You are on page 1of 109

Axial load capacity of drilled shafts in rock 215

In LRFD, the ultimate (factored) axial capacity of a drilled shaft can be calculated
using the expression for reinforced concrete columns:

(6.1)

where is the capacity reduction (resistance) factor=0.75 for spiral columns and 0.70
for horizontally tied columns (ACI, 1995); Qu is the nominal (computed) structural
capacity; β is the eccentricity factor=0.85 for spiral columns and 0.80 for tied columns;
is the specified minimum concrete strength; Ac is the cross-sectional area of the
concrete; fy is the yield strength of the longitudinal reinforcing steel; and As is the cross-
sectional area of the longitudinal reinforcing steel.
The Standard Specifications for Highway Bridges adopted by the American
Association of State Highway and Transportation Officials (AASHTO, 1989) stipulates a
minimum shaft diameter of 18 inches, with shaft sizing in 6-inch increments. Where the
potential for lateral loading is not significant, drilled shafts need to be reinforced for axial
loads only. The design of longitudinal and spiral reinforcement should conform to the
requirements of reinforced compression members.
Table 6.1 Allowable concrete stresses for drilled
shafts (after ASCE, 1993).
Uniform axial compression
Confined 0.33f′c
Unconfined 0.27f′c
Uniform axial tension 0
Bending (extreme fiber)
Compression 0.40f′c
Tension 0
Note: f′c is the specified minimum concrete strength.

6.3 CAPACITY OF DRILLED SHAFTS RELATED TO ROCK

Assuming that the shaft itself is strong enough, its load capacity depends on the capacity
of the rock to accept without distress the loads transmitted from the shaft. The required
area of shaft-rock interface (i.e., the size of drilled shaft) depends on this factor. The
ultimate axial load of a drilled shaft related to rock, Qu, consists of the ultimate side shear
load, Qus, and the ultimate end bearing load, Qub (see Fig. 6.1):
Qu=Qus+Qub
(6.2)
Drilled shafts in rock 216

The ultimate side shear load and the ultimate end bearing load are respectively calculated
as the average side shear resistance multiplied by the shaft side surface area and as the
end bearing resistance multiplied by the shaft bottom area, i.e.

Fig. 6.1 Axially loaded drilled shaft.


Qus=πBLτmax
(6.3)

(6.4)

where L and B are respectively the length and diameter of the shaft; and τmax and qmax are
respectively the average side shear resistance and the end bearing resistance.
The ultimate side shear resistance and the end bearing resistance are usually
determined based on local experience and building codes, empirical relations, or field
load tests. Methods based on local experience and building codes and empirical relations
are discussed in this chapter. The methods for conducting field load tests and
interpretation of test results will be discussed in Chapter 12.

6.3.1 Side shear resistance


The shear resistance mobilized at the shaft-rock interface is affected by many factors.
These include the shaft roughness, strength and deformation properties of the concrete
Axial load capacity of drilled shafts in rock 217

and the rock mass, geometry of the shaft, and initial stresses in the ground. The effect of
shaft roughness is emphasized by most investigators and considered in a number of
empirical relations for estimating the side shear resistance.

(a) Correlation with SPT N value


Standard Penetration Tests (SPT) are often carried out in weak or weathered rock. Table
6.2 shows the measured side shear resistances of drilled shafts and their corresponding
SPT N values in weathered sedimentary rocks. It can be seen that the τmax/N ratio is
generally smaller than 2.0 except the case reported by Toh et al. (1989). We can also see
that the τmax/N ratio tends to decrease as N increases.
Table 6.2 Side shear resistance and SPT N values in
weathered sedimentary rock.
Rock SPT N values τmax τmax/N Reference
(blows/0.3 m) (kPa) (kPa)
Highly weathered siltstone 230 >195– >0.87– Buttling (1986)
226 1.0
Highly weathered siltstone, silty 100–180 100– 1.0–1.8 Chang and Wong
sandstone and shale 320 (1987)
Very dense clayey/sandy silt to 110–127 80–125 0.63– Buttling and Lam
highly weathered siltstone 1.14 (1988)
Highly to moderately weathered 200–375 340 0.9–1.7
siltstone
Completely to partly weathered 100–150 – 1.2–3.7 Toh et al. (1989)
interbedded sandstone, siltstone and 150–200 – 0.6–2.3
shale/mudstone
Highly to moderately fragmented 400–1000 300– 0.5–0.8 Radhakrishnan and
siltstone/shale 800 Leung (1989)
Highly weathered sandy shale 150–200 120– 0.8–0.7 Moh et al. (1993)
140
Slightly weathered sandy shale and 375–430 240– ave. 0.65
sandstone 280

(b) Empirical relations between side shear resistance and unconfined


compressive strength of intact rock
Empirical relations between the side shear resistance and the unconfined compressive
strength of rock have been proposed by many researchers. The form of these empirical
relations can be generalized as
τmax= ασcβ
(6.5)
Drilled shafts in rock 218

where τmax is the side shear resistance; σc is the unconfined compressive strength of the
intact rock (if the intact rock is stronger than the shaft concrete, σc of the concrete is
used); and α and β are empirical factors.
The empirical factors proposed by a number of researchers have been summarized by
O’Neill et al. (1996) and are shown in Table 6.3. Most of these empirical relations were
developed for specific and limited data sets, which may have correlated well with the
proposed equations. However, O’Neill et al. (1996) compared the first nine empirical
relations listed in Table 6.3 with an international database of 137 pile load tests in
intermediate-strength rock and concluded that none of the methods could be considered a
satisfactory predictor for the database.
Kulhawy and Phoon (1993) developed a relatively extensive load test database for
drilled shafts in soil and rock and presented their data both for individual shaft load tests
and as site-averaged data. The results are shown in Figures 6.2 and 6.3, in terms of
adhesion factor, σc, versus normalized shear strength, cu/pa or σc/2pa (assuming cu≈ σc/2),
where pa is atmospheric pressure (≈0.1 MPa). It should be noted that Kulhawy and
Table 6.3 Empirical factors a and β for side shear
resistance (modified from O’Neill et al., 1996).
Design method α β
Horvath and Kenney (1979) 0.21 0.50
Carter and Kulhawy (1988) 0.20 0.50
Williams et al. (1980) 0.44 0.36
Rowe and Armitage (1984) 0.40 0.57
Rosenberg and Journeaux (1976) 0.34 0.51
Reynolds and Kaderbek (1980) 0.30 1.00
Gupton and Logan (1984) 0.20 1.00
Reese and O’Neill (1987) 0.15 1.00
Toh et al. (1989) 0.25 1.00
Meigh and Wolshi (1979) 0.22 0.60
Horvath (1982) 0.20–0.30 0.50

Phoon (1993) defined αc as the ratio of the side shear resistance τmax to the undrained
shear strength cu. Understandably, the results of individual load tests show considerably
greater scatter than the site-averaged data. On the basis of the site-averaged data,
Kulhawy and Phoon (1993) proposed the following relations for drilled shafts in rock:

(6.6a)

(6.6b)
Axial load capacity of drilled shafts in rock 219

(6.6c)

Equation (6.6) can be rewritten in a general form as

(6.7)

This leads to a general expression for the side shear resistance


τmax=Ψ[paσc/2pa]−0.5
(6.8)

It is very important to note that the empirical relations given in Equations (6.6b) and
(6.6c) are bounds to site-averaged data, and do not necessarily represent bounds to
individual shaft behavior. The coefficient of determination (r2) is approximately 0.71 for
the site-averaged data, but is only 0.46 for the individual data, reflecting the much greater
variability of the individual test results (Seidel & Haberfield, 1995).

Fig. 6.2 Adhesion factor


αc(=τmax/0.5σc) versus normalized
shear strength for site-averaged data
(after Kulhawy & Phoon, 1993).
Drilled shafts in rock 220

Fig. 6.3 Adhesion factor


αc(=τmax/0.5σc) versus normalized
shear strength for individual test data
(after Kulhawy & Phoon, 1993).

(c) Empirical relations considering roughness of shaft wall


The roughness of the shaft wall is an important factor controlling the development of side
shear resistance. Depending on the type of drilling technique and the hardness of the
rock, a drilled shaft will have a certain degree of roughness. Research has shown that the
benefits gained from increasing the roughness of a shaft wall can be quite significant,
both in terms of peak and residual shear resistance. Studies by Williams et al. (1980) and
others showed that smooth-sided shafts exhibit a brittle type of failure, while shafts
having an adequate roughness exhibit ductile failure. Williams and Pells (1981)
suggested that rough shafts generate a locked-in normal stress such that there is
practically no distinguishing difference between peak and residual side shear resistance.
Classifications have been developed so that roughness can be quantified. One such
classification proposed by Pells et al. (1980) is based on the size and frequency of
grooves in the shaft wall (see Table 6.4). Based on this classification, Rowe and
Armitage (1987b) proposed the following relation for shafts with different roughness:
τmax=0.45(σc)0.5 for shafts with roughness R1, R2 or R3
(6.9a)
τmax=0.60(σc)0.5 for shafts with roughness R4
(6.9b)

where both τmax and σc are in MPa.


Axial load capacity of drilled shafts in rock 221

Horvath et al. (1980) also developed a relation from model shaft behavior using
various roughness profiles. They found that as shaft profiles go from smooth to rough, the
roughness factor increases significantly, as does the peak side shear resistance. These
findings were confirmed in a later study by Horvath et al. (1983), and the following
equation was proposed for the roughness factor (RF):

(6.10)

where hm is the average roughness (asperity) height of the shaft; Lt is the total travel
length along the shaft wall profile; R is the nominal radius of the shaft; and L is the
nominal length of the shaft (see Fig. 6.4). Using Equation (6.10), the following relation
was developed between the side shear resistance and RF:
τmax= 0.8σc(RF)0.45
(6.11)

Kodikara et al. (1992) developed a rational model for predicting the relationship of τmax to
σc based on a specific definition of interface roughness, initial normal stress on the
interface and the stiffness of the rock during interface dilation. The parameters needed to
define interface roughness in the model are also shown in Figure 6.4. The model accounts
for variability in asperity height and angularity, assuming clean, triangular interface
discontinuities. Figure 6.5 shows the predicted adhesion factor, α(=τmax/σc), for
Melbourne Mudstone with the range of parameters and roughnesses as given in Table
6.5. The adhesion factor is presented as a function of Em/σc, σc/σn and the degree of
roughness, where Em is the elastic modulus of the rock mass and σn is the initial normal
stress on the shaft-rock interface. It can be seen that the adhesion factor is affected not
only by the interface roughness, but also by Em/σc and σc/σn.

Table 6.4 Roughness classes after Pells et al.


(1980).
Roughness Description
Class
R1 Straight, smooth-sided shaft, grooves or indentation less than 1.00 mm deep
R2 Grooves of depth 1–4 mm, width greater than 2 mm, at spacing 50 to 200 mm.
R3 Grooves of depth 4–10 mm, width greater than 5 mm, at spacing 50 to 200 mm.
R4 Grooves or undulations of depth greater than 10 mm, width greater than 10 mm,
at spacing 50 to 200 mm.
Drilled shafts in rock 222

Fig. 6.4 Parameters for defining shaft


wall roughness (after Horvath et al.,
1980 and Kodikara et al., 1992).
Axial load capacity of drilled shafts in rock 223

Fig. 6.5 Simplified design charts for


adhesion factor α(=τmax/σc) for
Melbourne Mudstone (after Kodikara
et al., 1992).
Drilled shafts in rock 224

Table 6.5 Definition of borehole roughness and


range of parameters for Melbourne Mudstone (after
Kodikara et al., 1992).
Range of values for shafts in Melbourne Mudstone
Parameter Smooth Medium Rough
im(degrees) 10–12 12–17 17–30
isd(degrees) 2–4 4–6 6–8
hm(mm) 1–4 4–20 20–80
hsd/hm 0.35
B(m) 0.5–2.0
σc(MPa) 0.5–10.0
σn(MPa) 50–500
Em(MPa) 50–500
Notes: 1) Refer to Figure 6.4 for the definitions of im, isd, hm and hsi
2) B=diameter of the shaft.
3) σc=unconfined compressive strength of the intact rock.
4) σn=initial normal stress on the shaft-rock interface.
5) Em=deformation modulus of the rock mass.

Seidel and Collingwood (2001) introduced a nondimensional factor called Shaft


Resistance Coefficient (SRC) to reflect the influence of shaft roughness and other factors
on the shaft side shear resistance. The SRC is defined as follows:

(6.12)

where hm is the mean roughness height (either assessed directly by estimation or


measurement, or computed as the product of asperity length, la, and the sine of the mean
asperity angle); B is the shaft diameter; ηc is the construction method reduction factor as
shown in Table 6.6; n is the ratio of rock mass modulus to the unconfined compressive
strength of the rock (Em/σc), known as the modulus ratio; and ν is the Poisson’s ratio of
the rock.
Using SRC, Seidel and Collingwood (2001) have created shaft resistance charts as
shown in Figures 6.6 and 6.7. These charts are based on results of a parametric study
using a computer program called ROCKET. To develop these charts, the intact rock
strength parameters were related to the unconfined compressive strength using the Hoek-
Brown strength criteria described in Chapter 4. Mohr-Coulomb strength parameters
adopted in the analyses were determined after the method of Hoek (1990) using the
unconfined compressive strength of the rock and appropriate values of parameters s and
m.
Axial load capacity of drilled shafts in rock 225

(d) Estimation of roughness height of shaft wall


Application of the empirical relations considering shaft wall roughness in design requires
estimation of likely shaft wall roughness height. A small number of studies have
produced actual roughness profiles which enable quantitative analysis. Detailed studies
have been carried out into shafts in Melbourne Mudstone (Williams, 1980; Holden, 1984;
Kodikara et al., 1992; Baycan, 1996). The results show that shaft wall roughness in this
low- to medium-strength argillaceous rock can vary considerably and appears to be
influenced by rock discontinuities, drilling techniques, and rate of advance. Shaft wall
roughness profiles in medium-strength shale were also recorded by Horvath et al. (1983),
but most of their shafts were artificially roughened by grooving. O’Neill & Hassan
(1994) and O’Neill et al., (1996) recorded measurements of roughness profiles of shafts
in clay shale, argillite and sandstone.
Table 6.6 Indicative construction method reduction
factor ηc (after Seidel & Collingwood, 2001).
Construction method ηc
Construction without drilling fluid
Best construction practice and high level of construction control 1.0
(e.g., shaft sidewalls free of smear and remoulded rock)
Poor construction practice or low-quality construction control (e.g., 0.3–0.9
smear or remoulded rock present on shaft sidewalls)
Construction under bentonite slurry
Best construction practice and high level of construction control 0.7–0.9
Poor construction practice or low-quality construction control 0.3–0.6
Construction under polymer slurry
Best construction practice and high level of construction control 0.9–1.0
Poor construction practice or low-quality construction control 0.8
Drilled shafts in rock 226

Fig. 6.6 Adhesion factor α(=τmax/σc)


versus σc (after Seidel & Collingwood,
2001).

Fig. 6.7 Adhesion factor α(=τmax/σc)


versus SRC (after Seidel &
Collingwood, 2001).
Axial load capacity of drilled shafts in rock 227

Based on roughness heights back-calculated from load tests on shafts in rock, Seidel and
Collingwood (2001) developed the effective roughness height versus the unconfined
compressive strength plot as shown in Figure 6.8. The back-calculations were conducted
using Equation (6.12) and assuming ηc=1.0. In the case of a shaft for which the concrete-
rock interface is clean and unbounded, the roughness height back-calculated assuming
ηc=1.0 should provide a reasonable estimate of the roughness height magnitude.
However, if the shaft resistance is adversely influenced by construction procedures, the
roughness height would be underestimated if ηc is assumed to be 1.

