Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

3 BEARING CAPACITY

OF SHALLOW
FOUNDATIONS
ALEKSANDAR S. VESIC, D.Sc.
J. A. Jones Professor and Chairman
Departme~t of Civil Engineering, Duke University

3.1 INTRODUCTION cern in design: the structural failure of the foundation and
the bearing capacity failure of the supporting soils.
The designer of a shallow foundation must ensure that the The structural failure of the foundation may occur if the
foundation meets basic considerations of safety, depend- foundation itself is not properly designed to sustain the im-
ability, functional utility, and economy. Specifically, the posed stresses. For example, an improperly proportioned
foremost of these are the requirements of adequate depth, or inadequately reinforced footing or mat may fail in ten-
tolerable settlements, and safety against failure. sion, compression, or shear, as any other reinforced con-
The requirement of adequate depth centers around en- crete structural member. An improperly cast or inade-
vironmental influences which could affect adversely the quately reinforced concrete pile may be broken during
foundation performance. The foundation must be deep handling and driving; it may also be broken by excessive
enough with respect to the depth of frost penetration and lateral loads for which it was not designed. A free-standing
depth of seasonal volume changes in the soil to prevent steel or wooden pile may buckle just as any other column,
excessive movements resulting from these influences. The particularly if it is subjected to combined axial forces and
foundation must' be deep enough to exclude the possibility moments. All foundations must possess a conventional
of erosion and u'ndermining of the supporting soil by water safety against such failures. The analysis and design pre-
and wind currents. The foundation should also be ade- cautions needed to avoid structural failure are discussed
quately placed with respect to adjacent structures, existing later in Chapters 12 through 21.
or anticipated, to minimize the possibility of mutual dam- This chapter is devoted primarily to the discussion of the
age by construction operations or by transmission of addi- problem of bearing capacity failure of the soil. Consider
tional loads to the supporting soils. (For a detailed discus- the simplest case of a shallow footing (Fig. 3.1 a) subjected
sion of this subject see, for example, Terzaghi and Peck, to a vertical static or transient load Q. If the vertical dis-
1948; Tschebotarioff, 1951; Little, 1961). placement w of the foundation is recorded as the load Q
The requirement of tolerable settlements is concerned increases, a load-settlement curve similar in shape to a
with total and differential settlements of all foundations stress-strain curve may be obtained (Fig. 3.1 b ). The shape
under the planned structure. The differential settlements of this curve depends generally on the size and shape of the
must be limited in order not to cause structural distress or footing, the compo'sition of the supporting soil, and the
excessive tilting of the superstructure and they are also character, rate, and frequency of the loading. Normally,
often limited by the serviceability requirements of the the curve will indicate the ultimate load Q 0 that the foun-
superstructure; for example, crane tracks and many ma- dation can support. This can be either a peak load, as
chines have limits of tolerable tilt. The total settlements shown by curves 1 and 2 in Fig. 3.1 b, or a load at which a
must be limited because they invariably induce differential constant final rate of penetration is achieved, as shown by
settlements, even in apparently homogeneous soil condi- curve 3.
tions. They are also often limited by considerations of such The average pressure in ton/ft 2 or kg/cm 2 over the con-
factors as access to adjacent buildings, water and sewage tact area A of the foundation, corresponding to the ulti-
connections, etc. The methods of determination of settle- mate load Q 0 , is called ultimate pressure and will be de-
ments and the magnitude of settlement tolerances are dis- noted by q O • By definition
cussed in Chapter 4.
The requirement of safety against failure is centered (3.1)
around two principal kinds of failure that may be of con-

121
~~-~-------------------------=~·-··--··-
122 Foundation Engineering Handbook

Load

(a) General shear

(a)

Load Q/B 2 , ton/ft 2


0 5 10 15
l""lllllii;::=:;::::=--i----,---, (b) Local shear
...,
0::
0)
(.)

'"
0)
0.

~
"
i::
0)
5

s
0)

E
0)
Cl)

(c) Punching shear

(b)
Fig. 3.1 Load-settlement relationships of shallow footings.

All foundations should be built so as to possess a certain Fig. 3.2 Modes of bearing capacity failure. (After Vesic, 1963a).
safety against bearing capacity failure. The safe or allow-
able pressure qa is defined as
both sides of the footing, although the final soil collapse
occurs only on one side.
qa = qo (3.2) In contrast with the above-described failure mode,
Fs
punching shear failure is characterized by a failure pattern
where Fs is a safety factor. Similarly, the load Qa = Qo /Fs which is not easy to observe (Fig. 3.2c). As the load in-
is called allowable load. creases, the vertical movement of the footing is accompa-
The following sections are concerned mainly with de- nied by compression of the soil immediately underneath.
termination of the ultimate pressure q 0 . The procedures Continued penetration of the footing is made possible by
for selection of safety factors will be discussed toward the vertical shear around the footing perimeter. The soil out-
end of the chapter. side the loaded area remains relatively uninvolved and there
is practically no movement of the soil on the sides of the
footing. Both the vertical and the horizontal equilibrium of
3.2 MODES OF FAILURE the footing are maintained. Except for sudden small move-
ments ("jerks") of the foundation in the vertical direction,
It is known from observation of behavior of foundations there is neither visible collapse nor substantial tilting. Con-
subjected to load that bearing capacity failure occurs usu- tinuous increase in vertical load is needed to maintain foot-
ally as a shear failure of the soil supporting the footing. ing movement in the vertical direction.
The three principal modes of shear failure under founda- Finally, local shear failure is also characterized by a fail-
tions have been described in the literature as general shear ure pattern which is clearly defined only immediately be-
failure (Caquot, 1934; Buisman, 1935; Terzaghi, 1943), low the foundation (Fig. 3.2b). This pattern consists of a
local shear failure (Terzaghi, 1943; De Beer and Vesic, wedge and slip surfaces, which start at the edges of the
1958), and punching shear failure (De Beer and Vesic, footing just as in the case of general shear failure. There is
1958; Vesic, 1963a). visible tendency of soil bulging on the sides of the footing.
General shear failure is characterized by the existence of However, the vertical compression under the footing is sig-
a well-defined failure pattern consisting of a continuous slip nificant and the slip surfaces end somewhere in the soil
surface from one edge of the footing to the ground surface mass. Only after a considerable vertical displacement of the
(Fig. 3.2a). In stress-controlled conditions, under which footing (say up to one-half the width or diameter of the
most foundations operate (and, perhaps, fail) failure is sud- footing) may the slip surfaces appear at the ground surface.
den and catastrophic. Unless the structure prevents the Even then there is no catastrophic collapse or tilting of the
footings from rotating, the failure is also accompanied by footing which remains deeply embedded, mobilizing the
substantial tilting of the foundation (Fig. 3.3). In strain- resistance of deeper soil strata. Thus, local shear failure re-
controlled conditions (occurring, for example, when the tains some characteristics of both general shear and punch-
load is transmitted by jacking) a visible decrease of load ing modes of failure, representing truly a transitional mode.
necessary to produce footing movement after failure may A few photographs illustrating the characteristic failure
be observed (Fig. 3. 2a). A tendency for bulging of adjacent modes are shown in Figs. 3.3 through 3.9. Figure 3.3,
soil can be recorded through most of the loading process on taken from Tschebotarioff (1951 ), shows the view of a
Bearing Capacity of Shallow Foundations 123

Fig. 3.3 Bearing capacity failure of a silo foundation. (From Tschebotarioff, 1951.)

Fig. 3.4 General shear failure pattern under a rectangular footing Fig. 3.6 Punching shear failure pattern under a rectangular founda-
on dense sand (Dr= 100%). (From De Beer and Vesic, 1958.) tion on the surface of loose sand (Dr= 15%). (From De Beer and
Vesic, 1958.)

grain elevator after general shear failure of the firm clay


under its foundation. In a similar case described in detail in
the literature (White, 1953; Peck and Bryant, 1953), the
structure, with its foundation, tilted more than 60° at col-
lapse of the underlying soil. The latter bulged some 12 feet
above the initially horizontal ground surface. Thanks to
extreme rigidity of the concrete structure, there was rela-
tively little structural damage. (The elevator was later
straightened into a vertical position and put to normal use.)
However, another elevator which failed under somewhat
similar circumstances was practically destroyed during col-
lapse (Nordlund and Deere, 1970). Figures 3.4, 3.5, and
3.6 show the three characteristic modes of failure, as ob-
served on carefully prepared models of long rectangular
plates resting initially on the surface of dry sand of differ-
ent densities (De Beer and Vesic, 1958). Figures 3.7, 3.8
Fig. 3.5 Local shear failure pattern under a rectangular footing on and 3.9 show the punching shear failures under small plates
medium dense sand !Dr= 47%). (From De Beer and Vesic, 1958.) on dense, dry sand, as observed, respectively, in conditions
124 Foundation Engineering Handbook

Fig. 3.9 Punching shear failure of a rectangular footing on dense


sand underlain by soft clay. (From Vesic, 1970.)

failure mode depends on the relative compressibility of the


soil in the particular geometrical and loading conditions. If
Fig. 3.7 Punching shear failure under a deep rectangular founda-
the soil is practically incompressible and has a finite shear-
tion in dense sand (Dr= 90%, B = 1.5 in, D = 15 in). (From Vesic,
ing strength, it will fail in general shear. On the other hand,
if a soil of given strength is very compressible, it will fail in
1963a.)
punching shear. Thus, as shown in Fig. 3,4, a footing on
the surface of very dense sand will normally fail in general
of deep embedment (Vesic, 1963a), very rapid loading shear, while the same footing on the surface of very loose
(Poplin, 1965) or presence of soft clay layer underneath sand (Fig. 3.6) will fail in punching shear. However, it is
(Vesic, 197 O). important to understand that the soil type alone does not
Which mode of failure is to be expected in any particular determine the mode of failure. For example, the mentioned
case depends on a number of factors that have been only footing on very dense sand can also fail in punching shear if
partially explored so far. It can be said generally that the the footing is placed at greater depth (Vesic, 1963a; see
Figs. 3. 7 and 3 .1 0) or if it is loaded by a transient, dynamic
load (Heller, 1964; Vesic, Banks, and Woodard, 1965; see
Fig. 3.8). * Similarly, the same footing will fail in punching
shear if the very dense sand below is underlain by any com-
pressible stratum such as loose sand or soft clay (Fig. 3.9).
Also, a footing on saturated, normally consolidated clay
will fail in general shear if it is loaded so that no volume
change can take place; while it may fail in punching shear if
it is loaded slow enough that all volume change can take
place in the soil under load.
While these differences in failure modes are now reason-
ably well understood, there are at present no general nu-
merical criteria that can be used for prediction of mode of
shear failure of soils loaded by footings. The only rational
parameter proposed so far for evaluation of relative com-
pressibility of soil masses under load is the rigidity index
Ir defined as

*There are reasons to believe that the mentioned footing on very


Fig. 3.8 Punching shear failure under a dynamically loaded 8-inch- dense sand would not fail in general shear if its diameter were in-
square footing on dense sand (Dr= 100%). Failure occurred in 12 creased beyond a certain limit. We shall return to this later in dis-
msec at a displacement of 0.3 in. (From Poplin, 1965.) cussing scale effects.
Bearing Capacity of Shallow Foundations 125

