧}= [ 1 − 𝜈]
𝑀𝑦𝑧 12(1 − 𝜈 2 ) 𝜕𝑧
0 0
2 𝜕𝜑𝑦 𝜕𝜑𝑧
+
{ 𝜕𝑧 𝜕𝑦 }
(3.68)
𝜕𝑢𝑥
𝑄𝑦 𝐸ℎ 𝜑𝑦 +
1 0 𝜕𝑦
{ }= [ ]
𝑄𝑧 (1
2 𝜂 + 𝜈) 0 1 𝜕𝑢𝑥
{ 𝜑𝑧 + 𝜕𝑧 }
(3.69)
Figure 3.13 Finite element model for membrane effects. (Ibrahimbegovic, 1993)
A straight beam element of length L is considered. The element has 6 degrees of freedom in each
nod, three displacements and three rotations. Please refer to (Przemieniecki, 1968) for more
details about the element.
In order to perform the analysis of the pile group connected using a flexible pile-cap, a series of
iterations are carried-out by using the successive-displacement-correction method until convergence is
achieved. It is an iterative numerical procedure which comprises of the following sequence of steps.
2. For each individual pile i, compute the induced displacement of the pile top.
3. For each individual pile i, compute the pile reaction vector for the given pile-top displacement.
4. Compute all the reaction at the flexible pile cap due to pile reaction at each pile head.
5. Compute the total force correction vector by taking the difference of the applied load and the
total pile reaction vector in the flexible pile-cap.
6. For each individual pile i, compute the stiffness matrix at each pile head (to do that, one has to
take small increments for each degree of freedom of the pile head.
7. Compute all the stiffnesses at the flexible pile cap due to pile-stiffness at each pile head.
8. Compute the stiffness matrix for the flexible pile cap, considering the stiffnesses of all the plate
and/or beam elements that belong to the flexible pile cap and the stiffnesses due to the pile
heads.
9. Compute the flexibility matrix by taking the inverse of the stiffness matrix.
10. Get the displacement correction vector by applying the flexibility matrix to the force correction
vector.
11. Correct the pile-cap displacement by adding the displacement correction vector computed in
(10).
Repeat steps (2) through (11) until the displacement correction vector becomes very small (less
than the defined tolerance).
The analysis of the laterally loaded pile by the finite difference method has been undertaken
extensively by Reese and Matlock since1960 (Reese & Matlock, 1960); (Matlock & Reese, Generalized
solution for laterally loaded piles, 1962); (Matlock, 1963); (Matlock & Ingram, 1963); (Reese L. C., 1966);
(Reese L. C., 1971); (Matlock, 1970); (Parker & Reese, 1971); (Georgiadis, 1983); (Reese, Cox, & Koop,
1974); (Reese, Cox, & Koop, 1975).
Their work proved the versatility and the theoretical unequivocability of the finite difference
method in dealing with highly nonlinear soil-pile interaction systems with any arbitrary change in soil
formation and pile properties.
Figure 3.16 shows an element of a beam-column. The basic differential equation is derived by
examining this element (Timoshenko & Gere, 1961). The equilibrium for forces in the y direction gives
or,
𝑑𝑄
=𝑞−𝑝
𝑑𝑥
(3.71)
where:
Q = shear force,
𝑝 = 𝐸𝑠 𝑦 (3.72)
𝑑𝑥 𝑑𝑦
𝑀 + 𝑞𝑑𝑥 − (𝑄 + 𝑑𝑄)𝑑𝑥 − (𝑀 + 𝑑𝑀) + 𝑃 𝑑𝑥 = 0
2 𝑑𝑥
(3.73)
𝑑𝑦 𝑑𝑀
𝑄=𝑃 −
𝑑𝑥 𝑑𝑥
(3.74)
where
Standard textbooks on the strength of material give an expression for the moment-curvature
relationship of a beam when shear and axial deformation are neglected.
𝑑2 𝑦
𝐸𝐼 = −𝑀
𝑑𝑥 2
(3.75)
The basic differential equation for a beam-column is obtained by differentiating Equation 3.74
with respect to x, then substituting in Equation 3.72. The basic differential equation for a beam-column
is obtained.
𝑑2 𝑀 𝑑 𝑑𝑦
2
= 𝐸𝑠 𝑦 − 𝑞 + (𝑃 )
𝑑𝑥 𝑑𝑥 𝑑𝑥
(3.76)
𝑑4 𝑦 𝑑2 𝑦
𝐸𝐼 − 𝑃 − 𝐸𝑠 𝑦 + 𝑞 = 0
𝑑𝑥 4 𝑑𝑥 2
(3.77)
The beam column is divided into n discrete elements of length h as shown in Figure 3.16c.
Stations −1, and n+1 through n+3 are imaginary or fictitious where the actual beam column does not
exist. These imaginary stations are necessary for technical reasons to apply the central difference
equations at all stations. The flexural rigidity at the imaginary stations is considered zero and deflections
are set in such a way as to satisfy the boundary conditions at the top and bottom of the pile.
Applying the finite difference approximation to the moment M, Equation 3.76 is now
Substituting Equations 3.78 and 3.79 into the above equation, we obtain the general finite
difference expression for a beam column.
𝑎𝑖 = 𝐸𝐼𝑖−1 (3.82)
𝑒𝑖 = 𝐸𝐼𝑖+1 (3.86)
𝑓𝑖 = 𝑞ℎ4 (3.87)
Assuming that the deflection yi−1 is expressed in terms of the deflections at two subsequent
stations:
The deflection at station i is obtained by substituting Equation 3.88 into Equation 3.81
𝐵𝑖 = 𝐷𝑖 (𝐶𝑖−1 𝐸𝑖 + 𝑑𝑖 ) (3.91)
𝐶𝑖 = 𝐷𝑖 𝑒𝑖 (3.92)
and
𝐸𝑖 = 𝑎𝑖 𝐵𝑖−2 + 𝑏𝑖 (3.94)
for −1 ≤ i ≤ n+1
It is shown in the following section that the pile-top boundary conditions are expressed by finite
difference equations similar to Equation 3.88 at stations −1 and 0. Then we compute the continuity
coefficients Ai, Bi, Ci for stations 1 through n+1. The deflection at the imaginary station yn+2 and yn+3
are assumed to be zero. Therefore, the round-trip path of the recursive solution for a set of banded
simultaneous equations is completed.
The analysis of a grouped pile foundation requires the pile-top reactions of a laterally loaded
pile. The general finite difference equation (Equation 3.81) must be solved in such a way as to satisfy the
displacement boundary condition imposed by the pile-cap displacement at the pile-top. The pile-top
reactions are subsequently computed from the pile deflection at discrete stations along the pile.
The lateral deflection at the pile top coincides with the lateral displacement of the pile cap in
the member coordinate system. However, the slope or rotation of the pile head is dependent on the
kind of pile connection to the pile cap, and not always equal to the rotation of the pile cap. In the
following, the displacement boundary conditions are expressed in terms of the lateral deflection at the
finite difference station near the pile top for pinned, fixed, and elastically restrained connections.
Usually a laterally loaded single pile is given a lateral load and a moment at the pile head. In
order to facilitate the analysis of a single laterally loaded pile, it is convenient to solve for the
displacement of a pile subjected to the force boundary conditions. The finite difference expressions of
the force boundary conditions are added in this section after the displacement boundary conditions.
A pinned pile head cannot carry any moment. The following equation gives the finite difference
expression for this condition (Figure 3.17a):
𝑦0 = 𝑦𝑡 (3.96)
For a pile which is perfectly fixed to an infinitely-stiff bent cap, the slope at the top of the pile is
equal to the rotation angle of the bent cap. Therefore, from the central difference expression of the
slope (Figure 3.17b):
𝑦𝑖 − 𝑦𝑖−1
=𝛼
2ℎ
(3.97)
where α is the rotation angle of pile cap. The other condition is the deflection at the pile top
from before (Equation 3.96).
A pile may have its top elastically restrained by a rotational spring force which is proportional to
the deviation angle q from the structural line (Figure 3.17c). The structural line is fixed to the pile cap
and tangent to the pile before loading, so one of the boundary conditions is written:
𝑀𝑡
=𝐶
𝜃
(3.98)
where
The sum of the deviation angle θ and the slope at the pile head St makes the pile cap rotation
angle α:
𝛼 = 𝜃+ 𝑆𝑡 (3.99)
Applying a central difference expression for the slope, Equations 3.98 and 3.99 give
𝐶ℎ 𝐶ℎ
(𝐸𝐼0 − ) 𝑦1 − 2𝐸𝐼0 ℎ2 𝑦0 + (𝐸𝐼0 + ) 𝑦−1 + 𝑐𝛼ℎ2 = 0
2 2
(3.100)
The lateral deflection at the pile also constitutes a boundary condition (Equation 3.96).
If the boundary conditions are given in terms of the forces of the lateral load and the moment at
the pile top, these forces are simulated by imaginary forces Zt and -Zt at station 0 and at the fictitious
station −1. A couple equal to the applied moment is formed, and a lateral load exists at station 0 of the
discretized model (Figure 3.18).
These transverse forces are taken into the solution by modifying the continuity coefficient given
by Equation 3.90.It is soon observed that the effect of the transverse distributed load q appears only in fi
(Equation 3.87). Therefore, the continuity coefficient A is rewritten for stations −1 and 0 as follows
(Matlock & Haliburton, 1966):
where
Zt = Mt/h
Qt = horizontal load at pile top. The concentrated load Qt is equivalent to the product of
increment h and the distributed load q.
At the pile tip a laterally loaded pile is subjected to neither a lateral load nor a bending moment.
These force boundary conditions may be applied to the pile tip by deriving the finite difference
representations of these conditions. However, the same effect is obtained by providing the additional
fictitious station n+2 and n+3 at the tip of the pile. These two fictitious stations assume no lateral
deflection and no flexural rigidity. The latter method is preferred in the numerical method because it
eliminates the special treatment of the continuity coefficients (Equations 3.90, 3.91, and 3.92) at the pile
tip. Thus, the path of the recursive computation, which starts from the pile top and makes a return path
at station n+1, is streamlined.
Once the lateral deflections at all discrete stations are calculated, the slope S, moment M, shear
force Q, distributed horizontal reaction q, and distributed horizontal spring force p are calculated using
the following finite difference equations.
𝑦𝑖+1 − 𝑦𝑖−1
𝑆𝑖 =
2ℎ
(3.103)
𝑀𝑖+1 − 𝑀𝑖−1
𝑄𝑖 = 𝑆𝑖
2ℎ
(3.105)
𝑝𝑖 = 𝐸𝑠,𝑖 𝑦𝑖
(3.107)
The lateral reaction at the top of the pile is given by q0 h. The lateral reaction should be equal to
the summation of soil reactions along the pile.
𝑞0 ℎ = ∑ 𝐸𝑠,𝑖 𝑦𝑖
𝑖=0
(3.108)
Checking the computations using a computer program shows that the inherent error to the
finite difference approximation always amounts to several percent for the distributed load q0, while the
summation of the soil resistance or spring force coincided well with the shear force Q immediately
below the pile top. Therefore, the lateral pile reaction is calculated by summing the soil resistance along
the pile shaft.
There are basically two analytical methods for calculating the load-versus-settlement curve of an
axially loaded pile. One method takes the theory-of-elasticity approach. The theories suggested by
(D’Appolonia & Romualdi, 1963); (Thurman & D’Appolonia, 1965); (Poulos & Davis, 1968); (Poulos &
Mattes, 1969); (Mattes & Poulos, 1969); (Poulos & Davis, 1980) belong to that method. All of these
theories use the Mindlin’s Equation, which gives a solution for the vertical deformation at any point due
to a downward force in the interior of a semi-infinite, elastic, anisotropic solid. The displacement of the
pile at a certain point is calculated by superimposing the influences of the load transfer (skin friction)
along the pile and the pile-tip resistance on that point. The compatibility of those forces and the
displacement of a pile are obtained by solving a set of simultaneous equations. This method takes the
stress distribution within the soil into consideration; therefore, the elasticity method presents the
possibility of solving for the behavior of a group of closely-spaced piles under axial loadings, (Poulos,
1968), (Poulos & Davis, 1980).
The drawback to the elasticity method lies in the basic assumptions which must be made. The
actual ground condition rarely satisfies the assumption of uniform and isotropic material. Additionally,
in spite of the highly nonlinear stress-strain characteristics of soils, the only soil properties considered in
the elasticity method are Young’s modulus E and Poisson’s ratio υ. The use of only these two constants
in representing soil characteristics is too much of an oversimplification. In actual field conditions, the
parameter υ may be relatively constant, but the parameter E can vary by several orders of magnitude.
The other method that calculates the load-versus-settlement curve for an axially loaded pile is the
finite-difference method. Finite-difference equations are employed to achieve compatibility between
pile displacement and load transfer along a pile, and between displacement and resistance at the tip of
the pile. This method was first used by (Seed & Reese, 1957); other studies have been done by (Coyle &
Reese, Load transfer for axially loaded piles in clay, 1966); (Coyle & Sulaiman, 1967); (Kraft, Focht, &
Amerasinghe, 1981). The finite-difference method assumes the Winkler concept, which states that the
load transfer at a certain pile section and the pile tip resistance are independent of the pile
displacement elsewhere. The close agreement between prediction and loading-test results in clays by
(Coyle & Reese, Load transfer for axially loaded piles in clay, 1966) and the scattering of prediction for
loading tests in sands; (Coyle & Sulaiman, 1967) may possibly be explained by the relative sensitivity of a
soil to changes in patterns of stress. Admitting the deficiency in the displacement-shear-force criteria of
sand, the finite-difference method is still a practical method because it can deal with complex soil layer
compositions where the displacement and shear force are nonlinearly related. Furthermore, the method
can accommodate improvements in soil criteria with no modifications to the basic theory.
Figure 3.19 shows the free body of a pile in equilibrium, where the applied load QT is balanced
by a tip load QB plus side loads R. The mechanism is shown in Figure 3.19b illustrates the deformations
in the pile. The pile has been replaced by an elastic spring. Representing the soil is a set of nonlinear
springs spaced along the pile, with one spring depicting the soil behavior beneath the pile tip. The
ordinate f of the curves is the load transfer, and the abscissa wz is the shaft movement. No load is
transferred from shaft to soil unless there is a downward movement of the shaft. This downward
movement is dependent on the applied load, the position along the pile, the stress-strain characteristics
of the pile material, and the load transfer-movement curves along the pile shaft and at the pile tip. To
solve the problem of the distribution of load along the pile for a given applied load and the
determination of downward movement at any point along the pile, a nonlinear differential equation
must be solved.
The differential equation can be obtained by considering an element from the shaft as shown in Figure
3.20. The unit strain is
𝑑𝑤𝑧 𝑄𝑧
=
𝑑𝑧 𝐸𝐴
(3.109)
where
𝑑𝑤𝑧
𝑄𝑧 = 𝐸𝐴
𝑑𝑧
(3.110)
𝑑𝑄𝑧 𝑑2 𝑤𝑧
= 𝐸𝐴
𝑑𝑧 𝑑𝑧 2
(3.111)
If the load transfer from the pile to the soil at point z, in force per unit of area, is defined as fz, then
𝑑𝑄𝑧 = 𝑓𝑧 𝐶𝑑𝑧
(3.112)
𝑑𝑄𝑧
= 𝑓𝑧 𝐶
𝑑𝑧
(3.113)
𝑑2 𝑤𝑧
𝐸𝐴 = 𝑓𝑧 𝐶
𝑑𝑧 2
(3.114)
The load transfer can be expressed as a function of the pile movement wz as follows:
𝑓𝑠 = 𝛽𝑧 𝑤𝑧
(3.115)
Equation 3.115 is substituted into Equation 3.114 to obtain the desired differential equation
𝑑2 𝑤𝑧
− 𝛽𝑧 𝑤𝑧 = 0
𝑑𝑧 2
(3.116)
where
𝐶
=
𝐸𝐴
(3.117)
If η and β are constants, a closed-form solution can be obtained for Equation 3.117. However,
since β cannot normally be a constant, the closed-form solution is of little importance and will not be
presented.
Referring to Figure 3.20, a convenient solution to the nonlinear differential equation, 3.116, is
obtained by writing the equation in finite-difference form and using numerical techniques. Equation
3.116 becomes
∆𝑤 ∆𝑤
( 𝑧) − ( 𝑧)
∆𝑧 𝑚+1 ∆𝑧 𝑚−1
= 𝛽𝑚 𝑤𝑚
2ℎ
(3.118)
∆𝑤𝑧 𝑄𝑖
( ) =
∆𝑧 𝑖 (𝐸𝐴)𝑖
(3.119)
Substituting the expressions from Equations 3.119 and 3.117 into Equation 3.118, the following
expression is obtained assuming a constant EA:
Equations 3.119 and 3.120 are elementary of course, but are sufficient to give a solution to the
problem of the axially loaded pile.
Assuming that curves showing load transfer as a function of pile movement are available, a
suggested procedure for computing the load-settlement curve and a family of load-distribution curves is
as follows:
1. Assume a slight downward movement of the pile tip and refer to the corresponding load-
transfer curve to obtain the resulting load on the pile tip.
2. Select the number of segments into which the pile is to be divided (some experimenting will
indicate the number required for acceptable accuracy) and consider the behavior of the bottom
segment.
3. Assume a load at the top of the bottom segment and compute the elastic compression in that
segment, using Equation 3.119 written for that location. Employ the appropriate value of pile
stiffness under axial load (EA) for the segment.
4. Use the assumed tip movement and the results of the computation in Step 3 to obtain the
downward movement at the mid height of the bottom segment.
5. Refer to the appropriate curve showing load transfer versus movement to compute the resulting
unit load transfer (βmwm in Equation 3.120).
6. Use Equation 3.120 to compute the load at the top of the bottom segment.
7. Repeat Steps 3 through 6 until the difference between successive computations of movement at
that segment is less than the tolerance that was selected.
8. Compute the pile loads and movements for the other segments until the top of the pile is
reached. This will yield one point on the load-settlement curve and one curve from the family of
load-distribution curves.
9. Select other assumed tip movements and repeat computations to produce the entire load-
settlement curve and the whole family of load-distribution curves.
The axial stiffness EA in a pile segment is assumed to be constant. However, EA can be different
from segment to segment in the case of composite or tapered piles, for example. The computation
scheme described in this document allows a different EA for each segment. Because the convergence is
checked on the basis of each individual segment, there is no convergence problem during computation
with this scheme.
This procedure described has been used successfully in technical literature, and limitations on use of
the method involve the accuracy with which load-transfer curves can be predicted. Thus, one of the
principal aims of further research is the development of additional information on load-transfer curves
for deep foundations.
As noted earlier, most piles are installed in groups and this paper is directed principally at the
analysis of groups. The nonlinear models for isolated piles under axial loading and lateral loading were
presented early because they are employed in the analysis of groups. For a similar reason, the analysis
of an isolated pile subjected to torque is presented here. While the models for single piles under axial
and lateral loading have numerous independent applications, the model for torque has little utility
except in the analysis of a group that rotates about a vertical axis.
The increment in Figure 3.21 represents a free-body element taken from a cylindrical pile
subjected to torque. The soil around the element is assumed to exhibit a resistance to rotation as a
function of the angle of twist. The torque at the top of the increment is T, and T+ dT at the bottom. The
change in torque dT along the length Δx is due to transfer of torque to the soil, as shown in the
following equation:
where
θ = angle of rotation.
The angle of twist over the increment is computed from the standard equation from mechanics
𝑇𝑥 ∆𝑥
𝜃𝑥 =
𝐺𝐽
(3.122)
where
Equations 3.121 and 3.122 may be solved in a number of ways. For example, one could select
values for the angle of twist at the bottom of the pile and then solve the two equations for the
corresponding torques. Even if the stiffness of the soil is nonlinear with the angle of twist, iteration will
not usually be necessary if the length of the increments is chosen to be relatively small.
(Stoll, 1972) pointed out that the torsion-load test on piles is simpler and more economical to
perform than the axial load test. The back-analysis of the torsion response of test piles yields values of
shear modulus for the soil which are in good agreement with those deduced from t-z curves in axial load
tests. (Seed & Reese, 1957) reached a similar conclusion in their research which used the torsion vane to
derive the load-transfer curve for side resistance under axial loading.
4.1 Introduction
As was indicated in Chapter 3, the solution of the problem of a pile under lateral loading
requires that values of the soil resistance p be computed as a function of pile deflection y.
Figure 4.1 shows a definition of p and y that should be useful to the engineer (Reese L. C.,
Behavior of piles and pile groups under lateral load, 1983).
Figure 4.1a is an elevation view of a pipe pile that has been driven without bending. An element
of soil at depth x1 is to be considered. The unit stresses after driving at depth x1 are shown in
Figure 4.1b. As may be seen, there is no unbalanced force on the pile at point x1. If the pile is
caused to deflect an amount y1, the unit stress will increase on the front side of the pile and decrease
on the backside, as shown in
Figure 4.1c. If the unit stresses are integrated, an unbalanced force p will be found which acts in
opposition to the deflection. The units of p will be force per unit length.
It would obviously be desirable if the soil reaction p could be found analytically for any depth
below the ground surface and for any pile deflection. Factors to be considered are pile geometry, soil
properties, and the methods of loading, whether static, cyclic, sustained, or dynamic. Unfortunately, at
the current time the methods of analysis are inadequate for solving the entire problem. However,
approximate solutions can be obtained for the ultimate soil resistance pu that can develop at any depth.
Figure 4.2 shows a model that can be used to develop equations for pu at points near the ground
surface. The surface CDEF represents the side of the pile. If the pile moves laterally, the assumption is
made that sliding occurs along the slant surface ABFE and along the vertical surfaces ADE and BCF. The
sliding is assumed to cause the maximum soil resistance to develop on the surface on which the force Fp
acts. The differentiation of the force Fp with respect to depth yields the ultimate soil resistance pu.
At points some distance below the ground surface, the value of pu is found by assuming that the
soil particles move horizontally; that is, at some depth it is easier for the soil to flow from the front of
the pile to the back of the pile rather than to move upward. Models for the “flow-around” failure can be
shown, but as an example let the value of pu for clay be approximately twice the value one would get for
the bearing stress of a long-strip footing that is resting on the surface of the ground. For the long-strip
footing on saturated clay, the bearing stress at failure is about 5c where c is the undrained shear
strength. In the sections that follow, it can be seen that pu for saturated clay is taken at 9c.
As was indicated, elementary theory can be employed to obtain some numerical values that
help define a p-y curve. For the most part, however, p-y curves are derived from the results of full-scale
experiments. These experiments are described and step-by-step procedures for formulating p-y curves
for several situations are given. Here we employ numerical p-y curves to generate curves giving pile-
head deflection as a function of applied load; these curves are then compared to ones obtained from
experiment. The results have generally been acceptable (Reese L. C., Behavior of piles and pile groups
under lateral load, 1983).
In order to analyze a pile under lateral loading, a family of p-y curves is necessary. Such a set of
curves is shown in Figure 4.3a. The figure shows that the curves vary with depth, with the initial slope
and ultimate resistance usually increasing with depth. The branches of the curves are in the second and
fourth quadrants, showing that p is opposite in sign to y. The curves in Figure 4.3a are at some distance
apart. GROUP uses equations computed point-by-point along the pile for each of the increments
employed in the difference-equation solution.