Example 6.1
A drilled shaft of diameter 1.0 m is to be socketed 3.0 meters in rock. The rock properties
are as follows:
Unconfined compressive strength of intact rock, σc=15.0 MPa
Deformation modulus of intact rock, Er=10.6 GPa
RQD=76

Determine the side shear resistance.

Fig. 6.8 Effective roughness height


versus σc (after Seidel & Collingwood,
2001).
Drilled shafts in rock 228

Solution:
Method of Kulhawy and Phoon (1993)—Equations (6.6) to (6.8)
Lower bound
τmax=1.0[paσc/2]0.5=1.0[0.1×15.0/2]0.5=0.87 MPa

Upper bound
τmax=3.0[paσc/2]0.5=3.0[0.1×15.0/2]0.5=2.60 MPa

Method of Seidel and Collingwood (2001)


From Figure 6.8, the mean roughness height hm=1.64 mm (lower bound) and 6.19 mm
(upper bound).
Using Equation (4.24), the rock mass modulus:
αE=0.0231(RQD)−1.32=0.297
Em=αEEr=0.297×10.6=3.15 GPa

The modulus ratio n=Em/σc=210.


The Poisson’s ratio of the rock is simply assumed to be ν=0.25.
Using ηc=1.0, SRC can be obtained from Equation (6.12) as:

From Figure 6.6, the adhesion factor a can be obtained as


α=0.102 (lower bound)
α=0.225 (upper bound)

So the side shear resistance can be obtained as


τmax=ασc=0.102×15.0=1.53 MPa (lower bound)
τmax=ασc=0.225×15.0=3.37 MPa (upper bound)

The results show that the shaft wall roughness (reflected by the roughness height) has
a great effect on the side shear resistance.

(e) Factors affecting side shear resistance


As stated above, the shaft wall roughness, which is an important factor controlling the
development of side shear resistance, has been studied extensively. Other factors such as
the discontinuities in the rock mass and the shaft geometry have also been studied by
some researchers. Williams et al. (1980) suggested that the existence of discontinuities in
Axial load capacity of drilled shafts in rock 229

the rock mass reduces the side shear resistance by reducing the normal stiffness of the
rock mass. They developed the following empirical relation that considers the effect of
discontinuities on the side shear resistance:
τmax=αwβwσc
(6.13)

where αw is a reduction factor reflecting the strength of the rock, as shown in Figure 6.9;
and βw is the ratio of side shear resistance of jointed rock mass to side shear resistance of
intact rock. βw is a function of modulus reduction factor, j, as shown in Figure 6.10, in
which
βw=f(j), j=Em/Er
(6.14)

where Em is the elastic modulus of the rock mass; and Er is the elastic modulus of the
intact rock. When the rock mass is such that the discontinuities are tightly closed and
seatns are infrequent, βw is essentially equal to 1.0. Comparing Equation (6.13) with
Equation (6.5), it can be seen that αwβw is just the adhesion factor, a, for β=1. Since αw is
derived from field test data, the effect of discontinuities is already included in αw. If αw is
multiplied by βw which is obtained from laboratory tests (Williams et al., 1980), the effect
of discontinuities will be considered twice. So Equation (6.13) may be too conservative.
Pabon and Nelson (1993) studied the effect of soft horizontal seams on the behavior of
laboratory model shafts. The study included four instrumented model shafts in
manufactured rock, three of which have soft seams. They concluded that a soft seam
significantly reduces the normal interface stresses generated in the rock layer overlying it.
Consequently the side shear resistance of shafts in rock with soft seams is much lower
than that of shafts in intact rock.

The effect of shaft geometry on side shear resistance was studied by Williams and Pells
(1981). They tested 15 shafts in Melbourne Mudstone, with diameters ranging from 335
mm to 1580 mm, and 27 shafts in Hawkesbury Sandstone, with diameters ranging from
64 mm to 710 mm. The results of these tests indicated that the shaft length, L, does not
have a discernible effect on the side shear resistance. They argued that the interface
dilation creates a locked-in normal stress with the result that the shear displacement
behavior exhibits virtually no peak or residual behavior. They also reported that the shaft
diameter has a negligible effect on the side shear resistance. On the other hand, tests by
Horvath et al. (1983) indicated that the side shear resistance decreases as the shaft
diameter increases. Williams and Pells (1981) explained this phenomenon by referring to
the theory of expansion of an infinite cylindrical cavity, which suggests that cylinders
with smaller diameters develop higher normal stresses for a given absolute value of
dilation. However, they offered no physical explanation why the shaft diameter does not
affect their own test results.
Drilled shafts in rock 230

Fig. 6.9 Side shear resistance reduction


factor αw [Equation (6.13)] (after
Williams & Pells, 1981).

6.3.2 End bearing resistance

(a) End bearing behavior of drilled shafts


The typical bearing capacity failure modes for rock masses depend on discontinuity
spacing with respect to foundation width (or diameter), discontinuity orientation,
discontinuity condition (open or closed), and rock type. Table 6.7 illustrates typical
failure modes according to rock mass conditions (ASCE, 1996). Prototype failure modes
may actually consist of a combination of modes.
The failure modes shown in Table 6.7 are for foundations with the base at or close to
the ground surface. The depth of shaft embedment may change the end bearing failure
modes of drilled shafts. As shown in Figure 6.11, when the base of the shaft is at or close
to the ground surface, a wedge type of failure is developed and the shaft undergoes both
vertical settlement and rotation. When the depth of embedment is greater than twice the
diameter of the shaft, a punching type of failure occurs and a truncated conical plug of
fractured rock is formed below the base (Williams et al., 1980).
Axial load capacity of drilled shafts in rock 231

Fig. 6.10 Side shear resistance


reduction factor βw [Equation (6.13)]
(after Williams & Pells, 1981).
In a study by Johnston and Choi (1985), stereo photogrammetric techniques were used
to study the process of failure of a model pile socketed into simulated rock. As shown in
Figure 6.12, the study suggests that failure progresses from initial radial cracking to a fan
shaped wedge. These observations were compared to typical load displacement curve
where four points are identified as: 1) at the end of elastic deformation; 2) a little before
major yielding; 3) a little after major yielding; and 4) failure.

(b) End bearing resistance based on local experience and codes

Peck et al. (1974) suggested a correlation between the allowable bearing pressure and
RQD for footings supported on level surfaces in competent rock (Fig. 6.13). This
correlation can be used as a first crude step in determination of the end bearing resistance
of drilled shafts in rock. It need be noted that this correlation is intended only for
unweathered jointed rock where the discontinuities are generally tight. If the value of
allowable pressure exceeds the unconfined compressive strength of intact rock, the
allowable pressure is taken as the unconfined compressive strength.
In Hong Kong design practice, for large diameter drilled shafts in granitic and
volcanic rocks, the allowable end bearing resistance may be used as specified in Table
6.8. The presumptive end bearing resistance values range from 3.0 to 7.5 MPa, depending
Drilled shafts in rock 232

on the rock category which is defined in terms of the rock decomposition grade, strength
and total core recovery.
Table 6.7 Typical bearing capacity failure modes
associated with various rock mass conditions (after
ASCE, 1996).
Rock mass conditions Failure
Joint Joint Illustration Mode
dip spacing
Brittle rock:
Local shear failure
caused by localized
brittle fracture

N/A s»B
Ductile rock:
General shear
failure along well
defined failure
surfaces

Open joints:
Compressive failure
of individual rock
columns. Near
70°<β vertical joint set(s)
<90°

s<B
Closed joints:
General shear
failure along well
defined failure
surfaces. Near
vertical joint(s)
Axial load capacity of drilled shafts in rock 233

s>B Open or closed


joints:
Failure initiated by
splitting leading to
general shear
failure. Near
vertical joint set(s)

20°<β s<B or s>B General shear


<70° if failure failure with
wedge can potential for failure
develop along joints.
along joints Moderately dipping
joint set(s)

Rock mass conditions Failure


Joint Joint Illustration Mode
dip spacing
Thick rigid upper layer:
Failure is initiated by
tensile failure caused by
flexure of the thick rigid
upper layer

Limiting
0°<β value of H
<20° with respect
to B Thin rigid upper layer:
Failure is initiated by
punching tensile failure
of the thin rigid upper
layer

General shear failure


with irregular failure
surface through rock
mass. Two or more
N/A s«B closely spaced joint sets
Drilled shafts in rock 234

The Standard Specifications for Highway Bridges adopted by the American


Association of State Highway and Transportation Officials (AASHTO, 1989) also
provide presumptive allowable bearing pressures for spread footing foundations in rock
(see Table 6.9). These presumptive values can be used as a first crude step in
determination of the end bearing resistance of drilled shafts in rock.

(c) End bearing resistance from pressuremeter test results


The Canadian Foundation Engineering Manual (CGS, 1985) proposed a method for
determining the end bearing resistance of drilled shafts based on in situ pressuremeter test
results:
qmax=Kb(Pl−Po)+σo
(6.15)
Axial load capacity of drilled shafts in rock 235

Fig. 6.11 Typical failure mechanism


for end bearing shafts: (a) Base of shaft
bearing at ground surface; and (b)
Shaft with length/diameter>2 (after
Williams et al., 1980).
Drilled shafts in rock 236

Fig. 6.12 Observed progressive failure


modes: (a) Typical load-displacement
curve; and (b) Failure modes
corresponding to the points in (a) (after
Johnston & Choi, 1985).
Axial load capacity of drilled shafts in rock 237

Fig. 6.13 Allowable bearing pressure


of jointed rock (after Peck et al., 1974).
Table 6.8 Presumed safe vertical bearing stress for
foundations on horizontal ground in Hong Kong
[simplified from PNAP 141 (BOO, 1990)].
Category Granitic and volcanic rock Presumed
bearing stress
(MPa)
1(a) Fresh to slightly decomposed strong rock of material weathering 7.5
grade II or better, with total core recovery>95 and minimum uniaxial
compressive strength of rock material σc not less than 50 MPa
(equivalent point load index strength PLI50a not less than 2 MPa)
1(b) Slightly to moderately decomposed moderately strong rock of 5.0
material weathering grade II or III or better, with total core
recovery>85% and minimum unconfined compressive strength of
rock material σc not less than 25 MPa (equivalent point load index
strength PLI50a not less than 1 MPa)
1(c) Moderately decomposed moderately strong to moderately weak rock 3.0
of material weathering grade III or IV or better, with total core
recovery>50%
a
Point load index strength PLI50 of rock quoted is equivalent value for 50-mm-diameter cores
(ISRM, 1979a).

where pl is the limit pressure as determined from pressuremeter tests in the zone
extending two shaft diameters above and below the shaft base; po is the at rest horizontal
Drilled shafts in rock 238

stress in the rock at the elevation of the shaft base; σo is the total overburden stress at
elevation of the shaft base; and Kb is an empirical non-dimensional coefficient, which
depends on the depth and shaft diameter ratio as shown in Table 6.10.

(d) Empirical and Semi-Empirical Relations


Unlike the side shear resistance, numerous theories have been proposed for estimating the
end bearing resistance. According to Pells and Turner (1980), the theoretical approaches
fall into three categories:
1. Methods which assume rock failure to be plastic.
2. Methods which idealize the zone of failure beneath the base in a form which allows
either the brittleness strength ratio or the brittleness modulus ratio to be taken into
account.
3. Methods based on limiting the maximum stress beneath the loaded area to a value less
than required to initiate fracture. These methods assume essentially that once the
maximum strength is exceeded at any point in a brittle material, total collapse will
occur.
There is a significant variation in the end bearing resistance predicted from different
theories. For example, the predicted end bearing capacity of rock with an internal friction
angle ranges from 4.9σc using the incipient failure theory (Category 3) based
on the modified Griffith theory to 56σc using the classical plasticity theory (Category 1),
where σc is the unconfined compressive strength of intact rock (Poulos
Table 6.9 Presumptive allowable bearing pressures
for spread footing foundations, modified after Navy
(1982) (simplified from AASHTO, 1989).
Range of Allowable bearing pressure
σc (MPa) (MFa)
Type of bearing material Consistency in Ordinary Recommended
place range value for use
Massive crystalline igneous and Very hard, >250 6–10 8
metamorphic rock: granite, sound rock
diorite, basalt, gneiss,
thoroughly cemented
conglomerate (sound condition
allows minor cracks)
Foliated metamorphic rock: Hard sound rock 100–250 3–4 3.5
slate, schist (sound condition
allows minor cracks)
Sedimentary rock: hard Hard sound rock 50–100 1.5–2.5 2
cemented shales, siltstone,
sandstone, limestone without
cavaties
Axial load capacity of drilled shafts in rock 239

Weathered or broken bedrock Medium hard 25–50 0.8–1.2 1


of any kind except highly rock
argillaceous rock (shale)
Compaction shale or other Medium hard 25–50 0.8–1.2 1
highly argillaceous rock in rock
sound condition
Notes:
1. Variations of allowable bearing pressure for size, depth, and arrangement of
footings must be determined by analysis.
2. Presumptive values for allowable bearing pressures obtained from building codes
and charts developed by various agencies based on local experience with
satisfactory and unsatisfactory performance; usually the pressure that will limit
total and differential settlements to 1 inch. Presumptive values are not based on
thorough engineering analysis.
3. Allowable bearing pressure for rock is controlled by rock mass discontinuities, and
should not exceed the unconfined compressive strength.