Relative density of sand. l),


(3.4)
o~----o~.2--~-o~A____o~.6--~-~o.s____~1.o
It is known that the rigidity index varies with stress level
and with the character of loading (Vesic, 1965a). A high
value of Irr, say over 250, definitely implies a relatively
incompressible soil mass, whereas a low value, say 10, im-
plies a relatively compressible soil mass. Nevertheless, in
the absence of theoretical solutions for an elastic-plastic
solid, there is no way yet, other than semiempirical, to
make use of index Ir in predicting the failure mode of shal-
low footings. Some possibilities along these lines are out-
Local shear lined later in this chapter, in discussions of influence of soil
,.2
compressibility.
0.
.g 3 3.3 UL Tl MATE LOAD CRITERIA
c;
> Punching shear

" From the preceding discussion it is apparent that the "fail-


~ ure" of a loaded footing is clearly defined only in the case
of general shear failure. In such a case the peak, ultimate
4
load is reached simultaneously with the appearance of slip
lines at the ground surface, which is followed by founda-
tion collapse and a considerable bulging of the soil mass on
the side of the footing. In contrast with this, in the case of
5'------'------'------'----~---~ two other failure modes (local and punching shear) the
B*= B for a square or circular footing
B*= BL/2(B+L) for a rectangular footing point of failure is less clearly defined and often difficult to
establish.
Fig. 3.10 Modes of failure of model footings in sand. (After Vesic, In the case of punching or local shear failure of footings
1963a, as modified by De Beer, 1970.) on sand surface it has been observed (De Beer and Vesic,
1958) that a "first failure" characterized by a sudden, large
plastic deformation of the soil under the footing, may oc-
I = G (3.3) cur rather early in the loading stage. However, to observe
r (c + q tan ef>)
this "first failure" it is essential to have a stress-controlled
where G is the shear modulus and c and ef> are strength pa- loading. As most loading tests are performed by the use of
rameters of the soil (Vesic, 1963b, 1965b). This index, ap- hydraulic jacks, this "first failure" cannot be noticed with
pearing in solutions of the problem of expansion of cavities any certainty and thus has limited practical value.
in an infinite solid, is associated with the assumed ideal A very versatile ultimate load criterion that can be rec-
elastic-plastic behavior of soil. To take care also of the aver- ommended for general use defines the ultimate load as the
age volumetric strain Ll in the plastic zone it has been sug- point where the slope of the load-settlement curve first
gested (Vesic et al., 1965) that the value given by Eq. 3.3 reaches zero or a steady, minimum value (Vesic, 1963a; see
should be reduced to Irr= rvir, where also Fig. 3.11 ). Another very consistent ultimate load cri-

Base pressure, lb/ in 2

0.5

Ultimate load
1.0
~
<I)
..c::
u
.s
-;::; l.5
<I)

E
E
Cl)
<I)

2.0 Circular plates on sand


Base diameter, 6 in
Tests at surface
Test No. Yd, lb/ft 3 D,
2.5 61 96.2 0.78
61
62 93.0 0.66
62 63 91.7 0.61
. .___ __,._6_4_ __,.__ _-=- 64-'----s-s......o__o_.3_2~
63:....i...._ _ _..___ _- ' - -_ _ _
30
Fig. 3.11 Ultimate load criterion based on minimum slope of load-settlement curve. (After Vesic, 1963a.)

L.
126 Foundation Engineering Handbook

pears to be a general tendency of increase of ultimate settle-


Q/(AyB) ments with increased size of footings (De Beer, 1965a,
2
1~_ _ _2,__ _~4--6~~8~10_ _ _2T0_ _ _4~0-~6~0_;;,80;;__;.;!00 1967).
Considering these facts, unless a clearly defined ultimate
load can be observed earlier, it is advisable to carry load
tests on footings and plates on loose and compressible soils
4
to settlements equal to at least 25 percent of the footing
width. In cases in which a peak load cannot be established
with certainty, it is expedient to adopt a limit of critical
settlement, such as 10 percent of the footing depth. (The
same limit has been often proposed and used for driven
c piles; see Vesic, 1967.)
E
E
~ 20
V
~ 3.4 COMPUTATION OF UL Tl MATE LOAD
~

~
40 The computation of ultimate load for a shallow footing
resting on soil represents a problem of elastic-plastic equi-
60 All tests: Mo! sand librium which can, in principle, be solved for at least the
Footing dia. 3.8 cm
80 q=O.lkg/cm 2
plane-strain and the axially symmetric cases. The foremost
I00 '-----'-----'----''----'-_,__ __.__ _ _u__1--....1_-J difficulty in finding acceptable solutions lies undoubtedly
in the selection of a mathematical model of soil behavior or
Fig. 3.12 Ultimate load criterion based on plot of log load vs. log its constitutive (stress-strain-time) relationships. In spite of
settlement. (After De Beer, 1967.) greatly improved capabilities for solution of boundary value
problems of this kind, the theory of bearing capacity is still
terion, suggested in recent years by Christiaens (De Beer, limited almost exclusively to solutions developed for the
1967), defines the ultimate load at the point of break of rigid-plastic solid of the classical Theory of Plasticity. As
the load-settlement curve in a log/log plot (Fig. 3.12). Both known, this solid is assumed to exhibit no deformation
criteria require, however, that the loading test be carried to whatsoever prior to shear failure and a plastic flow at con-
very large displacements, preferably of the order of 50 per- stant stress after failure. Thus, the capabilities of theoreti-
cent of the foundation size. Thus, from the practical point cal prediction of the ultimate load are, strictly speaking,
of view, it may be preferable to establish some other cri- limited at present to relatively incompressible soils or to the
terion of critical settlement. Such a criterion is no doubt general shear failure mode. However, it is rather common
justified by the basic philosophy of foundation design, in practice to use the available solutions for compressible
which considers excessive settlement as failure of the soils as well, with possible reduction for the effects of
foundation. compressibility.
It is thus of special interest to know the magnitude of The problem is generally posed as follows (Fig. 3.14a).
settlements of footings needed to mobilize ultimate loads. Consider a rectangular foundation of width B and length L,
Observations in saturated clays (Skempton, 1951) indicate resting in a soil mass at a depth D. The soil mass is of semi-
that these settlements may be about 3 to 7 percent of the infinite extent and is homogeneous. It has an effective unit
footing width for surface footings, increasing up to 15 per- weight 'Y and shear strength properties defined by a straight-
cent for deep footings. For footings in sand somewhat line Mohr envelope, with strength characteristics c and ¢,
higher values are found, ranging from 5 to 15 percent for and a stress-strain curve of a rigid-plastic body, shown in
surface footings and as high as 25 percent for deep footings Fig. 3.14b. To be determined is the maximum unit load
(Muhs and Kahl, 1954; De Beer and Vesic, 1958; Vesic, q 0 = Q 0 /BL which this foundation can support.
1963a; De Beer, 1967). As shown in Fig. 3.13, there ap- To solve this problem, the following simplifications are
usually made:
yB*, kg/cm 2 a) the shearing resistance of the overburden soil (along
0.005 0.0 IO 0.01 5 0.020 0.025 0.030 be, Fig. 14a) is neglected;
b) the friction between the overburden soil and the
¢ 3.8 cm
foundation (along ad, Fig. 3.14a) as well as between the
overburden and supporting soil (along ab, Fig. 3.14a) is
neglected;
c) the length L is assumed to be large in comparison
t 10
with the width B of the foundation.
In other terms, the overburden soil is replaced by a uni-
formly distributed surcharge q = "(D. At the same time
15
plane strain conditions are assumed.
Simplifications (a) and (b) are justified in most cases and
are always on the safe side. The overburden soil is usually
weaker or cracked, while the foundation is placed by exca-
vation and backfilling. Simplification (c), equivalent to
assuming the foundation to be an infinite strip of width B,
~s~--~---~---~--~---~--~ is justified, strictly speaking for L/B > 10 and practically
Circular footings: H* = fJ for L/B > 5. The corrections to be introduced forL/B < 5
Rl'ctangular footings: lJ* = IJL/2 (/J+L ): L/8 6
and shapes other than a rectangle will be discussed later.
Fig. 3.13 Ultimate settlements of surface footings. (After De Beer, The problem, formulated as shown in Fig. 14c, has been
1967.) solved by the methods of the Theory of Plasticity. The
Bearing Capacity of Shallow Foundations 127

Q
TABLE 3.1. BEARING CAPACITY FACTORS

<I> tan</>

0 5.14 1.00 0.00 0.20 0.00


1 5.38 1.09 0.07 0.20 0.02
2 5.63 1.20 0.15 0.21 0.03
3 5.90 1.31 0.24 0.22 0.05
--'~!""~-~- - - - - - _ __.'.' 9' (a)
4 6.19 1.43 0.34 0.23 0.07
r c.4>,y
5 6.49 1.57 0.45 0.24 0.09
6 6.81 1.72 0.57 0.25 0.11
7 7.16 1.88 0.71 0.26 0.12
8 7.53 2.06 0.86 0.27 0.14
9 7.92 2.25 1.03 0.28 0.16
10 8.35 2.47 1.22 0.30 0.18
er

~l:::=-· t=.
T
11 8.80 2.71 1.44 0.31 0.19
12 9.28 2.97 1.69 0.32 0.21
(b) 13 9.81 3.26 1.97 0.33 0.23
14 10.37 3.59 2.29 0.35 0.25
0 15 10.98 3.94 2.65 0.36 0.27
16 11.63 4.34 3.06 0.37 0.29
17 12.34 4.77 3.53 0.39 0.31
18 13.10 5.26 4.07 0.40 0.32
Q, 19 13.93 5.80 4.68 0.42 0.34
20 14.83 6.40 5.39 0.43 0.36