As will be evident from examination of the equations to be later presented, the “continuum”
effect is not explicitly reflected in the p-y equations. Thus, there has been some criticism of the p-y
method because no interaction expressions showing that the deflection at one point along the pile
influences the deflection at some other point are employed. However, the equations that were
developed for the p-y curves are based directly on field experiments for which the continuum effect was
present. Thus, the e p-y curves that are employed implicitly reflect the interaction that takes place
between the elements in the soil mass.
Figure 4.3b shows a p-y curve that is plotted in the first quadrant for convenience. Two dashed
lines are shown in the figure. The slope of the lines indicates the soil modulus. It can be seen that the
upper bound of the soil modulus is Esi, the slope of the initial portion of the curve. As the deflection of
the pile increases at the depth represented by the p-y curve, the slope of a secant to the curve gets
smaller; thus, the value of the soil modulus decreases with increasing deflection. For large deflections,
the soil modulus can take on a very small value. Thus, Es is a nonlinear function of pile deflection and
depth, and solutions using the p-y method require iteration, so a computer program is necessary.
Four types of loading are recognized and each of these types is discussed in the following
paragraphs. The four types of loading are static, repeated, sustained, and dynamic. The importance of
taking the type of loading into account when analyzing a pile under lateral loading cannot be
overemphasized.
The load tests that are described later in this chapter were usually performed by applying a
lateral load in increments, holding that load for a few minutes, and reading all the instruments that gave
the response of the pile. These data allowed p-y curves to be computed. Then, analytical expressions
were developed from the experimental results and these expressions yield “static” p-y curves. Repeated
loadings were applied as well, and this case will be discussed in the next section.
The static p-y curves can be thought of as backbone curves that are to some extent correlated
with soil properties. Thus, the curves are useful for the purpose of providing some theoretical basis to
the p-y method.
From the standpoint of design, the static p-y curves have application in the following cases:
where loadings are short-term and not repeated (probably not encountered); and for sustained
loadings, as in earth-pressure loadings, where the soil around the pile is not susceptible to consolidation
and creep (over-consolidated clays, soft rocks, and clean sands).
As will be noted later in this chapter, the use of p-y curves for repeated loading, a type of
loading that is frequently encountered in practice, will often yield significant increases in pile deflection
and bending moment. The engineer may wish to make computations with both the static curves and
with the repeated (cyclic) curves so that the influence of the loading on pile response can be seen
clearly.
The full-scale field tests that were performed included cyclic loading, as well as the static loading
described above. In the tests an increment of load was applied, the instruments were read, and then the
load was repeated a number of times. In some instances both a forward and backward load was applied,
while in other cases only a forward load was used. The instruments were read after a given number of
cycles and the cycling was continued until there was no obvious increase in ground line deflection or in
bending moments. Another increment was applied and the procedure was repeated. The final load that
was applied brought the maximum bending moment close to the moment that would cause the steel to
yield plastically.
Four specific sets of recommendations for p-y curves for cyclic loading are described in this
chapter. For three of the sets, the recommendations that are given are for the “lower bound” case. That
is, the data that were used to develop the p-y curves were from cases where the ground line deflection
had ceased with repetitions in loading. In the stiff clay case where there was no free water at the ground
surface, the recommendations for p-y curves are based on the number of cycles of load application,
among other factors.
The presence of free water at the ground surface for clay soils can be significant with regard to
the loss of soil resistance due to cyclic loading (Long & Reese, 1983). After a deflection based on the
“elastic” response of the soil is exceeded, a space develops between the pile and the soil when the load
is released. Free water moves into this space, and on the next load application the water is ejected
bringing soil particles with it. This erosion causes a loss of soil resistance, in addition losses due to cyclic
strains remolding the soil. Here we emphasize judgment in designing piles under lateral load. For
example, if the clay is below a layer of sand, or if provision could be made to supply sand around the
pile, the sand will settle around the pile, possibly restoring the soil resistance that was lost due to cyclic
loading.
There are many instances in which pile-supported structures are subject to cyclic loading. Some of the
cases are: wind load against overhead signs and high-rise buildings, traffic loads on bridge structures,
wave loads against offshore structures, impact loads against docks and dolphin structures, and ice loads
against locks and dams. The nature of the loading must be considered carefully. Factors to be
considered are frequency, magnitude, duration, and direction. The engineer will be required to use a
considerable amount of judgment in the selection of the soil parameters and response curves.
If the soil that is effective in resisting the lateral deflection of a pile is an over-consolidated clay,
the influence of sustained loading would probably be small. The maximum lateral stress from the pile
against the clay would probably be less than the previous lateral stress; thus, the additional deflection
due to consolidation and creep in the clay should be small or negligible.
If the soil that is effective in resisting lateral deflection of a pile is a granular material that is freely-
draining, the creep would be expected to be small in most cases. However, if the pile is subjected to
vibrations, there could be densification of the sand and a considerable amount of additional deflection.
Thus, the judgment of the engineering making the design is important
If the soil that is effective in resisting lateral deflection of a pile is a soft, saturated clay, the
stress from the pile to the soil could cause a considerable amount of additional deflection due to
consolidation and creep. Convergence can be achieved by making an initial solution that employed the
properties of clay to estimate the amount of additional deflection. The p-y curves could then be
modified to reflect the additional deflection, and a second solution obtained with the computer. The
writers know of no appropriate way to solve the three-dimensional, time-dependent problem of the
additional deflection that would occur, so as before, the judgment of the engineer will play an important
role in obtaining an acceptable solution.
Two types of problems involving dynamic loading are frequently encountered in design:
machine foundations and earthquakes. The deflection from the vibratory loading of machine
foundations is usually quite small, and the problem can be solved using the dynamic properties of the
soil. Equations yielding the response of the structure under dynamic loading would be employed instead
of the p-y method described herein.
With regard to earthquakes, a rational solution should proceed from the definition of the free-
field motion of the near-surface soil due to the earthquake. Thus, the p-y method described herein
could not be used directly. In some cases, an approximate solution to the earthquake problem has been
made by applying a horizontal load that is assumed to reflect the effect of the earthquake on the
superstructure. In this case, the p-y method can be used, but solutions would only be approximations at
best.
Four sets of detailed recommendations for p-y curves are given in the following sections. These
recommendations are coded as subroutines to the computer program so it will not be necessary for the
user to solve the equations by hand. However, the equations and procedures are given in detail so that
the engineer can check the computer output if desired.
The chapter also presents a method of dealing with layered soil. That method is coded and will
be actuated automatically if the user indicates that the soil profile is layered.
A section in the chapter presents a small amount of information on the response of soft rock to
lateral loading. Because the information on rock is so meager, the p-y curves were not coded as a
subroutine so that the engineer must use some judgment in preparing p-y curves in a numerical form.
The following procedure is for short-term static loading and is illustrated by Figure 4.4. As noted
earlier in this chapter, the presence of free water at the ground surface has a significant influence on the
behavior of a pile in clay under cyclic loading. If the clay is soft, the assumption can be made that there
is free water, otherwise the clay would have dried and become stiff. However, a question arises about
whether or not to use these recommendations if there is a thin stratum of stiff clay above the soft clay,
and if the water table is at the interface of the soft and stiff clay. The recommendations in this section
can be used in the cyclic loading case with the recognition that the results are likely conservative.
1. Obtain the best possible estimate of the variation of undrained shear strength c and submerged
unit weight γ with depth. Also obtain the value of 𝜖50 , the strain corresponding to one-half the
maximum principal stress difference. If no stress-strain curves are available, typical values of 𝜖50
are given in Table 4.1.
Figure 4.4 Characteristic shapes of the p-y curves for soft clay in the presence of free
water: (a) static loading, (b) cyclic loading (after Matlock, 1970)
2. Compute the ultimate soil resistance per unit length of pile, using the smaller of the values given
by the equations below.
𝛾′ 𝐽
𝑝𝑢 = [3 + 𝑥 + 𝑥] 𝑐𝑏
𝑐 𝑏
(4.1)
𝑝𝑢 = 9𝑐𝑏
(4.2)
where
(Matlock, 1970) states that the value of J was determined experimentally to be 0.5 for a soft clay
and about 0.25 for a medium clay. A value of 0.5 is frequently used for J. The value of pu is
computed at each depth where a p-y curve is desired, based on shear strength at that depth.
3. Compute the deflection, y50, at one-half the ultimate soil resistance from the following
equation:
𝑦50 = 2.5𝜖50 𝑏
(4.3)
4. Points describing the p-y curve are now computed from the following relationship:
𝑝 𝑦 1/3
= 0.5 ( )
𝑝𝑢 𝑦50
(4.4)
The following procedure is for cyclic loading and is illustrated in Figure 4.4b.
1. For values of p less than 0.72 pu, construct the p-y curve in the same manner as for short-term
static loading.
2. Solve Equations 4.1 and 4.2 simultaneously to find the depth, xr, where the transition occurs. If
the unit weight and shear strength are constant in the upper zone, then
6𝑐𝑏
𝑥𝑟 =
𝛾′𝑏+ 𝐽𝑐
(4.5)
If the unit weight and shear strength vary with depth, the value of xr should be computed with
the soil properties at the depth where the p-y curve is desired.
3. If the depth to the p-y curve is greater than or equal to xr, then p is equal to 0.72 pu for all
values of y greater than 3y50.
4. If the depth of the p-y curve is less than xr, then the value of p decreases from 0.72 pu at y=3y50
to the value given by the following expression at y=15y50.
𝑥
𝑝 = 0.72𝑝𝑢 ( )
𝑥𝑟
(4.6)
To determine the various shear strengths of the soil required in constructing p-y curves,
(Matlock, 1970) recommended the following tests in order of preference.
Tests must also be performed to determine the unit weight of the soil.
An example set of p-y curves was computed for a pile in soft clay with a diameter of 24 inches.
The soil profile is shown in Figure 4.5. The submerged unit weight was assumed to 40 lb/ft3. In the
absence of a stress-strain curve for the soil, 𝜖50 was taken as 0.020 for the full depth of the soil profile.
The loading was assumed to be static.
Figure 4.5 Shear strength profile used for example p-y curves for soft clay
The p-y curves were computed for the following depths below the mudline: 5, 10, 20 and 40 feet. The
plotted curves are shown in Figure 4.6.
Figure 4.6 Example p-y curves for soft clay with presence of free water
GROUP has two versions of the soft clay criteria. One version uses a value of J equal to 0.5 by
default. This is the version used by most users. The second version is identical in computations as the
first, but the user may enter the value of J at the top and bottom of the soil layer. GROUP does not
perform error checking on the input value of J.
If the p-y curve with variable J (API soft clay with user-defined J) is selected, the user should
consider the advice by Matlock for selecting the J value discussed on page 4-12.
The net effect of using a J value less than 0.5 is to reduce the strength of the p-y curve. An
example of the effect of J on a p-y curve at a depth of 5 feet for a 36-inch diameter pile in soft clay with c
= 1,000 psf and = 55 pcf is shown in Figure 4.7.
1,200
1,000
800
p, lbs/inch
600
400
J = 0.5
200 J = 0.25
0
0 1 2 3 4 5 6 7 8
y, inches
Figure 4.7 Example p-y curves for soft clay showing effect of J
The p-y formulation for soft clay was described above. The p-y curve for soft clay is based on
Equation 4-4, which does not contain an initial stiffness parameter k. Although the formulation for soft
clay has been used successfully for many years, it may generate some numerical issues for cases with
small lateral deflections. To smooth the convergence in some cases, the modified soft clay model with
an initial stiffness parameter can be used.
For this model, the initial straight-line portion of the p-y curve is defined by:
𝑝 = (𝑘𝑥)𝑦
(4.7)
The user may select an initial stiffness k based on Table 4.2 or from a site-specific lateral load
test. GROUP will use the lower of the p values computed using Equation 4-4 or Equation 4-7 for pile
response as a function of lateral pile displacement.
(Reese, Cox, & Koop, 1975) performed lateral load tests employing steel-pipe piles that were 24
inches in diameter and 50ft long. The piles were driven into stiff clay at a site near Manor, Texas. The
clay had an undrained shear strength rangingfromabout1 ton/ft2 at the ground surface to about 3
ton/ft2 at a depth of 12 feet.
Static Loading
The following procedure is for short-term static loading and is illustrated by Figure 4.8. As can be
seen by studying the p-y curves recommended for cyclic loading, the results of the Manor site showed a
very large loss of soil resistance. The data from the tests have been studied carefully, and the
recommended p-y curves for cyclic loading accurately reflect the behavior of the soil at the site.
Nevertheless, the loss of resistance due to cyclic loading at Manor is much more than has been observed
elsewhere. Therefore, the use of the recommendations in this section for cyclic loading will yield
conservative results for many clays. The work of (Long & Reese, 1983) was unable to show precisely why
the loss of resistance during cyclic loading occurred. One clue was that the clay from Manor was found
to lose volume by slaking when a specimen was placed in fresh water; thus, the clay was quite
susceptible to erosion from the hydraulic action of the free water as the pile was pushed back and forth.
1. Obtain the values of undrained shear strength c, soil submerged unit weight γ’, and pile
diameter b.
3. Compute the ultimate soil resistance per unit length pile using the smaller of the values given
by the equation below:
𝑝𝑐𝑑 = 11𝑐𝑏
(4.9)
𝑝 = (𝑘𝑥)𝑦
(4.10)
Use the appropriate value of ks or kc from Table 4.2 for k. Note that the average shear strength
should be computed from the shear strength of the soil to a depth of 5 pile diameters and
defined as half the total maximum principal stress difference in an unconsolidated undrained
triaxial test.
* The average shear strength should be computed from the shear strength of
the soil to a depth of 5 pile diameters. It should be defined as half the total
maximum principal stress difference in an unconsolidated undrained triaxial
test
Figure 4.8 Characteristic shape of p-y curve for static loading in stiff clay in the presence of
free water
𝑦50 = 𝜖50 𝑏
(4.11)
Use an appropriate value of 𝜖50 from results of laboratory tests or, in the absence of laboratory
test, from Table 4.3.
7. Using the following equation, establish the first parabolic portion of the p-y curve and obtain pc
from Equations 4.8 or 4.9.
𝑦 0.5
𝑝 = 0.5𝑝𝑐 ( )
𝑦50
(4.12)
Equation 4.12 should define the portion of the p-y curve from the point of the intersection with
Equation 4.10 to a point where y is equal to As y50 (see note following Step 10).
Equation 4.13 should define the portion of the p-y curve from the point where y is equal to As
y50 to a point where y is equal to 6As y50 (see note following Step 10).
0.0625
𝑝 = 0.5𝑝𝑐 (6𝐴𝑠 )0.5 − 0.411𝑝𝑐 − ( ) 𝑝𝑐 (𝑦 − 6𝐴𝑠 𝑦50 )
𝑦50
(4.14)
Equation 4.14 should define the portion of the p-y curve from the point where y is equal to 6As
y50 to a point where y is equal to 18As y50 (see note following Step 10).
Equation 4.15 should define the portion of the p-y curve from the point where y is equal to
18As y50 and for all larger values of y (see following note).
Note: The step-by-step procedure is outlined, and Figure 4.8 is drawn, as if there is an
intersection between Equations 4.10 and 4.12. However, there may be no intersection of Equation 4.10
with any of the other equations defining the p-y curve. Equation 4.10 defines the p-y curve until it
intersects with one of the other equations or, if no intersection occurs, equation 4.10 defines the
complete p-y curve.
Cyclic Loading
The following procedure is for cyclic loading and is illustrated in Figure 4.10.
Steps 1, 2, 3, 5, and 6 are the same as for the static case, so we only consider Steps 4, 7, 8, and
9.
4. Choose the appropriate value of Ac from Figure 4.9 for the particular non-dimensional depth.
Compute the following:
𝑦𝑝 = 4.1 𝐴𝑐 𝑦50
(4.17)
𝑦 − 0.45𝑦𝑝
𝑝 = 𝐴𝑐 𝑝𝑐 [1 − | | 0.25]
0.45𝑦𝑝
(4.18)
Equation 4.18 should define the portion of the p-y curve from the point of the intersection with
Equation 4.10 to where y is equal to 0.6yp (see note following Step 9).
Figure 4.10 Characteristic shape of p-y curve for cyclic loading in stiff clay in the presence of
free water
0.085
𝑝 = 0.936𝐴𝑐 𝑝𝑐 − ( ) 𝑝𝑐 (𝑦 − 0.6𝑦𝑝 )
𝑦50
(4.19)
Equation 4.19 should define the portion of the p-y curve from the point where y is equal to
0.6yp to the point where y is equal to 1.8yp (see note following Step 9).
0.102
𝑝 = 0.936𝐴𝑐 𝑝𝑐 − ( ) 𝑝𝑐 𝑦𝑝
𝑦50
(4.20)
Equation 4.20 should define the portion of the p-y curve from the point where y is equal to
1.8yp and for all larger values of y (see following note).
Note: The step-by-step procedure is outlined, and Figure 4.10 is drawn, as if an intersection
between Equations 4.10 and 4.18 exists. There may be no intersection of Equation 4.10 with any of the
other equations defining the p-y curve. If there is no intersection, the equation that gives the smallest
value of p for any value of y should be employed.
An example set of p-y curves was computed for a pile in stiff clay with a diameter of 24 inches.
The soil profile that was used is shown in Figure 4.11. The submerged unit weight of the soil was
assumed to be 50 lb/ft3 for the entire depth.
In the absence of a stress-strain curve, 𝜖50 was taken as 0.005 for the full depth of the soil
profile. The initial slope of the p-y curve was established by assuming a k value of 500 lb/in3. The
loading was assumed to be static.
Figure 4.11 Soil profile used for example p-y curves for stiff clay
The p-y curves were computed for 2, 5, 10, and 40 feet below the mudline. The plotted curves
are shown in Figure 4.12.
A lateral load test was performed at a site in Houston with a drilled shaft foundation36 in. in
diameter. A pipe with a diameter of 10 in. instrumented with electrical-resistance-strain gauges in
intervals along its length was positioned along the axis of the shaft before concrete was placed. The
embedded length of the shaft was 42 feet. The average undrained shear strength of the clay in the
upper 20 ft was approximately 2,200 lb/ft2 . The experiments and their interpretation are discussed in
detail by (Welch & Reese, 1972); (Reese & Welch, 1975).
Static Loading
The following procedure is for short-term static loading and is illustrated in Figure 4.13.
1. Obtain values for undrained shear strength c, soil unit weight γ, and pile diameter b.
Also obtain the values of 𝜖50 from stress-strain curves. If no stress-strain curves are available,
use a value of 𝜖50 from 0.010 to 0.005 as given in Table 4.1, keeping in mind that the larger
values are more conservative.
2. Compute the ultimate soil resistance per unit length of pile, pu, using the smaller of the values
given by Equations 4.1 and 4.2. (In the use of Equation 4.1, the shear strength is taken as the
average shear strength from the ground surface to the depth being considered and J is taken as
0.5. The unit weight of the soil should reflect the position of the water table.)
3. Compute the deflection y50 at one-half the ultimate soil resistance from Equations 4.3.
Figure 4.13 Characteristic shape of p-y curve for static loading in stiff clay with no free water
4. Points describing the p-y curve may be computed from the relationship below:
𝑝 𝑦 0.5
= 0.5 ( )
𝑝𝑢 𝑦50
(4.21)
Cyclic Loading
The following procedure is for cyclic loading and is illustrated in Figure 4.14.
1. Determine the p-y curve for short-term static loading by the procedure previously given.
2. Determine the number of times the lateral load will be applied to the pile.
3. For several values of p/pu obtain the value of C, the parameter describing the effect of
repeated loading on deformation, from a relationship developed by laboratory tests (Welch &
Reese, 1972), or in the absence of tests, from the following equation:
𝑝 4
𝐶 = 9.6 ( )
𝑝𝑢
(4.22)
4. At the value of p corresponding to the values of p/pu selected in Step 3, compute new values
of y for cyclic loading from the following equation:
𝑦𝑐 = 𝑦𝑠 + 𝑦50 𝐶 𝑙𝑜𝑔𝑁
(4.23)
where
y50 = deflection under short-term static load at one-half the ultimate resistance, and
Figure 4.14 Characteristic shape of p-y curve for cyclic loading in stiff clay with no free water
5. The p-y curve defines the soil response after N-cycles of load.
An example set of p-y curves was computed for a pile in stiff clay above the water table with a
diameter of 24 inches. The soil profile that was used is shown in Figure 4.11.Theunit weight of the soil
was assumed to be 125 lb/ft3 for the entire depth. In the absence of a stress-strain curve, 𝜖50 was
taken as 0.005. Equation 4.23 was used to compute values for the parameter C, and assuming 100 cycles
of load application.
The p-y curves were computed for the following depths below the ground line: 2, 5, 10, and 40 feet. The
plotted curves are shown in Figure 4.15.
Figure 4.15 Example curves for stiff clay with no free water, cyclic loading
4.6.5 Modified p-y Formulation for Stiff Clay with no Free Water
The p-y formulation for stiff clay with no free water was described above. The p-y curve for stiff
clay with no free water is based on Equation 4.20 which does not contain an initial stiffness parameter k.
Although the formulation for stiff clay without free water has been used successfully for many years,
there have been cases reported from the Southeastern United States where load tests on full-size piles
have found that the initial slope of the load-deformation response modeled using the original
formulation is too stiff.
The ultimate load-transfer resistance pu used in the p-y formulation is consistent with the theory
of plasticity and has also correlated well with the results of load tests. However, the soil resistance at
small deflections is influenced by factors such as soil moisture content, clay mineralogy, clay structure,
possible desiccation, and pile diameter. (Brown, 2002) has recommended using a field-calibrated k value
to modify the initial portion of the p-y curves if one has the results of lateral load test for calibration of
the initial stiffness k. Judicious use of this modified p-y formulation enables one to obtain improved
predictions with experimental readings that may be used later for design computations.
For this model, the initial straight-line portion of the p-y curve is defined by:
𝑝 = (𝑘𝑥)𝑦
(4.24)
The user may select an initial stiffness k based on Table 4.2 or from a site-specific lateral load
test. GROUP will use the lower of the p values computed using Equation 4.24 or Equation 4.20 for pile
response as a function of lateral pile displacement.
One drawback of the modified p-y formulation for stiff clay with no free water is that p-values
for a p-y curve computed at the ground surface will always be zero. This is not the case for the
unmodified formulation.
Also, this model can be used to smooth the convergence in some cases with small lateral
deflections.
An extensive series of tests were performed at a site on Mustang Island, near Corpus Christi
(Cox, Reese, & Grubbs, 1974). Two steel-pipe piles of 24 inches in diameter each were driven into sand
in a manner simulating the driving of an open-ended pipe, and then were subjected to lateral loading.
The embedded length of the piles was 69 feet. One of the piles was subjected to short-term loading and
the other to repeated loading.
The soil at the site was a uniformly graded, fine sand with an angle of internal friction of 39 degrees. The
submerged unit weight was 66 lb/ft3. The water surface was maintained a few inches above the mudline
throughout the test program.
Figure 4.16 Characteristic shape of a family of p-y curves for static and cyclic loading in sand
The following procedure is for short-term static loading and for cyclic loading (Figure 4.16)
(Reese, Cox, & Koop, 1974).