Table 6.10 Kb as fimction of depth and shaft


diameter ratio (CGS, 1985).
Depth/Diameter 0 1 2 3 5 7
Kb 0.8 2.8 3.6 4.2 4.9 5.2

& Davis, 1980). Because of the wide variation of theoretical results, empirical and semi-
empirical relations have been developed. Since they are more commonly used than the
theoretical methods, only the empirical and semi-empirical relations are discussed in the
following.
Analogous to the side shear resistance, many attempts have been made to correlate the
end bearing capacity, qmax, to the unconfined compressive strength, σc, of intact rock.
Some of the suggested relations are:
Coates (1967): qmax=3.0σc
(6.16)
Rowe and Armitage (1987b): qmax=2.7σc
(6.17)
ARGEMA (1992): qmax=4.5σc ≤10 MPa
(6.18)
Findlay et al. (1997): qmax=(1−4.5)σc
(6.19)

The bearing capacity of foundations on rock is largely dependent on the strength of the
rock mass. Discontinuities can have a significant influence on the strength of the rock
mass depending on their orientation and the nature of material within discontinuities
(Pells & Turner, 1980). As a result, relations have been developed to account for the
Drilled shafts in rock 240

influence of discontinuities in the rock mass. The Standard Specifications for Highway
Bridges adopted by the American Association of State Highway and Transportation
Officials (AASHTO, 1989) suggests that the end bearing capacity be estimated using the
following relationship:
qmax=Nmsσc
(6.20)

where Nms is a coefficient relating qmax to σc. The value of Nms is a function of rock mass
quality and rock type (Table 6.11), where rock mass quality, in essence, expresses the
degree of jointing and weathering. Rock mass quality has a much stronger effect on Nms
than rock type. For a given rock type, Nms for excellent rock mass quality is more than
250 times higher than Nms for poor quality. For a given rock mass quality, however, Nms
changes little with rock type. For example, for a rock mass of very good quality, the
values of Nms are 1.4, 1.6, 1.9, 2.0 and 2.3 respectively for rock types A, B, C, D and E
(see Table 6.11). It should be noted however that rock type is implicitly related to the
unconfined compressive strength. Equation (6.20) may thus represent a non-linear
relation between qmax and σc.
Although it is not explicitly mentioned in AASHTO (1989), Equation (6.20) and
coefficient Nms can be simply derived from the lower bound solution suggested by Carter
and Kulhawy (1988) (see Fig. 6.14):
qmax=[s0.5+(mbs0.5+s)0.5]σc
(6.21)

in which the expression in the brackets is simply the coefficient Nms in Equation (6.20);
and mb and s are the strength parameters for the Hoek-Brown strength criterion as
discussed in Chapter 4. Values of mb and s for the rock categories in Table 6.11 are
shown in Table 6.12. The values of Nms in Table 6.11 can be simply obtained by
inserting the corresponding values of mb and s from Table 6.12 in the expression in the
brackets of Equation (6.21).
Equation (6.21) does not consider the influence of the overburden soil and rock (i.e.,
overburden stress qs=0 is assumed). Zhang and Einstein (1998a) derived an expression
for the end bearing capacity that considers the influence of the overburden
Table 6.11 Values of Nms for estimating the end
bearing capacity of drilled shafts in broken or
jointed rock (after AASHTO, 1989).
Rock General RMR(1) Q(2) RQD(3) Nms(4)
Mask Description Rating Rating Rating
A B C D E
Quality
Excellent Intact rock with 100 500 95–100 3.8 4.3 5.0 5.2 6.1
joints spaced > 10
feet apart
Very Tightly interlocking, 85 100 90–95 1.4 1.6 1.9 2.0 2.3
Good undisturbed rock
Axial load capacity of drilled shafts in rock 241

with rough
unweathered
discontinuities
spaced 3 to 10 feet
apart
Good Fresh to slightly 65 10 75–90 0.28 0.32 0.38 0.40 0.46
weathered rock,
slightly disturbed
with discontinuities
spaced 3 to 10 feet
apart
Fair Rock with several 44 1 50–75 0.049 0.056 0.066 0.069 0.081
sets of moderately
weathered
discontinuities
spaced 1 to 3 feet
apart
Poor Rock with numerous 23 0.1 25–50 0.015 0.016 0.019 0.020 0.024
weathered
discontinuities
spaced 1 to 20
inches apart with
some gouge
Very Poor Rock with numerous 3 0.01 <25 Use qult for an equivalent soil
highly weathered
discontinuities
spaced<2 inches
apart
(1) Geomechanics rock mass rating (RMR) system (Bieniawski, 1988)—See Chapter 2
(2) Rock mass quality (Q) system (Barton et al., 1974)—See Chapter 2
(3) Range of RQD values provided for general guidance only; actual determination of rock
mass quality should be based on RMR or Q rating systems
(4) Value of Nms as function of rock type; refer to Table 2.8 for typical range of values of
σc for different rocks in each category

Table 6.12 Values of mb and s based on rock mass


classification (modified from Carter & Kulhawy,
1988).
Rock General RMR(1) Q(2) RQD(3) s mb
Mass Description Rating Rating Rating
A B C D E
Quality
Excellent Intact rock with 100 500 95–100 1 7 10 15 17 25
joints spaced
>10 feet apart
Drilled shafts in rock 242

Very Tightly 85 100 90–95 0.1 3.5 5 7.5 8.5 12.5


Good interlocking,
undisturbed rock
with rough
unweath-ered
discontinuities
spaced 3 to 10
feet apart
Good Freshtoslightly 65 10 75–90 0.004 0.7 1 1.5 1.7 2.5
weathered rock,
slightly
disturbed with
discontinuities
spaced 3 to 10
feet apart
Fair Rock with 44 1 50–75 10−4 0.14 0.2 0.3 0.34 0.5
several sets of
moderately
weathered
discontinuities
spaced 1 to 3
feet apart
Poor Rock with 23 0.1 25–50 10−5 0.04 0.05 0.08 0.09 0.13
numerous
weathered
discontnuities
spaced 1 to 20
inches apart
with some
gouge
Very Rock with 3 0.01 <25 0 0.007 0.01 0.015 0.017 0.025
Poor numerous highly
weathered
discontinuities
spaced <2
inches apart

stress (see Fig. 6.15):

(6.22)

where

(6.23)
Axial load capacity of drilled shafts in rock 243

Fig. 6.14 Lower-bound solution for


bearing capacity (after Kulhawy &
Carter, 1992).

Fig. 6.15 Assumed failure mode of


rock mass below the shaft base (after
Zhang & Einstein, 1998a).
Drilled shafts in rock 244

Kulhawy and Goodman (1980) presented the following relationship originally proposed
by Bishnoi (1968):

qmax=JcNcr
(6.24)

where J is a correction factor depending on normalized spacing of horizontal


discontinuities (spacing of horizontal discontinuities/shaft diameter) (see Fig. 6.16); c is
the cohesion of the rock mass; and Ncr is a modified bearing capacity factor, which is a
function of the friction angle of the rock mass and normalized spacing of vertical
discontinuities (spacing of vertical discontinuities/shaft diameter) (see Fig. 6.17).

Fig. 6.16 Correction factor for


discontinuity spacing (after Kulhawy
& Goodman, 1980).
Axial load capacity of drilled shafts in rock 245

Fig. 6.17 Bearing capacity factor for


open discontinuities (after Kulhawy &
Goodman, 1980).
Table 6.13 Suggested design values of strength
parameters c and (from Kulhawy Goodman,
1987).
Rock mass properties
RQD (%) Unconfined Compressive strength Cohesion c
Angle of friction
0–70 0.33σc 0.1σc 30°
70–100 0.33σc–0.8σc 0.1σc 30°–60°
σc=unconfined compressive strength of intact rock

As indicated in the preceding text, the strength parameters c and are rock mass properties.
Kulhawy and Goodman (1987) provided a table relating the rock mass properties c and
to intact rock properties and RQD (Table 6.13). The correction factor J considers the
effect of horizontal discontinuities and the variation of Jwith the discontinuity spacing is
shown in Figure 6.16, where H is the spacing of horizontal discontinuities. For the value
of Ncr the authors considered the discontinuities being either open or closed. According to
Goodman (1980), the presence of open discontinuities would allow failure to occur by
Drilled shafts in rock 246

splitting (because the discontinuities are open, there is no confining pressure and failure
is likely to occur by uniaxial compression of the rock columns), and this mode of failure
needs to be included in the calculation of the end bearing capacity. Several charts are
given by Kulhawy and Goodman (1980), following the method of Bishnoi (1968), to
determine Ncr for both open and closed discontinuities. Figure 6.17 shows Ncr for open
discontinuities.
The Canadian Foundation Engineering Manual (CGS, 1985) proposed that the end
bearing pressure be calculated using the following equation:
qmax=3σcKspD
(6.25)

where Ksp=[3+s/B]/[10(1+300g/s)0.5] is an empirical factor; s is the spacing of the


discontinuities; B is the shaft width or diameter; g is the aperture of the discontinuities;
D=1+0.4(L/B)≤3.4 is the depth factor; L is the shaft length. In general the method will
apply only if s/B ratios lie between 0.05 to 2.0 and the values of g/s between 0 and 0.02
(CGS, 1985).
Zhang and Einstein (1998a) developed a database of 39 shaft load tests about the
ultimate end bearing capacity (see Table 6.14). This database represents rocks of
relatively low strength. Table 6.14 lists, in addition to shaft dimensions, the unconfined
compressive strength of intact rock σc, the end bearing capacity qmax, and the end bearing
capacity factor Nc(=qmax/σc). The ratio of the shaft base displacement sb at qmax to the
shaft diameter B is also included in Table 6.14. A number of issues that need be
considered when studying the relationship between the end bearing capacity and the
unconfined compressive strength of intact rock are as follows:
1. Different interpretations of the load test data will give different capacities. For the test
shafts in Table 6.14, different interpretation methods were used. For example, Goeke
and Hustad (1979) took the load at plunging failure as the ultimate capacity of the
shaft (plunging failure is defined as the point at which additional load cannot be
applied to the shaft without experiencing continuous movement), while Jubenville and
Hepworth (1981) defined the ultimate capacity of the shaft as the load at which the
shaft head displacement reached 10% of the shaft diameter. Therefore, some
uncertainties and variabilities are likely to be included in the database. However, the
general trend reflected by the database will be useful.
2. The unconfined compressive strength is a property of the intact rock, not of the rock
mass. Clearly rock mass discontinuities must affect the end bearing
Table 6.14 Summary of database of shaft load tests
(Zhang and Einstein, 1998a).
No. Rock Diameter Depth σc qmax Nc=qmax/σc Sb/Ba Reference
description B (mm) to (MPa) (MPa) (%)
base L
(m)
1 Mudstone, 670 6 1.09 6.88 6.31 7.0 Wilson (1976)
weak, clayey
Axial load capacity of drilled shafts in rock 247

cretaceous
2 Clayshale, 762 8.8 0.81 4.69 5.79 6.2 Goeke and
with Hustad(1979)
occational thin
limestone
seams
3 Shale, thinly 457 13.7 3.82 10.8 2.83 >10.0 Hummert and
bedded with Cooling (1988)
thin sandstone
layers
4 Shale, 305 2.4 1.08 3.66 3.39 10.0 Jubenville and
unweathered Hepworth(1981)
5 Gypsumb 1064 4.20 2.1 6.51 3.1 15– Leung and Ko
20 (1993)
6 Gypsumb 1064 4.20 4.2 10.9 2.6 15– Leung and Ko
20 (1993)
7 Gypsumb 1064 4.20 5.4 15.7 2.9 15– Leung and Ko
20 (1993)
8 Gypsumb 1064 4.20 6.7 16.1 2.4 15– Leung and Ko
20 (1993)
9 Gypsumb 1064 4.20 8.5 23 2.7 15– Leung and Ko
20 (1993)
10 Gypsumb 1064 4.20 11.3 27.7 2.5 15– Leung and Ko
20 (1993)
11 Tillc 762 ** 0.7 4 5.71 ~1.3 Orpwoodetal.
(1989)
12 Tillc 762 ** 0.81 4.15 5.12 ~4.6 Orpwood et al.
(1989)
13 Tillc 762 ** 1 5.5 5.5 ~1.4 Orpwood et al.
(1989)
14 Diabase, 615 12.2 0.52 2.65 5.1 >4.0 Webb (1976)
highly
weathered
15 Hardpan (hard 1281 18.3 1.38 5.84 4.23 ~4.0 Baker (1985)
bearing till)c
16 Tillc 1920 20.7 0.57 2.29 4.04 ~1.9 Baker (1985)
17 Hardpan (hard 762 18.3 1.11 4.79 4.33 ~7.3 Baker (1985)
bearing till)c
18 Sandstone, 610 15.6 8.36 10.1 1.21 >1.7 Glos and Briggs
horizontally (1983)
bedded,
Drilled shafts in rock 248

shaley,
RQD=74%
19 Sandstone, 610 16.9 9.26 13.1 1.41 >1.7 Glos and Briggs
horizontally (1983)
bedded,
shaley, with
some coal
stringers,
RQD=88%
20 Mudstone, 300 2.01 0.65 6.4 9.8 6.4 Williams (1980)
highly
weathered
21 Mudstone, 300 1 0.67 7 10.5 5.7 Williams (1980)
highly
weathered
22 Mudstone, 1000 15.5 2.68 5.9 2.2 1.1 Williams (1980)
moderately
weathered
23 Mudstone, 1000 15.5 2.45 6.6 2.7 0.7 Williams (1980)
moderately
weathered
24 Mudstone, 1000 15.5 2.45 7 2.9 0.6 Williams (1980)
moderately
weathered
25 Mudstone, 1000 15.5 2.68 6.7 2.5 0.7 Williams (1980)
moderately
weathered
26 Mudstone, 600 1.8 1.93 9.2 4.8 14.1 Williams (1980)
moderately
weathered
27 Mudstone, 1000 3 1.4 7.1 5 10.9 Williams (1980)
moderately
weathered

No. Rock Diameter Depth σc qmax Nc= Sb/Ba Reference


description B (mm) to (MPa) (MPa) qmax/σc (%)
base L
(m)
28 Shale ** ** 34 28 0.82 ** Thorne (1980)d
29 Sandstone ** ** 12.5 14 1.12 ** Thorne (1980)d
30 Sandstone, fresh, ** ** 27.5 50 1.82 ** Thorne (1980)d
defect free
31 Shale, occational ** ** 55 27.8 0.51 ** Thorne (1980)d
Axial load capacity of drilled shafts in rock 249

recemented
moisture
fractures and
thin mud seams,
intact core
lengths 75 to
250 mm
32 Clayshale 740 7.24 1.42 5.68 4 ~8.8 Aurora and
Reese (1977)
33 Clayshale 790 7.29 1.42 5.11 3.6 ~8.9 Aurora and
Reese (1977)
34 Clayshale 750 7.31 1.42 6.11 4.3 ~6.0 Aurora and
Reese (1977)
35 Clayshale 890 7.63 0.62 2.64 4.25 ~6.6 Aurora and
Reese (1977)
36 Siltstone, 705 7.3 9 13.1 1.46 ~12.0 Radhakrishnan
medium hard, and Leung
fragmented (1989)
37 Marl, intact, 1200 18.5 0.9 5.3 5.89 ** Carrubba
RQD=100% (1997)
38 Diabase Breccia, 1200 19 15.0 8.9 0.59 ** Carrubba
highly fractured, (1997)
RQD=10%
39 Limestone, 1200 13.5 2.5 8.9 3.56 ** Carrubba
intact, (1997)
RQD=100%
a
Sb is the shaft base displacement at qmax.
b
Gypsum mixed with cement is used as pseudo-rock in centrifuge tests. The and depths are the
equivalent prototype dimensions corresponding to 40 g in the centrifuge tests. The equivalent
prototype depths to the shaft base range 4.04 m to 4.35 m with an average of 4.20 m.
c
Till is not a rock. It is used here because its σc is comparable to that of some rocks.
d
These tests were not conducted by Thorne (1980). He only reported the data other references

capacity. Unfortunately, relevant information on this factor is unavailable for


most of the cases in Table 6.14.
3. The conditions below the base of the shaft also influence the end bearing capacity. If
the base of the drilled hole cannot be cleaned, little or no end bearing support will be
developed. For all the test shafts in Table 6.14, the base of the drilled hole was
cleaned.
4. Different methods are used to separate the side shear resistance from the end bearing
capacity in load tests.
5. Clearly it would be interesting to have a relatively narrowly defined shaft base
displacement which one can associate with the end bearing capacity. However, the
values of sb/B in Table 6.14 indicate that the base displacement at qmax ranges from
0.6 to 20% of the shaft diameter, i.e., 6 to 210 mm. It is thus difficult to say at this
Drilled shafts in rock 250

point what typical base displacements at qmax are. [For comparison, the displacement
at ultimate side shear resistance is smaller; examination of more than 50 load-
displacement curves for large-diameter drilled shafts showed that an average
displacement of only 5 mm was necessary to reach initial failure of side shear
resistance (Horvath et al., 1983)].

Fig. 6.18 qmax versus σc (after Zhang &


Einstein, 1998a).
All the load test data in Table 6.14 are plotted in Figure 6.18. A log-log plot is used. It
can be seen that there is a strong relation between qmax and σc. Using linear regression, the
relationship between qmax and σc is as follows:
qmax=4.83(σc)0.51
(6.26)

The coefficient of determination, r2, is 0.81.

Example 6.2
A drilled shaft of diameter 1.0 m is to be socketed 3.0 meters in siltstone. The rock
properties are as follows:
Unconfined compressive strength of intact rock, σc=15.0 MPa
Axial load capacity of drilled shafts in rock 251

The rock mass is heavily jointed and the average discontinuity spacing near
the base of the shaft is 0.5 m
The discontinuities are moderately weathered and filled with debris with
thickness of 3 mm
Deformation modulus of intact rock, Er=10.6 GPa
RQD=45

Determine the end bearing resistance.