if 21 15.82 7.07 6.20 0.45 0.38


22 16.88 7.82 7.13 0.46 0.40

'
Ill
E

(c)
23
24
25
18.05
19.32
20.72
8.66
9.60
10.66
8.20
9.44
10.88
0.48
0.50
0.51
0.42
0.45
0.47
26 22.25 11.85 12.54 0.53 0.49
c.<P. y 27 23.94 13.20 14.47 0.55 0.51
28 25.80 14.72 16.72 0.57 0.53
Fig. 3.14 The problem of bearing capacity of shallow footings. 29 27.86 16.44 19.34 0.59 0.55
30 30.14 18.40 22,40 0.61 0.58
31 32.67 20.63 25.99 0.63 0.60
basic solution available (Prandtl, 1921; Reissner, 1924) in- 32 35.49 23.18 30.22 0.65 0.62
dicates that the failure pattern should consist of three 33 38.64 26.09 35.19 0.68 0.65
zones: I, II, and III. Zone I is an active Rankine zone, 34 42.16 29.44 41.06 0.70 0.67
which pushes the radial Prandtl zone II sideways and the 35 46.12 33.30 48.03 0.72 0.70
passive Rankine zone III in an upward direction. The lower 36 50.59 37.75 56.31 0.75 0.73
boundary ACDE of the displaced soil mass is composed of 37 55.63 42.92 66.19 0.77 0.75
two straight lines AC and DE, inclined at 45° + <P/2 and 38 61.35 48.93 78.03 0.80 0.78
45° - </)/2, respectively, to the horizontal. The shape of the 39 67.87 55.96 92.25 0.82 0.81
connecting curve CD depends on the angle <P and on the 40 75.31 64.20 109.41 0.85 0.84
ratio rB/q. For rB/q ~ 0 ("weightless soil") the curve be- 41 83.86 73.90 130.22 0.88 0.87
comes a logarithmic spiral which for 'Y = 0 degenerates into 42 93.71 85.38 155.55 0.91 0.90
a circle. In the general case ('YB =f. 0) the curve lies between 43 105.11 99.02 186.54 0.94 0.93
a spiral and a circle, as long as <P =f. 0. For a frictionless soil 44 118.37 115.31 224.64 0.97 0.97
(</) = 0) the curve is always a circle. All these findings have 45 133.88 134.88 271.76 1.01 1.00
been confirmed experimentally (De Beer and Vesic, 1958), 46 152.10 158.51 330.35 1.04 1.04
though the angle i/1 may be slightly larger than 45° + ¢/2, at 47 173.64 187.21 403.67 1.08 1.07
least for long rectangular footings on the surface of sand. 48 199.26 222.31 496.01 1.12 1.11
A closed analytical solution of this problem, as posed, 49 229.93 265.51 613.16 1.15 1.15
has not yet been found and probably will not be found, ex- 50 266.89 319.07 762.89 1.20 1.19
cept for special cases. For weightless soil ('Y = O), Prandtl
and Reissner have found that:
can be shown that:
Qo = cNe + qNq (3.5)
(3.7)
where Ne and Nq are dimensionless bearing capacity factors,
defined by where N'Y is again a dimensionless bearing capacity factor
which can be evaluated only numerically. This factor varies
Nq = etr tan <I> tan2 (rr/4 + ¢/2) sharply with angle i/1. The numerical values shown by
Ne = (Nq - 1) cot¢ (3.6) dashed lines in Fig. 3.15 are taken from an analysis made
by Caquot and Kerisel (1953) under assumption that i/1 =
The numerical values of these factors are given in Table 3 .1 45° + </)/2. It can be shown (Vesic, 1970) that these values
and shown graphically in Fig. 3.15. of N'Y · can be approximated with an error on the safe side
For cohesionless soil without overburden (c = 0, q = O) it (not exceeding 10 percent for 15° < <P < 45° and not ex-
128 Foundation Engineering Handbook

unsettled, because of difficulties in selecting a representa-


BOO..---~~~~~~~~~~~~~~~~~~~~-..
tive value of angle of shearing resistance </> for the bearing
Nq = e "''"°'tan (45° +q,/2)
2

600 capacity computations. Some authors (Meyerhof, 1963;


Nc=(Nq -1) cot</>
Brinch Hansen, 1970) use the plane-strain value of</>, which,
N~= 2(Nq +!)tan <f,
400 according to their views, may be up to IO percent higher
than the corresponding conventional triaxial test value.
300
This can help explain the results of tests with long rectangu-
200
lar plates on soil surface; however, it only contributes to
difficulties in interpreting the results of tests with circular
plates placed at a certain depth. (At any rate, there remains
the question as to what degree the conditions of soil ele-
100 ments along a slip surface under a circular footing are closer
80 to a2 a 3 than to plane-strain conditions.)
60 Much stronger is the argument that a shear failure in soil
under the footing is a phenomenon of progressive rupture
< 40 at variable stress levels (Muhs, 1963; De Beer, 1965a, b,
i
t 30
1967). Consequently when the slip line ACDE in Fig. 3.14
,.';: reaches E, just mobilizing the peak shearing strength at that
c point, the soil strength at the beginning of the slip line
·~ 20
B (point A) must be well below the peak. In addition, the
00
stress level at A is higher than at E. Thus, in view of known
·~ curvature of Mohr's envelope in the low stress range, the
0:, IO
</>-angle at A must be lower than at E. A representative
8
</>-value must be sought with due consideration of these
6 facts. Working along these lines, it has been suggested (De
Beer, 1965a) that the bearing capacity evaluations should
4 be made using strength characteristics corresponding to an
average mean normal stress equal to

2
ao =! (qo + 3q)(I - sin</>) (3.10)
While the discussions and investigations of these and
other questions associated with evaluation of the bearing
1.0 capacity continue, there is an increased trend among both
0.8 practicing engineers and researchers to retain the Prandtl-
0.6
Reissner and Caquot-Kerisel factors given in Table 3.1 as
the most reliable factors available at present. The widely
0.4 used Terzaghi factors, though not substantially different
0 IO 15 20 25 30 35 40 45 50 numerically, are being gradually abandoned, as they are
Angle of shearing resistance <p, degrees based on obviously incorrect failure patterns.
Fig. 3.15 Bearing capacity factors for shallow footings.

3.5 EFFECT OF FOUNDATION SHAPE


ceeding 5 percent for 20° < </> < 40°) by the analytical
expression
For foundation shapes other than the long rectangle, the
N'Y ~ 2(Nq + 1) tan</> (3.8) mathematical difficulties in obtaining solutions are consid-
erable. Only the axially-symmetrical case of a circular foot-
The N'Y-values according to 3.8 are given in Table 3.1 and ing has been solved so far (Ishlinskii, 1944; Berezantsev,
also presented graphically in Fig. 3.15. 1952; Mizuno, 1953; Eason and Shield, 1960; Cox, Eason,
For all intermediate cases, where c =I= 0, q =I= 0, and "f =I= and Hopkins, 1961 ). The proposed solutions make use of
0, Eqs. 3.5 and 3.7 are combined into some assumptions about soil behavior that remain to be
proved experimentally ("complete plasticity") and their re-
(3.9) sults are, at least in part, at some variance with observations
known as Buisman-Terzaghi equation (Buisman, 1940; (cf. Hansen and Christensen, 1969).
Terzaghi, 1943 ). This superposition is not strictly correct; In view of these facts, the engineering approach to eval-
however, it leads to errors which are on the safe sid~, not uation of effect of foundation shape has been mostly semi-
exceeding 17 to 20 percent for </> = 30° to 40°, while equal empirical. On the basis of comparative loading tests with
to O for </> = 0° (Lundgren and Mortensen, 1953; Hansen footings of different shapes, including long rectangles, the
and Christensen, 1969). following modification of Eq. 3.9 has come into general
It should be mentioned here that there exists in the liter- use:
ature a great variety of proposed solutions to this problem. (3.11)
While the variations in N c- and N q-values proposed remain
relatively insignificant, the differences in N 'Y-values, coming In this expression Ne, Nq, and N'Y are, as before, bearing
primarily from the mentioned sharp variation of N'Y with capacity factors for an infinite strip or long rectangle and
i/1, are substantial, ranging from about one-third to double tc, tq, hare dimensionless parameters called shape factors.
the values shown in Table 3.1. Shape factors also depend on the angle of shearing re-
In spite of the intensified experimental work on this sub- sistance </> of the soil, as well as on some other parameters.
ject, the question of "correct" N'Y-values remains somewhat However, many of them are taken to be constants or simply

L
Bearing Capacity of Shallow Foundations 129
I
TABLE 3.2. SHAPE FACTORS FOR SHALLOW' saturated unit weight of 118 lb/ft 3 and an average moist
FOUNDATIONS. unit weight above the water table of 100 lb/ft 3 . Drained
(After De Beer, 1967, as modified by Vesic, 1970). triaxial tests on sand samples show that the angle </) of
shearing resistance of sand varies with mean normal stress
Shape of a 0 according to the equation
the Base te l;q I;'Y </> = </>1 - (5.5°) log10 (a/a1)
Strip 1.00 1.00 1.00 where a 1 = 38° is the angle of shearing resistance at mean
Rectangle 1 + (BIL )(NqlNe) 1 + (BIL) tan <I> 1 - 0.4BIL
normal stress a 1 = 1 ton/ft 2 •
Circle and
Square 1 + (NqlNe) 1 +tan</> 0.60 Solution:
Submerged unit weight of sand: r' = 118- 62 = 56 lb/ft 3
Overburden stress: q = (8)(100) + (2)(56)/(2000) = 0.456
functions of the geometrical form of the foundation. Rec- ton/ft 2
ommended expressions of shape factors based primarily on To find the mean normal stress, according to Eq. 3.10, a pre-
extensive experiments at Ghent (De Beer, 1967) are given liminary estimate of bearing capacity is needed. It is as-
in Table 3.2. Numerical values of Nq/Ne and tan</> appear- sumed for this preliminary analysis that </) = 34 °.
ing in these expressions are given in Table 3 .1. Bearing capacity factors (Table 3.1): Nq = 29.44; N'Y =
41.06
EXAMPLE 3.1: A rectangular footing 28 feet wide and Shape factors: tq = 1 + (1/3)(0.67) = 1.22; t'Y = 1 -
84 feet long is to be placed at a depth of 10 feet in a deep 0.4(1/3) = 0.87
stratum of soft, saturated clay (bulk unit weight 105 Ultimate bearing pressure (Eq. 3.11):
Ib/ft 3 ). The water table is at 8 feet below ground surface.
Find the ultimate bearing capacity under the following two q0 = (0.456)(29.44)(1.22)
conditions: + (1/2)(56)(28) ( 41.06)(0.87)/(2000)
(a) assuming that the rate of application of dead and
live loads is fast in comparison with the rate of dissipation = 16.4 + 14.0 = 30.4 ton/ft 2
of excess pore-water pressures caused by loads, so that un-
Mean normal stress along the slip surface (Eq. 3.10):
drained conditions prevail at failure;
(b) assuming, as the other extreme, that the rate of a 0 = (1/4)[30.4+(3)(0.456)](1- 0.559)=3.5 ton/ft 2
loading is slow enough that no excess pore-water pressures
are introduced in the foundation soil. Representative angle of shearing resistance:
The strength parameters of the soil, obtained from un-
consolidated, undrained tests are Cu = 0.22 ton/ft 2 , <f>u = 0.
</> = 38° - (5.5°)(0.544) = 35°
Consolidated, drained tests give cd = 0.04 ton/ft 2 , <!>d = 23°. The analysis is now repeated with</)= 35°:
Condition (a): Nq = 33.3; N-y = 48.0; tq = 1 + (1/3)(0.70)
Submerged unit weight of soil: r' = 105 - 62 = 43 lb/ft 3
Overburden stress: q = [(8)(105) + (2)(43)] /(2000) = = 1.23; r.,, = 0.87;
0.463 ton/ft 2
Bearing capacity factors (Table 3.1): Ne = 5.14,Nq= 1,
q0 = (0.456)(33.3)(1.23)
N-y = 0 + (1 /2)(56)(28)( 48.0)(0.87)/(2000)
Shape factors (Table 3.2): te = 1 + (1/3)(0.20) = 1.067,
tq = 1.00 = 18.7 + 16.4 = 35.1 ton/ft 2
Ultimate bearing pressure (Eq. 3 .11):
In view of small change in mean normal stress from the pre-
qo = (0.22)(5 .14 )( 1.067) + (0.463)(1 )( 1.00) viously found value, this answer is retained.
= 1.21 + 0.46 = 1.67 ton/ft 2 Remarks: ( 1) The analysis of compressibility effects for
this case is presented in Example 3.8.
Condition (b) (2) Because of high value of ultimate bearing capacity
Bearing capacity factors: Ne= 18.05; Nq = 8.66;N'Y = it is possible that the allowable bearing pressure may be
8.20 controlled by maximum tolerable settlement for the
Shape factors: te = 1 + (1/3)(0.48) = 1.16; tq = 1 + structure in question.
0/3)(0.42) = 1.14; r'Y = 1 - co.4)(1/3) = o.87
Ultimate bearing pressure:
qo = (0.04)(18.05)(1.16) + (0.463)(8.66)(1.14) 3.6 EFFECT OF INCLINATION AND ECCENTRICITY
OF THE LOAD
+ (1/2)(43)(28)(8.20)(0.87)/(2000)
= 0.72 + 4.57 + 2.14 = 7.43 ton/ft 2 The preceding discussions were all concerned with a footing
loaded by a central, vertical load. If the load is inclined or
Remark: The computed values of q 0 represent the upper eccentric, or, as it most often happens, both inclined and
limit of bearing capacity under the assumption of incom- eccentric, the problem is somewhat more complicated be-
pressible soil. The analyses of effects o.f compressibility cause of the presence of the horizontal component P of the
for this case are presented in Example 3 .7. footing reaction (Fig. 3. l 6a). Failure can occur either by
sliding of the footing along its base AB, or by general shear
EXAMPLE 3.2: Solve the problem described in Example of the underlying soil.
3.1 if the footing is placed at the same depth (10 feet) in a At the verge of sliding the horizontal component P is
deep stratum of medium dense sand. Assume for sand a related to the vertical component Q of the footing reaction
130 Foundation Engineering Handbook

hof, 1953).* At the same time, it may be found convenient


to introduce the influence of the load inclination by multi-
plying the individual terms of the bearing capacity equation
3.11 by the inclination factors r ei, r qi, r'YdSchultze, 1952;
Brinch Hansen, 1961 ). Thus the bearing capacity equation
for the general case of eccentric and inclined loading can be
written in the form:

Qo = Bql, = cNererei + qNqrqrqi + }yBN'Yr'Yr'Yi (3.13)

where L' = L - 2e L is the effective length of the footing, in-


troduced to take into accourit possible eccentricity eL of
(a)
the load in the longitudinal direction.
Based on rigorous analyses for the plane-strain problem
of a footing on weightless soil loaded by a central, inclined
load (Fig. 3.16b ), Brinch Hansen (1961) proposed the fol-
lowing approximate expression for inclination factor rq(

rqi =[1 - Q + B':,c cot (/>r (3.14)

From the theorem of correspondence one finds also (De


Beer and Ladanyi, 1961)
1 - rqi
rci = rqi - Ne tan 1> (3.15)
(b)
Fig. 3.16 Theoretical slip patterns under eccentric and inclined It can be easily shown (Vesic, 1970) that, for 1> = 0, the lat-
loads. ter expression becomes
2P
1 1
(3.16)
by the expression: B L c Ne
P max = Q tan q> +A' Ca (3.12) Similar analyses for a footing on the surface of a soil
with weight (Sokolovskii, 1960) suggest the following val-
where A' is the effective bearing area of the foundation, ues for the load inclination factor r'Y(
while Ca and 1> represent, respectively, the adhesion and the
angle of friction between the soil and the footing. (It
should be noted that the presently available evidence- hi=[l-Q+B':ccotq>r (3.17)
Schultze and Horn, 1967-indicates that fi = 1>t and that
the adhesion in soft clays is equal to their undrained shear These values are considerably higher than those pro-
strength.) posed by Brinch Hansen (1961) and only slightly higher
To find the ultimate vertical component Q 0 that will than the values resulting from the latest computations at
cause a general shear failure, an analysis similar to that per- the Danish Geotechnical Institute (Brinch Hansen, 1970).
formed for vertical, central load must be performed. Such They are in fair agreement with available experimental data
an analysis discloses, as before, the existence of three zones on this subject (Giraudet, 1965).
in the soil under the footing, the size of which may be ap- Since expressions 3.14 through 3.17 have been derived
preciably reduced with the increase of inclination and ec- from plane-strain analyses, they should be applied, strictly
centricity of the load (Fig. 3.16a, b). Zone I is an elastic speaking, only to the case of a very long foundation acted
wedge, ABC, which is triangular in shape as long as the load upon by inclined loads in the direction of the shorter side B
is central (Schultze, 1952; Sokolovskii, 1960). For eccen- of the foundation. However, the case of loads inclined in
tric loads, the AC side of the wedge assumes the shape of a the direction of longer side L of the foundation is of equal,
circle whose center coincides with the center of rotation of if not greater practical interest. The experimental evidence
the footing (Fig. 3.16a, b) (Brinch Hansen, 1953, 1955). on this subject, coming largely from DEGEBO tests with
As long as the load eccentricity e is smaller than B/4 the large-scale models of shallow footings on sand (Muhs and
rotation center remains on the side of the footing opposite Weiss, 1969) suggests that there is a distinct difference in
the load (Fig. 3.16a). Fore = B /4 the center is exactly under the load inclination effects in the two cases. Thus, the di-
the footing edge, moving, for larger eccentricities, further rection of load inclination, as well as the ratio B /L of the
toward the axis of the footing and. causing uplift of the less sides of the footing have effect on inclination factor. Pend-
loaded side of the footing (Fig. 3. l 6b ). (The latter condi- ing more detailed investigations, it is suggested (Vesic,
tion is, for obvious reasons, to be avoided in design. To 1970) that expressions 3.14 and 3.17 be replaced by
provide adequate safety against lifting of the footing edge,
it is normally recommended that the eccentricity e not ex- rqi : 1- I
PI ]m (3.14a)
[ Q +BL c cot 1>
ceed B/6).
Theoretical and experimental investigations show that it
is on the safe side to take account of the eccentricity by *The newest, large-scale model experiments indicate that this width
1
introducing a fictitious effective width B = B - 2e of the reduction is somewhat conservative (Muhs and Weiss, 1969;
footing, instead of its actual width (De Beer, 1949; Meyer- Dorken, 1969).
Bearing Capacity of Shallow Foundations 131

p ]m+l (3.17a)
Inclination factors (Eqs. 3.14 and 3.16a):
hi= [ l- Q+B 1L 1ccot<p Sqi = 1; Sci= 1 - [(1.75)(203.3)/(406.6)(5.14)1 = o.83

where the exponent m is equal to Ultimate bearing pressure (see figures from Example 3.1):

2 +B/L q 0 = (1.21)(0.83) + (0.46)(1) = 1.00 + 0.46 = 1.46 ton/ft 2


mB = 1 +B/L
(3.18a)
Condition (b):
Assume tan 8 = tan <pd = 0.42, Ca = 0.
as long as the inclination of the load is in the direction of P = (0.5)(0.42)Q = 0.21Q
shorter side B of the foundation. In the contrary case, 75
Sqi = (1- 0.21)1. = 0.66 (Eq. 3.14a)
when the -load inclination is in the direction of longer side L Sci= 0.66 - [(1- 0.66)/(18.05)(0.42)] = 0.62 (Eq. 3.15)
of the foundation, the exponent in expressions 3 .14a and S-yi = (1 - 0.21) 2 · 75 = 0.52 (Eq. 3.17a)
3.17a becomes q 0 = (0.72)(0.62) + (4.57)(0.66) + (2.14)(0.52)
2 +L/B = 0.45 + 3.02 + 1.11 = 4.58 ton/ft 2
mL = 1 + L/B
(3.18b)
As in Example 3 .1 this value represents the upper limit of
bearing capacity under the assumption of an incompressible
Consistent with 3.14a and 3.15, expression 3.16, valid soil.
for <p = 0, then becomes:
mP EXAMPLE 3.4. For the same footing find the ultimate
Sci= 1 - B'L'c Ne (3.16a) bearing capacity if the reaction acts 6.5 feet off center in
the direction of the long side, and if the inclination is in the
The Scrfactors for other values of <p can be found, as be- same direction. Assume that the horizontal component is
fore, from Eq. 3.15. equal to the ultimate value given by Eq. 3.12.
If the load inclination is in the direction n, making an Condition (a)
angle 8 n with the direction of the long side L of the foun- Effective length of the footing: L' = 84 - (2)(6.5) = 71 ft
dation it is suggested to interpolate between the exponents P = Pmax = (28)(71)(0.22) = 437.4 ton
mL and mB and select for analysis an exponent determined Exponent mL (Eq. 3.18b): mL = (2 + 3)/(1 + 3) = 1.25;
from
sc = 1 [(1.25)(437.4)/(437.4)(5.14)1 = o.76
(3.18c)
qo = (1.21)(0.76) + (0.46)(1) = 1.38 ton/ft 2
For shapes other than rectangle, the effective founda-
Condition (b)
tion area may be determined as that of the equivalent rec-
tangle, constructed so that its geometric center coincides
P = 0.42 Q; sq o -
= 0.42)1. 25 = o.51
sc = o.51 - [0 - o.51 )/(18.05)(0.42) J = 0.45
with the load center and that it follows as closely as possi-
s-y = o - o.42) 2 ·25 = 0.29
ble the adjacent contour of the actual base area. A few ex- q 0 = (0.72)(0.45) + (4.57)(0.51) + (2.14)(0.29)
amples, after Brinch Hansen (1961), are shown in Fig. 3.17. = 0.32 + 2.33 + 0.62 = 3.27 ton/ft 2
EXAMPLE 3.3: For the footing discussed in Example As in Example 3.1 this value represents the upper limit of
3.1, find the ultimate bearing capacity in conditions (a) and bearing capacity under the assumption of an incompressible
(b) if the footing reaction acts 3 feet off center in the direc- soil.
tion of the short side B (eB = 3 ft) and if the inclination of
the reaction is in the same direction. Assume that the hori-
zontal component of the reaction is equal to half of the 3.7 EFFECT OF BASE Tl LT AND GROUND SURFACE
ultimate value given by Eq. 3 .12. SLOPE
Condition (a): There are situations in engineering practice where the foun-
Effective width of the footing: B' = 28 - (2)(3) = 22 ft dation base may be inclined to facilitate transmission of
Horizontal reaction: P= 0.5 Pmax = (0.5)(22)(84)(0.22) = larger horizontal reactions. Also, fairly often the ground
203.3 ton surface below which the shallow foundation is placed is
Exponent mB (Eq. 3.18a): mB = [2 + (1/3)] /[1 + (1/3)] = inclined with respect to the horizontal. Both situations are
1.75 shown irt Fig. 3.18, where the angle of base tilt is denoted

s--l I
I
L- ----

B
Fig. 3.17 Equivalent and effective foundation areas. (After Brinch
Hansen, 1961.) Fig. 3.18 Foundation with tilted base and sloped ground surface.
132 Foundation Engineering Handbook