1. Obtain values for the angle of internal friction , the soil unit weight , and pile diameter b
(Note: use buoyant unit weight for sand below the water-table and total unit weight for sand
above the water table).
𝛼=
2
(4.25)
𝛽 = 45 +
2
(4.26)
𝐾𝑎 = tan2 (45 − )
2
(4.27)
3. Compute the ultimate soil resistance per unit length of pile using the smaller of the values given
by the equation that follows:
4. In making the computation in Step 3, find the depth t where Equations 4.28 and 4.29 intersect.
Above this depth use Equation 4.28. Below this depth use Equation 4.29.
𝑝𝑢 = 𝐴𝑠 𝑝𝑠 𝑜𝑟 𝑝𝑢 = 𝐴𝑐 𝑝𝑠
(4.30)
Use the appropriate value of As or Ac from Figure 4.17 for the particular non-dimensional depth
given static or cyclic loading, respectively. Use the appropriate equation for ps by referring to
the computation in Step 4.
𝑝𝑚 = 𝐵𝑠 𝑝𝑠 𝑜𝑟 𝑝𝑚 = 𝐵𝑐 𝑝𝑠
(4.31)
Use the appropriate value of Bs or Bc from Figure 4.18 for the particular non-dimensional depth
given static or cyclic loading, respectively. Use the appropriate equation for ps . The two
straight-line portions of the p-y curve beyond the point where y is equal to b/60 can now be
established.
𝑝 = (𝑘𝑥)𝑦
(4.32)
Table 4.5 Representative Values of k for Sand Above Water (Static and Cyclic Loading)
𝑝 = 𝐶̅ (𝑦)1/𝑛
(4.33)
𝑝𝑢 − 𝑝𝑚
𝑚=
𝑦𝑢 − 𝑦𝑚
(4.34)
𝑝𝑚
𝑛=
𝑚𝑦𝑚
(4.35)
𝐶̅ = 𝑝𝑚 (𝑦𝑚 )−1/𝑛
(4.36)
𝐶 𝑛/(𝑛−1)
𝑦𝑘 = ( )
𝑘𝑥
(4.37)
(e) Compute appropriate number of points on the parabola by using Equation 4.33.
Note: The step-by-step procedure is outlined and Figure 4.16 is drawn as if an intersection
between the initial straight-line portion of the p-y curve and the parabolic portion of the curve at point k
exists. However, in some instances an intersection with the parabola does not exist. Equation 4.32
defines the p-y curve until there is an intersection with another branch of the p-y curve. If no
intersection occurs, Equation 4.32 defines the complete p-y curve. This completes the development of
the p-y curve for the desired depth. Any number of curves can be developed by repeating the above
steps for each desired depth.
Triaxial compression tests are recommended for obtaining the angle of internal friction of the
sand. The reader should use confining pressures close or equal to those at the depths being considered
in the analysis. Tests must be performed to determine the unit weight of the sand. However, it may be
impossible to obtain undisturbed samples so the angle of internal friction is frequently estimated from
results of some type of in-situ test.
An example set of p-y curves was computed for sand below the water table for a pile with a
diameter of 24 inches. The sand is assumed to have an angle of internal friction of 35◦and a submerged
unit weight of 62.4 lb/ft3. The loading was assumed to be static.
The p-y curves were computed for depths of 5, 10, 20, and 40 feet below the mudline. The
plotted curves are shown in Figure 4.19.
Figure 4.19 Example p-y curves for sand below the water table, cyclic loading
The procedure in the previous section can be used for sand above the water table if appropriate
adjustments are made in the unit weight and the angle of internal friction of sand. Small-scale
experiments were performed by (Parker & Reese, 1971) and recommendations for the p-y curves for dry
sand were developed from those experiments. The results from the Parker and Reese experiments
should be useful as a check for solutions made using results from the test program using full-scale piles.
4.8 Response of Sand Above and Below the Water Table (API RP2A
Recommendation)
4.8.1 Background
The method presented here is described by the American Petroleum Institute (API) in its official manual
on recommended practice (RP2A). The API procedure for p-y curves in sand was justified by a number of
field experiments. Generally, ultimate resistance pu calculations using the criteria of Reese, et al. are the
same as calculations using API criteria. However, the latter uses a more convenient trigonometric
equation for computation. The main difference between those two criteria will be the initial modulus of
subgrade reaction and the shape function of the curves.
The following procedure is for short-term static loading and for cyclic loading as described in API
RP2A (1987).
1. Obtain values for the angle of internal friction , the soil unit weight γ, and the pile diameter b.
2. Compute the ultimate soil resistance at a selected depth x. The ultimate lateral bearing capacity
(ultimate lateral resistance pu) for sand has been found to vary from a value at shallow depths
determined by 4.34 to a value at deep depths determined by 4.35. At a given depth the
equation giving the smallest value of pu should be used as the ultimate bearing capacity.
𝑝𝑢𝑑 = 𝐶3 𝑏𝛾𝑥
(4.39)
where
x = depth,
3. Develop the load-deflection curve based on the ultimate soil resistance pu which is the smallest
value of pu calculated in Step 2. The lateral soil resistance-deflection (p-y) relationships for sand
are nonlinear, and in the absence of more definitive information may be approximated at any
specific depth x by the following expression:
𝑘𝑥
𝑝 = (𝐴)(𝑝𝑢 )(tan ℎ) [ 𝑦]
(𝐴)(𝑝𝑢 )
(4.40)
where
Determine k from Figure 4.21 as function of the angle of internal friction, '
x = depth.
Very Very
Loose Medium Dense Dense
Loose
300
80
60
200
50
k, MN/m3
k, lb/in3
150 40
30
100
50
10
0 0
0 20 40 60 80 100
Relative Density, %
Figure 4.21 Initial modulus of subgrade reaction, k, used for API sand criteria
An example set of p-y curves was computed for sand above the water table using API criteria.
The soil properties are: unit weight (γ’) = 0.07 lbs/in3, and internal-friction angle (') = 35◦. The sand
layer exists from the ground surface to a depth of 40 feet. The pile is of reinforced concrete; the
geometry and properties are: pile length = 300 in., diameter = 36 in., moment in inertia = 82450 in4, and
the modulus of elasticity = 3.6106 lb/in2. The loading is assumed to be static. The p-y curves are
computed for the following depths: 20 in. (1.7 ft.), 40 in. (3.33 ft.), and 100 in. (8.33 ft.). The plotted
curves are shown in Figure 4.22.
Figure 4.22 Example of p-y curves computed based on API sand criteria
A hand calculation for p-y curves at a depth of 20 in. was made to check the computer solution,
as shown in the following:
' = 35◦
b= = 36 in.
𝑝𝑢𝑑 = 𝐶3 𝑏𝛾𝑥
𝑝𝑢𝑑 = (55)(36)(0.07)(20)
𝐴 = 2.55
If y =0.1 in. then k (above water table) = 140 lb/in3 (Figure 4.21) and
𝑘𝑥
𝑝 = (𝐴)(𝑝𝑢 )(tan ℎ) [ 𝑦]
(𝐴)(𝑝𝑢 )
(140)(20)
𝑝 = (2.55)(258)(tan ℎ) [ (0.1)]
(2.55)(258)
𝑝 = 264 𝑙𝑏/𝑖𝑛
If y =1.35 in then
𝑘𝑥
𝑝 = (𝐴)(𝑝𝑢 )(tan ℎ) [ 𝑦]
(𝐴)(𝑝𝑢 )
(140)(20)
𝑝 = (2.55)(258)(tan ℎ) [ (1.35)]
(2.55)(258)
𝑝 = 653 𝑙𝑏/𝑖𝑛
The check by hand computations yielded exact values for the two values of deflection that were
considered.
A survey of the available information on p-y curves for sand was made by (O’Neill & Murchison, 1983),
and some changes were suggested to the procedure given above. Their suggestions were submitted to
the American Petroleum Institute and modifications were adopted by the API review committee.
(Bhushan, Lee, & Grime, 1981) reported on lateral load tests of drilled piers in sand. A procedure for
predicting p-y curves was suggested.
A number of authors have discussed the use of the pressuremeter in obtaining p-y curves. The method
that is proposed is described in some detail by (Baguelin, Jezequel, & Shields, 1978).
4.9.1 Background
With regard to soil conditions, methods exist for soils that can be characterized as either
cohesive or cohesionless (clay or sand, for example). There are currently no generally accepted
recommendations for developing p-y curves for c− soils.
Soil characteristics are limiting for several reasons. For example, in foundation design, where p-y
analysis is most commonly used, soils are often characterized by a value of only c, or only .
Additionally, p-y predictions have been based on major experiments performed in soils that can be
described either by c or . However, there are now numerous occasions when it is desirable, and
perhaps necessary, to more carefully describe the characteristics of the soil.
For example, when piles are employed to improve the stability of a slope, it is necessary to have
predictions for p-y curves for c− soils. The analysis will generally conform to the following procedures:
analyze the slope with no piles present and find a factor of safety against slope failure; select the type of
pile to be used in the slope and find the driving forces above an estimated sliding surface and the
resisting forces below the surface; consider the number of piles in a slope; re-evaluate the factor of
safety; and modify the position of the sliding surface and the forces on the pile to achieve compatibility.
It is well known that most of the currently accepted methods of analyzing slope stability characterize the
soils in terms of both c and for long-term or drained analysis. Therefore, it is inconsistent to assume in
pile analysis that a soil with both cohesion and internal friction is characterized by only c or only .
There are other instances in the design of piles under lateral loading where it is desirable to
have methods for predicting p-y curves for c− soils. The shear strength of unsaturated, cohesive soils is
generally represented by strength components of both c and . However, in many practical cases there
is the likelihood that the deposit might become saturated because of rainfall and a rise of the ground
water table. But there could well be times when the ability to design for dry seasons is critical.
Cemented soils are frequently found in subsurface investigations. It is apparent that cohesion
from cementation will increase soil resistance significantly, especially for soils near the ground surface.
The following procedure is for short-term static loading and for cyclic loading and is illustrated in
Figure 4.23. The suggested procedure closely follows recommendations made earlier for sand.
Conceptually, the ultimate soil resistance (pu) is taken as the passive soil resistance acting on the face of
the pile in the direction of the horizontal movement, plus any sliding resistance on the sides of the piles,
less any active earth pressure force on the rear face of the pile. The force from active earth pressure and
the sliding resistance will generally be small compared to the passive resistance, and will tend to cancel
each other out. (Evans & Duncan, 1982) recommended an approximate equation for the ultimate
resistance of c− soils as:
𝑝 = 𝜎𝑝 𝑏 = 𝐶𝑝 𝜎ℎ 𝑏
(4.41)
where
where
c = cohesion
The modifying factor Cp can be divided into two terms: Cp to modify the frictional term of
Equation 4.41 and Cpc to modify the cohesion term of Equation 4.41. We can then write:
The relatively straightforward derivation of equations for developing p-y curves for c− soil
proceeds by using the concept proposed by (Evans & Duncan, 1982).
𝑝𝑢 = 𝐴̅ 𝑝𝑢 + 𝑝𝑢𝑐
(4.44)
where 𝐴̅ can be found from Figure 4.17. The friction component pu will be the smaller of the
values given by the equations below:
The cohesion component puc will be the smaller of the values given by the equations below:
𝛾′ 𝐽
𝑝𝑢𝑐 = [3 + 𝑥 + 𝑥] 𝑐𝑏
𝑐 𝑏
(4.47)
𝑝𝑢𝑐 = 9𝑐𝑏
(4.48)
Furthermore, it is assumed that the contribution of cohesion due to cementation is lost as the
load-transfer curve transitions from the peak value to the residual value, so the strength of the residual
curve is due to the frictional component only and is
𝑝𝑢 = 𝐴̅ 𝑝𝑢
(4.49)
To develop the p-y curves, the procedures described earlier for sand by (Reese, Cox, & Koop,
1974) will be used because the stress-strain behavior of c− soils are believed to be closer to the stress-
strain curve of cohesionless soil than for cohesive soil. The following procedures are used to develop the
p-y curves.
𝑝𝑢 = 𝐴̅𝑠 𝑝𝑢 or
𝑝𝑢 = 𝐴̅𝑐 𝑝𝑢
(4.50)
Use the appropriate value of 𝐴̅𝑠 or 𝐴̅𝑐 from Figure 4.17 for the particular non-dimensional
depth and loading case(static or cyclic).
𝑝𝑢 = 𝐵𝑠 𝑝𝑢 + 𝑝𝑢𝑐 or
𝑝𝑢 = 𝐵𝑐 𝑝𝑢 + 𝑝𝑢𝑐
(4.51)
Use the appropriate value of Bs or Bc from Figure 4.18 for the particular non-dimensional depth
and loading case. Use the appropriate equation for ps . The two straight-line portions of the p-y
curve beyond the point where y is equal to b/60 can now be established.
𝑝 = (𝑘𝑥)𝑦
(4.52)
The value of k for Equation 4.52 may be found from the following equation and by reference to
Figure 4.24.
𝑘𝑢 = 𝑘𝑐 + 𝑘
(4.53)
For example, if c is equal to 0.2 t/ft2 and is equal to 35 degrees for a layer of c− soil above
the water table, the recommended kc is 350 lb/in+ and k is 80 lb/in3 , yielding a k value of 430 lb/in3.
𝑝 = 𝐶̅ (𝑦)1/𝑛
(4.54)
𝑝𝑢 − 𝑝𝑚
𝑚=
𝑦𝑢 − 𝑦𝑚
(4.55)
𝑝𝑚
𝑛=
|𝑚|𝑦𝑚
(4.56)
𝐶̅ = 𝑝𝑚 (𝑦𝑚 )−1/𝑛
(4.57)
𝐶 𝑛/(𝑛−1)
𝑦𝑘 = ( )
𝑘𝑥
(4.58)
(e) Compute appropriate number of points on the parabola by using Equation 4.54.
Note: The step-by-step procedure is outlined as if an intersection between the initial straight-
line portion of the p-y curve and the parabolic portion of the curve at point k exists. However, in some
instances there may be no intersection with the parabola. Equation 4.52 defines the p-y curve until
there is an intersection with another branch of the p-y curve or if no intersection occurs, Equation 4.52
defines the complete p-y curve. This completes the development of the p-y curve for the desired depth.
Any number of curves can be developed by repeating the above steps for each desired depth.
4.9.3 Discussion
Example p-y curves were computed for a pile with a diameter of 12inches (0.3 meters) in c−
soils. The c value is 0.2 t/ft2 (20 kPa) and a value is 35 degrees. The unit weight of soil is 115 lb/ft3 (18
kN/m3). The p-y curves were computed for depths of39 in. (1 m), 79in. (2 m), and 118 inches (3 m). The
p-y curves computed by using the simplified procedure are shown in Figure 4.25. It can be seen that the
ultimate resistance of the soil is higher when based on the model procedure rather than the simplified
procedure. Both of the p-y curves show an initial peak strength before dropping to a residual strength at
a large deflection, as is expected. Because of a lack of experimental data with which to calibrate the soil
resistance, using the model procedure is currently recommended.
Data from experiments aimed at learning about the response of c− soils is unavailable. These
experiments would have used fully instrumented piles. Further, little information is
availableintheliteratureontheresponseofpilesunderlateralloadingin soils where response is given
principally by deflection of the pile at the point of loading.
The reader will note that the procedure presented above does not reflect the severe loss of soil
resistance characteristic for clays below a free-water surface under cyclic loading. Rather, the
procedures are for a granular material that principally does not reflect such loss of resistance. Therefore,
if a c− soil has a very low value of and a relatively large value of c, the user is advised to ignore the
and to use the recommendations for p-y curves for clay. Additionally, a relatively large factor of
safety is always recommended, and a field program for the testing of prototype piles is certainly in order
for jobs that involve any large number of piles.
4.10.1 Background
The lateral resistance of deep foundations in liquefied sand is often critical to the design of
bridges and buildings. Although reasonable methods have been developed to define p-y curves for non-
liquefied sand, considerable uncertainty remains regarding how much the lateral resistance can be
provided by liquefied sand. In some cases, liquefied sand is assumed to have no lateral resistance during
analysis.
When the sand is liquefied under undrained condition, some suggest that it behaves similar to
soft clay. (Wang & Reese, 1998) have studied the behavior of piles in liquefied soil by treating the
liquefied sand as soft clay. The p-y curves were generated using soft clay criteria and the cohesive
strength (c) is based on the residual strength of liquefied sand. The strain-control parameter (𝜖50 ) of
0.05 is used in their studies.
Because the laboratory method cannot predict the residual shear strength of liquefied soil with
reasonable accuracy. It is necessary to examine some case histories for further research. Recognizing
this, (Seed & Harder, 1990) examined documented cases where major sliding has occurred due to
liquefaction and where some conclusions can be drawn concerning the strength and deformation
resistance of the liquefied soil. Unfortunately, such available cases are rare. However, a small number of
such cases do exist, for which the residual strengths of liquefied sans and silty sand can be determined
with a reasonable degree of accuracy. The residual strength which is about 10% of the effective
overburden stress has been used for liquefied sand.
Although simplified methods have been used based on some engineering judgment, full scale
field tests are needed to develop a full range of p-y curves in liquefied sand. (Rollins, Gerber, Lane, &
Ashford, 2005) have presented their recent studied on lateral resistance of a full-scale pile group in
liquefied sand. Based on their studies, p-y curves in liquefied sand are characterized by a concave-up
load-displacement shape where the slope of the curve increases as displacement increases as shown in
Figure 4.26. This characteristic concave-up load displacement shape appears to result primarily from
dilative behavior during shearing, although gapping effects may also contribute to the observed
response. They also found that p-y curves for liquefied sand stiffen with depth (or initial confining
pressure). With increasing depth, smaller displacements are required to develop significant resistance
and the rate at which resistance develops as a function of displacement also increases.
Following liquefaction, p-y curves in sand become progressively stiffer as excess pore water
pressure dissipates. The p-y curve shape appears to transition from concave-up to concave down as
pore water pressure decreases. An empirically derived equation has been developed by Rollins et al.
(2005) to mathematically describe the observed load-displacement response of fully liquefied sand as a
function of depth. Based on their studies, p−y curves developed using this equation provide reasonable
estimates of the measured response of full-scale piles in liquefied sand.
The expression developed by (Rollins et al. 2005) for p-y curves in liquefied sands at different
depths is shown below is based on their fully-instrumented load tests. Coefficients for these equations
were fit to the test data using a trial and error process in which the errors between the target p-y curves
and those predicted by the equations were minimized. The resulting equations were then compared,
and the equation that produced the most consistent fit was selected
where
A = 3x10−7 (z+1)6.05
B = 2.80 (z+1)0.11
C = 2.85 (z+1)−0.41
P0.3m = soil resistance in kN/m for a reference pile with a diameter of 0.3 meters
y = lateral deflection of pile in millimeters (mm)
z = depth in meters (m), and
A, B, C are functions of the depth in meters
Note that the engineering units of pile diameter is in meters, pile displacement is in millimeters,
depth is in meters, and computed values lateral load transfer are in kilonewtons per meter.
The end of the upward curve is at a displacement of 150 mm only if p0.3 m is less than 15 kN/m at
a y value of 150 mm. If p0.3 m reaches 15 kN/m at a value of y smaller than 150 mm, the upward curve is
ended at that value of y.
If the pile diameter differs from 0.3 m, the value of p is scaled by a diameter modification factor.
The diameter modification factor is discussed below.
(Rollins et al. 2005) studied the diameter effects for different sizes of piles and recommended
using a modification factor for adjusting equation 4.59. The modification factor for pile diameters
between 0.3 and 2.6 meters is
where b is the diameter of the pile or drilled shaft in meters. The p value for the reference diameter of
0.3 meters can be multiplied by Pd to obtain values for p values for piles of varying diameters using the
equation below.
𝑝(𝑦) = 𝑃𝑑 𝑝0.3𝑚
(4.61)
Note that the diameter modification factor has been experimentally validated for pile diameters
ranging from 0.3 to 2.6 meters.
For pile diameters smaller than 0.3 meters, the procedure is to compute p0.3 m for a diameter of
0.3 meters then multiply by the ratio of the pile diameter over 0.3 meters. Thus, for pile diameters less
than 0.3 meters, the diameter modification factor is computed from
𝑏 𝑏
𝑃𝑑 = ( ) ( 3.81 ln(0.3 𝑚) + 5.6) = 1.0129 ( )
0.3 𝑚 0.3 𝑚
(4.62)
For pile diameters greater than 2.6 meters, the value of Pd is constant and equal to 9.24.
Application of equation 4.59 should generally be limited to conditions comparable to those from
which it was derived. These conditions are:
In some cases, the liquefying layer may not be at the surface. In such cases, the depth variable
(z) may be modified to a depth equal to the initial vertical effective stress divided by an effective unit
weight of 10 kN/m3, which is generally representative of the unit weight of the sand at the site.
(Franke & Rollins, 2013) developed the simplified hybrid p-y spring model that combines the
features of the Rollins et al. (2005) model and the residual strength model suggested by (Wang & Reese,
1998) . In this model, two p-y curves are computed. One curve is the curve computed using the Rollins et
al. (2005) method. The second curve is based on the Matlock method for p-y curves in soft clay under
static loading conditions in which the cohesive strength of the soil is based on the (Seed & Harder, 1990)
curves of residual strength of liquefied soils. This model is referred to as the hybrid model for liquefied
sands.
The concept used to formulate the hybrid model is to compute the p-y curves for the dilative
behavior based on the equations developed by Rollins, et al. (2005a) and for the residual strength
behavior based on the soft clay equations for static loading conditions developed by (Matlock, 1970).
The hybrid model uses the lowest p-value computed using either model for a given y-value.
All input values need to be converted to SI units before computation of the hybrid p-y curve. The
sole exception to this is that the correlation from SPT blowcount to residual strength outputs residual
strength in units of psf.
This model combines the Matlock soft clay model for static conditions using residual strengths
correlated to SPT blowcount and the Ashford and Rollins dilative liquefied sand model. The model uses
the lowest p-value computed using either model for a given y-value.
The equations for the dilative model are the same as those presented by Rollins, et al. (2005)
before.
𝑝 = 𝐴 (𝐵𝑦)𝑐
(4.63)
where
A = 3x10−7 (z+1)6.05
B = 2.80 (z+1)0.11
C = 2.85 (z+1)−0.41
𝑝0.3𝑚 = min [𝐴 (150𝐵)𝑐 , , 15 𝑘𝑁/𝑚]
𝑃𝑑 = 3.81 ln(𝑏) + 5.6, for diameters between 0.3 m and 2.6 m.
𝑃𝑑 = 𝑏/0.3 𝑚, for diameters less than 0.3 m.
𝑃𝑢 = 𝑃𝑑 𝑝0.3𝑚
𝑝(𝑦) = 𝑃𝑑 𝐴 (𝐵𝑦)𝑐 ≤ 𝑃𝑢 , for y values up to 150 mm
A, B, and C are dimensionless functions of the depth of the curve z in meters
p0.3m is the maximum dilative resistance for a 0.3 m diameter pile
Pd is a dimensionless factor to adjust for pile diameter
pu is the peak lateral load intensity, and
50 is the value required to compute the Matlock soft clay curve.