Solution:
Method of AASHTO (1989)—Equation (6.20)
From Table 2.8, the rock is classified as Type B.
From Table 6.11, the rock quality is classified as Fair and the value of Nms is 0.056.
Using Equation (6.20), the end bearing resistance can be obtained as
qmax=Nmsσc=0.056×15.0=0.84 MPa

Method of Zhang & Einstein (1998a)—Equations (6.22) & (6.23)


From Table 2.8, the rock is classified as Type B.
From Table 6.12, the rock quality is classified as Fair and the values of s and mb are
respectively 10−4 and 0.2.
Assuming that the effective unit weight of the rock mass is 13.0 kN/m3 and ignoring
the weight of the soil above the rock, the end bearing resistance can be obtained from
Equations (6.22) and (6.23) as

Method of CGS (1985)—Equation (6.25)


Empirical factor Ksp=[3+s/B]/[10(1+300g/s)0.5]=(3+0.5/1.0)/[10(1+300×0.003/0.5)0.5]
=0.21
Depth factor D=1+0.4(L/B)=1+0.4(3.0/1.0)=2.2.
Drilled shafts in rock 252

The end bearing resistance can be calculated from Equation (6.25) as

Method of Zhang and Einstein (1998)—Equation (6.26)


The end bearing resistance can be simply calculated from Equation (6.26) as

The results clearly show the wide range of the estimated end bearing capacity from
different methods. It is therefore important not to rely on a single method when
estimating the end bearing capacity.

6.4 CAPACITY OF DMLLED SHAFT GROUPS

In many cases, drilled shaft foundations will consist not of a single drilled shaft, but of a
group of drilled shafts. The drilled shafts in a group and the soil/rock between them
interact in a very complex fashion, and the axial capacity of the group may not be equal
to the axial capacity of a single isolated drilled shaft multiplied by the number of shafts.
One way to account for the interaction is to use the group efficiency factor η, which is
expressed as:

(6.27)

where QuG is the ultimate axial load of a drilled shaft group; N is the number of drilled
shafts in the group; and Qu is the ultimate axial load of a single isolated drilled shaft,
which can be determined using the methods described in Section 6.3. The group
efficiency for axial load capacity depends on many factors, including the following:
•The number, length, diameter, arrangement and spacing of the drilled shafts.
•The load transfer mode (side shear versus end bearing).
•The elapsed time since the drilled shafts were installed.
•The rock type.
Katzenbach et al. (1998) studied the group efficiency of a large drilled shaft group in
rock. For the 300 m high Commerzbank tower in Frankfurt am Main, 111 drilled shafts
are used to transfer the building load through the relatively weak Frankfurt Clay to the
stiffer underlying Frankfurt Limestone. Of the 111 drilled shafts, 30 were instrumented
Axial load capacity of drilled shafts in rock 253

and monitored during the 2-year construction period. The measurements give a detailed
view into the interaction between the drilled shafts in the group. Figure 6.19 shows the
variation of the group efficiency factor with the shaft head settlement. At service loads of
the building the value of the group efficiency factor is about 60%.
When drilled shafts are closely spaced, the shafts in a group may tend to form a
“group block” that behaves like a giant, short shaft (see Fig. 6.20). In this case, the
bearing capacity of the drilled shaft group can be obtained in a similar fashion to that for
a single isolated drilled shaft, by means of Equation (6.2), but now taking the shaft base
area as the block base area and the shaft side surface area as the block surface area. It
should be noted that the deformation required to mobilize the base capacity of the block
will be larger than that required for a single isolated shaft.

6.5 UPLIFT CAPACITY

In many cases, drilled shafts in rock may be required to resist uplift forces. Examples are
drilled shaft foundations for structures subjected to large overturning moments such as
tall chimneys, transmission lines, and highway sign posts. Drilled shafts through
expansive soils and socketed into rock may also subject to uplift forces due to the
swelling of the soil.
Drilled shafts can be designed to resist uplift forces either by enlarging or belling the
base, or by developing sufficient side shear resistance. Belling the base of a shaft is
common in soils, but this can be an expensive and difficult operation in rock. Moreover,
since large side shear resistance can be developed in drilled shafts socketed into rock, it is
usually more economical to deepen the socket than to construct a shorter, belled socket.
For drilled shafts subject to uplift forces, it is important to check the structural
capacity of the shaft. This can be done using the methods presented in Section 6.1. The
ultimate uplift resistance of a straight-sided drilled shaft related to rock can be
determined by
Quu=πBLτmax+Ws
(6.28)

where Quu is the ultimate uplift resistance; L and B are respectively the length and
diameter of the shaft; τmax is the average side shear resistance along the shaft; and Ws is
the weight of the shaft.
Drilled shafts in rock 254

Fig. 6.19 Variation of group efficiency


factor with shaft head settlement (after
Katzenbach et al., 1998).

Fig. 6.20 Treating the drilled shaft


group as a group block.
Uplift loading does not produce the same stress conditions in the shaft or rock mass as
those produced by compression loading. Compression loading compresses the shaft,
Axial load capacity of drilled shafts in rock 255

causing outward radial straining in the concrete (positive Poisson effect), which results in
higher frictional stresses at the interface with the rock mass; simultaneously it adds total
vertical stress to the rock mass around the shaft through the process of load transfer,
which consequently adds strength to rock masses that drain during loading. Uplift
loading, however, produces radial contraction of the concrete (negative Poisson effect)
and reduces the total vertical stresses in the rock mass around the shaft. Because of the
different stress conditions, the average side shear resistance for uplift loading should
usually be lower than that for compression loading.

Fig. 6.21 Measured side shear


resistance from compression tests and
pull-out tests.
Figure 6.21 shows the variation of measured side shear resistance with the unconfined
compressive strength of intact rock respectively from the compression load tests and the
pull-out load tests. The data are collected from the published literature. We can see that
the measured side shear resistances from the pull-out load tests are about the same as or
even higher than those from the compression load tests. One of the reasons for this might
be that the pull-out test shafts have rougher wall surfaces than the compression test
shafts. However, we are not sure about this at this point since no information on the wall
roughness is available for most of the test shafts shown in Figure 6.21.
Drilled shafts in rock 256

For preliminary design, the side shear resistance for uplift loading can be simply taken
to be the same as that for compression loading and estimated using the methods presented
in Section 6.3.1.
Where vertical drilled shafts are arranged in closely-spaced groups the uplift resistance
of the complete group may not be equal to the sum of the resistance of the individual
shafts. This is because, at ultimate-load conditions, the block of rock enclosed by the
shafts may be lifted. The uplift resistance of the block of rock may be determined by (see
Fig. 6.20)

(6.29)

where QuuG is the total ultimate uplift resistance of the shaft group; B1 and B2 are
respectively the overall length and width of the group (see Fig. 6.20); and WB is the
combined weight of the block of rock enclosed by the shaft group plus the weight of the
shafts.
7
Axial deformation of drilled shafts in rock

7.1 INTRODUCTION

Predicting the axial load-displacement response of drilled shafts is in some cases as


important as, or possibly more critical than, predicting the ultimate bearing capacity.
Many methods are available for predicting the axial displacement of drilled shafts in
rock. While the most reliable means for predicting the axial displacement of drilled shafts
is probably to carry out an axial loading test of the prototype shaft (which will be
discussed in Chapter 12), theoretical analyses may also be usefully employed. The main
three theoretical methods used to predict the axial load-displacement response of drilled
shafts in rock are the load-transfer (t-z) method, the continuum approach and the finite
element method.
The general load-displacement curve for a drilled shaft under axial loading can be
simply illustrated in Figure 7.1. The whole curve can be described in three stages:
1. As load is first applied to the head of the shaft, a small amount of displacement occurs
which induces the mobilization of side shear resistance from head to base. During this
initial period, the shaft behaves essentially in a linear manner, and the displacement
can be computed using the theory of elasticity. This linear behavior is illustrated in
Figure 7.1 as the line OA. The side shear stress along the shaft is smaller than the
ultimate side shear resistance (Fig. 7.2a).
2. As load is increased to point A in Figure 7.1, the shear stress at some point along the
interface will reach the ultimate side shear resistance (Fig. 7.2b), and the shaft-rock
‘bound’ will begin to rupture and relative displacement (slip) will occur between the
shaft and the surrounding rock. As the loading is increased further (beyond point A),
this process will continue along the shaft, more of the shaft will slip, and a greater
proportion of the applied load will be transferred to the end of the shaft (Fig. 7.2c). If
loading is continued, eventually the side shear stress everywhere will reach the
ultimate side shear resistance and the entire shaft will slip (point B in Fig. 7.1).
3. Beyond point B, a greater proportion of the total axial load will be transmitted directly
to the end of the shaft. When both side shear resistance and end bearing resistance are
fully mobilized (point C), any increase of load may produce significant displacement.
This indicates that the ultimate bearing capacity of the drilled shaft has been reached.
Axial deformation of drilled shafts in rock 259

Fig. 7.1 Generalized load-displacement


curves for drilled shafts under
compressive loading.

7.2 LOAD-TRANSFER (t-z CURVE) METHOD

The load-transfer method models the reaction of soil/rock surrounding the shaft using
localized springs: a series of springs along the shaft (the t-z or τ-w curves) and a spring at
the tip or bottom of the shaft (the q-w curve). τ is the local load transfer or side shear
resistance developed at displacement w, q is the base resistance developed at
displacement w, and w is the displacement of the shaft at the location of a spring. The
physical drilled shaft is also represented by a number of blocks connected by springs to
indicate that there will be compression of the drilled shaft due to the applied compressive
load. The mechanical model is shown in Figure 7.3. The displacement of the shaft at any
depth z can be expressed by the following differential equation:

(7.1)

where Ep is the composite Young’s modulus of the shaft (considering the contribution of
both concrete and reinforcing steel); A and B are respectively the cross-sectional area and
diameter of the shaft; w is the displacement of the shaft at depth z; and τ is the side shear
resistance developed at displacement w at depth z.
Equation (7.1) can be solved analytically or numerically depending on the τ-w and q-w
curves (linear or nonlinear), which is discussed in the sections below.
Drilled shafts in rock 260

7.2.1 Linear analysis


For linear analysis, the relationship between τ and w at any depth z and that between q
and w are assumed to be linear, i.e.,

Fig. 7.2 Shear stress at different values


of applied load (QA is the applied load
corresponding to point A in Fig. 7.1).

(7.2a)
Axial deformation of drilled shafts in rock 261

(7.2b)

where ks and kb are spring constants respectively of the side springs and the base spring.
Substitution Equation (7.2a) into Equation (7.1) gives

(7.3)

where

Fig. 7.3 Load-transfer (t-z curve)


model of axially loaded drilled shaft.

(7.4)

The general solution to Equation (7.3) is


Drilled shafts in rock 262

(7.5)

where C1 and C2 are integration constants.


The axial force at any depth is proportional to the first derivative of the displacement
with respect to depth:

(7.6)

If a load Qt is applied at the top of the shaft (z=0) and the force transferred to the base of
the shaft (z=L) is Qb, we have, from Equation (7.6),

(7.7a)

(7.7b)

From Equations (7.2b) and (7.5), We have

(7.8)

Solving Equations (7.7) and (7.8), constants C1 and C2 can be obtained as

(7.9a)

(7.9a)

The displacement at the top of the shaft (z=0) is then obtained from Equations (7.5) and
(7.9) as

7.2.2 Nonlinear analysis


In general, the τ-w and q-w curves are nonlinear. In this case, a convenient way to solve
differential Equation (7.1) is to use the finite difference method (Desai & Christian,
1977). Computer programs can be easily written to do the computations. The main issue
for the nonlinear analysis is the determination of the τ-w and q-w curves.
There are several techniques for determining the load transfer curves in soils
(Vijayvergiya, 1977; Kraft et al., 1981; Castelli et al., 1992) and rock masses (Baguelin et
Axial deformation of drilled shafts in rock 263

al., 1982; O’Neill & Hassan, 1994). However, research has not advanced to the point that
the load transfer curves (τ-w and q-w curves) can be determined for all conditions with
confidence (O’Neill & Reese, 1999). Construction practices and the particular response
of a given formation to drilling and concreting will affect the load transfer curves. For
major projects, therefore, it is advisable to measure the load transfer curves using full-
scale loading tests of instrumented shafts. Chapter 12 will show how to obtain the
experimental load transfer curves from the results of an axial loading test of an
instrumented shaft.
Based on measured load displacement curves, Carrubba (1997) conducted numerical
analyses to evaluate the side shear resistance and the end bearing capacity and obtained
the load transfer curves for five rock-socketed shafts. The model is based on a hyperbolic
transfer function approach and solves the equilibrium of the shaft by
means of finite element discretization. The interaction at the shaft-soil and shaft-rock
interfaces is described by the following function

(7.11)

where f(z) is the mobilized resistance along a shaft portion (τ) or at the shaft base (q); and
w(z) is the corresponding displacement (see Fig. 7.3). In the transfer function, parameters
a and b represent the reciprocals of initial slope and limit strength, respectively:

(7.12a)

(7.12b)

where flim is the end bearing capacity (qmax) in rock or the side shear resistance in soil or
rock (τmax).
Numerical analyses are carried out by selecting three transfer functions for each shaft:
one representative of overall friction in soil, one for overall friction in rock, and the last
one for end bearing resistance in rock. The friction transfer functions in soils, once
selected, are maintained constant throughout the analyses. Transfer function parameters
for rock, both along the shaft and at the base, are first estimated and then modified with
an iterative process until the actual load displacement curve is reproduced. Figure 7.4
shows the comparison between the test results and the numerical simulations for the shaft
in marl. Since the side and base strengths are not mobilized at the same time and the
numerical model used cannot simulate this event, two different ideal shaft behaviors are
examined. The first neglects the base reaction; the second takes into account the
contemporary mobilization of side and base resistances from the beginning of the test.
The rock properties and the transfer function parameters obtained for the five rock-
socketed shafts are shown in Table 7.1.
O’Neill and Hassan (1994) proposed an interim criterion for a hyperbolic τ-w curve in
most types of rock until better solutions become accepted:
Drilled shafts in rock 264

(7.13)

where B is the diameter of the shaft; and Em is the deformation modulus of the rock mass.
This model is based on the fact that the interface asperity pattern is regular and the
asperities are rigid, even though in most cases the interface asperity pattern is not regular,
some degree of smear exists, and asperities are deformable, which results in ductile,
progressive failure among asperities. Equation (7.13) is a special form of Equation (7.11)
with a=2.5B/Em.
Axial deformation of drilled shafts in rock 265

Fig. 7.4 Comparison between test


results and numerical simulations for
the drilled shaft in Marl. Curve a
neglects base reaction; curve b takes
into account cotemporary mobilization
of side and base resistances (after
Carrubba, 1997).
Table 7.1 Rock properties and transfer function
parameters (Carrubba, 1997).
Shaft side in rock Base in rock
Rock type σc RQD Em 1/b 1/a 1/b 1/a
(MPa) (%) (MPa) (MPa) (MN/m3) (MPa) (MN/m3)
Marl 0.90 100 200a 0.14 100 5.30 220
b
Diabasic 15.00 10 200 0.49 70 8.90 300
Breccia
Gypsum 6.00 60 2,000a 0.47 200 – –
a
Diabase 40.00 50 10,000 1.20 500 – –
b
Limestone 2.50 100 500 0.40 500 8.90 3,000
a
From compression tests on specimens
b
From plate bearing tests

The q-w curve is usually assumed to have an initial elastic response given by

where Eb and νb are respectively the deformation modulus and Poisson’s ratio of the rock
below the shaft base. Nonlinear response is usually assumed to initiate between 1/3 and
1/2 of qmax. This response can be simply modeled using an equation similar to Equation
(7.13).