by a and the ground surface slope by w, positive down- base is tilted 1 (vertical) to 4 (horizontal). Assume, as in Ex-
wards. As in the case of load inclination, it has been found ample 3.3, that the reaction is 3 feet off center in the di-
convenient to introduce these two effects by multiplY.ing rection of the short side B, inclined in the same direction,
the individual terms of the bearing capacity equation 3.11 with a horizontal reaction equal to one-half the ultimate
with base tilt factors St and/or ground slope factors Sg, anal- value from Eq. 3.12.
ogous to factors Si in Eq. 3.13 (Hvorslev, 1970; Brinch Han-
Condition (a):
sen, 1970). On the basis of analyses performed by Meyer-
Angle of base tilt: a= tan- 1 (1/4) = 0.253
hof (1953) for weightless soil and Brinch Hansen (1970) for
Base tilt factor (Eq. 20): Set= 1 - (2)(0.253)/(3.14 + 2) =
surface footings on soil with weight, the following joint ex-
pression for tilt factors may be proposed (Vesic, 1970): 0.90
Ultimate bearing pressure (see figures from Example 3.3):
Sqt = S'Yt = (1 - a tan ¢)2 (3.19)
q 0 = (1.00)(0.90) + 0.46 = 0,90 + 0.46 = 1.36 ton/ft 2
To find Set one can use, as before, expression 3.15. The
s
limiting value of ct for¢ = 0 is Condition (b):
Base tilt factors (Eqs. 3.19 and 3.15):
Set= 1 - [2a/(7T + 2)] (3.20)
Sqt = s'Yt = l 1 - (0.253)(0.42)1 2 = o.79
where a is expressed in radians. Expression 3.19 is very set= o.79 - lO - o.79)/08.05)(0.42)1 = o.76
accurate for ht and somewhat on the safe side for Sqt· It q 0 = (0.45)(0. 76) + (3.02)(0. 79) + (1.11 )(0. 79)
has the advantage of yielding the correct limit value for = 0.34 + 2.38 + 0.88 = 3.60 ton/ft 2
Set for¢= 0 (Eq. 3.20).
For ground slope factors Brinch Hansen (1970) points EXAMPLE 3.6: For the footing discussed in Example 3.1
out that Sqg varies with tan w in exactly the same manner find the ultimate bearing capacity if the ground slopes
1 1
as the load inclination factor Sqi varies with P/(Q + B L c 5 (horizontal) to 1 ( vertical). The load is assumed to remain
cot ¢). It is also possible to show that, for all practical pur- central and vertical.
poses, S'Yg c::,. Sqg· We can thus adopt Condition (a):
Sqg = S'Yg = [1 - tan w] 2 (3.21) Angle of ground slope: w = tan -I (1/5) = 0.201 = 11.5°;
sin w = 0.19; cos w = 0.98
The cohesion factor Scg can be found, as before, from ex- Ground slope factor (Eq. 22): Scg = 1 - [(2)(0.201)/
pression 3.15. The limiting value ·of this factor for¢= 0 is (3.14 + 2)] = 0.92
Bearing capacity factor (Eq. 3.24): N'Y = -2(0.19) = -0.38
Scg =1- [2 w/(7T + 2)] (3.22) Ultimate bearing pressure (see figures from Example 3 .1 ):
It should be noted, however, that the existence of ground q0 = (1.21)(0.92) + (0.46)(0.98)
slope in the case of frictionless soil (¢ = 0) requires the ad-
dition of the third (weight) term in the bearing capacity - [ (1/2)( 43)(28)(0.38)(0.87)] /(2000)
equation. It can be shown (Vesic, 1970) that the N'Y-value
= 1.02 + 0.45 - 0.10 = 1.37 ton/ft 2
for this term is negative and equal to
Condition (b):
N'Y = - 2 sin w (3.23) Ground slope factors (Eqs. 3.21 and 3.15):
Expressions 3.19 through 3.24 can be used, theoreti- Sqg = s'Yg = o - 0.201) 2 = o.64
cally, as long as scg = o.64 - o - o.64)/l08.o5)(0.42)l = o.59
a< 45° and w < 45° (3.24) q 0 = (0.72)(0.59) + (4.57)(0.98)(0.64) + (2.14)(0.64)
= 0.42 + 2.87 + 1.37 = 4.66 ton/ft2 ·
It is also required that
w<¢ (3.25)
3.8 EFFECT OF SHEARING RESISTANCE OF THE
However, one should keep in mind that the analyses of OVERBURDEN
slope effects, from which expressions 3.21 and 3.22 for
slope factors were proposed, do not take into consideration In the discussion of computation of ultimate load it was
the existing shearing stresses in the ground. The effect of mentioned that the analyses presented neglect the shearing
these stresses may be negligible as long as O < w < ¢/2. It is resistance of the overburden. This is normally justified by
advised, for slopes steeper than ¢/2, to perform also an the fact that the overburden soil is weaker than the bearing
analysis of slope stability, using one of the methods de- stratum. In some cases, however, the expected increase of
scribed in Chapter 10. ·-- bearing capacity due to shearing resistance of the overbur-
It is also important to remember that the analyses lead- den cannot be neglected. The problem may be formulated,
ing to adoption of factors given in Eqs. 3.19 through 3.23 as before, as the plane-strain problem of general shear fail-
are all based on assumption of plane strain conditions. ure of a rigid-plastic solid, with the difference that the solid
Thus, they are, strictly speaking, valid only for long rec- extends above the level of the foundation base (Fig. 3.14a).
tangular footings, with main axis parallel to the slope. Con- The exact solution of this problem is, again, not known.
sidering the mentioned similarity between slope factors and Approximate solutions have been found by Meyerhof
load inclination factors, it is expected that there must be (1951) and others. The results are often presented for
variation of slope and tilt factors with the foundation shape analysis in the form of "depth factors" Sd (Skempton,
similar to those presented in the preceding section. There 19 51; Brin ch Hansen, 1961 ). These are dimensionless
are, however, still no experimental data on this subject. parameters, analogous to factors Si in Eq. 3.13, indicating
the increase in individual terms of the bearing capacity due
EXAMPLE 3.5: For the footing discussed in preceding to the shearing strength of the overburden. Their values are
examples find the ultimate bearing capacity if the footing given by the following approximate formulas proposed by
Bearing Capacit:y of Shallow Foundations 133

Brinch Hansen (1970) and valid for D/B ~ 1: and Johnson, 1963). For sands, a flat reduction of¢ in the
case of local and punching shear failures is probably too
rqd = 1 + 2 tan¢ (1 - sin ¢)2 D/B conservative. It also suggests a jump in bearing capacity on
r-yd = 1 (3.26) transition to general shear failure, a phenomenon which,
of course, does not occur. Observations of failure loads of
The quantity red can be found again from the correspon- small footings on at least four sands (Vesic, 1970) suggest
dence formula 3.15. From 3.15 and 3.26 one finds, for that the factor 0.67 in Eq. 3.30 should be replaced by a
¢= 0. correction factor varying with relative density Dr, such as
red = 1 + 0.4 D/B (3.27) 0.67 + Dr - 0.75D;, applicable in the range O ~Dr~ 0.67.
Proposals of this kind may be useful in practice; how-
For D/B > 1 the calculation of depth factors is fraught with ever, their ultimate value is quite limited as they are based
uncertainty; requiring rather arbitrary assumptions about on the doubtful premise that the relative compressibility of
stress conditions in the overburden soil. Experimental data a soil under different geometrical and loading conditions is
are difficult to interpret properly, as the scale and compres- related exclusively to its strength characteristics c and ¢. In
sibility effects (to be discussed in the next section) inter- other words, the philosophy of this approach ignores the
vene along with uncertainties about exact stress conditions existence of scale effects other than those expressed by
in the surrounding soil. To provide a transition to deep Eq. 3.9.
foundations Brinch Hansen (1970) tentatively proposed for Scale effects differing from those predicted by the classi-
D/B > 1 the following formulas: cal earth pressure theories have been known in bearing ca-
rqd = 1 + 2 tan¢ (1 - sin ¢) 2 tan- 1 (D/B) pacity and earth pressure phenomena for quite some time.
Yet, the understanding of the variety of reasons for their
r'Yd = 1 (3.26a) existence has come only in very recent years, mostly in
connection with studies of shallow and deep foundations
Combined with 3.15 the above formula yields for¢= 0 (De Beer, 1963, 1965b; Vesic, 1964, 1965a; Kerisel, 1967).
red = 1 + 0.4 tan- 1 (D/B) (3.27a) These studies indicate that, in case of shallow foundations,
the average shear strength mobilized along a slip line under
These expressions, combined with shape factors from Table the foundation decreases with foundation size.* They also
3.2 give for very deep square or circular foundations in satu- show clearly that the relative compressibility of soils, both
rated clay(¢= 0) the following well-known result: with respect to gravity forces and with respect to the soil
q 0 = (5.14)[1 +0.20+(0.4)(1.57)] c+q =9.28c+q (3.28) strength, increases with foundation size.
In view of these facts a decrease in apparent values of
In cohesionless soils these expressions give q O = 3 .18 qNq bearing capacity factors with size should to a certain degree
for¢= 30° and q 0 = 3.68qNq for¢= 45°, where Nq is be expected in all soils. Probably the most conspicuous of
given by expression 3.6. These results are in fair agreement all is the decrease in N-y-values with increased size of sur-
with observed point bearing capacities of driven piles in face footings on sand. Figure 3.19, taken from De Beer
sand in conditions where q could be determined with some (1965a) shows that this decrease has been apparent in all
certainty (cf. Vesic, 1965a, 1967). major experimental studies of the problem of bearing ca-
It should be noted, however, that this increase of bearing pacity of shallow footings. As the largest of these footings
capacity due to "depth effect" occurs in conditions where has been only one meter (3.3 feet) square, there is great
the method of placement of the foundation (driving) causes practical as well as theoretical interest in determining
significant lateral compression. There exists good evidence whether the N-y-values shown in Fig. 3.19 tend asymptoti-
that this effect is practically nonexistent if the foundations cally to some minimum.
are drilled in or buried and backfilled (Vesic, 1963a) or if Recent studies on this subject (Vesic, 1969) seem to
the overburden strata are relatively compressible. For this indicate that the N -y-values for arbitrarily large footings
reason, it is advised not to introduce depth factors in design may be much smaller than conventionally assumed. This is
of shallow foundations. illustrated in Fig. 3.20, which presents a comparison of
measured ultimate resistances of small surface footings with
those· of deep footings, showing also the predicted bearing
3.9 INFLUENCE OF SOIL COMPRESSIBILITY AND
capacities of large footings according to conventional the-
SCALE EFFECTS ory. It is postulated that the bearing capacity of large sur-
face footings cannot be greater than the resistance of deep
It has b~en emphasized earlier that all preceding analyses of footings on the same soil.** In other words, there should
ultimate load are based on the assumption of incompress- be an upper limit of bearing capacity of all footings which
ibility of soil and that they should be applied, strictly .speak- may be related to the void ratio of the soil at failure.
ing, only to cases in which general shear failure of the soil is To arrive at an adequate assessment of the influence of
expected. There exists a lack of rational methods for ana- soil compressibility and related scale effects, it would be
lyzing bearing capacity failures in the two other modes necessary to have a bearing capacity theory based on some
characteristic for compressible soil.
To satisfy the immediate needs of engineering practice *There are actually three independent reasons for this decrease of
Terzaghi (1943) proposed the use of the same bearing ca- strength with foundation size: (a) the curvature of Mohr envelope;
pacity equation and factors with reduced strength charac- (b) progressive rupture along the slip line; (c) presence of zones or
teristics c* and¢* defined as follows: seams of weakness ·in all soil deposits. The relative contribution
of each of the reasons varies with soil type and the range of foot-
c* = 0.67c (3 .29) ing size, their total effect being discernible in practically all soils.
**This postulate seems to imply that very large footings should fail
</)* = tan- 1 (0.67 tan¢) (3.30) exclusively in punching shear, as apparently all deep footings do.
This should not be surprising, if one considers the mentioned fact
Such an approach may give satisfactory answers in some that the relative compressibility of soils increases with footing
soils, although it is not always on the safe side ( cf. Vesic size.
134 Foundation Engineering Handbook