The parameter p0.3m is the lesser of the computed dilative resistance at a displacement of 150 mm or 15
kN/m.
The equations for the p-y curves using residual strengths are those developed by (Matlock,
1970) for soft clays under static loading conditions. The equations and p-y curve for soft clay are
presented in Figure 4.4. The peak load-transfer capacity is computed using
𝛾′𝑎𝑣𝑔 𝐽
𝑝𝑢 = 𝑚𝑖𝑛 {[3 + 𝑥 + 𝑥] 𝑐𝑏, 9𝑐𝑏}
𝑐 𝑏
The pile diameter is accounted for by the factor y50. This factor is computed using
𝑦50 = 2.5𝜖50 𝑏
𝑝 𝑦 1/3
= 0.5 ( ) for y ≤ 8 𝑦50
𝑝𝑢 𝑦50
The lateral load-transfer is constant and equal to pu for lateral pile deflections greater than 8y50.
The engineering units used in the three equations above are consistent units of force and
length. The units may be lbs and ft, lbs and inches, or kN and meters. See the hybrid model equations
for the correlation of SPT blowcounts to 50.
4.10.3.2 Correlation for 33% residual strength as function of SPT (N1)60-cs -value
The residual shear strength is taken as the 33rd percentile of the residual strength correlation
developed by (Seed & Harder, 1990). This correlation is illustrated in Figure 4.27. In the original
recommendations for the hybrid p-y curve, the residual undrained shear strength for soils less than 5
blows per foot was taken to be zero.
2,000
`
800 20
Recommended
400 residual strengths
10
for use in the
hybrid p-y model
0 0
0 4 8 12 16 20 24
Equivalent Clean Sand SPT Blowcount, (N1)60-cs
Figure 4.27 Method for Computing Residual Shear Strength of Liquefied Soil for Use in
Hybrid p-y Model
The 33rd percentile for residual strength as a function of SPT blowcount is shown as the dashed
gray line the figure above. The 33rd percentile of residual strength is determined by:
Note that the lower and upper limits on 50 are 0.0005 and 0.05. The correlation of 50 with SPT
blowcount is illustrated in Figure 4.28.
0.06
0.05
0.04
50 0.03
0.02
0.01
0
0 5 10 15 20 25 30
(N1)60-cs
There are four possible patterns for the over lapping of the dilative and residual p-y curves.
These patterns are illustrated in Figure 4.29, with the residual curves shown in red and the dilative curve
shown in blue. There are two patterns where there is one intersection point between the two curves.
These two patterns are indicated by points 1 and 2 in the graph. There is one pattern in which the curves
intersect at two points, indicated by points 3 and 4, and one pattern in which the dilative curve
underlies the entirety of the residual curve. GROUP computes the coordinates of the intersection points
and includes them in the output report for the generated p-y curves.
No intersections
4 Intersections at 2 and 3
3
Intersection at 2
2
Intersection at 1
1
Figure 4.29 Possible Intersection Patterns of Residual and Dilative p-y Curves in Hybrid p-y
Model
14
12
Pile Diam., b = 0.3 m
Depth, z = 4.0 m
Eff. Unit Wt., ’ = 10 kN/m3
10
(N1)60-cs = 7 bpf
Sur = 128.9 psf
8
p, kN/m
Residual p
4
Dilative p
Hybrid p
2
0
0 50 100 150 200 250 300
y, mm
16
14
Pile Diam., b = 0.3 m
Depth, z = 4.0 m
12 Eff. Unit Wt., ’ = 10 kN/m3
(N1)60-cs = 6 bpf
Sur = 104.9 psf
10
p, kN/m
8
y1i
6
Residual p
4 Dilative p
Hybrid p
2
0
0 100 200 300 400
y, mm
Figure 4.31 Example of Curves with One Intersection of Dilative and Residual Curves
10
8
Pile Diam., b = 0.3 m
7 Depth, z = 0 m
Eff. Unit Wt., = 10 kN/m3
6 (N1)60-cs = bpf
Sur = 99.8 psf
p, kN/m
4
Residual p
3 Dilative p
Hybrid p
2
0
0 50 100 150 200 250
y, mm
Figure 4.32 Example of Curve with One Intersection of Dilative Curve and Residual Plateau
18
16
Pile Diam., b = 0.3 m
14 Depth, z = 4.0 m
Eff. Unit Wt., ’ = 10 kN/m3
(N1)60-cs = 7 bpf
12
Sur = 128.9 psf
p, kN/m
10
6
Residual p
Dilative p
4
Hybrid p
2
0
0 50 100 150 200 250 300
y, mm
When liquefaction takes place in sandy deposits during earthquakes, the ground is often known
to undergo a large amount of permanent deformation as a result of the lateral flow of liquefied soils,
even though the ground is nearly flat. If the free-field soil movement is greater than the pile
displacement, then soils will apply extra driving pressure on the already loaded piles. The actual force on
the pile induced by the soil movement is dependent on the relative displacement between the pile and
soil. If the liquefied soil causes the upper layer to become unstable and move laterally, a model
recommended by (Wang & Reese, Design of pile foundations in liquefied soils, 1998)may be used to
solve for the behavior of the pile.
4.11.1 Background
A procedure was formulated by (Johnson, et al. 2006) for loess soil that includes degradation of
the p-y curves by load cycling.
The soil strength parameter used in the model is the cone tip resistance (qc) from cone
penetration (CPT) testing. The p-y curve for lateral resistance with displacement is modeled as a
hyperbolic relationship. Recommendations are presented for selection of the needed model
parameters, as well as a discussion of their effect.
Shafts were tested in pairs to provide reaction for each other. Both shafts used in the load test
were fully instrumented. Load tests were performed on one pair of 30-inch diameter loaded statically,
one pair of 42-inch diameter test shafts loaded statically, and one pair of 30-inch diameter test shafts
loaded cyclically. Lateral loads were maintained at constant levels for load increments without
inclinometer readings, and the hydraulic pressure supply to the hydraulic rams was locked off during
load increments with inclinometer readings to eliminate creep of the deflected pile shape with depth
while inclinometer readings were made.
13 and 15 load increments were used to load the 30-inch and 42 inch diameters pairs of static
test piles, respectively, while both sets of static test piles were unloaded in four decrements. Six sets of
inclinometer readings were performed for each static test pile, four of which occurred at load
increments. Load increments and decrements for the static test shafts were sustained for approximately
5 minutes, with the exception of the load increments with inclinometer readings where the duration
was approximately 20 minutes (this allowed for approximately 10 minutes for inclinometer
measurements for each of the two test shafts in the pair). Lateral loads were applied to the 30-inch and
42-inch diameter static test shafts in approximately 10-kip and 15-kip increments, respectively.
There were four load increments (noted as “A” through “D”) on the 30-inch diameter cyclic test
shafts, with ten load cycles (N = 1 through 10) performed per load increment. The lateral load for each
load cycle were sustained for only a few seconds with the exception of load cycles 1 and 10 which were
sustained for approximately 15 to 20 minutes to allow time for the inclinometer readings to be
performed. For load cycles 2 through 9, the duration for each load cycle was approximately 1 minute, 2
minutes, 3.5 minutes, and 6.5 minutes for load increments A though D, respectively, as a greater time
was required to reach the larger loads. The load was reversed after each load cycle to return the top of
pile to approximately the same location.
A back-fit model of the pile behavior using the available soil strength data obtained (from both
in-situ and laboratory tests) to the measured pile performance led to the conclusion that the CPT testing
provided the best correlation. Furthermore, CPT testing can be easily performed in the loess soils being
modeled and has become readily widely available.
Three cone penetration tests were performed by the Kansas Department of Transportation at
the test site location. A preliminary cone penetration test was performed in the general vicinity of the
test shafts (designated as CPT-1). Two additional cone penetration tests were performed subsequent to
the lateral load testing. A cone penetration test was performed between the 42-inch diameter static test
shafts (Shafts 1 and 2) shortly after on the same day the lateral load test was performed on these shafts.
A cone penetration test was performed between the 30-inch diameter static test shafts (Shafts 3 and 4)
two days after the completion of the load test performed on these shafts. The locations of the cone
penetration tests were a few feet from the test shafts. Given the nature of the soil conditions and the
absence of a ground water table, it is reasonable to assume that the cone penetration tests were
unaffected by any pore water pressure effects that may have been induced by the load testing.
An idealized profile of cone tip resistance with depth interpreted as an average from the cone
penetration tests performed between the static test shafts is shown in Figure 4.34. This profile is
considered representative of the subsurface conditions for all the test shaft locations. Note that it is
most useful to break the idealized soil profile into layers wherein the cone tip resistance is either
constant with depth or linearly varies with depth as these two conditions are easily accommodated by
most lateral pile analyses software.
The cone tip resistance is reduced by 50% at the soil surface, and allowed to increase linearly
with depth to the full value at a depth of two pile diameters, as shown in Figure 4.34. This is done to
account for the passive wedge failure mechanism exhibited at the ground surface that reduces the
lateral resistance of the soil between the ground surface and a lower depth (assumed at two shaft
diameters). Below a depth of two shaft diameters, the lateral resistance is considered as a flow around
bearing failure mechanism.
The idealized cone tip resistance values were correlated with depth with the ultimate lateral soil
resistance (pu0) at corresponding depths.
10
Depth Below Grade (ft)
15
20
40
0 20 40 60 80 100 120 140 160 180 200
qc (ksf)
Figure 4.34 Idealized Tip Resistance Profile from CPT Testing Used for Analyses.
Procedures are provided to produce a p-y curve for loess, shown generically in Figure 4.35. The
ultimate soil resistance (pu0) that can be provided by the soil is correlated to the cone tip resistance at
any given elevation. Note that to account for the passive wedge failure mechanism exhibited at the
ground surface, the cone tip resistance is reduced by 50% at the soil surface and allowed to return to
the full value at a depth equal to two pile diameters. The initial modulus of the p-y curve, Ei, is
determined from the ultimate lateral soil reaction expressed on a per unit length of pile basis, pu, for the
specified pile diameter, and specified reference displacement, yref. A hyperbolic relationship is used to
compute the secant modulus of the p-y curve, Es, at any given pile displacement, y. The lateral soil
reaction per unit pile length, p, for any given pile displacement is determined by the secant modulus at
that displacement. Provisions for the degradation of the p-y curve as a function of the number of cycles
loading, N, are incorporated into the relationship for ultimate soil reaction.
The model is of a p-y curve that is smooth and continuous. This model is similar to the lateral
behavior of pile in loess soil measured in load tests.
The ultimate unit lateral soil resistance, pu0, is computed from the cone tip resistance multiplied
by the cone bearing capacity factor, NCPT using
𝑝𝑢0 = 𝑁𝐶𝑃𝑇 𝑞𝑐
(4.65)
pu
Ei
Es
y
yref
Figure 4.35 Generic p-y curve for Drilled Shafts in Loess Soils
where NCPT is dimensionless, and pu0 and qc are in consistent units of (force/length2)
The value of NCPT was determined from a best fit to the load test data. It is believed that NCPT is
relatively insensitive to soil type as this is a geotechnical property determined by in-situ testing. The
value of NCPT derived from the load test data is
𝑁𝐶𝑃𝑇 = 0.409
(4.66)
The ultimate lateral soil reaction, pu, is computed by multiplying the ultimate unit lateral soil
resistance by the pile diameter, b, and dividing by an adjustment term to account for cyclic loading. The
adjustment term for cyclic loading takes into account the number of cycles of loading, N, and a
dimensionless constant, CN.
𝑝𝑢0 𝑏
𝑝𝑢 =
1 + 𝐶𝑁 log |𝑁|
(4.67)
where:
CN is a dimensionless constant,
pu is in units of (force/length).
CN was determined from a best fit of cyclic degradation for two 30-inch diameter test shafts
subjected to cyclic loading. CN is
𝐶𝑁 = 0.24
(4.68)
The cyclic degradation term (the denominator of equation 4.67 equals 1 for N = 1 (initial cycle,
or static load) and equals 1.24 for N = 10. The value of CN has a direct effect on the amount of cyclic
degradation to the p-y curve (i.e., a greater value of CN will allow greater degradation of the p-y curve,
resulting in a smaller pu). Note that the degradation of the ultimate soil resistance per unit length of
shaft parameter will also have the desired degradation effect built into the computation of the p-y
modulus values.
A parameter is needed to define the rate at which the strength develops towards its ultimate
value (pu0). The reference displacement, yref, is defined as the displacement at which the tangent to the
p-y curve at zero displacement intersects the ultimate soil resistance asymptote (pu), as shown in Figure
4.35. The best fit to the load test data was obtained with the following value for reference displacement.
Note that the suggested value for the reference displacement provided the best fit to the piles
tested at a single test site in Kansas for a particular loess formation. Unlike the ultimate unit lateral
resistance (pu0), it is believed that the rate at which the strength is mobilized may be sensitive to soil
type. Thus, re-evaluation of the reference displacement parameter is recommended when performing
lateral analyses for piles in different soil conditions because this parameter is likely to have a substantial
effect on the resulting pile deflections. The effect of the reference displacement is proportional to pile
performance that is a larger value of yref will allow for larger pile head displacements at a given lateral
load.
The initial modulus, Ei, is defined as the ratio of the ultimate lateral resistance expressed on a
per unit length of pile basis over the reference displacement.
𝑃𝑢
𝐸𝑖 =
𝑦𝑟𝑒𝑓
(4.70)
A secant modulus, Es, is determined for any given displacement, y, by the following hyperbolic
relationship of the initial modulus expressed on a per unit length of pile basis and a hyperbolic term ( y h
) which is in turn a function of the given displacement (y), the reference displacement (yref), and a
dimensionless correlation constant (a).
𝐸𝑖
𝐸𝑠 =
1 + 𝑦′ℎ
(4.71)
𝑦
𝑦 −( )
𝑦′ℎ = ( ) [1 + 𝑎 𝑒 𝑦𝑟𝑒𝑓 ]
𝑦𝑟𝑒𝑓
(4.72)
𝑎 = 0.10
(4.73)
The constant a was found from a best fit to the load test data. Note that the constant a
primarily affects the secant modulus at small displacements (say within approximately 1 inch or 25 mm),
and is inversely proportional to the stiffness response of the p-y curve (i.e., a larger value of a will
reduce the mobilization of soil resistance with displacement). Combining the two equations above, one
obtains
𝑦
𝑦 −( )
𝑦′ℎ = ( ) [1 + 0.1 𝑒 𝑦𝑟𝑒𝑓 ]
𝑦𝑟𝑒𝑓
(4.74)
The modulus ratio (secant modulus over initial modulus, Es/Ei) versus displacement used for p-y
curves in loess is shown in Figure 4.36. Note that the modulus ratio is only a function of the hyperbolic
parameters of the constant (a) and the reference displacement (yref), thus the curve presented is valid
for all pile diameters and cone tip bearing values tested.
1.0
0.9
a = 0.1
0.8
0.7
0.6
Es
0.5
Ei
0.4
0.3
0.2
0.1
0
0.001 0.01 0.1 1.0 10 100
y
yref
Both the initial modulus and the secant modulus are proportional related to the pile diameter
because the ultimate soil resistance is proportional to a given pile size, as was shown in Equation 3-
Error! Bookmark not defined.. It follows that the lateral response will increase in proportion to the pile
diameter.
For a given pile displacement, the lateral soil resistance per unit length of pile is a product of the
pile displacement and the corresponding secant modulus at that displacement.
𝑝 = 𝐸𝑠 𝑦
(4.75)
where:
Several p-y curves obtained from the model described above is presented in Figure 4.37 for the
30-inch diameter shafts, and Figure 4.38 for the 42-inch diameters shafts. Note that there are three sets
of curves presented for each shaft diameter which correspond to the cone tip resistance values of 11
ksf, 22 ksf, and 100 ksf (as was shown in Figure 3-Error! Bookmark not defined.). These p-y curves were
used in the analyses presented later.
9,000
8,000
7,000
6,000
11 ksf
p, lb/in. .
5,000 22 ksf
100 ksf
4,000
3,000
2,000
1,000
0
0 1 2 3 4 5 6 7
y , inches
14,000
12,000
10,000
11 ksf
p, lb/in. .
8,000
22 ksf
100 ksf
6,000
4,000
2,000
0
0 1 2 3 4 5 6 7
y , inches
Figure 4.38 p-y Curves and Secant Modulus for the 42-inch Diameter Shafts
The static p-y curves shown in Figure 4.37 and Figure 4.38 were degraded with load cycle
number (N) for use in the cyclic load analyses. Figure 4.39 presents the cyclic p-y curve generated for the
analyses of the 30-inch diameter shafts at the cone tip resistance value of 22 ksf.
2,000
1,800
1,600
1,400
N= 1
p, lb/in. .
1,200
N= 5
1,000 N = 10
800
600
400
200
0
0 1 2 3 4 5 6 7
y , inches
1. Develop an idealized profile of cone tip resistance with depth that is representative of the local soil
conditions. It is most useful to subdivide the soil profile into layers where the cone tip resistance is
either constant with depth or varies linearly with depth.
2. Reduce the cone tip resistance by 50% at the soil surface, and allowed the value to return to the full
measured value at a depth equal to two pile diameters. Linear interpolation may be used between
the surface and the depth of two pile diameters.
3. For each soil layer, compute the ultimate soil resistance from the cone tip resistance in accordance
with equation 4.65 for both the top and the bottom of each layer.
4. Multiply the ultimate soil resistance by the pile diameter to obtain the ultimate soil reaction per unit
length of shaft (pu). For cyclic analyses, pu may be degraded for a given cycle of loading (N) in
accordance with equation 4.67.
5. Select a reference displacement (yref) that will represent the rate at which the resistance will
develop.
6. Determine the initial modulus (Ei) in accordance with equation 4.70.
7. Select a number of lateral pile displacements (y) for which a representative p-y curve is to be
generated.
8. Determine the secant modulus (Es) for each of the displacements selected in Step 7 in accordance
with equations 4.71 and 4.72.
9. Determine the soil resistance per unit length of pile (p) for each of the displacements selected in
Step 7 in accordance with equation 4.75.
The p-y curve for static loading was based on best fits of data from full-scale load tests on 30-
inch and 42-inch diameter shafts installed in a loess with average cone tip resistance values ranging from
20 to 105 ksf (960 to 5,000 kPa).
Caution is advised when extrapolating the static model formulation for shaft diameters or soil
types and/or strengths outside these limits. In addition, the formulation for the cyclic degradation
model parameters are based on load tests with only ten cycles of loading (N = 1 to 10) obtained at four
different load increments on an additional two 30-inch diameter shafts. Caution is thus also warranted
when extrapolating the cyclic model to predict results beyond 10 cycles of load (N > 10), particularly as
the magnitude of loading increases.
It is hardly surprising that little information is available on the behavior of piles that have been
installed in rock since other types of foundations are normally used. However, a study has been done on
the behavior of an instrumented drilled shaft installed in vuggy limestone in the Florida Keys (Reese &
Nyman, 1978). The test was performed to gather information before designing foundations for highway
bridges in the Florida Keys.
Difficulty was encountered in obtaining properties of the intact rock. Cores broke during
excavation, and penetrometer tests were misleading (because of the vugs) or could not be run. It was
possible to test two cores from the site. The small discontinuities on the outside surface of the
specimens were covered with a thin layer of gypsum cement in an effort to minimize stress
concentrations. The ends of the specimens were cut with a rock saw and lapped flat and parallel. The
specimens were 5.88 inches in diameter with heights of 11.88 in. for Specimen 1 and 10.44 in. for
Specimen 2. The undrained shear strength of the specimens were taken as one-half the unconfined
compressive strength and were 17.4 and 13.6 ton/ft2 for Specimens 1 and 2, respectively.
The rock at the site was also investigated by in-situ-grout-plug tests under the direction of
(Schmertman, 1977). A 5.5-in.diameter hole was drilled into the limestone, a high strength steel bar was
placed to the bottom of the hole, and a grout plug was cast over the lower end of the bar. The bar was
pulled until failure occurred. Then the grout was examined and it was seen that failure occurred at the
interface of the grout and limestone. Tests were performed at three borings and the results are
presented in Table 4.6. The average of the eight tests was 16.3 ton/ft2. However, the rock was stronger
in the zone where the deflections of the drilled shaft were most significant and a shear strength of 18
ton/ft2 was selected for correlation.
The drilled shaft was 48 inches in diameter and penetrated 43.7 ft into the limestone. The
overburden of fill was 14 ft thick and was cased. The load was applied about 11.5 ft above the
limestone. A maximum horizontal load of75 tons was applied to the drilled shaft. The maximum
deflection at the point of load application was 0.71 in. and at the top of the rock (bottom of casing) it
was 0.0213 inches. While the curve of load versus deflection was nonlinear, there was no indication of
rock failure.
A single p-y curve, shown in Figure 4.40, was proposed for the design of piles under lateral
loading in the Florida Keys. There is insufficient data to create a family of curves that would reflect any
increased resistance with depth. Cyclic loading caused no measurable decrease in resistance of the rock.
As shown in the figure, load tests are recommended if deflection of the rock (and pile) are greater than
0.0004b, and brittle fracture is assumed if the lateral stress (force per unit length) against the rock
becomes greater than the diameter times the shear strength Su of the rock.
The p-y curve shown in Figure 4.40 should be employed with considerable caution because of
the limited amount of experimental data available and the great variability in rock. The behavior of rock
at a site could be very well determined by joints, cracks, and the secondary structure of rock, so that the
strength of intact specimens is not representative.
Figure 4.40 Recommended p-y curve for design of drilled shafts in vuggy limestone
The use of deep foundations in rock is frequently required for support of transmission towers
and other structures that sustain lateral loads of significant magnitude. Because the rock must be drilled
in order to make the installation, drilled shafts (called caissons or bored piles) are frequently used.
However, a steel pile could be grouted into the drilled hole. In any case, the designer must use
appropriate mechanics to compute the ultimate bending moment and the variable bending stiffness EI.
Experimental results show conclusively that the EI must be reduced as the bending moment increases in
order to achieve a correct result (Reese L. C., 1997).
In some applications, the axial load is negligible, so the penetration is controlled by the lateral
load. The designer will wish to initiate computations with a relatively large penetration of the pile into
the rock. After finding a suitable geometric section, the factored loads are employed and computer runs
are made with penetration being gradually reduced. The ground-line deflection is plotted as a function
of penetration, and a penetration that provides adequate security against a sizable deflection of the
bottom of the pile is selected.
Concepts presented in the following section form the basis of computing the response of piles in
rock. The background for designing piles in rock is given and then two sets of criteria are presented, one
for strong rock and another for weak rock. Much of the presentation follows Reese (1997) and the
reader can refer there for more detail.
The secondary structure of rock is an overriding feature with respect to its response to lateral
loading. Thus, a subsurface investigation should be undertaken prior to making any design. The Rock
Quality Designation (RQD) and the compressive strength of intact specimens should be taken with
appropriate tools. If possible, enough data should be taken to allow the computation of the Rock Mass
Rating (RMR). Sometimes, the RQD is so low that no specimens can be obtained for compressive tests.
The performance of pressuremeter tests in such instances is needed.
If investigation shows that there are soil-filled joints or cracks in the rock, the procedures
suggested here in should not be used and full-scale testing at the site is instead recommended.