7.3 CONTINUUM APPROACH

The continuum approach assumes the soil/rock to be a continuum. Mattes and Poulos
(1969) are among the first to investigate the load-displacement behavior of rock-socketed
shafts by integration of Mindlin’s equations. Carter and Kulhawy (1988) provide a set of
approximate analytical solutions to predict the load-displacement response of drilled
shafts in rock by modifying the solutions of Randolph and Wroth (1978) for piles in soil.
Drilled shafts in rock 266

The majority of the theoretical continuum solutions for predicting the displacement of
drilled shafts in rock, however, have been developed using finite element analyses (e.g.,
Osterberg & Gill, 1973; Pells & Turner, 1979; Donald et al., 1980; Rowe & Armitage,
1987a). Most of the techniques proposed for calculating the vertical displacements of
drilled shafts in rock are based on the theory of elasticity. It has been usual to assume that
the drilled shaft is essentially an elastic inclusion within the surrounding rock mass and
that no slip occurs at the interface between the shaft and the rock mass, although the
solutions of Rowe and Armitage (1987a) and Carter and Kulhawy (1988) can consider
the possibility of slip.

7.3.1 Linear continuum approach

(a) Solutions based on finite element results


As stated in Chapter 6, axially loaded drilled shafts in rock are designed to transfer
structural loads in one of the following three ways (CGS, 1985):
1. Through side shear only;
2. Through end bearing only;
3. Through the combination of side shear and end bearing.
The following presents the elastic solutions based on the finite element results for
estimating the axial deformation of the above three types of shafts.

Side shear only shaft


Based on finite element analysis, Pells and Turner (1979) presented the following general
equation for calculating the axial deformation of side shear only shafts in a single elastic
half space:

(7.15)

where wt is the axial deformation of the shaft at the rock surface; Qt is the applied load at
the top of the shaft; Em is the deformation modulus of the rock mass; B is the diameter of
the shaft; and I is the axial deformation influence factor given in Figure 7.5. The values
of I given in Figure 7.5 have been calculated for a Poisson’s ratio of 0.25. It has been
found that variations in the Poisson’s ratio in the range 0.1–0.3 for the rock mass and
0.15–0.3 for the concrete have little effect on the influence factors.
The values of the influence factor shown in Figure 7.5 are for drilled shafts that are
fully bonded from the rock surface. In many cases, the drilled shaft is recessed by casing
the upper part of the drilled hole or for conditions where the shaft passes through a layer
of soil or weathered rock where little or no side shear resistance will be developed.
Recessment of the shaft will result in a decrease in axial deformation of the shaft at the
head of the socket. This reduction can be expressed in terms of a reduction factor RF
such that the axial deformation of the shaft at the ground surface is given by
Axial deformation of drilled shafts in rock 267

(7.16)

Fig. 7.5 Axial deformation influence


factors for side shear only drilled
shafts (after Pells & Turner, 1979).
where Qt is the applied load at the top of the shaft; D and Bl are respectively the length
and diameter of the recessed shaft; Ep is the composite Young’s modulus of the shaft
(considering contributions of both concrete and reinforcing steel); RF is a reduction
factor for the effect of recessment; B is the diameter of the socketed shaft; Em is the
deformation modulus of the rock mass; and I is the influence factor for shaft with no
recessment (see Fig. 7.5). The first portion of Equation (7.16) simply represents the
elastic compression of the shaft over the length D. The second portion of Equation (7.16)
gives the axial deformation of the socketed portion of the shaft. The reduction factor RF
is given in Figure 7.6 for a range of situations.
Drilled shafts in rock 268

Fig. 7.6 Reduction factors for


calculation of axial deformation of
recessed drilled shafts (after Pells &
Turner, 1979).
End bearing only shaft
Axial deformation of drilled shafts in rock 269

An end bearing only shaft can be considered a shaft that is wholly recessed (See Fig. 7.7).
The axial deformation of an end bearing only shaft at the ground surface consists of the
elastic compression of the shaft and the axial deformation of the shaft base:

(7.17)

where Qt is the applied load at the top of the shaft; D and Bl are respectively the length
and diameter of the shaft; Ep is the composite Young’s modulus of the shaft (considering
contributions of both concrete and reinforcing steel); Em and νm are respectively the
deformation modulus and Poisson’s ratio of the rock mass; Cd is the shape and rigidity
factor equal to 0.85 for a flexible footing and 0.79 for a rigid footing; and RF′ is a
reduction factor for an end bearing only shaft as shown in Figure 7.7.

The axial deformation of the shaft base is calculated in a similar manner to that of a
footing on the surface. However, because the rock mass below the base of the shaft is
more confined than surface rock mass, the axial deformation of the shaft base will be
smaller than that of a footing at the surface. The effect of this confinement if accounted
for by applying the reduction factor RF′ to the deformation equation as shown in
Equation (7.17). The value of the reduction factor depends on the ratio of the shaft length
D to the shaft diameter B1, and the relative stiffness of the shaft and the rock mass. Figure
7.7 shows the values of the reduction factor RF′ obtained by Pells and Turner (1979).

Side shear and end bearing shaft


For side shear and ending bearing shafts, the axial deformation at the rock surface can be
calculated using Equation (7.15). Considering the interaction between the side shear and
end bearing, the influence factors given in Figure 7.8 should be used. These factors have
been developed for elastic behavior without slip along the side walls by Rowe and
Armitage (1987a).
Drilled shafts in rock 270

Fig. 7.7 Reduction factors for


calculation of axial deformation of end
bearing only drilled shafts (after Pells
& Turner, 1979).
Comparison of Figure 7.8(a) (for Eb/Em=1) with Figure 7.5 shows that the influence
factor for a side shear and end bearing shaft is smaller than that for a side shear only
shaft, which demonstrates that a shaft with both side shear and end bearing will settle less
than a shaft with side shear only. Figure 7.9 shows the percentage of the load carried in
the end bearing.

(b) Analytical solutions of Carter and Kulhawy (1988)


Carter and Kulhawy (1988) provide a set of approximate analytical solutions to predict
the load-displacement response of drilled shafts in rock. Two layers of rock mass as
shown in Figure 7.10 are considered in the solutions. The solutions are for a shaft without
slip or with full slip. The following presents the solution for a shaft without slip while the
solution for a shaft with full slip will be presented in Section 7.3.2.
Axial deformation of drilled shafts in rock 271

Under an applied axial load, the displacements in the rock mass are predominantly
vertical, and the load is transferred from the shaft to the rock mass by vertical shear
stresses acting on the cylindrical interface, with little change in vertical normal stress in
the rock mass (except near the base of the shaft). The pattern of deformation around the
shaft may be visualized as an infinite number of concentric cylinders sliding inside each
other (Randolph & Wroth, 1978). Randolph and Wroth (1978) have shown that, for this
type of behavior, the displacement of the shaft w may be described adequately in terms of
hyperbolic sine and cosine functions of depth z below the surface, as given below:
w=A1 sinh(µz)+A2 cosh(µz)
(7.18)

in which, A1 and A2 are constants which can be determined from the boundary conditions
of the problem. The constant µ is given by

(7.19)

where ζ=ln[2.5(1−νm)L/R]; R=B/2 is the radius of the shaft; λ=Ep/Gm; Ep is the Young’s
modulus of the shaft; Gm=Em/[2(1+νm)] is the shear modulus of the rock mass
surrounding the shaft; and Em and νm are respectively the deformation modulus and
Poisson’s ratio of the rock mass surrounding the shaft.

For side shear and end bearing shafts as shown in Figure 7.10(a), the shaft base can be
approximated as a punch acting on the surface of an elastic half-space with Young’s
modulus Eb and Poisson’s ratio νb. Using the standard solutions for the displacement of a
rigid punch resting on an elastic half-space as the boundary condition at the base of the
shaft, the elastic displacement at the head of the shaft can be obtained by (Randolph &
Wroth, 1978):

(7.20)
Drilled shafts in rock 272

Fig. 7.8 Axial deformation influence


factors for side shear and end bearing
drilled shafts (after Rowe & Armitage,
1987a).
Axial deformation of drilled shafts in rock 273

Fig. 7.9 Load distribution curves for


side shear and end bearing drilled
shafts (after Rowe & Armitage,
1987a).
Drilled shafts in rock 274

where ξ=Gb/Gm; Gb=Eb/[2(1+νb)] is the shear modulus of the rock mass below the shaft
base; and Eb and νb are respectively the deformation modulus and Poisson’s ratio of the
rock mass below the shaft base. The proportion of the applied load transmitted to the
shaft base is

(7.21)

For side shear only shafts as shown in Figure 7.10(b), the boundary condition at the shaft
base is one of zero axial stress. For this case, the elastic displacement at the head of the
shaft can be obtained by

Fig. 7.10 Axially loaded drilled shafts


in rock (after Carter & Kulhawy,
1988).
Axial deformation of drilled shafts in rock 275

Fig. 7.11 Comparison of analytical


solution with finite element solution
for predicting axial elastic
displacement (after Carter & Kulhawy,
1988).

(7.22)

The solution given by Equations (7.20) and (7.22) are in general agreement with the finite
element solutions by Pells and Turner (1979) and Rowe and Armitage (1987a) as
presented in last sections (Fig. 7.11).

Example 7.1
A drilled shaft of 3.0 meters long and 1.0 meter in diameter is to be installed in siltstone.
The rock properties are as follows:
Unconfined compressive strength of intact rock, σc=15.0 MPa
Deformation modulus of intact rock, Er=10.6 GPa
RQD=70

Determine the settlement of the drilled shaft at a work load of 10.0 MN.
Drilled shafts in rock 276

Solution:
For simplicity, the Young’s modulus of the drilled shaft is simply assumed to be Ep=30
GPa. The Poisson’s ratio of 0.25 is selected for both the drilled shaft and the rock.
Using Equation (4.24), the rock mass modulus:
αE=0.0231×70−1.32=0.297
Em=0.297×10.6=3.15Gpa

Using solutions based on finite element method

L/B=3.0/1.0=3.0
Ep/Em=30/3.15=9.52

If the drilled shaft is side shear resistance only (i.e., the shaft base cannot be cleaned),
from Figure 7.5, the axial deformation influence factor is I=0.462. Using Equation (7.15),
the settlement of the drilled shaft at the rock surface is

If the drilled shaft has both side shear and end bearing resistance, from Figure 7.8, the
axial deformation influence factor is I=0.417 for Eb/Em=1.0. Using Equation (7.15), the
settlement of the drilled shaft at the rock surface is

From Figure 7.9, it can be seen that about 15% of the load is transmitted to the shaft base.
Using analytical solutions of Carter and Kulhawy (1988)
Axial deformation of drilled shafts in rock 277

ξ=Gb/Gm=1.0 for Eb/Em=1.0

If the drilled shaft is side shear resistance only (i.e., the shaft base cannot be cleaned), the
settlement of the drilled shaft at the rock surface can be calculated from Equation (7.22)
as

If the drilled shaft has both side shear and end bearing resistance, the settlement of the
drilled shaft at the rock surface can be calculated from Equation (7.20) as
Drilled shafts in rock 278

The percentage of the load transmitted to the shaft base can be calculated from Equation
(7.21) as

The results from the solutions based on the finite element method are in good agreement
with those from the analytical solutions of Carter & Kulhawy (1988).

7.3.2 Nonlinear continuum approach

(a) Solutions based on finite element results


Rowe and Armitage (1987a) performed an elastic-plastic finite element analysis that
accounts for slip along the interface based on the technique developed by Rowe and Pells
(1980). Two layers of rock are considered in the analyses. The interface behavior is
established in terms of the Coulomb failure criterion. The roughness of the interface is
modeled implicitly through the use of an angle of interface dilatancy that produces
additional normal stress on the interface as the shaft deflects vertically due to the applied
load. The contribution of the interface dilatancy commences once slip occurs at the
interface. The results of this study are presented in three sets of design charts respectively
for Eb/Em=0.5, 1.0 and 2.0. Although the analysis is carried out considering the behavior
of a cohesive-frictional-dilative interface, the design charts are developed only for
nondilative-cohesive interfaces. The procedure for using the design charts is described in
Rowe and Armitage (1987b).

(b) Analytical solutions of Carter and Kulhawy (1988)


The case of slip along the entire length of the shaft has also been considered in detail by
Carter and Kulhawy (1988). For this case, the shear strength of the interface is given by
the Coulomb criterion:

(7.23)
Axial deformation of drilled shafts in rock 279

where c is the interface cohesion; is the interface friction angle; and σr is the radial
stress acting on the interface.
As relative displacement (slip) occurs, the interface may dilate, and it is assumed that
the displacement components follow the dilation law:

(7.24)

where ∆u and ∆w are the relative shear and normal displacements of the shaft-rock
interface; and ψ is the angle of dilation defined by Davis (1968).
To determine the radial displacements at the interface, the procedure suggested by
Goodman (1980) and Kulhawy and Goodman (1987) is followed, in which conditions of
plane strain are assumed, as an approximation, independently in the rock mass and in the
slipping shaft. The rock mass is considered to be linear elastic, even after full slip has
taken place, and the shaft is considered to be an elastic column. These assumptions,
together with the dilatancy law, allow one to derive an expression for the variation of
vertical stress in the compressible shaft. The distribution of the shear stress acting on the
shaft can then be calculated from equilibrium conditions, and the vertical displacement
can be determined as function of depth z by treating the shaft as a simple elastic column.
The ‘full slip’ solution for the displacement of the shaft head is derived as

(7.25)

in which
F3=a1(λ1BC3−λ2BC4)−4a3
(7.26)

(7.27)

C3,4=D3,4/(D4−D3)
(7.28)

(7.29)

(7.30)

(7.31)
Drilled shafts in rock 280

(7.32)

a1=(1+νm)ς+a2
(7.33)

(7.34)

(7.35)

All other parameters in Equations (7.25) to (7.35) are as defined before. The adequacy of
the closed-form expressions is demonstrated by comparing them with the finite element
solution of Rowe and Armitage (1987a, b). The overall agreement between the closed-
form solutions and the finite element results is good (Fig. 7.12).
It must be noted that the closed-form solutions of Carter and Kulhawy (1988) just
consider “no slip” (presented in Section 7.3.1) and “full slip” conditions. They cannot
predict the load-displacement response between the occurrence of first slip and full slip
of the shaft. However, the finite element results indicate that the progression of slip along
the shaft takes place over a relatively small interval of displacement. Therefore it seems
reasonable, at least for most practical cases, to ignore the small region of the curves
corresponding to the progressive slip and to assume that the load-displacement
relationship is bilinear, with the slope of the initial portion given by Equation (7.20) and
the slip portion by Equation (7.25) (Carter & Kulhawy, 1988).
Axial deformation of drilled shafts in rock 281

Fig. 7.12 Axial displacement of a


drilled shaft in rock considering full
slip (after Carter & Kulhawy, 1988).