- - - - - - Circular plates
800
- · - · - · - Square plates
0 - - - - Rectangular plates
\ : : ' · , , 1.67 4',.,/m'
600 G) Gent,yk= 1.619 ton/m 3
\_ ',, Vesic, ® Gent,yk= 1.509 ton/m 3
"\ ' , , / Yk = 1.538 ton/m 3 @ Meyerhof, yk = 1.62 ton/m 3
400 Meyerhof,
....... ' , _ @) Meyerhof, yk = 1.485 ton/m 3
\::k = 1.70 ton/m' Golder~;,;-j' @ Vesic,yk= 1.440 ton/m 3
'°'=":{L / / yk = I.76 ton/m 3

,,, ~- & .L-Vandeperre, 1950,yk= 1.647 ton/m 3


200 ~ -=::::::·-·-·-
\
® ...
.. _ ·-·---:::-:.:::=.·=:....:=:=
. . . . .....··-:--<J;:t--------,,\5
....... - - · - -1..W Me1sche1der, 1940, Yk= J.788 ton/m' Muhs
· - · - · - · •-@
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 100
yB, kg/cm 2
Fig. 3.19 Effect of size on bearing capacity of surface footings in sand. (After De Beer, 1965.)

tqc = exp {[(-4.4 + 0.6B/L) tan¢]


600
400 Circular footings
+ [(3.07 sin¢) (log10 2Ir)/(I + sin cp)J} (3.31)
Chattahoochee sand (vibrated)
200 Dry unit weight, 96.4 lb/ft 3 The quantity tee can be found again from correspondence
Relative density, Dr= 0.79 formula 3.15. Applying L'Hopital's rule one finds for¢= 0
Standard triaxial, <p = 39°
100
,:: 80 tee= 0.32 + O.I2B/L + 0.60 log10 Ir (3.32)
? 60
2

~
40 Considering expression 3.8 for all practical purposes we can
.~ unite
;:: 20
00
.5 (3.33)
0
.£: 10
0

.§ The use of expressions 3.31 through 3.33 makes sense,


5 obviously, only as long as the compressibility factors re-
4
main smaller than unity. Numerical values of the compres-
sibility factors tqc for two extreme cases, B/L = 0 (infinite

0.2
------
'l l':,,t plall' siLc:-.

0.4 0.6 0.8 1.0 4 6


Usual footing sizes

8 IO 20 40 60 80 I 00 200
strip) and B/L = I (square) are given in Table 3.3 and shown
graphically in Fig. 3 .21. Table 3 .4 gives the factors t cc ac-
cording to Eq. 3.32, for the same two extreme cases.
Foo tin!! size, ft From expression 3.31 one can find the magnitude of the
Fig. 3.20 Variation of ultimate resistance of footings with size. rigidity index for any angle ¢ and any particular founda-
(After Vesic, 1969.) tion shape below which it becomes necessary to reduce the
bearing capacity because of compressibility effects. This
more realistic soil models (such as, for example, elastic-
c;
plastic solid). In the absence of needed exact solutions for !=---
r c + q tan¢
this class of problems, it may be proposed to use the pres-
ently available theory based on solutions for a rigid-plastic
solid, along with some compressibility factors tc, analogous
to t;-factors in Eq. 3.13.
To find tentative expressions for compressibility factors
one can use the assumption that the ultimate normal pres-
sure o.n the sides of the wedge under the footing (AC and
CB, Fig. 3.14c) is equal to the pressure needed to expand a
cavity in the same soil mass. (This assumption, first used by
Skempton, Yassin and Gibson, 1953, was found to be rea-
sonable for deep foundations, at least under certain condi-
tions.) Combining this assumption with solutions for cavity
expansion in an elastic-plastic solid (Vesic, 1963b) one can Square or circle
L=B
obtain bearing capacity factors for comparison with those
given by Tables 3.1 and 3.2. In this way the following ex- 0 10 20 30 40 50 0 10 20 30 40 so
pression for compressibility factor tq is obtained (Vesic, Angle of shearing resistance,¢

1970): Fig. 3.21 Theoretical compressibility factors. (After Vesic, 1970.)


Bearing Capacity of Shallow Foundations 135

TABLE 3.3. VALUES OF COMPRESSIBILITY FACTOR tqc·


BIL = 1 (Square)

~ 00
2.5 5 10 25 50 100 250 500

1.000 1.000 1.000 (1.039)


50 0.772 0.852 0.917 0.988 (1.090)
10° 0.587 0.703 0.806 0.924 (1.107)
15° 0.437 0.562 0.679 0.821 (1.056)
20° 0.317 0.433 0.548 0.694 0.948 (1.199)
25° 0.224 0.322 0.423 0.557 0.801 (1.054)
30° 0.152 0.228 0.310 0.422 0.634 0.863 (1.175)
35° 0.098 0.153 0.214 0.300 0.468 0.655 0.918 (1.433)
40° 0.059 0.096 0.137 0.197 0.317 0.456 0.654 (1.055)
45° 0.033 0.054 0.080 0.117 0.194 0.284 0.417 0.692 (1.015)
50° 0.016 0.027 0.041 0.061 0.104 0.155 0.231 0.393 0.587

BIL = 0 (Strip)

~00
2.5 5 10 25 50 100 250 500

1.000 1.000 1.000 1.000


50 0.733 0.808 0.870 0.937 (1.034)
10° 0.528 0.632 0.725 0.831 0.996 (1.142)
15° 0.372 0.478 0.578 0.699 0.899 (1.087)
20° 0.255 0.348 0.441 0.558 0.762 0.964 (1.220)
25° 0.169 0.243 0.320 0.421 0.605 0.796 (1.048)
30° 0.107 0.161 0.219 0.299 0.449 0.610 0.831 (1.248)
35° 0.064 0.100 0.141 0.197 0.307 0.431 0.603 0.941 (1.318)
40° 0.036 0.058 0.083 0.119 0.192 0.275 0.395 0.638 0.916
45° 0.018 0.030 0.044 0.064 0.107 0.156 0.229 0.380 0.557
50° 0.008 0.013 0.020 0.030 0.051 0.076 0.113 0.192 0.287

In area marked by dots take tqc = 1.

3.4. VALUES OF COMPRESSIBILITY FACTOR tee TABLE 3.5. VALUES OF CRITICAL RIGIDITY INDEX.
FOR¢= 0.
Angle of

~
Critical Rigidity Index for:
Shearing
2.5 5 10 25 50 100 250 Resistance Strip foundation Square foundation
</J BIL =O BIL= 1
1 0.440 0.679 0.859 (1.039)
0 13 8
0 0.320 0.559 0.739 0.919 (1.157) 5 18 11
10 25 15
In area marked by dots take tee= 1. 15 37 20
20 55 30
25 89 44
critical rigidity index is given by 30 152 70
35 283 120
Ur)crit =! exp ((3.30 0.45B/L) cot (45 - ¢/2)] (3.34) 40 592 225
45 1442 486
Numerical values of the critical rigidity indexes corre- 50 4330 1258
sponding to different angles of shearing resistance ¢ are
given, for the two extreme cases B/L = 0 (strip) and B/L = 1
(square), in Table 3.5. soils by numerical methods such as finite-element technique
It should be understood that all the information pro- will allow a more accurate prediction of compressibility ef-
duced by this approximate analysis must be treated as ten- fects (cf. Desai and Reese, 1970).
tative and, in some sense, qualitative in nature. The pur-
pose of publishing Eqs. 3.31 through 3.34 at this time is to EXAMPLE 3.7: Investigate the effect of compressibility
allow the designer, in absence of any other rational method, on ultimate bearing capacity of the footing in clay discussed
to assess numerically the order of magnitude of expected in Example 3 .1. The modulus of deformation of the soil in
reduction of bearing capacity caused by the compressibility undrained condition is Eu = 24. l ton/ft 2 . The modulus of
effects. It is hoped that further development of our ability confined compression (in drained conditions) increases with
to treat analytically the complex elastic-plastic behavior of pressure q according to M v = 12.6 q.
136 Foundation Engineering Handbook

Condition (a): where E 1 = 364 ton£ft2 is the modulus at mean normal


Critical rigidity index (Eq. 3.34): stress of a1 = 1 ton/ ft .
The average mean normal stress in the expansion zone is
Ur)crit =!exp {[3.30- (0.45)(1/3)] (1.00)} = 12
taken as initial mean normal stress at a depth B/2 below the
Rigidity index (Eq. 3.3): Ir= (24.1)/2(1 + 0.5)(0:22) base of the footing. With <P = 38° for the sand in the "elas-
= 37 > 12 tic" zone the coefficient of earth pressure at rest is
The assumption of soil incompressibility is justified. The K 0 = 1 - sin (1.2)(38°) = 0.29
computed yalue of ultimate bearing pressure in Example
Mean normal stress at depth of 24 feet:
3.1 can be used without reduction.
Condition (b): a= {[ 1 + (2)(0.29)] /3} {[(8)(100) + (16)(56) l /(2,000)}
The average overburden pressure in the expansion zone is = 0.45 ton/ft2
taken as pressure at the depth B/2 below the base of the
Modulus of deformation:
footing. In this way one obtains:
q = [(8)(105) + (2 + 14)(43)]/(2,000) = 0.764 ton/ft 2 E = (364)-v'oAs = 244 ton/ft 2
Mu= (12.6)(0.764) = 9.62 ton/ft 2
Poisson's ratio v = [ 1 - sin ( 1.2)(23°)] / Poisson's ratio:
[2 ~ sin ( 1.2) (23°)] = 0.35 V = 0.29/(1 + 0.29) = 0.23
Modulus of deformation in drained condition: Representative angle of shearing resistance for the plastic
Ed = (9.62)[ 1 - 0.35 - (2)(0.35 2
)] /(1 - 0.35) zone is taken again to be 35°.
Critical rigidity index (Eq. 3.34):
= 5.99 ton/ft 2
Ur)crit = ! exp {3.30 - (0.45)(1/3)
Critical rigidity index:
· cot [45 0 - (0.5)(35 0 )] }
= 278
Ur)crit = ! exp {[3.30 - (0.45)(1/3)]
Neglecting the volume change in the plastic zone one
· cot [ 45° - (0.5)(23°)] } = 59 finds a rigidity index of
Actual rigidity index: Ir= (244)/[2(1 + 0.23)(0.45)(0.70)] = 316 > 278
Ir = (5 .99)/ { 2(1 + 0.35)[0.04 + (O. 764 )(0.42)]} ,.; 6 <° 59 Thus, the assumption of incompressibility appears to be
justified.
Thus, the assumption of soil incompressibility is not justi-
fied. Compressibility factors (Eqs. 3.31, 3.33 and 3.15): Remark: The assumption of an average volumetric strain
of 1 percent in the plastic zone would reduce the rigidity
tqe = exp {[-4.4 + (0.6)(1/3)] (0.42)] index to (Eqs. 3.3 and 3.4)
+ [(3.07)(0.391)(1.09)/(1 + 0.391)]} = 0.44
Irr= 316/[ 1 + (316)(0.01)] = 152
tee = 0.44 - (1- 0.44)/(18.05)(0.42) = 0.37
The corresponding compressibility factor would be (Eq.
t'Ye = t qe = 0.44 3.31 and 3.33)
Ultimate bearing pressure (see figures in Example 3 .1): tqe = t'Ye = exp {[-4.4+(0.6)(1/3)] 0.70
qo = (0.72)(0.37) + (4.57)(0.44) + (2.14)(0.44) + [(3.07)(0.574 )(2.18)/(1 + 0.574) l} = 0.607
2
= 0.27 + 2.10 + 0.94 = 3.22 ton/ft The ultimate bearing capacity would be reduced to
It would be of interest to compare this value with that qo = (35.1)(0.607) = 21.3 ton/ft 2
obtainetl by the Terzaghi approach. The reduced strength
characteristics are (Eqs. 3.29 and 3.30): However, since the sand in question is of medium density,
the assumption of average volume compression of 1 per-
c* = (0.67)(0.04) = 0.027 ton/ft 2 cent in the plastic zone is unfavorable. The actual ulti-
<P* = tan -t (0.67)(0.42) = 16° mate bearing capacity of this footing on sand would be
close to the upper value of 35 .1 ton/ft 2 . As mentioned
Bearing capacity factors (Table 3.1): in the remarks after Example 3.2, the allowable bearing
pressure may be controlled by maximum tolerable settle-
Ne= I l.62;Nq = 4.34;N'Y = 3.06
ment for the structure in question.
Ultimate bearing pressure:
qo = (0.027)(11.63)(1.16) + (0.463)(4.34)(1.14)
3.10 INFLUENCE OF ROUGHNESS OF THE
+ (1/2)(43)(28)(3.06)(0.87)/(2,000) FOUNDATION BASE
= 0.36 + 2.29 + 0.80 = 3.45 ton/ft 2
It has often been contended in the literature that the failure
EXAMPLE 3.8: Investigate the effect of compressibility pattern of the Prandtl solution (Fig. 3.14b) and subsequent
on ultimate bearing capacity of the footing in sand discussed extensions require perfect roughness of the foundation
base. The argument has been that the original Prandtl
in Example 3.2. In the low and elevated pressure range the
modulus of deformation of sand increases with mean nor- solution does not satisfy the strain rate compatibility equa-
mal stress according to the equation tions and that the Hencky (1923) pattern shown in Fig.
3.22 should be used instead, at least for smooth footings.
E = E1 v'a{a; The latter pattern, used by Meyerhof (1955) to evaluate the
Bearing Capacity of Shallow Foundations 137