Furthermore, full-scale testing may be economical if a large number of piles are to be installed at a
particular site. Field testing will add to the data bank and lead to improvements in the recommendations
shown below, which are preliminary because of the meager amount of experimental data that is
available.
In most designs, the deflection of the drilled shaft (or other kind of pile) will be so small that the
ultimate resistance pur of the rock is not reached. However, the ultimate resistance of the rock should
be predicted for computation of the lateral loading that would cause pile failure. Contrary to p-y curves
for soil, where unit weight is a significant parameter, the unit weight of rock is neglected in developing
the prediction equations that follow. While a pile may only move laterally a small amount under working
loads, the prediction of the early portion of the p-y curve is still needed because small deflections may
be critical in some designs.
Most specimens of intact rock are brittle and will develop shear planes under low amounts of
strain. This fact leads to an important concept about intact rock. The rock is assumed to fracture and lose
strength under small values of pile deflection. If the RQD of a stratum of rock has a low value, the rock is
assumed to be fractured and consequently will deflect without significant loss of strength. The above
concept leads to the recommendation of two sets of criteria for rock, one for strong rock and the other
for weak rock. For our purposes, strong rock is assumed to have a compressive strength of 6.9 MPa
(1,000 psi) or above.
The methods of predicting the response of rock are strongly based on a limited number of
experiments, and on correlations that have been presented in technical literature. Some of the
correlations are imprecise; for example, if the engineer enters the figure for correlation between
stiffness and strength with a value of stiffness from the pressuremeter, the resulting strength can vary
by an order of magnitude, depending on the curve that is selected.
Imprecision and lack of experimental data contribute to the need for judgment and caution
when analyzing piles in rock. Full-scale testing in the field is needed when justified by a particular design
and when the owner of the project would make a contribution to the technical literature and the
profession.
Islamorada
San Francisco
Two borings were made into the rock and sampling was done using a NWD4 core barrel in a
cased hole with a diameter of 102 mm (4 in.). A 98-mm (3.88-in.) tricone roller bit was used to advance
the casing and clean the bore hole. The sandstone was medium to finely grained with grain sizes from
0.1 to 0.5 mm (0.004 to0.02in.). It was also well sorted and thinly bedded withthicknessof25to75 mm (1
to 3in.). Recovery wasgenerally100%. The reported values of RQD ranged from zero to 80, with an
average of 45. The sandstone was described by Speer (1992) as very intensely to moderately fractured
with bedding joints, joints, and fracture zones.
Pressuremeter tests yielded scattered results. The plotted values for the rock moduli are shown
in Figure 4.41. The dashed lines show the averages that were used for analysis. Figure 4.42 was
employed in developing the correlation between the initial stiffness and the compressive strength. The
obtained values are shown in Table 4.7.
Two piles 2.25 m (7.38 ft) in diameter, with penetrations of 12.5 m (41 ft) and 13.8 m (45 ft),
were simultaneously tested. Lateral loading was accomplished using hydraulic rams acting on high-
strength steel bars that were passed through tubes that were transverse and perpendicular to the axes
of the piles. Load was measured by load cells and deflection was measured by transducers. The slope
and deflection of the pile tops were obtained from slope indicator readings.
The load was applied in increments of 1.41 m (4.6 ft) above the ground line for Pile A and 1.24 m
(4.1 ft) for Pile B. The pile-head deflection was measured at slightly different points above the rock line,
but the results were adjusted to yield equivalent values for each of the piles. Other details about the
loading-test program are shown in the case studies that follow.
Note: The rock below 8.8m is in the range of strong rock, but
the rock above that depth will control the behavior.
Figure 4.41 Initial Moduli of Rock from Pressuremeter, San Francisco Test (After
Reese,1997)
Figure 4.42 Engineering properties for intact rock, after (Horvath & Kenney, 1979); (Peck,
1976); and (Deere, 1968)
4.12.4 Interim Recommendations for Computing p-y Curves for Strong Rock
The p-y curve recommended for strong rock, with compressive strength qur of intact specimens
larger than 6.9 MPa (1,000 psi), is shown in Figure 4.40.If the rock increases in strength with depth, the
strength at the top of the stratum will normally control. Cyclic loading is assumed to cause no loss of
resistance.
As shown in the Figure 4.40, load tests are recommended if deflection of the rock (and pile) is
greater than 0.0004b and brittle fracture is assumed if the lateral stress (force per unit length) against
the rock becomes greater than half the diameter times the compressive strength of the rock.
The p-y curve shown in Figure 4.40 should be employed with considerable caution because of
the limited amount of experimental data available and the great variability in rock. The behavior of rock
at a site could be very well determined by joints, cracks, and the secondary structure of rock, so that the
strength of intact specimens is not representative.
4.12.5 Interim Recommendations for Computing p-y Curves for Weak Rock
The p-y curve that is recommended for weak rock is shown in Figure 4.43. The expression for the
ultimate resistance pur of rock is derived from the mechanics for the ultimate resistance of a wedge of
rock at the of the rock.
𝑥𝑟
𝑝𝑢𝑟 = 𝛼𝑟 𝑞𝑢𝑟 𝑏 (1 + 1.4 ) 𝑓𝑜𝑟 0 ≤ 𝑥𝑟 ≤ 3𝑏
𝑏
(4.76)
where
We assume that fracturing will occur at the surface of the rock under small deflections, so the
compressive strength of intact specimens is reduced by a factor αr. The value of αr is assumed to be 1/3
for RQD of 100 and to linearly increase to unity at RQD of zero. If RQD is zero, the compressive strength
may be obtained directly from a pressuremeter curve, or approximately from Figure 4.41, by entering
the value of the pressuremeter modulus.
Figure 4.43 Sketch of p-y curve for weak rock (after Reese, 1997)
If one were to consider a strip from a beam resting on an elastic, homogeneous, and isotropic
solid, the initial modulus Ki (pi /yi) in Figure 4.43 may be shown to have the following value(using the
symbols for rock):
where
The following equations for kir are derived from experimental data and reflect the assumption
that a rock surface will have a similar effect on kir as on pur for ultimate resistance.
400𝑥𝑟
𝑘𝑖𝑟 = (100 + ) 𝑓𝑜𝑟 0 ≤ 𝑥𝑟 ≤ 3𝑏
3𝑏
(4.79)
Using the guidelines for computing pur and Kir we present the following equations for the three
branches of the family of p-y curves for rock shown in Figure 4.43:
𝑝 = 𝐾𝑖𝑟 𝑦 𝑓𝑜𝑟 𝑦 ≤ 𝑦𝐴
(4.81)
𝑝𝑢𝑟 𝑦 0.25
𝑝= ( ) 𝑓𝑜𝑟 𝑦 > 𝑦𝐴 , 𝑝 ≤ 𝑝𝑢𝑟
2 𝑦𝑟𝑚
(4.82)
𝑦𝑟𝑚 = 𝑘𝑟𝑚 𝑏
(4.84)
where krm is a constant ranging from 0.0005 to 0.00005 that serves to establish the overall
stiffness of the curves.
The value of yA is found by solving for the intersections of Equations 4.81 and 4.82, and is
shown by the following equation:
𝑝𝑢𝑟 1.333
𝑦𝐴 = ( )
2 ( 𝑦𝑟𝑚 )0.25 𝑘𝑖𝑟
(4.85)
As shown in the case studies, the equations from weak rock predict with reasonable accuracy
the behavior of single piles under lateral loading for the two cases that are available. An adequate factor
of safety should be employed in all cases.
The equations are based on the assumption that p is only a function of y, an idea that appears to
be valid under static loading and resistance due only to lateral stresses. However, O’Neill points out that
“in large diameter drilled shafts, moment is resisted in the push-pull resistance produced by the axial
shears caused by the rotation of the pile. In rocks, this effect could be significant, especially for small
deflections, if the diameter of the pile is large” (O’Neill, Townsend, Hassan, Buller, & Chan, 1996).
Islamorada
The bored pile was 1.22 m (48 in.) in diameter and penetrated 13.3 m (43.7 ft) into the
limestone. A layer of sand over the rock was retained by a steel casing, and a lateral load was applied at
3.51 m (11.5 ft) above the surface of the rock. A maximum lateral load of 667kN (150 Kips) was applied,
yielding a nonlinear curve of load versus deflection.
Details on the strengths of the concrete and steel were unavailable and the bending stiffness of
the gross section was used for the initial solutions. The following values were used in the equations for
p-y curves:
αr =1.0,
krm =0.0005,
Figure 4.44 presents pile-head deflection from experiment and analysis. There is excellent
agreement between experimental and analytical results until about 350 kN (78.7 kips), when a sharp
change in the load-deflection curve occurs.
Curves giving deflection and bending moment as a function of depth were computed for a
lateral load of 334 kN (75 kips), one-half of the ultimate lateral load (Figure 4.45). The plotting is shown
for limited depths because the values to the full length are too small to plot. The stiffness of the rock,
compared to the stiffness of the pile, is reflected by a total of 13 points of zero deflection over the
length of the15.2 meters (50 ft) pile. However, for this data, the pile behaved as a long pile through the
full range of loading.
Figure 4.44 Comparison of experimental and computed values of pile head deflection,
Islamorada Test ( after Reese, 1997)
Values of EI were gradually reduced where bending moments were large in order to obtain
deflections that would agree fairly well with experimental values. Figure 4.45 shows that the largest
moment occurs in the zone of about2.5 m (8.2 ft) to 4.5 meters (14.8 ft). The following values of load
(KN) and bending stiffness (KN-m2) were used in the analyses: 350 and below 3.73 x 106; 400, 1.24 x 106;
467, 9.33 x 105; 534, 7.46 x 105; 601, 6.23 x 105; and 667, 5.36 x 105 . The computed bending-moment
curves were studied and reductions were only made where the bending stiffness was expected to be
nonlinear
Figure 4.45 Plot of Computed Curves of Deflection and Bending Moment versus Depth,
Islamorada Test, Lateral Load of334kN (after Reese, 1997)
The lowest value of EI that was used is believed to be roughly equal to the EI for the fully
cracked section. The decrease in the slope of yt versus Pt at Islamorada can be reasonably explained by
a reduction in the values of EI. Except for early loads, analyzing the tests at Islamorada gives little
guidance to the designer of piles in rock. Fortunately, a study of the tests done at San Francisco is more
instructive.
San Francisco
The value of krm used in the analyses was 0.00005. For the beginning loads, the value used for EI
was 35.15 x 106 kN-m2 (12.25 x 109 kip-in2); E=28.05x106 kPa (4.07 x 106 psi); I=1.253 m4(3.01 x 105 in4).
The ultimate bending moment Mult was computed to be 17,740 m-kN (1.57 x 105 in-kips) and values of
EI were computed as a function of bending moment. Data from (Speer, 1992) listed the following
properties of the cross section: compressive strength of the concrete was 34.5 MPa (5000 psi), tensile
strength of the rebars was 496MPa (72000 psi), there were 40 bars with a diameter of 43 mm (1.69 in),
and cover thickness was0.18 m(7.09 in.).
Figure 4.46 Comparison of Experimental and Computed Values of Pile-Head Deflection for
Different Values of EI, San Francisco Test
The data on deflection as a function of loads showed that the two piles behaved almost
identically for the beginning loads but the curve for Pile B exhibited a large increase in pile-head
deflection at the largest load. The experimental curve for Pile B shown by the heavy solid line in Figure
4.46 suggests that a plastic hinge developed at the ultimate bending moment of 17,740 m-kN (157,012
in-kips).
Consideration was given to the probable reduction in the values of EI when loads increased, and
three methods were used to predict the reduced values. The methods were: the analytical method as
presented in Part III, the approximate method of the American Concrete Institute (1989) which does not
account for axial load and can be used here, and the experimental method where EI is found by trial
computations that yields the observed deflections. The plots of the three methods are shown in Figure
4.47, and all three curves show a sharp decrease in EI with increase in bending moment. For
convenience, the value of EI was changed for the entire length of the pile, since the errors from using
constant values of EI in regions with low values of M are thought to be small.
The computed pile-head deflections are shown in Figure 4.46. The EI from experimental values
yields deflections that agree with computations, but the computed deflections from the other methods
are less than those from experiment. However, if load factors of 2.0 and higher are selected, the
computed deflections would be about 2 or 3 mm (0.078 to 0.118 in.) with the experiment showing
about 4 mm(0.157 in). Thus, the differences are not important in the range of service loading.
As shown in Figure 4.48, the values of I in Figure 4.47 along with I of the gross section were
used to compute the maximum bending moment as a function of the applied load. The good agreement
between all the methods is striking. The curve that results from using the gross value of I is reasonably
close to the curves resulting from adjusted values of EI, indicating that the computation of bending
moment for this particular example is not sensitive to the selected values of bending stiffness. The value
of the ultimate bending moment Mult computed by the analytical method is shown in Figure 4.48.
Assuming Mult to be correct, all of the methods predict bending failure with good accuracy.
4.13.1 Introduction
(Liang et. Al., 2009) developed a criterion for computing p-y curves for drilled shafts in massive
rock. This criterion is based on both full-scale load tests and three-dimensional finite element modeling.
𝑦
𝑝=
1 𝑦
𝐾𝑖 + 𝑝𝑢
(4.86)
where pu is the ultimate lateral resistance of the rock mass and Ki is the initial slope of the p-y curve. A
drawing of the p-y curve for massive rock is presented in Figure 4.49. Both of these parameters, Ki and
pu, are computed using the properties of the rock mass. The ultimate lateral resistance pu is computed
for two conditions; near the ground surface and at great depth. The lower of the two values of pu is used
in computing the p-y curve.
p
pu
y
p=
1 y
+
K i pu
Ki
𝑛
𝜎′3
𝜎′1 = 𝜎′3 + 𝜎𝑐𝑖 (𝑚𝑏 + 𝑠)
𝜎𝑐𝑖
(4.87)
where 1 and 3 are the major and minor principal stresses at failure, ci is the uniaxial compressive
strength of intact rock, and the parameters mb, s, and a are material constants that depend on the
characteristics of the rock mass; s = 1 for intact rock, and a = 0.5 for most rock types. The parameters mb
and s can be determined for many types of rock using the recommendations of (Marinos & Hoek, 2000).
Parameter mb can be computed using the Hoek-Brown material index mi and the Geologic Strength
Index, GSI, and blast damage factor Dr using
𝐺𝑆𝐼−100
( )
𝑚𝑏 = 𝑚𝑖 𝑒 28−14𝐷𝑟
(4.88)
Representative values for the Hoek-Brown material index are presented in Table 4.8. For deep
excavations like drilled shaft or bored piles, the blast damage factor Dr is assumed equal to zero.
(Hoek E. , Estimating Mohr-Coulomb Friction and Cohesion Values from the Hoek-Brown Failure
Criterion, 1990) provided a method for estimating the Mohr-Coulomb failure parameters c and of the
rock mass from the principal stresses at failure. These parameters are:
2𝜏
𝜑′ = 90𝑂 − 𝑎𝑟𝑐𝑠𝑖𝑛 ( )
𝜎′1 − 𝜎′3
(4.89)
𝑐 ′ = 𝜏−𝜎𝑛 tan 𝜑′
(4.90)
1 can be found from equation 4.87, and n and are found from
(𝜎′1 − 𝜎′3 )2
𝜎′𝑛 = 𝜎′3 +
2(𝜎 ′1 − 𝜎 ′ 3 ) + 0.5𝑚𝑏 𝜎𝑐𝑖
(4.91)
𝑚𝑏 𝜎𝑐𝑖
𝜏 = (𝜎 ′1 − 𝜎 ′ 3 ) √1 +
2(𝜎 ′1 − 𝜎 ′ 3 )
(4.92)
Table 4.8 Values of Material Index mi for Intact Rock, by Rock Group, from (Hoek E. , 2001)
Two methods for evaluating the modulus of the rock mass are recommended by Liang et al. One
method is to compute rock mass modulus by multiplying the intact rock modulus measured in the
laboratory by the modulus reduction ratio E m / Ei , computed from the geological strength index, GSI,
using equation 4.93.
𝐸𝑚 𝑒 𝐺𝑆𝐼/21.7
𝐸𝑚 = 𝐸𝑖 ( ) = 𝐸𝑖
𝐸𝑖 100
(4.93)
The experimental data and correlation for the modulus reduction ratio are shown as a function
of GSI in Figure 4.50.
100
Bieniawski (1978)
Serafin and Pereira (1983)
80 Ironton-Russell
Regression Line
Modulus Reduction Ratio
GSI / 21.7
E e
%=
m
E i
100
60
Em/Ei, (%)
40
20
0
0 20 40 60 80 100
Geologic Strength Index
Figure 4.50 Equation for Estimating Modulus Reduction Ratio from Geological Strength
Index.
The second method recommended by Liang et al. for determining the modulus of the rock mass
is to perform an in-situ rock pressuremeter test. The difficulty in using this approach is that many
pressuremeter testing devices are not capable of reaching sufficiently large pressures to deform the
rock. If this is the case, interpretation of test results may be restricted because of the limited range of
expansion pressures achievable.
Values for Poisson’s ratio are also required to compute p-y curves in massive rock. Values of
Poisson’s ratio vary with the quality of the rock mass. Typical values for Poisson’s ratio and other
properties for rock masses reported by (Hoek E. , 2001) are shown in Table 4.9.
Values of Poisson ratio for the rock mass can be estimated by interpretation measurements of
in-situ stress wave velocities. Poisson’s ratio can be computed from the compression and shear wave
velocities using equation 4.94, (Zhang, 2004). This relationship is drawn in Figure 4.51, , (Zhang, 2004).
(𝑉𝑃 − 𝑉𝑆 )2 − 2
𝜐=
2[(𝑉𝑃 − 𝑉𝑆 )2 − 1]
(4.94)
Table 4.9 Typical Properties for Rock Masses from (Hoek E. , 2001)
0.5
0.4
0.3
0.2
0.1
0
0 1 2 3 4 5 6
Vp/Vs
For a passive wedge type failure near the ground surface, as shown in Figure 4.52, the ultimate
lateral resistance per unit length, pus of the drilled shaft at depth H near the ground surface is
where = 45 + and = .
2 2
Fs
Fs H
Fnet
W
Fa
Fn
(Liang et. Al., 2009) note that the value of 3 can be taken as the effective overburden pressure
at a depth of H/3 for estimating and c using equations 4.87 and 4.88.
The following equations are used to compute parameters C1 through C5 with c = effective
cohesion, = effective friction angle, and, = effective unit weight respectively of the rock mass.
K a = tan 2 45 −
2
K 0 = 1 − sin
2c
z0 = − v0
Ka
H
C1 = H tan sec c + K 0 v0 tan + K 0 tan
2
D tan ( v0 + H ) + H tan 2 tan (2 v0 + H ) + c(D + 2H tan tan ) + 2C1 cos cos
C3 =
sin − tan cos
H
C4 = K 0 H tan sec v0 + , and
2
Equation 4.93 is valid for homogeneous rock mass. For layered rock mass, representative
properties can be computed by a weighted method based on the volume of the failure wedge. Methods
for obtaining the rock properties c and are given in Section 4.13.2 on page 4-101.
The passive wedge failure mechanism is not likely to form if the overburden pressure is
sufficiently large. Studies of rock sockets using three-dimensional stress analysis using the finite element
method have concluded that at depth the rock failure first in tension, followed by failure in friction
between the shaft and rock, followed finally by failure of the rock in compression. Therefore, the
expression for ultimate resistance at depth is a function of the limiting pressure, pL, and the peak
frictional resistance max. The ultimate resistance at depth can be computed using
2
pud = pL + max − pa D (4.96)
4 3
The effective overburden pressure, v, at the depth under consideration includes the pressure from
overburden soils.
The limiting normal pressure of the rock mass, pL, is taken as the compressive strength of the
rock mass, 1, computed using equation 4.87 and equating 3 equal to v,.
a
p L = v + ci mb v + s (4.98)
ci
The limiting shear stress, max, is the maximum axial side resistance of the rock-shaft interface, proposed
by (Rowe & Armitage, 1987).
where both max and ci are in units of megapascals. Values of max in units of kPa and psi are computed
by GROUP using the following equations:
For max and ci in units of kPa: max = 14.23025 ci (4.100)
For max and ci in units of psi: max = 5.4194 ci (4.101)
The initial tangent stiffness of the p-y curve, Ki, is computed rock mass modulus, Poisson’s ratio,
pile diameter, and mobilized bending stiffness of the pile using
D − 2 E p I p
0.284
K i = Em e
E D 4 (4.102)
D
ref m
where Em is the modulus of the rock mass, is Poisson’s ratio of the rock mass, D is the diameter of the
drilled shaft, Dref is a reference shaft diameter equal to 0.3048 m or 12 inches, and E p I p is the bending
stiffness of the drilled shaft.
1. Obtain the values of the intact rock strength ci and the intact rock modulus Ei.
2. Obtain values for the rock mass modulus, Em, by either use of equation 4.93 if pressuremeter data
are unavailable or from interpretation of pressuremeter testing results. If equation 4.93 is used,
obtain values of GSI and mi according to the recommendations of (Marinos & Hoek, 2000).
3. Obtain the value of Poisson’s ratio of the rock mass from in-situ measurements or estimated from
Table 4.9.
4. Select a shaft diameter, compressive strength of concrete, and reinforcing details.
5. Compute the bending stiffness and nominal moment capacity of the drilled shaft. Set the value of
bending stiffness equal to the cracked section bending stiffness at a level of loading where the
reinforcement is in the elastic range.
6. Compute Ki using equation 4.102.
7. Compute pus at shallow depth using equation 4.95 with 3 equal to the vertical effective stress at
H/3 when computing the values of and c using equations 4.89 and 4.90.
8. Compute pud at great depth using equation 4.96 with pL computed using Equation 4.98 and equating
3 equal to v.
9. Compute pu as the smaller of pus computed using equation 4.95 or pud computed using equation
4.96.
10. The values of the p can then be computed as a function of y, Ki, and pu using Equation 4.86 as a
function of pile displacement y.
The Piedmont residual soils are found east of the Appalachian ridge in a region extending from
southeastern Pennsylvania south through Maryland, central Virginia, eastern North Carolina, eastern
South Carolina, northern Georgia, into Alabama. It is a weathered in-place rock, underlain by
metamorphic rock. In general, the engineering behavior of Piedmont residual soil is poorly understood,
due to difficulties in obtaining undisturbed samples for laboratory testing and relatively wide variability.
The degree of weathering varies with local conditions. Weathering is greatest at the ground
surface and decreases with depth until the unweathered, parent rock is found. The residual soil profile is
often divided into three zones: an upper zone of red, sandy clays, an intermediate zone of micaceous
silts, and a weathered zone of gravelly sands mixed with rock. Often the boundaries of the zones are
indistinct or inclined. Weathering is greatest near seepage zones.
The method for computing p-y curves in Piedmont residual soils was developed by (Simpson &
Brown, 2006). This method was developed to use correlations for estimates of soil modulus measured
using four field testing methods: dilatometer, Menard pressuremeter, Standard Penetration Test, and
cone penetration tests. The basic method is described in the following paragraphs.