7.4 FINITE ELEMENT METHOD (FEM)

The finite element method is probably the most powerfiil and the most widely used
numerical method currently available to engineers. Suitable elements can be used to
simulate not only linearly elastic materials, but also nonlinear materials with different
failure criteria, including rock discontinuities and shaft-rock interfaces (see Sections 4.3.4
and 4.4.3 for discussion of joint elements). However, the finite element method is time
consuming and needs sophisticated soil or rock constitutive relations whose parameters
are often difficult if not impossible in design practice to obtain. Therefore, the finite
element method is, in general, used for analysis of important structures and for generation
of parametric solutions for the load-displacement relations of axially loaded drilled
shafts, such as the charts presented in Sections 7.3.1 and 7.3.2.
Typical of many geotechnical problems, the analysis of drilled shafts in rock involves
an unbounded domain. It is a common practice in finite element modeling of these
problems to truncate the finite element mesh at a distance deemed far enough so as not to
influence the near field solutions. These truncations are usually determined by trail and
Drilled shafts in rock 282

error until an acceptable solution is obtained. Such a method places a heavy demand on
computer resources, both memory and time, as solutions for the far field which are of no
interest are generated as well. In the last decade or so, several methods have been
developed to model unbounded domains. Of these methods, the use of infinite elements
with finite elements appears to be the most popular. Leong and Randolph (1994)
successfully used finite elements and infinite elements in the modeling of axially loaded
shafts in rock.

7.5 DRILLED SHAFT GROUPS

Numerous methods exist for analyzing axially loaded pile groups in soil (Poulos, 2001),
some of which can be applied to drilled shaft groups in rock and are briefly described in
the following.

7.5.1 Settlement ratio method


In the settlement ratio method, the group settlement is related to the single-shaft
settlement as follows:

(7.36)

where wtG is the settlement of the shaft group; wtav is the settlement of a single shaft at the
average load of a shaft in the group; and Rw is the settlement ratio. wtav can be estimated
using the methods presented in the previous sections or from the results of load test on a
prototype drilled shaft.
Theoretical values of Rw for various pile groups in soil have been presented by Poulos
and Davis (1980) and Butterfield and Douglas (1981). A particularly useful
approximation for the settlement ratio has been derived by Fleming et al. (1992):

(7.37)

where n is the number of piles in the group; and e is an exponent depending on pile
spacing, pile proportions, relative pile stiffness and the variation of soil modulus with
depth. For typical pile proportions and pile spacings, Poulos (1989) suggested the
following approximate values: e≈0.5 for piles in clay, and e≈0.33 for piles in sand.
For drilled shafts in rock, the e values suggested by Poulos (1989) for soils may be
used for the very preliminary design. For the final design of major projects, it is desirable,
when feasible, to conduct axial load tests on groups of two or more drilled shafts in rock
in order to confirm the e values of Poulos (1989) or to derive new, site-specific values.

7.5.2 Equivalent pier method


The equivalent pier method, frequently used for pile groups in soils, treats the pile group
as an equivalent pier consisting of the piles and the soil between them (Poulos & Davis,
Axial deformation of drilled shafts in rock 283

1980; Randolph, 1994). For closely spaced drilled shafts in rock, the shaft group may
also be analyzed using the equivalent pier method.
Consider the drilled shaft group as an equivalent pier (Fig. 7.13), the diameter of the
equivalent pier Beq can be taken as (Randolph, 1994).

Fig. 7.13 Equivalent pier method


treating drilled shaft group as a group
block.

(7.38)

where Ag is the plan area of the drilled shaft group as a block.


Deformation modulus of the equivalent pier Eeq is then calculated as
Eeq=Em+(Ep−Em)Apt/Ag
(7.39)

where Ep is the Young’s modulus of the drilled shafts; Em is the deformation modulus of
the rock mass; and Apt is the total cross-sectional area of the drilled shafts in the group.
The load-settlement response of the equivalent pier can be calculated using the
solutions as described in the previous sections for the response of a single drilled shaft.
Based on the equivalent pier method and the load-transfer (t-z curve) approach,
Castelli and Maugeri (2002) presented a simplified nonlinear analysis for settlement
prediction of pile groups in soil. To take into account the group action due to pile-soil-
pile interaction, load-transfer functions are modified to relate the behavior of a single pile
to that of a pile group. The bearing capacity of the equivalent pier can be evaluated using
the procedure in Section 6.4. The initial stiffness of the equivalent pier is estimated by
Drilled shafts in rock 284

(7.40)

where Kgi is the initial stiffness of the equivalent pier and β is an empirical parameter. To
take into account the increase of pile group head settlements with respect to the case of a
single pile, the following expression is used

(7.41)

where wg is the average settlement of the equivalent pier and ε is an empirical parameter.
The empirical parameters β and ε can be derived on the basis of numerical analysis of
field tests. Castelli and Maugeri (2002) derived values of 0.30 and 0.15 respectively for β
and ε based on analysis of field test piles and pile groups in soils. For drilled shafts in
rock, similar values of β and ε can be obtained from field tests of shafts and shaft groups.

7.5.3 Finite element method (FEM)


The finite element method has been used to analyze axially loaded pile groups in soil by
simplifying the group to an equivalent plane strain or axisymmetric system. If necessary,
it can also be used to analyze drilled shaft groups in rock.
8
Lateral load capacity of drilled shafts in rock

8.1 INTRODUCTION

In the design of drilled shafts subjected to lateral forces, two criteria must be satisfied:
first, an adequate factor of safety against ultimate failure, second, an acceptable
deflection at working loads. This chapter discusses the prediction of ultimate load of
drilled shafts and drilled shaft groups. The calculation of lateral deflection will be
discussed in Chapter 9.
As the axial load capacity, the lateral load capacity of a drilled shaft in rock is
determined by the smaller of the two values: the structural strength of the shaft itself, and
the ability of the rock to support the loads transferred by the shaft.

8.2 CAPACITY OF DRILLED SHAFTS RELATED TO


REINFORCED CONCRETE

The structural capacity of a drilled shaft under lateral loading is controlled by the bending
capacity and the shear capacity. The bending capacity is usually checked by considering
the interaction between axial load and bending moment. Figure 8.1 shows the normalized
axial load-moment intersection diagrams for fy=10f′c and fy=15f′c, where fy is the yield
strength of the longitudinal reinforcing steel and f′c is the specified minimum concrete
strength. The factored axial load ΣγiQi is normalized by dividing by the factored nominal
axial capacity , where γi is the load factor for axial load i, Qi is the nominal value
of axial load i, and is the resistance factor for the nominal (computed) structural axial
load capacity Qu. The factored moment ΣγmMm is similarly normalized by dividing by the
factored nominal moment capacity , where γm is the load factor for moment m, Mm
is the nominal value of moment m, and is the resistance factor for the nominal
(computed) structural moment capacity Mu. The factored axial capacity is estimated from
Equation (6.1). Normalized axial load-moment interaction diagrams may be developed
for any fy/f′c ratios and cage diameters other than 0.6B.
With the axial load-moment interaction diagrams available, the structural capacity can
be checked as follows:
1. Estimate the combined axial load ΣγiQi.
2. Compute the factored nominal axial capacity from Equation (6.1).
Drilled shafts in rock 286

Fig. 8.1 Normalized axial load-


moment interaction diagrams for
drilled shafts for (a)fy=10f′c and
(b)fy=15f′c.
Lateral load capacity of drilled shafts in rock 287

Table 8.1 Nominal structural moment capacity Mu


of drilled shaft.
Mu/f′cBAg
As/Ag fy−10f′c fy=15f′c
0.01 0.037 0.050
0.02 0.067 0.092
0.03 0.088 0.119
0.04 0.107 0.147
0.05 0.126 0.172
0.06 0.144 0.197
0.07 0.161 0.208
0.08 0.176 0.244
As is the cross-sectional area of the longitudinal reinforcing steel
Ag is the gross cross-sectional area of the shaft
fy is the yield strength of the longitudinal reinforcing steel
f′c is the specified minimum concrete strength
B is the diameter of the shaft

3. Estimate the factored (required) moment ΣγmMm.


4. Estimate the nominal structural moment capacity Mu of the drilled shaft. This may be
based on complete analysis, or it may be obtained from design aids such as Table 8.1.
5. Compute the factored nominal moment capacity where for
reinforced concrete.
6. Determine the ratios , and and with these values locate
an appropriate point on the axial load-moment interaction diagram. If the point falls
inside the area defined by the interaction curve, the shaft capacity is adequate. If this is
not the case, the shaft size should be increased and the analysis repeated until the shaft
capacity is adequate.
The factored nominal shear capacity of a drilled shaft without special shear reinforcement
can be calculated by (O’Neill & Reese, 1999):

(8.1)

where is the capacity reduction (resistance) factor for shear=0.85; Vu is the nominal
(computed) shear resistance; νc is the limiting concrete shear stress; and Av is the area of
the shaft cross section that is effective in resisting shear, which can be taken as
B(0.5B+0.5756rls) for a circular drilled shaft, where r1s is the radius of the ring formed by
the centroids of the longitudinal reinforcing steels. The limiting concrete shear stress νc
can be evaluated from:
Drilled shafts in rock 288

(8.2a)

(8.2b)

where both f′c and νc are in kPa.


If the factored shear load is greater than the factored nominal shear resistance
determined above, two options are available. The first and simplest solution is to increase
the shaft diameter to increase the shear capacity. The second alternative is to provide
properly designed shear reinforcement (O’Neill & Reese, 1999).
O’Neill and Reese (1999) provide detailed discussion and examples on checking the
structural load capacity of drilled shafts.

8.3 CAPACITY OF DRILLED SHAFTS RELATED TO ROCK

8.3.1 Method of Carter and Kulhawy (1992)


Carter and Kulhawy (1992) presented a method to determine the lateral load capacity of
drilled shafts related to rock. When a lateral load is applied at the rock surface, the rock
mass immediately in front of the shaft will be subject to zero vertical stress, while
horizontal stress is applied by the leading face of the shaft. Ultimately, the horizontal
stress may reach the uniaxial compressive strength of the rock mass and, with further
increase in the lateral load, the horizontal stress may decrease as the rock mass softens
during postpeak deformation. Large lateral deformations may be required for the rock
mass at depth to exert a maximum reaction stress on the leading face of the shaft.
Therefore, Carter and Kulhawy (1992) assumed that the reaction stress at the rock mass
surface, in the limiting case of loading of the shaft, is zero or very nearly zero as a result
of the postpeak softening. Along the sides of the shaft, some shearing resistance may be
mobilized. The shearing resistance varies along the perimeter and the average can be
chosen as τmax/2, where τmax is likely to be approximately the same as the maximum unit
side resistance under axial compression. Therefore, at the rock surface, the ultimate force
per unit length resisting the lateral loading is Bτmax. At greater depth, Carter and Kulhawy
(1992) assume that the stress in front of the shaft increases from the initial in situ
horizontal stress to the limit stress, pL, reached during the expansion of a long cylindrical
cavity, i.e., a plane strain condition will apply. Behind the shaft, the horizontal stress will
decrease, and after tensile rupture of the bond between the concrete and the rock mass,
the horizontal stress will reduce to zero. At the sides of the shaft, some shearing
resistance may also be mobilized. Therefore, at depth, the ultimate force per unit length
resisting the lateral loading is B(pL+τmax). To determine the depth at which the limit stress
is mobilized, the result of Randolph and Houslby (1984) in a cohesive material is
adopted, i.e., the depth is about three times shaft diameter. Therefore, the distribution of
ultimate force per unit length resisting the shaft is as shown in Figure 8.2.
Lateral load capacity of drilled shafts in rock 289

The ultimate lateral force that may be applied can be obtained from the horizontal
equilibrium as:
For L<3B

Fig. 8.2 Distribution of ultimate lateral


force per unit length (after Carter &
Kulhawy, 1992)

(8.3a)

For L>3B

(8.3b)

where τmax is the shearing resistance along the sides of the shaft, which is assumed to be
the same as the maximum side resistance under axial loading; pL is the limit stress
reached during the expansion of a long cylindrical cavity. Closed-form solutions have
been found for the limit stresses developed during the expansion of a long cylindrical
cavity in an elasto-plastic, cohesive-frictional, dilatant material (Carter et al., 1986). This
limit stress pL can be determined from the following parametric equation in the
nondimensional quantity ρ (Carter et al., 1986):

(8.4)
Drilled shafts in rock 290

with

(8.5)

in which

(8.6)

(8.7)

(8.8)

(8.9)

(8.10)

(8.11)

(8.12)

(8.13)

(814)

and σhi is the initial in situ horizontal stress; Gm is the elastic shear modulus; νm is the
Poisson’s ratio; cm is the cohesion intercept; φm is the friction angle; and ψm is the dilation
angle, all of the rock mass. The rock mass is assumed to obey the Coulomb failure
criterion, and dilatancy accompanies yielding according to the following flow rule

(8.15)

in which dε1p and dε3p are the major and minor principal plastic strain increments,
respectively. For convenience, solutions for the limit pressures pL have been plotted in
Figure 8.3 for selected values of νm, φm, ψm. The central vertical axis on each plot
Lateral load capacity of drilled shafts in rock 291

indicates the ratio of the plastic radius at the limit condition R to the cavity radius a.
These charts may be used by entering with a value of Gm/(σhi+cmcotφm) and working
clockwise around the figure, determining in turn values of R/a, then
ρL=(pL+cmcotφm)/(σR+ cmcotφm), and thus, determining the limit pressure pL.

8.3.2 Method of Zhang et al. (2000)


Zhang et al. (2000) presented an approximate method for calculating the ultimate lateral
resistance of drilled shafts in rock. As shown in Figure 8.4(a), the total reaction of the
rock mass consists of two parts: the side shear resistance and the front normal resistance.
So the ultimate resistance pult can be estimated by (Briaud & Smith, 1983; Carter &
Kulhawy, 1992):

(8.16)
Drilled shafts in rock 292

Fig. 8.3 Limit solution for expansion


of cylindrical cavity (after Carter &
Kulhawy, 1992)
Lateral load capacity of drilled shafts in rock 293

Fig. 8.4 (a) Components of rock mass


resistance; and (b) Calculation of
normal limit stress pL (after Zhang et
al., 2000).
Drilled shafts in rock 294

Fig. 8.5 Distribution of ultimate lateral


resistance with depth (after Zhang et
al., 2000).
where B is the diameter of the shaft; τmax is the maximum shearing resistance along the
sides of the shaft; and pL is the normal limit resistance.
For simplicity, τmax is assumed to be the same as the maximum side resistance under
axial loading and can be determined using the methods presented in Chapter 6.
To determine the normal limit stress pL, the strength criterion for rock masses
developed by Hoek and Brown (1980, 1988) is used. Assuming that the minor principal
effective stress σ′3 is the effective overburden pressure γ′z and the limit normal stress pL
is the major principal effective stress σ′1 [see Fig. 8.4(b)], we have, from Equation (4.64),
the following

(8.17)

where γ′ is the effective unit weight of the rock mass; z is the depth from the rock mass
surface; and mb, s and a are rock mass parameters as described in Chapter 4.
With the distribution of pult along the depth determined, the ultimate lateral load that
may be applied can be approximated by (see Fig. 8.5)

(8.18)
Lateral load capacity of drilled shafts in rock 295

Fig. 8.6 Comparison of estimated pult


and that from field shaft tests (after
Cho, 2002).
Cho (2002) used the method of Zhang et al. (2000) to estimate pult of field test shafts in
rock. The estimated values agree well with the field test data (see Fig. 8.6).

Example 8.1
A drilled shaft of diameter 1.0 m is to be installed 3.0 meters in siltstone. The rock
properties are as follows:
Unconfined compresive strength of intact rock, σc=15.0 Mpa
RMR=55

Determine the ultimate lateral load capacity of the shaft.

Solution:
From Table 4.5, mi=9 for siltstone.
Drilled shafts in rock 296

Using Equation (4.68),

Using Equation (6.8) and choosing Ψ=1.0 (lower bound), the side shear resistance can
be obtained as

Assuming that the effective unit weight of the rock mass is 13.0 kN/m3, the limit
normal stress pL can be obtained from Equation (8.17) as follows

Using Equation (8.18), the ultimate lateral load capacity can be obtained as

It need be noted that the ultimate lateral load capacity obtained above does not
consider the moment equilibrium of the shaft. The structural strength of the shaft should
also be checked when using the ultimate lateral load capacity obtained above in design.