¢ = 30°

Fig. 3.22 Slip pattern under a perfectly smooth footing postulated


by Hencky (1923).
'.,, = 45° +1>/2
10.__~--'---'~-'---'---'-~'----'----'
90° 80° 70° 60° 50° 40° 30° 20° 10° 0
effects of base smoothness, suggests that the bearing capac- Base angle, if;
ity of a smooth footing on the surface of a cohesionless soil (a)
(b)
should be only one-half the capacity of a rough footing.
However, the experiments performed to verify these con- Fig. 3.23 Bearing capacity of wedge-shaped strip footings. (After
cepts (De Beer and Vesic, 1958; Biarez, Burel, and Wack, Meyerhof, 1961.)
l 96 l) showed practically no effect of foundation rough-
ness on bearing capacity. It is significant to add that it was
never possible to reproduce experimentally the double
wedge with the slip surfaces starting in the middle of the
footing. A single wedge, very similar to that shown in Fig.
3.14c, was formed even in cases where the footing was split
longitudinally into two or more footings, free to move lat-
erally in opposite directions; see De Beer and Vesic, 19 5 8.
Nadai ( l 963) reports a similar experience with punching of
metals. Thus, patterns such as that shown in Fig. 3.22 are
fictitious, and should not be used in bearing capacity com-
putations. It may be concluded that the stress and defor-
mation pattern under compressed areas is such that it al-
R
ways leads to formation of single wedges, such as that

\
shown in Fig. 3.14a. The foundation roughness has little
effect on bearing capacity as long as applied external loads
remain vertical.
In the case of inclined loads, the foundation roughness
may limit the maximum horizontal component P of the Fig. 3.24 Footing with a concave contact area. (After Szechy,
load to be transmitted across the contact surface of the 1967.)
base (Eq. 3.12). However, experience seems to indicate
that most cast-in-place concrete foundations, by the way be explained by taking the depth of the sharp edge of the
they are constructed, possess roughnesses defined by fric- concave footing as D. It may be concluded that the vertical
tion angles equal to or greater than the angles of internal profile of the contact area has no effect on bearing capacity
friction <f>t of the underlying soil (see Schultz~ and Horn, as long as D is taken as depth ·of the footing edges and the
1967). (The latter angles vary from about 32 for quartz profile remains contained within the wedge defined by I/; =
sands to about l 0° for some colloidal clays of very high 45° +</J/2 (wedge I, Fig. 3.14c).
plasticity.)

3.12 INFLUENCE OF ADJACENT FOOTINGS


3.11 INFLUENCE OF THE VERTICAL PROFILE OF
THE FOUNDATION BASE The preceding discussions were concerned with the bearing
capacity of an isolated footing. In all analyses it is assumed,
The preceding considerations were concerned with footings essentially, that the soil mass affected by the footing
having plane contact areas. The influence of the vertical (ACDEA in Fig. 3.14c) is under the action of gravity forces
profile of contact areas was investigated theoretically by alone and is not under the influence of any other footing.
Biarez, Burel, and Wack (1961), Meyerhof (1961), Szechy There are, however, conditions in engineering practice
(1967), and others. Some experimental data are reported where footings are placed so close to each other that their
by Szechy. The first two of these papers were concerned zones of action overlap.
with bearing capacity of wedge-shaped long rectangular The problem of interaction of footings has been stud-
footings (Fig. 3.23a). The results (Fig. 3.23b) show that ied by Stuart (1962), Mandel (1965), and West and Stuart
there is practically no difference between bearing capacity (1965). These authors considered the problem of bearing
of wedges and that of ordinary flat footings of the same capacity of simultaneously loaded parallel strips of width B
size, as long as the base angle I/; (Fig. 3.14c) remains smaller placed at a spacings from each other, center-to-center (Fig.
than 45° + </J/2. Wedges with steeper angles I/; should have a 3.25a). The results indicate that the effects of adjacent
larger bearing capacity, at least if the comparison is made footings may vary considerably with the angle of shearing
using the depth of the wedge base (Fig. 3.23a) as D. The resistance </>. For low ¢-values they are negligible; however,
study by Szechy was concerned with the effects of curved for high ¢-values they appear to be significant (cf. Fig.
contact areas (Fig. 3.24 ); there was no apparent difference 3.25b), particularly if a footing is surrounded by others on
in bearing capacities of flat and convex areas. At the same both sides.
time, a slight concavity of the contact areas resulted in a It should be noted, however, that these effects are con-
measurable increase of bearing capacity. The difference can siderably reduced as L/B ~ l. Similarly, the compressibility
138 Foundation Engineering Handbook

Q Q

----B--~""'

(a)

W.T.
- - - Analysis A sz
------ Analysis B
a - - , Experiment
c;
2 200
y'
"0.
~ Fig. 3.26 Influence of groundwater table.

·c"
"
8 150 unit weight of soil, corresponding to the minimum moisture
w content of the soil above the water table (Meyerhof, 1955).
If the water table is permanently below the depth Zw = B,
'Y should be taken equal 'Ym. For water table at or above
2.0 3.0 4.0 the level of the foundation base, submerged unit weight
s/B -y' should be used.
All preceding considerations are based on the assump-
(b)
tion that the seepage forces acting on soil skeleton are
Fig. 3.25 Effect of interaction of long parallel strip footings. negligible. Should there be significant groundwater seepage
(After West and Stuart, 1965.) in any direction, it may have an effect on bearing capacity.
In addition to possible internal erosion of the soil (under-
of solids reduces and may completely eliminate the inter- mining, piping, and similar phenomena), the seepage force
ference effects. There are practically no such effects in the adds a component to the body forces caused by gravity.
case of punching shear failure. For these and other reasons This component, acting in the direction of stream lines, is
it is not recommended to consider interference effects in equal to 'Yw i, where i is the hydraulic gradient causing seep-
bearing capacity computations. A designer should be age. A simplified analysis of this effect of seepage force on
aware, however, of the possibility of their existence in some bearing capacity can be made by assuming that the seepage
special circumstances. through the soil zone directly involved in potential shear
failure is parallel and homogeneous. In such a case the
vectorial sum of effective soil weight and seepage force de-
3.13 INFLUENCE OF GROUNDWATER TABLE fines the direction and magnitude of a fictitious 'Y to be
used in Eqs. 3.7, 3.9 or 3.11, with possible base tilt and
The position of the groundwater table may have a signifi- ground-slope factors, as defined by Eqs. 3.19 through 3.23.
cant effect on bearing capacity of shallow foundations.
Generally the submergence of soils will cause loss of all ap-
parent cohesion, coming from capillary stresses or from 3.14 NONHOMOGENEOUS SOIL CONDITIONS
weak cementation bonds. At the same time, the effective
unit weight of submerged soils will be reduced to about In a discussion of effects of nonhomogeneity on bearing
one-half the weight of the same soils above the water table. capacity, distinction must be made between two basic kinds
Thus, through submergence, all three terms of the bearing of nonhomogeneity that can be encountered. In erratic soil
capacity equation may be considerably reduced. For this profiles different borings on the site disclose the presence
reason, it is essential that the bearing capacity analyses be of a variety of soils of different composition and character-
made assuming the highest possible groundwater level at the istics, appearing in lenses or in irregular strata of limited
particular location for the expected lifetime of the struc- extent. In such conditions the only reasonable assumption
ture in question. The assessment of this highest possible that can be made about the soil profile-if one can be made
level must be made taking into consideration the probabil- at all-is that a shear failure will occur in the soil mass along
ity of temporary high levels that could be expected in some the weakest soil encountered. One can question, however,
locations during heavy rainstorms or floods, although they whether it is technically possible and economically feasible
may not appear in the official records. to plan a soil exploration which will disclose all weaker
If the highest groundwater level is within the depth Zw ,( strata and lenses existing in the soil mass. Thus, taking and
B below the foundation level (Fig. 3.26) the effective unit testing samples from a few borings may not be justified. A
weight of the soil below the foundation base ('Y in Eq. 3. 7 crude estimate of bearing capacity on the basis of penetra-
or 3.9) should be taken equal to tion test results, combined with previous experience on sim-
ilar sites, may often be preferred in such cases.
(3.35) In contrast to this situation, the nonhomogeneity in reg-
ular soil profiles follows a definite pattern, such as a dis-
where -y' is the submerged unit weight and 'Ym is the moist cernible increase or decrease of soil strength with depth in
Bearing Capacity of Shallow Foundations 139

TABLE 3.6. MODIFIED BEARING CAPACITY FACTOR


Nm
Long Rectangular Footing (LIB ~ 5)

~
Cz/C1 - 2

5.14
4 6 8 10 20

5.14
1.0 5.14 5.14 5.14 5.14 5.14
1.5 5.14 5.31 5.45 5.59 5.70 6.14 7.71
2 5.14 5.43 5.69 5.92 6.13 6.95 10.28
3 5.14 5.59 6.00 6.38 6.74 8.16 15.42
4 5.14 5.69 6.21 6.69 7.14 9.02 20.56
(a) (b)
5 5.14 5.76 6.35 6.90 7.42 9.66 25.70
Fig. 3.27 Typical two-layer soil profiles. 10 5.14 5.93 6.69 7.43 8.14 11.40 51.40
5.14 6.14 7.14 8.14 9.14 14.14
all borings at the site. Increasing the number of borings and
testing a greater number of samples may be justified for im- Square or Circular Footing (LIB = 1)
portant structures, as this may bring the estimates of soil
characteristics closer to statistical averages. These averages,
combined with standard deviations, can then be used in
bearing capacity analyses in some meaningful way.*
~
Cz/C1