Given a shaft diameter b, and soil modulus Es, the relationship between p and y is
y/b
Es = Esi 1 − ln for 0.001 y/b 0.0375 (4.104)
0.001
pu = b y (1 − 3.624 ) (4.105)
pu = 1.834 Esi b y
Esi = Etest
Es/Esi
ln y/b
pu
y
0.001b 0.0375b
There are numerous cases where the soil near the surface of the ground is heterogeneous and
layered. If the layers are in the zone where the soil could move up and out as a wedge, some
modification to the way the ultimate soil resistance pu is computed would be needed. Consequently,
modifications to the p-y curves would also be needed.
The problem of layered soil was intensively studied by (Allen, 1985); however, Allen's
formulations require the use of several computer codes. Integrating the methods of Allen with the
methods shown herein must be delayed until his research can be put in a readily usable form.
The proposal of (Georgiadis, 1983) was used for the computer code that is presented here. His
method is based on determining the “equivalent” depth of all the layers existing below the upper layer.
The p-y curves of the upper layer are determined according to the methods presented here for
homogeneous soils. To compute the p-y curves of the second layer, the equivalent depth H2 to the top
of the second layer has to be determined by summing the ultimate resistances of the upper layer and
equating that value to the first summation as if the upper layer had been composed of the same
material as in the second layer. The values of pu are computed according to the equations given earlier.
Thus, the following two equations are solved simultaneously for H2.
𝐻1
𝐹1 = ∫ 𝑝𝑢1 𝑑𝐻
0
(4.106)
𝐻2
𝐹1 = ∫ 𝑝𝑢2 𝑑ℎ
0
(4.107)
The equivalent thickness H2 of the upper layer, along with the soil properties of the second
layer, are used to compute the p-y curves for the second layer.
The concepts presented above can be used to obtain the equivalent thickness of two or more
dissimilar layers of soil overlying the layer for which the p-y curves are desired. The equivalent depths
may be either smaller or greater than the actual depths of the soil layers and will depend on the relative
strengths of the layers in the soil profiles. This is illustrated in Figure 4.57.
h3
h1 Soft Soil (Layer 1)
h2
1
F2
Weaker Soil Below Stronger Soil
(behaves as if deeper)
F3
The equations used to compute lateral load transfer at failure are the ultimate values for flat
ground surfaces and vertical piles.
Layering corrections are applied only to the p-y curve models that have different expressions for
ultimate lateral load transfer for shallow and deep conditions. The p-y curve models that have different
expressions for shallow and deep conditions are:
The example problem demonstrating GROUP’s approach to layered soils is shown in Figure 4.56. As seen
in the sketch, a pile with a diameter of 24 in. is embedded in soil consisting of an upper layer of soft clay
overlying a layer of loose sand, which in turn overlays a layer of stiff clay. The water table is at the
ground surface and the loading is assumed to be static.
Four p-y curves for the case of layered soil are shown in Figure 4.57. The curve at a depth of 3 ft.
falls in the upper zone of soft clay, the curve at a depth of 6 ft. falls in the sand just below the soft clay,
and the curve for depths of 12 ft. and 24 ft. are in the lower zone of stiff clay.
Following the procedure suggested by (Georgiadis, 1983), the p-y curve for soft clay can be
computed as if the profile consists entirely of that soil. When dealing with sand, an equivalent depth is
found such that the value of the sum of the ultimate soil resistance for the equivalent sand and the soft
clay are equal at the interface. The equivalent depth of loose sand to substitute for the effect of the clay
was computed to be 6.82 ft. Thus, 5.50 ft. of soft clay is replaced by 6.82 ft. of loose sand, and point B,
which defines the position of the p-y curve in sand, is 7.32 ft. below the assumed ground surface that
has an actual depth of 6 ft. inches. Error! Reference source not found. shows a plot of the equivalent
thickness of the layer (computed as H2 by use of Equations 4.106 and 4.107) as shown in xEQ.
An equivalent depth of stiff clay was found so that the sum of the ultimate soil resistance for the
stiff clay is equal to the sum of the ultimate soil resistance of the loose sand and soft clay. That
equivalent depth was found to be 4.05 ft. and is indicated in Error! Reference source not found.. Thus,
the depths to the two p-y curves in the stiff clay are assumed to be 6.05 ft. and 18.05 ft. rather than the
actual depths of 12 ft. and 24 ft.
c = 500 psf
Soft Clay
5.5 ft. 50 = 0.02
’ = 50 pcf
c = 2000 psf
Stiff Clay 50 = 0.005
Without ’ = 60 pcf
Free Water
20 ft.
24 in.
Also interesting is that the p-y curve recommendations for stiff clay consider the presence of no
free water in the stiff clay. This decision is based on the assumption that the sand above the stiff clay
can move downward and fill any gap that develops between the clay and the pile. Furthermore, in the
stiff-clay experiment where free water was present, the free water moved upward along the face of the
pile with each cycle of loading. The presence of soft clay and sand to a depth of 10 ft. above the stiff clay
is believed to suppress the hydraulic action of free water even though the sand does not serve to close
the potential gaps in the stiff clay.
Stiff Clay
Without Point Actual Equivalent
Free Water Depth, ft Depth, ft
------------------------------------------------------------
A 3.0 3.0
20 ft. B 6.0 7.32
C 12.0 6.05
D 24.0 18.05
24 in.
Figure 4.58 Equivalent depths of overlying soil for use in computing p-y curves for a
layered soil
The effect of batter on the behavior of laterally loaded piles was investigated in a test tank. The lateral
soil-resistance curves of a vertical pile were modified by a constant to express the effect of the pile
inclination.
The values of the modifying constant as a function of the batter angle were deduced from
model tests in sand and also from full-scale, pile-loading tests. The criterion is expressed by a solid line
in Figure 4.59.
Plotted points in Figure 4.59 show the modification factors for the batter piles tested in
experiments by (Awoshika & Reese, 1971). The modification factors were obtained for two series of
tests independently.
Figure 4.59 Modification of p-y curves for battered piles, after (Kubo, 1964), and (Awoshika
& Reese, 1971)
The experimental modification factors were obtained after a few trial-and-error comparisons of
the horizontal pile-top displacements at the maximum load between a vertical pile and a battered pile.
Figure 4.59 indicates that for the out-batter piles, the agreement between the empirical curve
and the experiments is good, while for the in-batter piles the experiment did not show any effect due to
the batter.
The program does not consider automatically the effect of batter on the behavior of laterally
loaded piles. The user may consider this effect by using user-defined p-y curves or user-defined
reduction factors for lateral load.
(Focht & Koch, 1973) proposed another model that combined the well-documented p-y
approach of a single pile with the elastic-group effects from Poulos’ work. Focht and Koch’s modification
begins by introducing a term ρ into Poulos’ equation as:
𝜌𝑘 = 𝜌𝐹 ∑(𝐻𝑗 𝛼𝜌𝐻𝑗𝑘 + 𝐻𝑘 )
𝑗=1
𝑗≠𝑘
(4.108)
where
ρF = the unit reference displacement of a single pile under a unit horizontal load, computed
by using elastic theory,
αρHjk = the coefficient to get the influence of pile j on pile k in computing the deflection ρ (the
subscript H pertains to the case in which shear is applied at the pile head; there are also
influence coefficients for applied moment and for the fixed head case, αρFjk)
The above equation can be used to solve for group deflection, Yg, and loads on individual piles.
With the known group deflection, Yg, the p-y curves at each depth for a single pile can be multiplied by a
factor, termed the “Y” factor, to match the pile-head deflection of a single pile with the group
deflection, Yg, by repeated trials. The “Y” factor is a constant multiplier employed to increase the
deflection values of each point on each p-y curve; thus, generating a new set of p-y curves that include
the group effects. The modification of p-y curves as described above for piles in the group allow the
computation of deflection and bending moment as a function of depth.
From a theoretical viewpoint, group effects for the initial part of p-y curves can be obtained
from elastic theory. The ultimate resistance of soil on a pile is also affected by the adjacent piles due to
the interference of the shear-failure planes, called shadowing effects. Focht and Koch suggested a p-
factor may need to be applied to the p-y curves in cases where shadowing effects occur. The p-factor
should be less than one and the magnitude depends on the configuration of piles in a group.
Other approaches regarding the modification of the coefficient of subgrade reaction were also
used for pile-group analyses. The (Society, 1978) recommends that the coefficient of subgrade reaction
for pile groups be equal to that of a single pile if the spacing of the piles in the group is eight diameters.
For spacing smaller than eight diameters, the following ratios of the single-pile subgrade reaction were
recommended: six diameters, 0.70; four diameters, 0.40; and three diameters, 0.25.
The (Association, 1976) is less conservative. It suggested that a slight reduction in the coefficient
of horizontal subgrade reaction has no serious effect with regard to bending stress and that the use of a
factor of safety should be sufficient in design, except in the case where the piles are close together.
When piles are less than 2.5diameters apart, the following equation is suggested for computing a factor
to multiply the coefficient of subgrade reaction for the single pile.
𝐿
𝜇 = 1 − 0.2 (2.5 − ) , 𝐿 < 2.5𝐷
𝐷
(4.109)
where
D = pile diameter.
(Bogard & Matlock, 1983) present a method in which the p-y curve for a single pile is modified
to take into account the group effect. Excellent agreement was obtained between their computed
results and results from field experiments (Matlock, Ingram, Kelly, & Bogard, 1980).
After reviewing the common approaches for the analysis of pile groups, it seems that a hybrid
soil model that modifies the p-y curve for a single pile, with some factors to take into account group
effects, can be established for closely-spaced-pile groups.
The group reactions from elastic theory can be added to the p-y curves of a single pile. If
shadowing effects do not exist, the p-y curves with or without elastic-group reactions should reach the
same ultimate resistance (Figure 4.60). However, it is realized that to develop the same ultimate
resistance, a larger deflection is needed when group action is present, because the supporting soil is
forced to deform by the stresses from nearby piles. A y-multiplier can be used to increase the deflection
of p-y curves on each individual pile if the group effect is significant.
Figure 4.60 p-y curves for single pile and pile group without shadowing effect
When piles are in a closely spaced group, the shear-failure planes resulting from the movement
of each pile will overlap, and the ultimate resistance for piles in a group may be less than that of a single
pile. Generally, the ultimate resistance can be derived from limit analysis or by other approximate
methods. A p-multiplier can be used to reduce the ultimate soil resistance in cases where the shadowing
effect is an important factor.
With the p-multiplier (α1) and y-multiplier (α2) described above, complete p-y curves for piles in
a closely-spaced group can be constructed as shown in Figure 4.61.Specific recommendations for values
of α1 and α2 are not included herein because research is incomplete. The reader can refer to the
references for suggestions that may prove useful.
Figure 4.61 p-y curves for single pile and pile group with group effect
5.1 Introduction
As noted in Chapter3, pile-group analysis requires load-settlement curves for axial loads and
load-deflection curves for lateral loads to allow the conditions of equilibrium and compatibility to be
satisfied. The methods for predicting soil-response curves (p-y curves) and for computing a load-
deflection curve at the pile head have been discussed in previous chapters. Although the load-
settlement curve can be obtained directly from an axial loading test, it will not be available on many
occasions, especially during the preliminary design stage. This chapter aims to provide analytical
procedures for prediction of the load-settlement curve.
While the mechanics of load transfer from a deep foundation (pile) to the supporting soil is not
well understood and is the subject of many investigations, some general aspects of the mechanics are
known and presented here to provide a useful tool for analysis.
A typical load-settlement curve for a deep foundation is shown in Figure 5.1. The vertical dashed
line to the far right in the figure is the load which causes plunging; that is, the load which will cause
continued settlement of the pile with no increase in load.
If the load is increased from zero to some value Qp at point P, the gross settlement is
represented by the horizontal dotted line to point P. If the load is then released, there is some rebound
as indicated by the light solid line, with the net settlement being defined as the settlement at zero load
after unloading.
The manner in which the load is distributed from the pile to the supporting soil is of interest. A
typical curve of the distribution of load along the length of an axially loaded deep foundation is shown in
Figure 5.2. The slope of the curve indicates the rate of load transfer from the foundation to the soil.
Some insight into the problem of the mechanics of load transfer from a deep foundation may be
obtained by considering the idealized load-distribution and load-settlement curves in Figure 5.3. Figure
5.3a shows the results of loading a pile which rests on an unyielding surface in which all the load is
transferred by tip resistance and none by side resistance. The load-distribution curves for loads of QT1
and QT2 show a constant load in the pile regardless of depth. The load-settlement curve for such a case
is also shown. The settlement, or movement at the top of the pile, can be obtained by computing the
compression in the pile from basic principles of mechanics. The load-settlement curve will be a straight
line, as shown, if the effective modulus of elasticity of the shaft is linear.
Figure 5.2 Curve showing typical distribution of load along the length of an axially-loaded
pile
Figure 5.3b shows a similar pile, but in this instance its tip is assumed to be resting on an elastic
surface that yields linearly with the applied load. The load-distribution curves remain unchanged, as
shown. However, the settlement of the top of the pile is now made up of two quantities: (1)
compression in the pile due to the applied load, and (2) the settlement of the pile tip.
Figure 5.3c shows the case where the soil produces uniform shaft resistance and no tip
resistance. The load-distribution curves are triangular, as indicated. The load settlement is again linear
and is composed of the compression in the pile due to the triangular distribution of load, and the
settlement of the tip of the pile.
While of interest, none of these idealized models represent the true behavior of the axially
loaded deep foundation. A real pile has some combination of all three factors illustrated in Figure 5.3,
plus non uniform and nonlinear behavior. A more realistic model is shown in Figure 5.4. Figure 5.4a
shows the free body of a pile in equilibrium where the applied load QT is balanced by a tip load QB plus
side loads R. A mechanism is shown in Figure 5.4b that can be used to illustrate the deformations in the
pile. The pile has been replaced by an elastic spring. Representing the soil is a set of nonlinear springs
spaced along the pile, with one spring depicting the soil behavior beneath the pile tip. The ordinate f of
the curves is load transfer and the abscissa wz is the shaft movement. (Rather than fz -wz curves, the
present practice is to designate them as t-z curves.) No load is transferred from shaft to soil unless there
is a downward movement of the shaft. This downward movement is dependent on the applied load, the
position along the pile, the stress-strain characteristics of the pile material, and the load transfer-
movement curves along the pile shaft and tip. To solve the problem of load distribution along the pile
for a given applied load, along with the downward movement at any point along the pile, a nonlinear
differential equation must be solved.
The acquisition of load-transfer curves from a load-test requires that the pile be internally
instrumented for the measurement of axial load with depth. The number of such experiments is
relatively small, and in some cases the data are barely adequate; therefore, the amount of information
that can be used to develop analytical expressions is limited.
There will undoubtedly be additional studies reported in technical literature. Any improvements
that are made in load-transfer curves can be readily incorporated into the analyses.
(Coyle & Reese, 1966) examined the results from three instrumented field tests in addition to
results from rod tests in the laboratory before developing a recommendation for a load-transfer curve.
The curve was tested by using results of full-scale experiments with uninstrumented piles. The
computed load-settlement curves agreed well with experimental curves. Table 5.1 presents the
fundamental curve developed by Coyle and Reese and shows that the movement to develop full load
transfer is quite small. Furthermore, the curve is independent of soil properties and pile diameter.
(Reese & O’Neill, 1987) studied the results of several field-load tests of instrumented drilled
shafts to develop the curves shown in Figure 5.5, which shows that the maximum load transfer occurs at
approximately 0.6% of the diameter of a pile. Because the piles tested had diameters of 24 to 36 inches,
the movement at full load transfer would be around 0.2 inches, a much larger value than obtained by
Coyle and Reese.
Figure 5.5 Normalized curves showing load transfer in side resistance versus settlement for
drilled shafts in clay (after Reese and O’Neill, 1987)
(Kraft, Focht, & Amerasinghe, 1981) studied the theory related to the transfer of load in side
resistance, noting that pile diameter, axial pile stiffness, pile length, and distribution of soil strength and
stiffness along the pile are all factors that influence load-transfer curves. Equations for computing the
curves we represented. (Vijayvergiya, 1977) also presented a method for obtaining load-transfer curves.
The details of their methods are not presented here.
The work of (Skempton, 1951) was employed for developing a method to predict the end-
bearing load of a pile in clay as a function of the tip movement of the pile. The laboratory stress-strain
curve for the clay at the base of the pile was obtained by testing or estimating from 𝜖50 values given by
Skempton for laboratory strain at one-half of the ultimate compressive strength of the clay. Skempton
reported that 𝜖50 ranged from 0.005 to 0.02. Skempton further developed approximate equations for
the settlement of a footing (base of a pile) by using the theory of elasticity in conjunction with
laboratory stress-strain curves. His equations are as follows:
𝜎𝑓
𝑞𝑏 = 𝑁𝑐 ( )
2
(5.1)
𝑤𝑏
= 2 𝜖50
𝐵
(5.2)
where
Stress-strain curves from a number of laboratory tests have been found to plot as an
approximately straight line on log paper. The slopes are often approximately 0.5. Therefore, in the
absence of a laboratory stress-strain curve, a parabola can be used to yield a stress-strain curve up to
the failure stress. The value of 𝜖50 can be selected in consideration of whether the clay is brittle or
plastic.
As an example of the use of the Skempton relationships, the following data were selected.
c = 780 lb/ft2
𝜖50 = 0.01
The basic equation for load versus settlement at the tip of the pile is as follows.
𝑄𝑏 = 𝐾𝑏 (𝑤𝑏 )𝑛
(5.3)
where
Kb = a fitting factor.
Substituting,
In this example, it is assumed that the load will not drop as the tip of the pile penetrates the
clay.
(Reese & O’Neill, 1987) studied the results of many tests of drilled shafts in clay where
measurements yielded load in end bearing versus settlement. Figure 5.7 resulted from their studies. A
study of the mean curve for this example shows that a value of about 1.2 in. is obtained at the ultimate
bearing stress, rather than the value of 2.4 in. found by the method adapted from Skempton’s work.
However, the range of values shown by Wright encompasses the earlier result.
It is of interest to note that the movement of a pile that causes full load transfer in end-bearing
is several times that which is necessary to develop full load transfer in skin friction. The above concept is
easily demonstrated by considering the soil elements that are strained in end-bearing as compared to
those strained in skin friction.
Figure 5.6 Normalized curves showing load transfer in end bearing versus settlement for
drilled shafts in clay ( after Reese and O’Neill,1987)
Coyle & Sulaiman (1967) studied the load transfer in skin friction of steel piles driven into sand
to obtain the curves shown in Figure 5.7. The piles had diameters ranging from 13 in. to 16 in. and had a
penetration of about 50 feet. Examining the shape of the curves shows that they can be fitted with the
following equation:
𝑤 0.15
𝑓 = 𝐾𝑠 ( ) 𝑍/𝐵 ≤ 0.07
𝐵
(5.4)
where
f = load transfer,
w = movement of pile,
Ks = fitting factor.
Equation 5.4 can be used to obtain the approximate load-transfer in skin friction for a pile in
sand. The value of the maximum unit load transfer fmax would be obtained by use of appropriate
equations. The value of the fitting factor Ks can be found:
𝑓𝑚𝑎𝑥 = 𝐾𝑠 (0.07)0.15
(5.5)
The curves are assumed to be flat for movements larger than 0.07 B. While Equation 5.4 and
Figure 5.7 show that the load transfer increases until a movement of 0.07 B, the maximum load transfer
essentially occurs at a movement of about 0.03 B.
Figure 5.7 Load transfer-movement curves for piles in sand from field tests(after Coyle and
Sulaiman,1967)
(Reese & O’Neill, 1987) examined the results of load tests of a number of full-sized drilled shafts
that were instrumented for the measurement of axial load with respect to depth. The results of his
studies showed that the curves in Fig 5.8 can be used for cohesionless soil.
Figure 5.8 Normalized curves showing load transfer in side resistance versus settlement for
drilled shafts in cohesionless soil (after Reese and O'Neill, 1987)
(Mosher, 1984) studied load transfer problem in skin friction (side resistance) of axially loaded
piles in sand. He recommended the use of the following equation and table for obtaining t-z (load
transfer curves):
𝑤
𝑓=
1 1
𝐸𝑠 + 𝑓𝑚𝑎𝑥 𝑤
(5.6)
where
Prior to making use of Equation 5.6, the mixed dimensions should be noted.
(Vijayvergiya, 1977); (Mosher, 1984) proposed an equation for computing the load versus tip
settlement for piles in sand. (Vijayvergiya, 1977) gives the displacement function as
𝑧 1/3
𝑞 = ( ) 𝑞𝑚𝑎𝑥
𝑧𝑐
(5.7)
𝑞𝑚𝑎𝑥 = 𝜎̅𝑣 𝑁𝑞
(5.8)
and 𝜎̅𝑣 is the effective vertical stress and equal to the overburden pressure. The bearing
capacity factor Nq is computed based on Meyerhof's equation as,
𝑁𝑞 = 𝜀 𝜋 tan tan2 (45 + )
2
(5.9)
The greatest limitation of using the maximum bearing capacity is its dependence on overburden
pressure.
(Vijayvergiya, 1977) linked the critical displacement, zc, to the width of the pile tip. For
displacements, z, greater than the critical displacement, zc, the tip resistance is a constant. This is
contrary to early statements in his paper, that for sands the tip resistance continued to increase with
displacement.
(Mosher, 1984) proposed another approach for defining the maximum tip resistance. The soil
beneath the tip is assumed to behave similarly to an elastic-plastic material with strain hardening, as
shown in Figure 5.8. The yield point is then defined as the point of transition between elastic and plastic
strain hardening behavior. The yield capacity of a pile has been assumed to be reached when the tip
displacement is 0.25 inches. Substituting 0.25 inches in Vijayvergiya's expression for displacement yields
𝑞 = (4𝑧)1/3 𝑞𝑚𝑎𝑥
(5.10)
This displacement function showed better agreement with field data. It was further altered to
reflect the field data for different densities:
where the maximum yield tip resistance qmax is obtained from the method described in
Equation 5.8.
(Reese & O’Neill, 1987) studied the results of experiments with drilled shafts and developed
Figure 5.10. The information in Figure 5.10 was developed from a relatively small amount of data and, as
with other methods presented in this chapter, should be sued with appropriate discretion.
Figure 5.10 Normalized curves showing load transfer in end bearing versus settlement for
drilled shafts in cohesionless soil (after Reese and O'Neill, 1987)
A review of the technical literature on single piles and groups of piles under lateral loading
shows that piles exhibit distinctly nonlinear behavior. Nonlinear p-y curves have been in use for the
analysis of single piles under lateral loading for some time. The best method for analyzing groups of piles
appears to use p-y curves that have had their p-values reduced to account for the reduction of soil
resistance due to group effect. (Brown, Reese, & O’Neill, 1987) found that the behavior of the individual
piles was best modeled by using a family of p-y curves developed from lateral-loading tests of a single
pile at the test site and modified by reducing all the p values on each p-y curve by a factor fm, as
illustrated in Figure 6.1.
Three levels of spacing between piles in a group are indicated: piles far enough apart for no pile-
soil-pile interaction to be evident; piles that are an intermediate distance apart, so pile-soil-pile
interaction is relatively minor; and piles close together enough for the interaction to be substantial. For
piles that are an intermediate distance apart, elastic theory may be used to compute a value that simply
stretches the deflection y for a certain soil resistance p, as shown in Figure 6.2. The p-values, with or
without elastic group-action, should reach the same ultimate resistance. However, when group action is
present, a larger deflection is needed to develop the same ultimate resistance, because stresses from
nearby piles force the supporting soil to deform elastically.