8.4 CAPACITY OF DRILLED SHAFT GROUPS

The ultimate lateral load capacity of a drilled shaft group can be calculated in a similar
way to calculating the axial load capacity of a drilled shaft group, i.e.

(8.19)

where HultG is the ultimate lateral load of a drilled shaft group; N is the number of drilled
shafts in the group; Hult is the ultimate lateral load of a single isolated drilled shaft; and α
is the group efficiency factor. Table 8.2 lists the values of α recommended by the
American Association of State Highway and Transportation Officials (AASHTO, 1989).
Lateral load capacity of drilled shafts in rock 297

Equation (8.19) applies when the head boundary conditions of the single shaft and the
shaft group are the same. If the head boundary conditions of the single shaft and the shaft
group are different (e.g., the single shaft has a free-head boundary condition while the
shaft group has a flxed-head boundary condition because of the cap), a modification
factor should be added to Equation (8.19) to account for the difference in head boundary
conditions (Frechette et al., 2002):
HultG=αNHultR
(8.20)

where R is the modification factor to account for the difference in head boundary
conditions. For the case of a single shaft at a free-head boundary condition and a shaft
group at a fixed-head boundary condition, Frechette et al. (2001) recommended a R value
of 2.2 based on five case studies while Matlock and Foo (1976) recommended a R value
of 2.0 based on a single case study (Frechette et al., 2002).
If the drilled shafts are closely spaced, the drilled shaft group can be represented by a
group block and its ultimate load can be calculated using the methods described in
Section 8.3 by treating the group block as a big single shaft.

8.5 DISCONTINUUM METHOD

In Sections 8.3 and 8.4, the rock mass is treated as a continuum. Since most rocks contain
discontinuities, drilled shafts may fail due to the sliding of the rock blocks or wedges
along discontinuities (see Fig. 8.7). In such cases, the lateral resistance is only provided
by the shear resistance along the discontinuities and the weight of wedge bounded by the
shaft and the discontinuities. Obviously, the rock mass need be treated as a discontinuous
medium in order to obtain the lateral resistance provided by the wedges.
Table 8.2 Group efficiency factor a recommended
by AASHTO (1989).
Center-to-Center Shaft Spacing in Group Efficiency Factor
Direction of Loading α
3B 0.25
4B 0.40
6B 0.70
8B 1.00
Drilled shafts in rock 298

Figure 8.7 Sliding of rock blocks due


to laterally loaded shaft.
To (1999) developed a discontinuum method for determining the lateral load capacity of
drilled shafts in a jointed rock mass containing two or three discontinuity sets. The
method consists of two parts: a kinematic analysis and a kinetic analysis. In the kinematic
analysis, Goodman’s block theory (Goodman & Shi, 1985) is extended to analyze the
movability of a combination of blocks laterally loaded by a drilled shaft. Based on the
extended theory, a 2-dimensional (2D) graphical method was developed to select the
possible combinations of movable blocks. This 2D graphical method can be easily
implemented with CAD programs such as AutoCAD or with spreadsheet programs such
as Excel.
In the kinetic analysis, the stability of each kinematically selected movable
combination of blocks or wedges is analyzed with the limit equilibrium approach. This
analysis, similar to slope stability analysis, considers the axial and lateral forces exerted
by the drilled shaft in addition to the weight of the wedge and the shearing resistance
along the discontinuities. From the stability analysis, simple analytical relations were
developed to solve for the lateral load capacity of the drilled shaft.
The lateral load capacity can also be obtained by analyzing the load-displacement
response of a drilled shaft using the discrete element method (DEM) as described in
Section 9.5.
9
Lateral deformation of drilled shafts in rock

9.1 INTRODUCTION

For drilled shafts in rock to resist lateral loads, the design criterion in the majority of
cases is not the ultimate lateral capacity of the shafts, but the maximum deflection of the
shafts. Predicting the deformation of laterally loaded drilled shafts is, therefore, the most
important aspect in designing drilled shafts to withstand lateral loads.
To date, it has been customary practice to adopt the techniques developed for laterally
loaded piles in soil (Poulos, 1971a, b, 1972; Banerjee & Davies, 1978; Randolph, 1981)
to solve the problem of drilled shafts in rock under lateral loading (Amir, 1986; Gabr,
1993; Wyllie, 1999). However, the solutions for laterally loaded piles in soil do not cover
all cases for laterally loaded drilled shafts in rock in practice (Carter & Kulhawy, 1992).
Carter & Kulhawy (1992), therefore, developed a method for predicting the deformation
of laterally loaded drilled shafts in rock. This method treats the rock mass as an elastic
continuum and has been found to give reasonable results of predicted deflections only at
low load levels (20–30% capacity). At higher load levels, the predicted displacements are
too small (DiGioia & Rojas-Gonzalez, 1993). Reese (1997) developed a p-y curve
method for analyzing drilled shafts in rock under lateral loading. The major advantage of
the p-y curve approach lies in its ability to simulate the nonlinearity and nonhomogeneity
of the rock mass surrounding the drilled shaft. However, since it represents the rock mass
as a series of springs acting along the length of the shaft, the p-y curve approach ignores
the interaction between different parts of the rock mass. Also, the p-y curve approach
uses empirically derived spring constants that are not measurable material properties.
Advances in computer technology have made it possible to analyze laterally loaded piles
using three-dimensional (3D) finite element (FE) models. p-y curves (Hoit et al., 1997) or
sophisticated constitutive relations (Wakai et al., 1999) are usually used to represent the
soil or rock behavior in the 3D FE analyses. However, p-y curves have the limitations as
described above. As for sophisticated soil or rock constitutive relations, it is often
difficult if not impossible in design practice to obtain the parameters in the constitutive
relations.
Zhang et al. (2000) developed a nonlinear continuum method for analyzing laterally
loaded drilled shafts in rock. The method can consider drilled shafts in a continuum
consisting of a soil layer overlying a rock mass layer. The deformation modulus of the
soil is assumed to vary linearly with depth while the deformation modulus of the rock
mass is assumed to vary linearly with depth and then stay constant below the shaft tip.
The effect of soil and/or rock mass yielding on the behavior of shafts is considered by
assuming that the soil and/or rock mass behaves linearly elastically at small strain levels
and yields when the soil and/or rock mass reaction force exceeds the ultimate resistance.
Drilled shafts in rock 300

9.2 SUBGRADE-REACTION (p-y CURVE) APPROACH

Treating the rock as a series of springs along the length of the shaft (see Fig. 9.1), the
behavior of the shaft under lateral load can be obtained by solving the following
differential equation (Reese, 1997)

where Q is the axial load on the shaft; y is the lateral deflection of the shaft at a point z
along the length of the shaft; p is the lateral reaction of the rock; EpIp is the flexural
rigidity of the shaft; and W is the distributed horizontal load along the length of the shaft.
Equation (9.1) is the standard beam-column equation where the values of EpIp may
change along the length of the shaft and may also be a function of the bending moment.
The equation (a) allows a distributed load to be placed along the upper portion of a shaft;
(b) can be used to investigate the axial load at which a shaft will buckle; and (c) can deal
with a layered profile of soil or rock (Reese, 1997).
Computer programs, such as COM624P and LPILE, are available to solve equation
(9.1) efficiently. COM624P (version 2.0 and higher) and LPILEPLUS can also consider the
variation of EpIp with the bending moment (see O’Neill & Reese, 1999 for the detailed
procedure). To solve Equation (9.1), boundary conditions at the top and bottom of the
shaft also need be considered. For example, the applied shear and moment at the shaft
head can be specified, and the shear and moment at the base of the shaft can be taken to
be zero if the shaft is long. For short shafts, a base boundary condition can be specified
that allows for the imposition of a shear reaction on the base as a function of lateral base
deflection. Full or partial head restraint can also be specified. Other formula that are used
in the analysis are

(9.2)

(9.3)

(9.4)

where V, M and S are respectively the transverse shear, bending moment and deflection
slope of the drilled shaft.
The major difference between various methods lies in the determination of the
variation of p with y or the p-y curve, which are described below.
Lateral deformation of drilled shafts in rock 301

Fig. 9.1 Subgrade-reaction (p-y curve)


model of laterally loaded drilled shafts.

9.2.1 Linear analysis


For linear analysis, the relationship between rock reaction p and shaft deflection y at any
point along the shaft is assumed to be linear, i.e.,

(9.5)

where kh is the coefficient of subgrade reaction, in the unit of force/length3; and B is the
width or diameter of the shaft. Substituting Equation (9.5) into Equation (9.1) and
neglecting the influence of Q and W, the governing equation for the deflection of a
laterally loaded shaft with constant EpIp can be simplified as

(9.6)

Solutions to the above equation may be obtained analytically as well as numerically with
a computer program.
The analysis of the load-displacement behavior of a drilled shaft also requires
knowledge of the variation of kh along the shaft. A number of distributions of kh along the
Drilled shafts in rock 302

depth have been employed by different investigators, which can be described by the
following general expression proposed by Bowles (1996):
kh=Ah+Bhzn
(9.7)

where Ah, Bh and n are empirical constants which can be determined for a particular site
by working backward from the results of lateral shaft load tests.
If the rock is considered homogeneous with a constant kh down the length of the
drilled shaft, the deflection u (both y and u are used to denote lateral deflection in this
book) and rotation θ at the ground level due to applied load H and moment M can be
calculated by

(9.8a)

(9.8b)

where Lc is the critical length given by

(9.9)

It should be noted that Equation (9.8) is applicable only to flexible shafts, i.e., shafts
longer than their critical length defined by Equation (9.9). For non-flexible shafts,
solutions in closed-form expressions or in the form of charts are also available
(Tomlinson, 1977; Reese & Van Impe, 2001).

9.2.2 Nonlinear analysis


In general, the relationship between rock reaction p and shaft deflection y at any point
along a shaft is nonlinear. Kubo (1965) used the following nonlinear relationship for soil
between reaction p, deflection y, and depth z:
p=kzmyn
(9.10)

where k, m, and n are experimentally determined coefficients. Equation (9.10) can also be
used for rock if the corresponding coefficients k, m, and n can be determined.
Since Matlock (1970) developed a method for deriving the variation of p with y, or the
p-y curves, for soft clay, based on field test results, a number of methods for deriving p-y
curves for different soils have been developed. Some of them are listed below (for details,
the reader can refer to the listed references):
1.API RP2A (1982) or Reese et al. (1974) method for sand.
Lateral deformation of drilled shafts in rock 303

2. Bogard and Matlock (1980) method for sand.


3. API RP2A (1991) or O’Neill and Murchison (1983) method for sand.
4. API RP2A (1982) or Matlock (1970) method for soft clay.
5. API RP2A (1982) or Reese et al. (1975) method for stiff clay.
6. Integrated method for clay by Gazioglu and O’Neill (1984).
7. Pressuremeter methods for all soils (Robertson et al., 1982, 1986; Briaud & Smith,
1983; Briaud, 1986).
The method, developed by Reese (1997), specifically for calculating the p-y curves for
rock is described in the following section.

9.2.5 p-y curves for rock


Reese (1997) presented a p-y curve method for analyzing laterally loaded drilled shafts in
rock. The concepts and procedures for constructing the p-y curves for rock are as follows
(Reese, 1997):
(1) The secondary structure of rock, related to joints, cracks, inclusions, fractures, and
any other zones of weakness, can strongly influence the behavior of the rock and thus
need be taken into account when applying the method described in this section.
(2) The p-y curves for rock and the bending stiffness E0Ip for the shaft must both reflect
nonlinear behavior in order to predict loadings at failure.
(3) The initial slope Kmi of the p-y curves must be predicted because small lateral
deflections of shafts in rock can result in resistances of large magnitudes. For a given
value of compressive strength, Kmi is assumed to increase with depth below the ground
surface.
(4) The modulus of the rock Em, for correlation with Kmi, may be taken from the initial
slope of a pressuremeter curve. Alternatively, the correlations presented in Chapter 4
can be used to determine Em.
(5) The ultimate resistance pult for the p-y curves will rarely, if ever, be developed in
practice, but the prediction of pult is necessary in order to reflect nonlinear behavior.
(6) The component of the strength of rock from unit weight is considered to be small in
comparison to that from compressive strength, and therefore the weight of rock is
ignored.
(7) The compressive strength σc of the intact rock for computing pult may be obtained
from tests of intact specimens.
(8) The assumption is made that fracturing will occur at the surface of the rock under
small deflections; therefore, the compressive strength of intact rock specimens is
reduced by multiplication by αm to account for fracturing. The value of αm is assumed
to be 1/3 for RQD of 100 and to increase linearly to unity at RQD of zero. If RQD is
zero, the compressive strength may be obtained directly from a pressuremeter curve.

(a) Calculation of ultimate resistance pult of rock


The following expressions are used for calculating the ultimate resistance pult of rock
Drilled shafts in rock 304

(9.11a)

(9.11b)

where B is the diameter of the shaft; zm is the depth below the rock surface; σc is the
unconfined compressive strength of the intact rock; and αm is the strength reduction factor
considering that fracturing will occur at the surface of the rock under small deflections
and thus reducing the resistance of the rock.

(b) Calculation of the slope of initial portion of p-y curves


The slope of the initial portion of p-y curves, kmi, is estimated by
Kmi=kmiEm
(9.12)

where Em is the modulus of the rock (mass); and kmi is a dimensionless constant which
can be determined by

(9.13a)

kmi=500 zm≥3B
(9.13b)

Equation (9.13) is developed from experimental data and reflect the assumption that the
presence of the rock surface has a similar effect on kmi, as was shown for the ultimate
resistance pult.

(c) Calculation of p-y curves


Referring to Figure 9.2, the p-y curve consists of three portions. The initial and the third
portions are straight-lines and the second portion is a curve. The three portions can be
expressed by
First Portion: p=Kmiy; y≤yA
(9.14a)

(9.14b)

Third Portion p=pult


(9.14c)
Lateral deformation of drilled shafts in rock 305

in which ym=kmB, where km is a constant, ranging from 0.0005 to 0.00005, that serves to
establish overall stiffness of curves. The value of yA is found by solving the intersection
of Equations (9.14a) and (9.14b), and is shown by

(9.15)

Fig. 9.2 Sketch of p-y curve for rock


(after Reese, 1997).

(d) Comments
The equations described above for constructing the p-y curves for rock are based on
limited data and should be used with caution. An adequate factor of safety should be
employed in all cases; preferably, field tests should be undertaken on full-sized shafts
with appropriate instrumentation. If the rock contains joints that are filled with weak soil,
the selection of strength and stiffness must be site-specific and will require a
comprehensive geotechnical investigation. In those cases, the application of the method
presented in this section should proceed with even more caution than normal (Reese,
1997).
Cho et al. (2001) conducted lateral load tests on two drilled shafts embedded in
weathered Piedmont rock. These shafts were instrumented with inclinometers and strain
gauges. The field data obtained from the instrumented shafts were used to backcalculate
the p-y curves. A comparison of the back-calculated p-y curves with the p-y curves
predicted using the method of Reese (1997) shows that the method of Reese (1997)
significantly overestimates the resistance of the weathered rock.
Drilled shafts in rock 306

9.3 CONTINUUM APPROACH

The continuum approach assumes the soil and rock to be a continuum. Numerical
solutions were developed by assuming that the soil and rock are ideally elastic, first with
the boundary element method (Poulos, 1971a, b, 1972; Banerjee & Davies, 1978) and
second with the finite element method (Randolph, 1981). Most of these elastic solutions
were presented in the form of charts. Randolph (1981) published approximate but
convenient closed-form expressions for the response of flexible piles to lateral loading.
Considering the fact that the closed-form expressions of Randolph (1981) for the lateral
response of flexible piles in soils may not cover the ranges of material and geometric
parameters encountered in drilled shafts in rock, Carter and Kulhawy (1992) expanded
the solutions by Randolph (1981). The solutions of Carter and Kulhawy (1992) give a
reasonable agreement between measured and predicted displacements for drilled shafts in
rock at low load levels (20–30% capacity). At higher load levels, however, the predicted
displacements are too small (DiGioia & Rojas-Gonzalez, 1993). Zhang et al. (2000)
developed a nonlinear continuum approach for the analysis of laterally loaded drilled
shafts in rock. The approach can consider the effect of soil and/or rock mass yielding on
the behavior of shafts.