When piles are in a closely-spaced group with substantial interaction, the shear-failure planes
resulting from the movement of each pile will overlap, and the ultimate resistance for piles in a group
may be less than that of a single pile. Thus, a modification factor similar to fm is needed to reduce the
ultimate soil resistance, as shown in Figure 6.3 (Brown et al., 1987)
Figure 6.2 p-y curves for single pile and pile group without shadowing effect
Figure 6.3 p-y curves for single pile and pile group with shadowing effect
For closely-spaced piles, the influence of elastic behavior is present to a minor degree because
of the “shadow” or “edge” effects that dominate the behavior. These shadow effects have a far greater
influence on the efficiency of the individual piles. The reduction factor fm for various spacing of piles is
derived based on results from loading tests, which explicitly includes all effects. However, a simple
reduction factor fm, applied only on the soil resistance, will be presented herein.
Experimental results provide the main sources of information that engineers can rely upon to
develop design criteria for taking into account pile-soil-pile interaction for closely-spaced piles.
The main objective of experiments is to investigate the reduction ratio(group efficiency), which
is the ratio of the load carried by a pile in a group as compared to that of an isolated pile, under similar
conditions.
As revealed in the experimental studies, the capacity of a group of piles to sustain loading is
dependent on a number of parameters, but we focus on the positions of the individual piles. Reduction
factors for p-values are recommended for side-by-side piles, and line-by-line piles. A combined
reduction from side-by-side and line-by-line positions is presented. At present, is insufficient for deriving
the reduction factors of many soil types, so we give general values.
The following sections present information concerning the efficiency of piles in a single row,
either front to back or side by side. Information will also be given on groups in other configurations.
Some of the experiments described earlier yielded information on the distribution of load to the
individual piles in a group. In using the available data to obtain recommendations for design, two factors
guided the studies and are separately discussed.
First, the principal objective was to find the distribution of loading to the piles in the group when
the group was at the failure condition. In most instances, the failure of a single pile or an individual pile
in a group is considered to occur when the bending moment exceeds the moment-capacity of the pile.
However, in many of the studies that are used to derive the values of fm, information was unavailable
on bending moment, so failure was taken at some value of deflection. Nevertheless, the definition of
failures by bending moment is recommended for design, and means of computing the loading that will
develop the failure condition are presented.
Second, data allowing fm to be found directly only existed for two experimental studies, where
piles were instrumented along their lengths. In most experiments, available information only allowed
only the reduced load to be found. Therefore, we assume that fm is equal to the reduction factor for the
load of the individual piles in a group. We recognize that a reduction in fm may result in a distribution of
load to the piles in a group that agrees only roughly to the loads that were measured in an experiment.
However, we believe that the method that is presented is the best that can be developed from the data
that are available.
Experimental studies conducted by (Cox, Dixon, & Murphy, 1984); (Prakash, 1962); (Wang,
1986); (Lieng, 1988) have all included loading tests on side-by-side piles. The factor for reducing p is
defined as the ratio of the averaged capacity of individual piles in a group to the capacity of a single pile.
The p-reduction versus the pile spacing is in terms of s/b (s is the center to center spacing, b is the pile
diameter) and is presented in Figure 6.4.
It clearly shows that the reduction is negligible if the spacing is greater than three times the pile
diameter. According to the analytical study by Wang, the reduction for ultimate resistance is near 0.5
when piles are in contact edge to edge. While the scatter in Figure 6.4 is significant, the curve in the
figure currently represents the best estimate of the reduction factor as a function of pile spacing
regardless of the soil type.
The interaction of piles in the direction of loading is more complicated than that of piles in a
row. Many experiments have concluded that the interaction principally depends on the relative
positions of the piles. (Dunnavant & O’Neill, 1986) formalized the data of Cox et al. (1984) and
recommended reduction factors for leading piles and trailing piles as a function of pile spacing in the
direction of loading. An approach similar to that of Dunnavant and O’Neill is used herein to find the
reduction factor for p.
Figure 6.5 Curve giving reduction factors βbl for leading piles in a line
Although experiments were conducted in many different soil conditions, the influence of soil
properties on reduction factors is not quantifiable at present. Therefore, reduction factors are only
based on the relative positions of the piles in the group, though it is necessary to present separate
recommendations for leading piles and trailing piles.
Figure 6.6 Curve giving reduction factors βbt for trailing piles in a line
The reduction factors for the leading piles in a line may be found by referring to the curve in
Figure 6.5b. Regarding the group of three piles Figure 6.5a, Pile 1 is a leading pile relative to Piles 2 and
3, while Pile 2 is a leading pile relative to Pile 3. Generally, piles in the leading position are affected only
slightly by trailing piles in the same line. (Brown et al., 1987) indicate that the leading piles (first row in
their full-scale loading tests of a3-by-3group) sustain the largest loads. (Cox et al., 1984); (Schmidt,
Group action of laterally loaded bored piles, 1981) (Schmidt, 1985); (Lieng, 1988) all obtained results
from tests in the laboratory. Their results show that the load carried by the leading piles is only slightly
smaller than for a single pile.
The reduction factors for the trailing piles in a line may be found by referring to the curve in
Figure 6.6b. Referring to the group of three piles in Figure 6.6a, Pile 2 is a trailing pile relative to Pile 1,
and Pile 3 is a trailing pile relative to Piles 1 and 2. The study conducted by (Prakash, 1962) concluded
that the trailing pile reduction can only be ignored if s/b is equal to or greater than 8. Test data from
(Cox et al., 1984); (Schmidt, Group action of laterally loaded bored piles, 1981) (Schmidt, 1985); (Lieng,
1988) are presented in Figure 6.6b and shows that reduction can be ignored if s/b is about 6.
The effects of skewed piles (piles neither in line or side-by-side relative to the direction of
applied load) were not measured directly by experiment. However, a simple mathematical expression
for an ellipse in polar coordinates is suggested for use in obtaining the reduction factor for a skewed
pile. The expression is illustrated in Figure 6.7. Pile A and Pile B are skewed relative to the loading
direction, and the distances in the side-to-side and in the in-line positions are x and y, respectively. The
side-by-side reduction βa can be computed from Figure 6.4 based on the spacing r/b (b is the pile
diameter). The in-line reduction βb for loading piles can be computed from Figure 6.5 or Figure 6.6,
based on the spacing r/b. The reduction factor for one of the piles with respect to the other can be
expressed as:
1/2
𝛽𝑠 = (𝛽𝑏 2 cos2 + 𝛽𝑎 2 sin2 )
(6.1)
where is the angle between the direction of loading and the line connecting Piles A and B.
Figure 6.7 System for computing reduction factor for skewed piles
The reduction factors shown in Figure 6.4 through Figure 6.6 were derived from the results of
tests on groups with a small number of piles, but are suggested for use with any arrangement of piles in
a group. Example computations are shown here for the 2 x3 group as shown in Figure 6.8 (The pile
group has two rows, and each row consists of three piles.).
For Pile 4 in this example, the p reduction factor fm, is computed as:
then
β34 =(0.64)(2.0)0.34=0.8101
β24 =(0.48)(3.0)0.38=0.7287
then
For each Pile i in the group, the group reduction factor fm can be expressed as:
where βji is the group reduction factor due to the effect of Pile j on Pile i.
The group-reduction factor can be calculated for each pile in the group. Then, the analytical
procedure described for the widely-spaced pile group can be used to solve the pile-group behavior by
applying the group reduction factor to the soil-resistance curves for the single pile.
In a group of closely-spaced piles, the axial capacity of the group is influenced by variations in
the load-settlement behavior of individual piles because of pile-soil-pile interaction. The efficiency of a
pile is defined as the load it carries in the group divided by the load it would carry as a single, isolated
pile. The group effects for piles under axial loading, discussed in this chapter, will focus on proposals for
determining the efficiency of the individual piles.
The problem is complicated by the presence of the pile cap in two ways. First, if the cap is
perfectly rigid and the axial loading is symmetrical, all of the piles will settle the same amount. However,
if the cap is flexible, the settlement of the piles will not be the same. Secondly, if the cap rests on the
ground surface, some of the axial load will be sustained by bearing pressure on the cap. However, we
can conservatively assume that soil can settle beneath the cap, and that all of the load is taken by the
piles.
Unlike the behavior of a group of piles under lateral loading, the behavior of a group under axial
loading strongly depends on methods of installation, types of soils, and the stress-induced settlement.
The position of each individual pile in the group is less important for piles under axial loading.
Investigations determining group efficiency under various soil conditions (cohesive and cohesionless)
and piles spacing will be discussed here.
𝑄𝑔
𝐾=
𝑛𝑄𝑠
(7.1)
where
n = number of piles
Qs = capacity of a reference pile that is identical to a group pile but is isolated from the
group.
Figure 7.1 Results of model tests on groups of instrumented driven piles in granular soils
(fromVesic,1969(
Figure 7.1 shows the average efficiencies for driven piles as reported by (Vesic, 1969). Theory
predicts that the capacity of a pile in cohesionless soil increases with the effective stress. Thus, the
overlapping zones of stress at the base of a group of piles will cause an increase in end bearing. As
shown in Figure 7.1, the average values of efficiency are slightly greater than unity. Also, the lateral
compaction of the cohesionless soil during installation can cause an increase in effective stress along the
sides of a driven pile (the shaft). Figure 7.1 shows that the efficiency of the shaft can range from slightly
over 1.5 to almost 3.0. The figure shows that the overall efficiency of a group in cohesionless soil is well
above one.
While not reported in the figure, settlement frequently controls the capacity of piles in granular
soils. Where settlement is critical, special study of the data from Vesic and from other authors is
desirable because settlement cannot be reduced using small spacing as bearing capacity is being
increased.
(O’Neill, 1983) comprehensively studied the behavior of pile groups under axial loading. Figure
7.2 shows a compilation of results of model tests for inserted piles (similar to driven piles). The trends in
Figure 7.2 were described by O’Neill as follows: (a) K in loose sands always exceeds unity, with highest
values occurring at spacing-to-diameter ratios s/d of about 2, and a generally higher K occurring with
increased numbers of piles. Block failure (i.e., simultaneous failure of the piles and mass of soil within
the pile group) only affects K in 4-pile groups below s/d=1.5 and in 9 to 16-pile groups below s/d=2.
(b) K in dense sands maybe either greater or less than unity, although the trend is toward K>1 in groups
of all sizes for 2 <s/d<4. Efficiency of less than unity is probably a result of dilatency and would not
generally be expected in the field for other than bored or partially jetted piles, although theoretical
studies of interference suggest K slightly below 1 at s/d>4.
Little data exist concerning efficiency of piles in uplift. The results of studies on the resistance to
uplift of two 4-pile groups in loose sand are included in Figure 7.2. The results indicate sharp reductions
in K at all s/d ratios; however, these tests were conducted by molding sand around the piles and
undoubtedly represent lower bounds to field conditions for driven piles. More representative K values
can possibly be obtained by applying the theoretical frictional models of (Nishida, 1969) or (Hansen J. B.,
1968) which yield K only slightly less than unity for uplift loading.
O’Neill also summarized evidence on the efficiency of groups of full-scale piles in sands with
their caps suspended and in contact with soil, as shown in Figure 7.3.Thenotation Bav is the average
group width, dm is the mean pile diameter, p is the pile penetration, and s is the pile spacing. Based on
the discussion by O’Neill, the trends for free-standing pile groups are: (a)observations from model tests
confirm that efficiency in compression always exceeds one unless the piles are bored or partially jetted,
and(b) peak efficiency is achieved for groups having Bav/dm ratios of 2 to 7.
Figure 7.2 Vertical efficiencies for model groups ins and(After O’Neill,1983)
(Hansbo, 1993) pointed out that both large-scale and full-scale tests of groups of freestanding
piles in loose to medium dense sand have resulted in values of K>1 with a maximum of K=2 at values
of s/d of about 2 to 3. In dense sand results have been found where K>1 and K<1.
As noted earlier, the values of K greater than one in sand are apparently due to increased
effective stress at the walls of piles as a result of driving. A careful study of experimental results shows
that dense sand, with an angle of internal friction greater that 40◦, does not experience the increased
lateral stress. However, the driving of piles into sand with values of greater that 40◦is likely to be
impossible. Since the settlement of a group of piles in cohesionless soils is greater than that of an
isolated pile with the same load per pile, the stiffness of the pile cap may be a matter of concern.
Conventional practice generally assigns 1.0 to group efficiency for driven piles in cohesionless
soil, unless the pile group is founded on dense soil of limited thickness underlain by a weak soil deposit.
In such conditions, (Meyerhof, 1974)suggested that the efficiency of the group be based on the capacity
of an equivalent base that punches through the dense sand in block failure.
For bored and jetted piles, the group efficiency for a suspended pile cap is generally taken as
about 0.67 to 0.9. If the cap is in contact with the soil, the group efficiency is close to 1,as shown in
Figure 7.3; however, as noted earlier, a prudent approach could assume that the soil under the cap will
settle. (Liu, Yuan, & Shang, 1985) presented information on the group efficiencies for pile caps in contact
with soil, as given in Table 7.1.Thegroup efficiency for such conditions is in general greater than 1.
Table 7.1 Group efficiencies for bored piles according to Liu et al.(1985) Pile diameter Dp = 0.25 m;
loose, silty sand; pile cap in contact with soil.
In general, piles in cohesive soils sustain load principally in side resistance or behave as “friction”
piles. The load-carrying capacity of a group of friction piles in clay is the smaller of the following:
2. Load carried by an imaginary block encompassing the group where the load is the sum of the
load carried by the base and perimeter of the block.
In certain types of clay, especially highly sensitive clay, the efficiency of closely-spaced piles in a
group is less than one. However, there is insufficient field data from testing of full-sized piles to allow for
quantification of the efficiency of the piles as a function of the center-to center spacing. The use of the
imaginary block, described above, is the common way to investigate the efficiency of the group. Figure
7.4 was summarized by (Mello, 1969), and results show that block failure of a pile group does not
become pronounced until the pile spacing is less than 2 pile diameters.
(O’Neill, 1983) conducted a group efficiency study for clay that was similar to his study on group
efficiency for clay. Figure 7.5 shows a compilation of the model test results of pile groups in clay. Model
tests in clay, in contrast to those in sand, always yield K<1 and trend towards block failure in square
groups at s/d values of roughly less than 2. The group efficiency is higher in stiff clays than in soft clays.
Figure 7.6 presents the result from full-scale tests summarized by O’Neill (1983), where the caps
are suspended or in contact with soil. The efficiencies are one or less for the suspended caps, and may
be higher than one for the caps in contact with soil. However, the possibility of the soil settling
eliminates the effectiveness of the cap. An important factor with respect to a group of piles in saturated
clay is that the generation of pore water pressures during pile driving creates a large “bulb” of excess
pore pressure around the group of piles. Thus, the dissipation of pore pressure around pile groups will
be slower that the dissipation around single piles, a fact that may have been neglected addressed in
previous tests of full-scale piles.
The settlement of a group of piles in saturated clay is time-dependent and may be an over-riding
consideration with respect to the capacity of the group to sustain axial loading. The design of a group of
piles in clay depends on a large number of factors, including site-specific conditions, the nature of the
superstructure, as well as the center-to-center spacing of the piles.
Figure 7.4 Results of model tests on pile groups in clay under compression (After
Mello,1969)
The most reliable data concerning the efficiency of the piles in a group is derived from the
results of full-scale tests. However, the behavior of pile groups under axial loading depends on many
factors that require a large number of carefully controlled tests for thorough investigation. Such a
program is beyond the capabilities of the agencies currently interested in the behavior of piles. As an
example of current limitations, guidelines are unavailable for estimating the efficiency of groups of piles
in layered soils. The data presented in this chapter are useful, present an introduction to the efficiency
of piles in groups under axial loading, but are not meant to provide specific information for design, even
for the most routine problem. The judgment of the engineer should be made based, among other
things, on the type of piles, construction methods, distribution of friction and tip resistance, and the
dominant soil layers.
As a concluding comment, the design of piles in groups under axial loading should be based on
the concepts of soil-structure interaction. Curves showing the transfer of load in side resistance (t-z
curves) and in end bearing (t-z curves) can potentially be modified to account for efficiency. Then, both
capacity and settlement can be computed with a rational model. Such an approach must await the
results of further research.
8.1 Introduction
Recent developments on the seismic design of bridge foundations and abutments are focused
on determining the foundation stiffness and ultimate capacity of piles. In general, the rotational stiffness
of the pile cap is primarily controlled by the axial stiffness of the piles. The effect of the rotational
resistance due to soil being in contact with the base of the pile cap has been ignored, because this
resistance cannot be relied upon, since soil might settle during the service period. In regard to the
translational stiffness of the pile cap, the soil acting against the embedded portion of the cap may
provide accountable contribution. The sizes of pile caps for bridge foundations and abutments in recent
design may be much larger than those used before. The translational stiffness of the foundation used in
the analysis can lead to an incorrect result if the resistance from the pile cap is not included.
Lam, Martin, and Ibsen (1991) made recommendations for values of stiffness resulting from
base shear and passive resistance against the cap. Those values are based on standard soil properties
and can be used only for linear-elastic analysis. However, shear against the base could be negated due
to long-term settlement. Therefore, passive soil resistance against the pile cap is the only contribution to
be considered in this program.
The passive soil resistance is a highly-nonlinear function of the lateral movement of the pile cap;
thus, adding a constant value of stiffness to the whole foundation system is impractical. The soil-
structure interaction for the pile cap under translational movement is similar to the soil resistance (p-y
curves) on piles under lateral loading.
The approach adopted in the program is to derive the soil resistance for the pile cap using the
same p-y criteria for piles, but with the diameter equal to the width of the front side of the concrete cap
(Figure 8.1). The movements at the top and bottom of the pile cap are computed based on the
translation and rotation of the cap. The soil resistance generated by the program corresponds to the cap
movement for each iteration.
The failure pattern for soil in front of the pile cap is characterized principally as a wedge type
failure. The lateral resistance from the cohesion of the soil near the ground surface should not be
considered due to possible shrinkage in dry weather. Similar to the passive earth for cohesionless soil,
the resistance at the ground surface is assumed to be zero and to increase linearly to the base of the pile
cap (Figure 8.2). The program will add resistance on the front side of the pile cap to the global system
for force equilibrium and compatibility, which is achieved through iteration.
Figure 8.1 p-y curves for pile cap and individual pile
This new option was added to assist the user by providing a reasonable way for considering the
effect of an embedded pile cap. The user should exercise this option with care, especially for situations
where the soil in front of the pile cap may be removed due to erosion or other natural actions.
Figure 8.2 Distribution of soil resistance on the front side of pile cap
9.1 Introduction
The designer of deep foundations under lateral loading must make computations to ascertain that
three factors of performance are within tolerable limits: combined axial and bending stress, shear stress
and pile-dead deflection. The flexural rigidity, EI, of the deep foundation (bending stiffness) is an
important parameter that influences the computations. (Reese & Wang, 1988); (Isenhower, 1994)
In general, flexural rigidity of reinforced concrete varies nonlinearly with the level of applied
bending moment, and to employ a constant value of EI in the p-y analysis for a concrete pile will result in
some degree of inaccuracy in the computations.
The response of a pile is nonlinear with respect to load because the soil has nonlinear stress-strain
characteristics. Consequently, the load and resistance factor design (LRFD) method is recommended
when evaluating piles as structural members. This requires evaluation of the nominal (i.e. unfactored)
bending moment of the deep foundation.
The flexural behavior of a structural element such as a beam, column, or a pile subjected to
bending is dependent upon its flexural rigidity, EI, where E is the modulus of elasticity of the material of
which it is made and I is the moment of inertia of the cross section about the axis of bending. In some
instances the values of E and I remain constant for all ranges of stresses to which the member is
subjected, but there are situations where both E and I vary with changes in stress conditions because
the materials are nonlinear or crack.
Consider an element from a beam with an initial unloaded shape of abcd as shown by the
dashed lines in Figure 9.1. This beam is subjected to pure bending and the element changes in shape as
shown by the solid lines. The relative rotation of the sides of the element is given by the small angle d
and the radius of curvature of the elastic element is signified by the length measured from the center
of curvature to the neutral axis of the beam. The bending strain x in the beam is given by:
∆
𝜀𝑥 =
𝑑𝑥
(9.1)
where:
𝜌 𝜂
=
𝑑𝑥 ∆
(9.2)
where:
= radius of curvature.
∆ 𝜂𝑑𝑥 1 𝜂
𝜀𝑥 = = =
𝑑𝑥 𝜌 𝑑𝑥 𝜌
(9.3)
𝜎𝑥 = 𝐸 𝜀𝑥
(9.4)
where:
= Young’s modulus.
𝐸𝜂
𝜎𝑥 =
𝜌
(9.5)
𝑀𝜂
𝜎𝑥 =
𝐼
(9.6)
where:
𝑀𝜂 𝐸𝜂
=
𝐼 𝜌
(9.7)
𝑀 1
=
𝐸𝐼 𝜌
(9.8)
1 𝑑𝜃
= =𝜙
𝜌 𝑑𝑥
(9.9)
For convenience here, the symbol is substituted for the curvature 1/. The following equation
is developed from this substitution and Equations 9.8 and 9.9:
𝑀
𝐸𝐼 =
𝜙
(9.10)
And because = d and x = /dx, we may express the bending strain as:
𝜀𝑥 = 𝜙𝜂
(9.11)
The computation of the moment curvature for a given section follows the fiber method in which
the section is divided in a number of fibers. For longitudinal reinforcement and prestressing strands,
each bar is accounted individually. For a given value of curvature () , the position of the neutral axis is
estimated. The strain at points along the depth of the section can be computed by use of Equation
9.11, which in turn will lead to the forces in the concrete and/or steel. The stress-strain curves for
concrete and steel are those shown in the following section.
With the magnitude of the forces, both tension and compression, the equilibrium of the section
can be checked, taking into account the external compressive loading. If the section is not in
equilibrium, a revised position of the neutral axis is selected and iterations proceed until the neutral axis
is found.
Bending moment in the section is computed by integrating the moments of forces in the slices
times the distances of the slices from the centroid. The value of EI is computed using Equation 9.10. The
computations are repeated by incrementing the value of curvature until a failure strain in the concrete
or steel is developed.
9.3.1 Concrete
The program uses the strain stress curve for unconfined concrete as shown in Figure 9.2:
The following equations are used to compute the concrete stress. The value of concrete
compressive strength, f’c, in these equations is specified by the engineer.
𝜀 𝜀 2
𝑓𝑐 = 𝑓’𝑐 [2 − ( ) ] 𝑓𝑜𝑟 0 ≤ 𝜀 ≤ 𝜀0
𝜀0 𝜀0
(9.12)
𝜀 − 𝜀0
𝑓𝑐 = 𝑓’𝑐 [1 − 0.15 ( )] 𝑓𝑜𝑟 𝜀0 ≤ 𝜀 ≤ 0.0038
0.0038 − 𝜀0
(9.13)
The modulus of rupture, fr, is the tensile strength of concrete in bending. The modulus of
rupture for drilled shafts and bored piles is computed using:
𝑓𝑟 = 7.5√𝑓’𝑐 (psi)
𝑓𝑟 = 19.7√𝑓’𝑐 (kPa)
(9.14)
𝑓𝑟 = 4.0√𝑓’𝑐 (psi)
𝑓𝑟 = 10.5√𝑓’𝑐 (kPa)
(9.15)
𝐸𝑐 = 57000√𝑓’𝑐 (psi)
𝐸𝑐 = 151000√𝑓’𝑐 (kPa)
(9.16)
𝑓’𝑐
𝜀0 = 1.7
𝐸𝑐
(9.17)
𝑓𝑟
𝜀𝑡 = −𝜀0 (1 − √1 + )
𝑓’𝑐
(9.18)
The program uses the strain stress curve for confined concrete proposed by Mander et. Al.
(Mander, Priestley, & Park, 1988) as shown in Figure 9.3.
Figure 9.3 Stress-Strain relationship for confined concrete (Mander, Priestley, & Park, 1988).
𝑓’𝑐𝑐 𝑥𝑟
𝑓𝑐 =
𝑟 − 1 + 𝑥𝑟
(9.19)
𝜀
𝑥=
𝜀𝑐𝑐
(9.20)
𝐸𝑐
𝑟=
𝐸𝑐 − 𝐸𝑠𝑒𝑐
(9.21)
𝑓’𝑐𝑐
𝐸𝑠𝑒𝑐 =
𝜀𝑐𝑐
(9.22)
where f’CC and CC are the concrete stress and strain at peak stress.
7.94𝑓’𝑙 2𝑓’𝑙
𝑓’𝑐𝑐 = 𝑓’𝑐 (2.254√1 + − − 1.254)
𝑓’𝑐 𝑓’𝑐
(9.23)
𝑓’𝑐𝑐
𝜀𝑐𝑐 = 𝜀0 [1 + 5 ( − 1)]
𝑓’𝑐
(9.24)
For circular sections the effective lateral confining stress is estimated by:
1
𝑓’𝑙 = 𝑘𝑒 𝜌𝑠 𝑓𝑦ℎ
2
(9.25)
The ratio of the volume of transverse confining steel to the volume of concrete core, s, is
computed by:
4𝐴𝑠𝑝
𝜌𝑠 =
𝑑𝑠 𝑠
(9.26)
where Asp is the area of the transverse reinforcement bar, s is the center to center spacing or
pitch of spiral or circular hoop and ds is the diameter of spiral between bar centers.
𝑛
𝑠′
(1 − )
2𝑑𝑠
𝑘𝑒 =
1 − 𝜌𝑐𝑐
(9.27)
where n is equal to 1 for circular spirals and equal to 2 for circular hoops, s’ is the clear vertical
spacing between spiral or hoop bars and cc is the ratio of longitudinal reinforcement to area of core.
9.3.3 Steel
The program uses the strain stress curve for steel as shown in Figure 9.4:
The following equations are used to compute the steel stress. The value of yield strength, fy, in
these equations is specified by the engineer.
𝑓𝑠 = 𝐸𝑠 𝜀 𝑓𝑜𝑟 0 ≤ 𝜀 ≤ 𝜀𝑦
(9.28)
𝑓𝑠 = 𝑓𝑦 𝑓𝑜𝑟 𝜀𝑦 ≤ 𝜀
(9.29)
The program uses the strain stress curve proposed for Thompson et.al. (Thompson & Park,
1978) as shown in Figure 9.5.
Figure 9.5 Stress-Strain relationship for steel with hardening (Thompson & Park, 1978).
The following equations are used to compute the steel stress. The value of yield strength, fy, in
these equations is specified by the engineer.
𝑓𝑠 = 𝐸𝑠 𝜀 𝑓𝑜𝑟 0 ≤ 𝜀 ≤ 𝜀𝑦
(9.30)
𝑓𝑠 = 𝑓𝑦 𝑓𝑜𝑟 𝜀𝑦 ≤ 𝜀 ≤ 𝜀𝑠ℎ
(9.31)
𝑓𝑠𝑢
( ) (30𝑞 + 1)2 − 60𝑞 − 1
𝑓𝑦
𝑚=
15𝑞2
(9.33)
𝑞 = 𝜀𝑠𝑢 − 𝜀𝑠ℎ
(9.34)
The stress-strain relationships used in prestressed concrete is defined using the stress-strain
curves of concrete recommended by the Design Handbook of the Prestressed Concrete Institute
(PCI), as shown in Figure 9.6.
Figure 9.6 Stress-Strain relationship for prestressing strands recommended by PCI Design
Handbook, 5th Edition
0.04
𝑓𝑝𝑠 = 250 − (𝑘𝑠𝑖) 𝑓𝑜𝑟 0.0076 ≤ 𝜀
𝜀 − 0.0064
(9.36)
0.04
𝑓𝑝𝑠 = 270 − (𝑘𝑠𝑖) 𝑓𝑜𝑟 0.0086 ≤ 𝜀
𝜀 − 0.007
(9.38)
PCI does not have any recommendations for grade 300 strands, which are not widely
available. The above equations were used as a model to develop a stress-strain relationship for
grade 300 strands. The equations are
0.0835
𝑓𝑝𝑠 = 300 − (𝑘𝑠𝑖) 𝑓𝑜𝑟 0.0088846 ≤ 𝜀
𝜀 − 0.0071
(9.40)
For Cable Fiber Composite Cables (CFCC), the behavior is elastic, as presented in Figure 9.7
Figure 9.7 Load and elongation diagram for Carbon Fiber Composite Cables (CFCC) by Tokyo Rope
𝑓𝑝𝑠 = 𝐸𝑠 𝜀 𝑓𝑜𝑟 0 ≤ 𝜀 ≤ 𝜀𝑢
(9.41)
𝑓𝑢 is the ultimate stress of the strand that is estimated by capacity of the strand divided
by the area of the strand
If desired by the user, the program can estimate an Equivalent Elastoplastic Moment Curvature
following CALTRANS recommendations (CALTRANS, 2013).
Allen, J. (1985, May). p-y Curves in Layered Soils. Ph. D. Thesis, The University of Texas at Austin,
Department of Civil Engineering.
Asplund, S. O. (195). Generalized elastic theory for pile groups. Publications of the International
Association for Bridge and Structural Engineering, 16, pp. 1–22.
Association, J. (1976). Road bridge substructure design guide and explanatory notes. In Designing of Pile
Foundations.
Awoshika, K., & Reese, L. C. (1971, February). Analysis of foundation with widely-spaced batter piles.
Center for Highway Research at The University of Texas at Austin.
Baguelin, F., Jezequel, J. F., & Shields, D. H. (1978). The Pressuremeter and Foundation Engineering.
Trans Tech Publications.
Bhushan, K., Lee, L. J., & Grime, D. B. (1981). Lateral load test on drilled piers in sand. ASCE Annual
Meeting.
Bogard, D., & Matlock, H. (1983). Procedures for analysis of laterally loaded pile groups in soft clay.
Specialty Conference of Geotechnical Engineering in Offshore Practice, (pp. 499–535).
Brown, D. A. (2002). Specifying initial k for stiff clay with no free water. Personal Communication.
Brown, D. A., & Reese, L. C. (1985, May). Behavior of a large-scale pile group subjected to cyclic lateral
loading. Austin, Texas: Minerals Management Service.
Brown, D. A., Reese, L., & O’Neill, M. W. (1987, November). Cyclic lateral loading of a large-scale pile
group. Journal of Geotechnical Engineering , 113(11), 326–1343.
Bryant, L. M., & Matlock, H. (1977, December). Three-dimensional analysis of framed structures with
nonlinear pile foundations. The University of Texas at Austin.
Cox, W. R., Dixon, D. A., & Murphy, B. S. (1984). Lateral load tests of 25.4 mm diameter piles in very soft
clay in side-by-side and in-line groups. Laterally Loaded Deep Foundations: Analysis and
Performance. American Society for Testing and Materials.
Cox, W., Reese, L. C., & Grubbs, B. (1974). Field testing of laterally-loaded piles in sand. Offshore
Technology Conference.
Coyle, H. M., & Reese, L. C. (1966, March). Load transfer for axially loaded piles in clay. Journal of the Soil
Mechanics and Foundations Division , 92, 1–26.
Coyle, H. M., & Sulaiman, I. H. (1967, November). Skin friction for steel piles in sand. Journal of the Soil
Mechanics and Foundations Division, 93, 261.
D’Appolonia, E., & Romualdi, J. P. (1963, March). Load transfer in end bearing steel h-piles. Journal of
the Soil Mechanics and Foundations Division , 89, 1–25.
Deere, D. V. (1968). Geological considerations. In K. Stagg, & O. Zienkiewicz (Eds.), Rock Mechanics in
Engineering Practice (pp. 1–20). New York: Wiley.
Dunnavant, T. W., & O’Neill, M. W. (1986). Evaluation of design-oriented methods for analysis of vertical
pile groups subjected to lateral load. Numerical Methods in Offshore Piling (pp. 303–316. ).
Institut Francais du Petrole.
Evans, L. T., & Duncan, J. M. (1982). Simplified analysis of laterally loaded piles. University of California at
Berkeley.
Focht, J. A., & Koch, K. J. (1973). Rational analysis of the lateral performance of offshore pile groups.
Fifth Offshore Technology Conference, 2, pp. 701–708. Houston.
Francis, A. J. (1964, May). Analysis of pile groups with flexural resistance. Journal of the Soil Mechanics
and Foundations Division(3887), 1–32.
Franke, K., & Rollins, K. (2013). Simplified Hybrid p-y Spring Model for Liquefied Soils. Journal of
Geotechnical and Geoenvironmental Engineering, 139, 2013.
Georgiadis, M. (1983, April). Development of p-y curves for layered soils. Geotechnical Practice in
Offshore Engineering, 536–545.
Hansbo, S. (1993). Interaction problems related to the installation of pile groups. In V. Impe (Ed.), Deep
Foundations on Bored and Auger Pies. Balkema.
Hansen, B. (1959, September). Limit design of pile foundations. Bygningsstatiske Meddelelser Argang,
30(2).
Hansen, J. B. (1968). A theory for skin friction on piles. Danish Geotechnical Institute.
Hoek, E. (1990). Estimating Mohr-Coulomb Friction and Cohesion Values from the Hoek-Brown Failure
Criterion. International Journal of Rock Mechanics, Mining Sciences, and Geomechanics
Abstracts, 27(3), 227-229.
Hoek, E. (2001). Rock Mass Properties for Underground Mines. In W. Hustrulid, & R. Bullock (Ed.),
Underground Mining Methods: Engineering Fundamentals and International Case Studies.
Litleton: Society for Mining, Metallurgy, and Exploration.
Hoek, E., & Brown, E. (1980). Empirical Strength Criterion for Rock Masses. Journal of Geotechnical
Engineering, 106, 1013-1035.
Horvath, R. G., & Kenney, T. C. (1979). Shaft resistance of rock-socketed drilled piers. Symposium on
Deep Foundations, (pp. 182–214). Atlanta.
Hrennikoff, A. (1950). Analysis of pile foundations with batter piles. Transactions of the American Society
of Civil Engineers, 115(2401).
Ibrahimbegovic, A. (1993). Quadrilateral Elements for Analysis of Thick and Thin Plates. Computer
Methods in Applied Mechanics and Engineering, 110, 195-209.
Ibrahimbegovic, A., Taylor, R., & Wilson, E. (1990). A robust quadrilateral membrane finite element with
drilling degrees of freedom. International Journal for Numerical Methods in Engineering, 30,
445–457.
Isenhower, W. (1994). Improved Methods for Evaluation of Bending Stiffness of Deep Foundations. Intl.
Conf. on Design and Construction of Bridge Foundations, (pp. 571-575).
Johnson, R., Parsons, R., Dapp, S., & Brown, D. (2006). Soil Characterization and p-y Curve Development
for Loess. Kansas Department of Transportation, Bureau of Materials and Research.
Karol, R. H. (1960). Soils and Soil Engineering. . Englewood Cliffs, New Jersey: Prentice Hall.
Kraft, L. M., Focht, J. A., & Amerasinghe, S. F. (1981, November). Friction capacity of piles driven in to
clay. Journal of the Geotechnical Engineering Division, 107(11), 1521–1541.
Kubo, K. (1964). Experimental Study of the Behavior of Laterally Loaded Piles. Japan:
TransportationTechnology Research Institute.
Lam, I. P., Martin, G. R., & Imbsen, R. (1991, March). Modeling bridge foundations for seismic design and
retrofitting. Third Bridge Engineering Conference. Denver, Colorado.
Liang, R., Yang, K., & Nusairat, J. (2009). p-y Criterion for Rock Mass. Journal of Geotechnical and
Geoenvironmental Engineering, 135(1), 26-36.
Lieng, J. T. (1988). Behavior of Laterally Loaded Piles in Sand, Large Scale Model Tests. Norwegian
Institute of Technology, Department of Civil Engineering.
Liu, J. L., Yuan, Z. L., & Shang, K. P. (1985). Cap-pile-soil interaction of bored pile groups. 11th
A.C.S.M.F.E., 3, pp. 1433–1436. San Francisco.
Long, J. H., & Reese, L. C. (1983). An investigation of the behavior of vertical piles in cohesive soils
subjected to repetitive lateral loads. U.S. Army Corp of Engineers.
Mander, J., Priestley, M., & Park, R. (1988). Theoretical Stress-Strain Model for Confined Concrete. ASCE
Journal of Structural Engineering, 1804-1826.
Marinos, P., & Hoek, E. (2000). GSI – A Geologically Friendly Tool for Rock Mass Strength Estimation.
Proceedings, GeoEngineering 2000 Conference, (pp. 1,422-1,442). Melbourne.
Matlock, H. (1970). Correlations for design of laterally-loaded piles in soft clay. Second Annual Offshore
Technology Conference, 1, pp. 577–594. Houston, Texas.
Matlock, H., & Haliburton, A. T. (1966, September). Finite-element method of solution for linearly elastic
beam columns. Center for Highway Research at The University of Texas at Austin.
Matlock, H., & Ingram, W. B. (1963, July). Bending and buckling of soil supported structural elements.
Second Pan American Conference on Soil Mechanics and Foundation Engineering. Brazil.
Matlock, H., & Reese, L. C. (1962). Generalized solution for laterally loaded piles. Transactions of the
American Society of Civil Engineers, 127, 1220–1251.
Matlock, H., Ingram, W. B., Kelly, A. E., & Bogard, D. (1980). Field tests of latera lload behavior of pile
groups in soft clay. Twelfth Annual Offshore Technology Conference, (pp. 163–174). Houston,
Texas.
Mattes, N. S., & Poulos, H. G. (1969, January). Settlement of single compressible pile. Journal of the Soil
Mechanics and Foundations Division , 189–207.
Mello, V. F. (1969). Foundations of buildings on clay. Proceedings of the 7th International Congress on
Soil Mechanics and Foundation Engineering, (pp. 49–136).
Meyerhof, G. G. (1974, May). Ultimate bearing capacity of footing on sand layer overlaying clay. CGJ,
11(2).
Morrison, C. S., & Reese, L. C. (1988). A lateral-load test of a full-scale pile group in sand. Vicksburg,
Mississippi: U.S. Army Engineer Waterways Experiment Station.
Mosher, R. L. (1984). Load transfer criteria for numerical analysis of axially loaded piles in sand. U.S.
Army Corp of Engineers.
O’Neill, M. W. (1983, April). Group action in offshore piles. Specialty Conference on Geotechnical
Engineering in Offshore Practice. American Society of Civil Engineers.
O’Neill, M. W., & Murchison, J. M. (1983). An evaluation of p-y relationships in sands. American
Petroleum Institute.
O’Neill, M. W., Townsend, F. C., Hassan, K. H., Buller, A., & Chan, P. S. (1996, November). Load transfer
for drilled shafts in intermediate geomaterials. FHWA.
Parker, F. J., & Cox, W. R. (1969). A method of the analysis of pile supported foundations considering
nonlinear soil behavior. Center for Highway Research at The University of Texas at Austin.
Parker, F. J., & Reese, L. C. (1971, July). Lateral pile-soil interaction curves for sand. The International
Symposium on the Engineering Properties of Sea-Floor Soils and Their Geophysical Identification.
Seattle, Washington: The University of Washington.
Peck, R. B. (1976). Rock foundations for structures. Specialty Conference on Rock Engineering or
Foundations and Slopes, (pp. 1–21. ). Boulder, Colorado.
Poulos, H. G. (1968). Analysis of the settlement of pile groups. Geotechnique, 18, 449–471.
Poulos, H. G., & Davis, E. H. (1968, September). The settlement behavior of single axially loaded
incompressible piles and piers. Geotechnique , 18(3), 351–371.
Poulos, H. G., & Davis, E. H. (1980). Pile Foundation Analysis and Design. New York: Wiley.
Poulos, H. G., & Mattes, N. S. (1969, June). The behavior of axially loaded end bearing piles.
Geotechnique, 19(7), 285–300.
Prakash, S. (1962). Behavior of Pile Groups Subjected to Lateral Load. University of Illinois.
Radosavljevic, Z. (1957). Calcul et essais des pieux en groupe. Fourth International Conference on Soil
Mechanics and Foundation Engineering, 2, pp. 56-60. London.
Reese, L. C. (1966). Analysis of a bridge foundation supported by batter piles. Fourth Annual Engineering
and Geology and Soils Engineering Symposium, (p. 61). Moscow, Idaho.
Reese, L. C. (1971, July). The analysis of piles under lateral loading. Symposium on the Interaction of
Structure and Foundation (pp. 206–218). Birmingham, England : University of Birmingham.
Reese, L. C. (1983). Behavior of piles and pile groups under lateral load. Federal Highway Administration.
Reese, L. C. (1997, November). Analysis of piles in weak rock. Journal of the Geotechnical and
Geoenvironmental Engineering Division.
Reese, L. C., & Matlock, H. (1960). Numerical analysis of laterally loaded piles. Second Structural Division
Conference on Electronic Computation (p. 657). Pittsburgh, Pennsylvania: American Society of
Civil Engineers.
Reese, L. C., & Matlock, H. (1966, February). Behavior of a two-dimensional pile group under inclined
and eccentric loading. Offshore Exploration Conference. Long Beach, California.
Reese, L. C., & Nyman, K. J. (1978, February). Field load tests of instrumented drilled shafts at Isla
Morada, Forida. Girdler Foundation and Exploration Corporation.
Reese, L. C., & O’Neill, M. W. (1967, September). The analysis of three-dimensional pile foundations
subjected to inclined and eccentric loads. Conference on Civil Engineering (pp. 245–276).
American Society of Civil Engineers.
Reese, L. C., & O’Neill, M. W. (1987). Drilled shafts: Construction procedures and design methods . U.S.
Department of Transportation.
Reese, L. C., & Welch, R. C. (1975). Lateral loading of deep foundations in stiff clay. Journal of the
Geotechnical Division, 101(7), 633–649.
Reese, L. C., Cox, W. R., & Koop, F. (1975). Field testing and analysis of laterally loaded piles in stiff clay.
Seventh Annual Offshore Technology Conference. Houston, Texas.
Reese, L. C., Cox, W. R., & Koop, F. D. (1974). Analysis of laterally loaded piles in sand. Fifth Annual
Offshore Technology Conference. Houston, Texas.
Reese, L. C., O’Neill, M. W., & Smith, R. E. (1970, January). Generalized analysis of pile foundations.
Journal of the Soil Mechanics and Foundations Division, 96(SM1), 235–250.
Reese, L., & Wang, S. (1988). The Effect of Nonlinear Flexural Rigidity on the Behavior of Concrete Piles
Under Lateral Loading. Texas Civil Engineer, 17-23.
Robertson, R. N. (1961). The analysis of a bridge foundation with batter piles. Master’s thesis, The
University of Texas at Austin.
Rollins, K. M., Gerber, T. M., Lane, J. D., & Ashford, S. A. (2005). Lateral resistance of a full-scale pile
group in liquefied sand. Journal of the Geotechnical and Geoenvironmental Engineering Division,
131, 115-125.
Rowe, R., & Armitage, H. (1987). A Design Method for Drilled Piers in Soft Rock. Canadian Geotechnical
Journal, 24, 126-142.
Saul, W. E. (1968, May). Static and dynamic analysis of pile foundations. Journal of the Structural Division
, ST5(5936), 1077–1100. .
Schmertman, J. H. (1977). Report on development of keys limestone shear test for drilled shaft design.
Clearwater, Florida: Girdler Foundation and Exploration Corporation.
Schmidt, H. G. (1981). Group action of laterally loaded bored piles. Tenth International Conference, Soil
Mechanics and Foundation Engineering, (pp. 833–837. ). Stockholm.
Schmidt, H. G. (1985). Horizontal load tests on files of large diameter bored piles. Eleventh International
Conference on Soil Mechanics and Foundation Engineering, (pp. 569–1573). San Francisco.
Seed, H. B., & Reese, L. C. (1957). The action of soft clay along friction piles. Transactions of the
American Society of Civil Engineers, 122, 731–754.
Seed, R., & Harder, L. (1990). SPT-Based Analysis of Cyclic Pore Pressure Generation and Undrained
Residual Strength. H. Bolton Seed, Memorial Symposium. 2, pp. 351-376. BiTech Publishers Ltd.
Skempton, A. W. (1951). The bearing capacity of clays. Building Research Congress. London.
Speer, D. (1992). Shaft lateral load test terminal separation. California Department of Transportation
(unpublished).
Stoll, U. W. (1972). Torque shear test on cylindrical friction piles. Civil Engineering, 42 (4), 63–64.
Szilard, R. (1973). Theory and analysis of Plates, Classical and Numerical Methods. New Jersey: Prentice-
Hall, Inc., Englewood Cliffs.
Terzaghi, K. (1956). In Theoretical Soil Mechanics (pp. 363–366). New York: Wiley.
Thompson, K., & Park, R. (1978). Stress-Strain Model for Grade 275 Reinforcing Steel with Cyclic Loading.
Bulletin of the New Zealand National Society for Earthquake Engineering, 101-109.
Thurman, A. G., & D’Appolonia, E. (1965). Computed movement of friction and end bearing piles
embedded in uniform and stratified soils. Sixth International Conference on Soil Mechanics and
Foundation Engineering, 2, pp. 323–327.
Timoshenko, S. P., & Gere, J. (1961). Theory of Elastic Stability (Second ed.). McGraw-Hill.
Turzynski, L. D. (1960). Groups of piles under mono-planar forces. Structural Engineer, 38(9), 286.
Vesic, A. S. (1969). Experiments with instrumented pile groups in sand. Performance of Deep
Foundations, (pp. 172–222).
Wang, S.-T. (1986, August). Analysis of Drilled Shafts Employed in Earth-Retaining Structures. The
University of Texas at Austin.
Wang, S.-T., & Reese, L. C. (1998). Design of pile foundations in liquefied soils. Geotechnical Earthquake
Engineering and Soil Dynamics III, (pp. 1331–1343).
Weaver, T. J. (2001). Behavior of Liquefying Sand and CISS PIles During Full-Scale Lateral Load Tests.
Dissertation, Doctor of Philosophy, The University of California at San Diego, Department of
Structural Engineering.
Welch, R. C., & Reese, L. C. (1972, May). Laterally loaded behavior of drilled shafts. Texas Highway
Department and U.S. Department of Transportation.