9.3.1 Linear continuum approach

(a) Approach of Poulos (1971a, b, 1972) and Poulos and Davis (1980)
By modeling the soil as an elastic continuum and idealizing the pile as an infinitely thin
strip of the same width and bending rigidity as the prototype pile, Poulos (1971a, b, 1972)
and Poulos and Davis (1980) obtained the solutions for laterally loaded piles using the
boundary element method. The solutions are presented in the form of charts and can be
used to predict the deflection of drilled shafts in rock.
For a free head drilled shaft, the lateral deflection u and rotation θ under lateral force
H and overturning moment M at ground surface are given by

(9.16a)

(9.16b)

where L is the length of the shaft; EmL is the deformation modulus of the rock mass at the
level of shaft tip; and IuH, IuM, IθH and IθM (note that IuM=IθH) are deflection and rotation
influence factors which are a function of the drilled shaft flexibility factor KR and the
rock mass non-homogeneity η:

(9.17)
Lateral deformation of drilled shafts in rock 307

(9.18)

where Em0 is the deformation modulus of the rock mass at the ground surface. A
homogeneous rock mass is represented by η=1, whereas η=0 represents a rock mass with
zero modulus at the surface. The deflection and rotation influence factors are plotted in
Figures 9.3 to 9.5 for values of η of 0 and 1. If the shaft is partially embedded, the
deflection of the free-standing portion due to shaft rotation and bending can be added to
the groundline deflection to obtain the deflection at the shaft head.
If the drilled shaft is fixed-headed, the horizontal deflection can be obtained by putting
θ =0 in Equation (9.16b) and substituting for the obtained moment in Equation (9.16a), as

(9.19)

Fig. 9.3 Deflection influence factor IuH


(after Poulos & Davis, 1980)
Drilled shafts in rock 308

Fig. 9.4 Deflection and rotation


influence factors IuM and IθH (after
Poulos & Davis, 1980)
Lateral deformation of drilled shafts in rock 309

Fig. 9.5 Rotation influence factor IθM


(after Poulos & Davis, 1980).
For a single raking shaft, Poulos and Madhav (1971) have shown that the force acting on
the shaft head may be resolved into axial and normal components and the shaft then
treated as a vertical shaft subjected to these forces and the applied moment.

(b) Approach of Randolph (1981) and Carter and Kulhawy (1992)


Randolph (1981) conducted a parametric study of the response of laterally loaded piles
embedded in an elastic soil continuum. The study was conducted using the finite element
method and the results were fitted with closed-form expressions from which the lateral
response of piles may be readily calculated. Considering the fact that the closedform
expressions for the lateral response of flexible piles in soils may not cover the ranges of
material and geometric parameters encountered in drilled shafts in rock, Carter and
Kulhawy (1992) expanded the solutions by Randolph (1981). The expressions were
derived from the results of finite element studies of the behavior of laterally loaded
drilled shafts in rock. For a drilled shaft wholly embedded in rock [Fig. 9.6(a)], the shaft
response can be calculated in the following way (Carter & Kulhawy, 1992):
(1) The shaft is considered flexible when

(9.20)
Drilled shafts in rock 310

where Ee is the effective Young’s modulus of the shaft

(9.21)

in which B and EpIp are respectively the diameter and flexural rigidity of the shaft; and
G* is the equivalent shear modulus of the rock mass

(9.22)

in which Gm and νm are respectively the shear modulus and Poisson’s ratio of the rock
mass.
The shaft response can then be obtained by the closed-form expressions suggested by
Randolph (1981), i.e.,

(9.23a)

(9.23b)
Lateral deformation of drilled shafts in rock 311

Fig. 9.6 (a) Drilled shaft wholly


embedded in rock; and (b) Drilled
shaft embedded in soil and rock.
(2) The shaft is considered rigid when

(9.24)

The shaft response can then be obtained by the following closed-form expressions

(9.25a)

(9.25b)

(3) The shaft can be described as having intermediate stiffness whenever the slenderness
ratio is bounded approximately as follows
Drilled shafts in rock 312

(9.26)

The finite element results show that the displacements for an intermediate
case exceed the maximum of the predictions for corresponding rigid and
flexible shafts by no more than about 25%, and often by much less. For
simplicity, it is suggested that the shaft displacement in the intermediate
case be taken as 1.25 times the maximum of either: (a) The predicted
response of a rigid shaft with the same slenderness ratio L/B as the actual
shaft; or (b) the predicted response of a flexible shaft with the same
modulus ratio (Ee/G*) as the actual shaft. Values calculated in this way
should, in most cases, be slightly larger than those given by the more
rigorous finite element analysis for a shaft of intennediate stiffness.

If there exists a layer of soil overlying rock as shown in Figure 9.6(b), Carter and
Kulhawy (1992) assume that the complete distribution of soil reaction on the shaft is
known and that the socket provides the majority of resistance to the lateral load or
moment. The groundline horizontal displacement u and rotation θ can then be determined
after structural decomposition of the shaft and its loading, as shown in Figure 9.7. To
determine the distribution of the soil reaction, they simply assume that the limiting
condition is reached at all points along the shaft, from the ground surface to the interface
with the underlying rock mass, and then use the reaction distribution suggested by Broms
(1964a, b).
For shafts through cohesive soils (Fig. 9.8), the lateral displacement uAO and rotation
θAO of point A relative to point O are given by

(9.27a)
Lateral deformation of drilled shafts in rock 313

Fig. 9.7 Consideration of soil reaction:


(a) Loading and displaced shape; and
(b) Decomposition of loading (after
Carter & Kulhawy, 1992).

(9.27b)

where Ls is the thickness of the soil layer; and su is the undrained shear strength of the
soil. The shear force Ho and bending moment Mo at point O are determined by
HO=H−9su(Ls−1.5B)B
(9.28a)
2
MO=M−4.5su(Ls−1.5B) B+HLs
(9.28b)

The contribution to the groundline displacement from the loading transmitted to the rock
mass can then be computed by analyzing a fully rock-socketed shaft of embedded length
L, subject to horizontal force HO and moment MO applied at the level of the rock mass.
For shafts through cohesionless soils (Fig. 9.9), the lateral displacement uAO and
rotation θAO of point A relative to point O are given by

(9.29a)

(9.29b)
Drilled shafts in rock 314

where γ′ is the effective unit weight of the soil; and Kp is the Rankine passive earth
pressure coefficient. The shear force HO and bending moment MO at point O are
determined by

(9.30a)

Fig. 9.8 Idealized loading of socketed


shaft through cohesive soil (after
Carter & Kulhawy, 1992).
Lateral deformation of drilled shafts in rock 315

Fig. 9.9 Idealized loading of socketed


shaft through cohesionless soil (after
Carter & Kulhawy, 1992).

(9.30b)

Example 9.1
A drilled shaft of diameter 1.0 m is to be installed 3.0 meters in siltstone. The rock
properties are as follows:
Unconfined compressive strength of intact rock, σc=15.0 Mpa
Deformation modulus of intact rock Er=10.6 GPa
RQD=70

Determine the lateral displacement and rotation of the drilled shaft at the groundline
by a horizontal force of 2.6 MN at 2.5 m above the groundline.

Solution:
Drilled shafts in rock 316

For simplicity, the Young’s modulus of the drilled shaft is simply assumed to be
Ep=30 GPa. A Poisson’s ratio of 0.25 is selected for both the drilled shaft and the rock.
The flexural rigidity of the shaft is

Using Equation (4.24), the deformation modulus of the rock mass is


αE=0.0231×70−1.32=0.297
Em=0.297×10.6=3.15 Gpa

and the shear modulus of the rock mass is


Gm=3.15/(1+0.25)=1.26 Gpa

Using Equation (9.22), the equivalent shear modulus of the rock mass is
G*=1.26×(1+3×0.25/4)=1.50 Gpa

Since

the shaft is considered flexible and the lateral displacement and rotation of the drilled
shaft at the groundline can be obtained from Equation (9.23) as

9.3.2 Nonlinear continuum approach


Poulos and Davis (1980) presented an approximate nonlinear approach for calculating the
deflection of laterally loaded piles in soil. This approach uses the elastic solutions
presented in the last section, but introduces yield factors. The yield factors are a function
of relative flexibility and load level and allow for the increased deflection and rotation of
a pile due to the onset of local yielding of the soil adjacent to the pile. This approach can
also be used to calculate the nonlinear deflection of drilled shafts in rock.
Lateral deformation of drilled shafts in rock 317

For a drilled shaft subjected to a lateral load H at an eccentricity of e above the


groundline, the groundline deflection u and rotation θ can be expressed as follows:
(1) Uniform modulus with depth, i.e., η=1.0

(9.31a)

(9.31b)

(2) Linearly increasing modulus with depth, i.e., η=0

(9.32a)

(9.32b)

where uelastic and θelastic are respectively deflection and rotation from elastic solutions as
described in the previous section; and Fu, Fθ, F′u, and Fθ are yield deflection and rotation
factors which can be found from Poulos and Davis (1980). The yield factors are functions
of a dimensionless load level H/Hu, where Hu is the ultimate lateral load capacity of the
equivalent rigid shaft and can be estimated using the methods presented in Chapter 8.
Zhang et al. (2000) developed a nonlinear continuum approach for the analysis of
laterally loaded drilled shafts in rock. This approach adopts and extends the basic idea of
Sun’s (1994) work on laterally loaded piles in soil. Sun’s model treats soil as a
homogeneous elastic continuum with a constant Young’s modulus, which may apply to
stiff clay, and it does not consider yielding of the soil. In the nonlinear approach
developed by Zhang et al. (2000), drilled shafts in a soil and rock mass continuum (see
Fig. 9.10) are considered, and the effect of soil and/or rock mass yielding on the behavior
of shafts is included. For simplicity, the shaft is assumed to be elastic, while the soil/rock
mass can be either elastic or elasto-plastic. It is, nevertheless, possible to also check
whether the shaft concrete will yield or not using standard concrete design methods, as
will be briefly mentioned later.

(a) Method of analysis—elastic behavior

Governing equations of shaft and soil/rock mass system


Consider a drilled shaft of length L, radius R and flexural rigidity EpIp, embedded within
a
Drilled shafts in rock 318

Fig. 9.10 (a) Shaft and soil/rock mass


system; (b) Coordinate system and
displacement components; and (c)
Shear force V(z) and moment M(z)
acting on shaft at z (after Zhang et al.,
2000).
soil/rock mass system (Fig. 9.10). The deformation modulus of the soil varies linearly
from Es1 at the ground surface to Es2 at the soil/rock mass interface. The deformation
modulus of the rock mass varies linearly from Em1 at the soil/rock mass interface to Em2 at
the shaft tip and stays constant below the shaft tip. For convenience of presentation, non-
uniformity indices defined by

(9.33)

(9.34)

are introduced. The increase of the deformation moduli of the soil and the rock mass with
depth, z, can then be expressed, respectively, by

(9.35a)
Lateral deformation of drilled shafts in rock 319

(9.35b)

Em=Em2 (z>Ls+L)
(9.35c)

By adopting the basic idea of Sun (1994), the displacements usm, νsm and usm of the soil
and/or rock mass can be approximated by separable functions of the cylindrical
coordinates r, θ and z as

(9.36a)

(9.36b)
wsm(r,θ,z)=0
(9.36c)

where u(z) is the displacement of the shaft as a function of depth; and is a


dimensionless function representing the variation of displacements of the soil and/or rock
mass in the r-direction.
For the displacements of Equation (9.36), the governing equations for the shaft can be
obtained as

(9.37a)

(9.37b)

with boundary conditions

(9.38a)

(9.38b)

(9.38c)

us−um=0 (z=Ls)
(9.38d)
Drilled shafts in rock 320

(9.38e)

(9.38f)

(9.38g)

(9.38h)

(9.38i)

where us and um are the displacement components u of the shaft in the soil and in the rock
mass, respectively; and ts, ks and tm, km are parameters that can be expressed as

(9.39a)

(9.39b)

(9.39c)

(9.39d)

where m1 and m2 are parameters describing the behavior of the elastic foundations, which
can be obtained by

(9.40a)

(9.40b)

Function can be obtained by solving the following equation


Lateral deformation of drilled shafts in rock 321

(9.41)

where γ is a nondimensional parameter that can be expressed as

(9.42)

The solution to Equation (9.41) that satisfies the unit condition at and
the finite condition at can be obtained and the parameters m1
and m2 can then be expressed as (Sun, 1994)

(9.43a)

(9.43b)

where K0( ) is the modified Bessel function of the second kind of zero order; and K1( ) is
the modified Bessel function of the second kind of first-order.
The shear force V(z) acting on the shaft (see Fig. 9.10) can be obtained by

(9.44a)

(9.44b)

and the bending moment M(z) acting on the shaft (see Fig. 9.10) can be obtained by

(9.45a)
Drilled shafts in rock 322

(9.45b)

The governing differential equations and the shear force V(z) and bending moment M(z)
are solved using the classical finite difference method as described below. At this point it
is also possible to check if the shaft concrete yields (recall that the basic assumption is
non-yielding concrete). This can be done using the calculated shear force and moment
together with the axial force on the shaft and using standard concrete design methods.

Finite difference model


The classical finite difference method (Desai & Christian, 1977) is employed to solve the
governing differential Equation (9.37). By dividing the shaft in the soil into Ns equal
segments (see Fig. 9.11) and using the central difference operator, for an interior node i
(i= 0, 1, 2,…, Ns), the following equation is obtained:

Fig. 9.11 Dividing shaft into segments


for finite-difference analysis, and
estimating reaction force p of soil and
rock from shear force V (after Zhang et
al., 2000).
Lateral deformation of drilled shafts in rock 323

(9.46)

where

(9.47)

in which hs=Ls/Ns.
Similarly, by dividing the shaft in the rock mass into Nm equal segments (see Fig.
9.11), the following equation is obtained for an interior nodey j(j=0, 1, 2,…, Nm):

(9.48)

where

(9.49)

in which hm=L/Nm.
Equations (9.46) and (9.48) can be written recursively for each point i=0, 1, 2,…, Ns
and j=0, 1, 2,…, Nm(see Fig. 9.11), resulting in a set of simultaneous equations in u. To
solve the set of equations the boundary conditions must be introduced. By incorporating
the boundaiy conditions expressed by Equation (9.38), the following finite difference
equations can be obtained:
at z=0

(9.50a)

(9.50b)
Drilled shafts in rock 324

−us(−1)−us(1)=0 (fixed-head)
(9.50c)

at z=Ls

(9.50d)

(9.50e)

(9.50f)

(9.50g)

at z=Ls+L

(9.50h)

(9.50i)

The set of equations [Equations (9.46) and (9.48)] is modified by introducing the
boundary conditions given in Equation (9.50). The resulting equations are solved
simultaneously for u by using the Gaussian elimination procedure.
After the shaft displacement u is obtained, the shear force V acting on the shaft can be
obtained from Equation (9.44) as:

(9.51a)

(9.51b)

You might also like