1
4

6.17
8

6.17
12

6.17
16

6.17
20

6.17
40

6.17 6.17
A very common kind of soil nonhomogeneity is that of 1.5 6.17 6.34 6.49 6.63 6.76 7.25 9.25
distinct soil layers of different strength and approximately 2 6.17 6.46 6.73 6.98 7.20 8.10 12.34
constant thickness. The simplest situations that can be con- 3 6.17 6.63 7.05 7.45 7.82 9.36 18.51
sidered would be those of a two-layer profile in two char- 4 6.17 6.73 7.26 7.75 8.23 10.24 24.68
acteristic conditions: 5 6.17 6.80 7.40 7.97 8.51 10.88 30.85
a) bearing stratum softer than the underlying stratum 10 6.17 6.96 7.74 8.49 9.22 12.58 61.70
(Fig. 3.27a); 6.17 7.17 8.17 9.17 10.17 15.17
b) bearing stratum stiffer than the underlying stratum
(Fig. 3.27b ).
The first situation is often found when, for example, a where (3 = BL/[2(B + L)H] may be called the punching in-
rigid or flexible footing or an embankment is placed on a dex of the footing, while Ne* = tcNc represents the earlier
relatively thin layer of soft clay overlying stiff clay or rock. given bearing capacity factor of the foundation, corrected
The second situation occurs when the footing is placed on a
for shape. ((3 = B/4H and Ne*= 6.17 for a circular or square
stiff clay crust or a sand stratum on the top of a deposit of
foundation; (3 = B/2H and NJ = 5.14 for a strip founda-
relatively soft, normally consolidated clay. Button (1953)
tion.) Numerical values of the modified bearing capacity
analyzed both situations for saturated clays in undrained
factor Nm in this situation for square and long rectangular
condition (<p = O) assuming general shear failure along the
footings are given in Table 3.6. They are also shown graph-
cylindrical slip surfaces starting at the edges of the founda-
ically in Fig. 3.28. For absolutely rigid footings they are
tion.
probably on the safe side. However, caution is advised in
Later experimental research (Brown and Meyerhof,
applying these factors to very flexible footings.
1969) showed that the assumed failure modes were un-
For the second situation (stiff clay layer over soft clay
realistic and that the resulting bearing capacity factors were
layer, Fig. 3.27b), Brown and Meyerhof (1969) suggest that
on the unsafe side. It appears that the failure in the first
the analysis assuming simple shear punching around the
situation (Fig. 3.27a) occurs, at least in part, by lateral plas- footing perimeter would be appropriate. Such an analysis
tic flow similar to that occurring in a solid squeezed be- yields
tween two rough parallel plates (Hartmann, 1925). The
failure in the second situation (Fig. 3.27b) is basically a (3.38)
punching failure, with vertical slip surfaces similar to those
shown in Fig. 3.9. The bearing capacity of the footing in It should be noted that the tests reported by Brown and
both situations is given by Meyerhof indicate a reduction of effective strength of the
(3.36) upper stiff clay layer, which may be attributed to progres-
sive failure phenomena. It is thus suggested that the shear
where c 1 represents the undrained shear strength of the strength c 1 in Eq. 3.36 be reduced by an appropriate factor.
upper layer and Nm a modified bearing capacity factor In conditions of the mentioned tests, with a clay sensitivity
which depends on the ratio of the shear strengths of the of about 2 the factor appears to be 0.75.
two layers, K = c 2 /c 1 , the relative thickness of the upper Other contributions to the subject of bearing capacity of
layer, H/B, as well as on the foundation shape. By interpo- layered clays in undrained conditions have been made by
lation between known rigorous solutions of the related Suklje (19 54) who made extensive theoretical and experi-
problems one can obtain the following expression for Nm mental studies of the stability of a layer of soft clay resting
in the first situation (soft clay layer over stiff clay layer, on a firm base such as rock. Some of his solutions, verified
Fig. 3.27a; Vesic, 1970): by small-scale model experiments, are given in Fig. 3.29.
1<NJ(Nf +(3- l)[(K + l)NJ 2 +(l +1<(3)N/ +(3- 1]
(3.37)
Nm= [1<(1< + l)Nf +1< +(3- 1] [(NJ +{3)NJ +(3- 1] - (1<Nf +(3- l)(NJ+ 1)

*It is of interest to note that the variation of soil strength in an The studies of Siva Reddy and Srinivasan (1967) and James
otherwise regular soil profile may be the source of sometimes im- et al. (1969), both representing extensions of Button's
portant scale effects (cf. Kerisel, 1967). work, should be mentioned because they demonstrate the
140 Foundation Engineering Handbook
-----------~-~--r----,---.,,---, 13

12 II

II IO

Square or circular footing


Long rec~angular footing

8
B/H= 12

Bill= 8
7 Bili= 4

Bill,;;; 4 6.17 Bill,;;; 2 5.14


6
2 3 4 6 8 9 10 3 4 6 9 10
Undrained strength ratio, K = c 2 jc 1 Undrained strength ratio. K = c 2 /c 1
Fig. 3.28a Modified bearing capacity factor Nm for square or cir- Fig. 3.28b Modified bearing capacity factor Nm for long rectangu-
cular footings on two-layer cohesive soil in undrained conditions. lar footings on two-layer cohesive soil in undrained conditions.
(After Vesic, 1970.) (After Vesic, 1970.)

TABLE 3.7. COEFFICIENTS OF INCREASE OF BEARING CAPACITY


FACTORS DUE TO PRESENCE OF AN INFINITELY
STIFF LAYER AT DEPTH H BELOW THE STRIP
FOUNDATION OF WIDTH 8.
(After Mandel and Salen~on, 1969.)
Coefficients r C (upper number)
tq (lower number)

<P 8/H--+ 1 2 3 4 5 6 8 10

00 1.02 1.11 1.21 1.30 1.40 1.59 1.78


t=1 forB!H< 1.41
1.00 1.00 1.00 1.00 1.00 1.00 1.00

10° 1.11 1.35 1.62 1.95 2.33 3.34 4.77


t=1 forB!H< 1.12
1.07 1.21 1.37 1.56 1.79 2.39 3.25

20° r= 1 for 1.01 1.39 2.12 3.29 5.17 8.29 22.00 61.50
8/H < 0.86 1.01 1.33 1.95 2.93 4.52 7.14 18.70 51.90

30° S" = 1 for 1.13 2.50 6.36 17.40 50.20 150.00 1444 14 800
8/H < 0.63 1.12 2.42 6.07 16.50 47.50 142.00 1370 14 000

Coefficients 'Y t
<P 8/H--+ 2 3 4 5 6 8 10
00 S"-y = 1 for all 8/H
10° S"-y = 1 for 8/H < 4.07 1.01 1.04 1.12 1.36
20° S"-y = 1 for 8/H < 2.14 1.07 1.28 1.63 2.20 4.41 9.82
30° S"-y = 1 for 8/H < 1.30 1.20 2.07 4.23 9.9 24.8 178 1450
Bearing Capacity of Shallow Foundations 141

6=2.4!5
q 0 =21rc

{3H - - - - - - - i

6 = 2.4!5
q 1 = l.95{3c

6 = 2.4!5
Qo (21rc) / (1-0.0142(3 2 )
(] /3 )a?+ Ba 2 + [ (2/9) -· (/3/ 0 )] 0 3= O

-----{3H--------.~

0=2.415
18.3 {3 2 c
q=--------'-------
H l "4 /3a3 a2a 2 2 /3 3 . 2 /3 1
- + - - - - + ( - - - ) 6 a+6 4 ( - - - )
!2 3 2 9 6 96 4

6=2.415
O'. /3[(1),-1 )/t/1]
q 1 36.6 c/[8 2- {3 2 (1 l/t/1)]
q,<a21rc q 2 -q 1 <21rc
Q2 < [t/l/(t/1-1)] 21rc

Fig. 3.29 Bearing capacity of a layer of soft clay overlying rock. (After Suklje, 1954.)

fact that both the case of anistropic layered clays and the of shearing resistance </> and low H/B ratios. However,
case of a single clay layer with continuously variable it may be questioned to what extent these results, ob-
strength can be handled with sufficient accuracy by intro- tained under the assumption of rigid-plastic behavior,
ducing an average strength for the layer or soil zone in ques- apply to real soils. It is reasonable to expect that, in
tion. Desai and Reese ( 1970) presented an analysis of bear- cases of high </> and low H/B, footings will fail by vertical
ing capacity of a circular footing on a two-layer clay soil in compression of the bearing stratum, in which crushing of
undrained conditions, showing the potential of finite-ele- soil particles plays a predominant role. Also the strength of
ment techniques for this class of problems. the rigid stratum, which is always finite, will after a certain
The problem of bearing capacity of a layer of soil o flim- point play a role in the deformation process, limiting the
ited depth H resting over a lower layer of infinite rigidity bearing capacity below the indicated values.
and strength has been treated for the plane strain case (1011:g Of far greater interest is the general problem or bearing
rectangular footing) by Mandel and. Salern;on ( 1969). Their capacity of a stiff stratum underlain by a softer stratum,
solution, obtained by the method of characteristics, indi- when the soils involved possess both cohesion and friction.
cates that the presence of a rigid layer below the bearing One of the simplest and most frequent problems of this
stratum results in an increase of bearing capacity. Factors kind is that of a layer of sand of finite thickness underlain
of increase due to this effect, analogous to factors ti in Eq. by a. soft clay layer. Experimental studies (Tcheng, 1957;
3.11 are presented in Table 3. 7. It is seen that this effect Vesic, 1970) show that the mode of failure under these cir-
becomes aj5parent whenever the depth of the bearing cumstances is punching along essentially vertical slip lines
stratum is smaller than the foundation width, and may following the foundation perimeter (Fig. 3.9). According
become fairly significant for large values of the angle to Tcheng's analysis, the bearing capacity q 0 of a long rec-
142 Foundation Engineering Handbook

tangle on the top of the sand layer with shear strength EXAMPLE 3. 9: For the footing from Example 3 .1 find
characteristics c = 0 and ¢ should be related to the bearing the ultimate bearing capacity in undrained conditions if the
capacity q~ of the underlying (clay) layer by the expression bearing stratum of soft clay is underlain by a thick stratum
of stiff clay, (undrained shear strength 0.5 3 ton/ft 2 ) start-
q0 = q~/{1 - 2(H/B) tan¢ (1 +sin¢) ing from elevation 16 feet.
· exp [-(rr/4 - ¢/2) tan¢]} (3.39) Punching index (3 = [(84)(28)] /[2(84 + 28)(16 - 10))
= 1.75
Tcheng reported good agreement between his test results Shape factor (see Example 3.1): tc
= 1.067;
and the above expression at least in the domain H ~ 1. 5 B. NJ= (1.067)(5.14) = 5.48
For greater depths he proposed semiempirical formulas, Shear strength ratio: K = c 2 /c 1 = 0.53/0.22 = 2.41
suggesting also that for H),, 3.5 B the influence of the soft Modified bearing capacity factor (Eq. 3.37):
clay layer becomes negligible.
N = (2.41)(5.48)(5.48 + 0.75) {(3.41)(5.48) + [1 + (2.41)(1.75)] 5.48 + 0.75}
m [(2.41)(3.41)(5.48) + 2.41 + 0.75] [(5.48 + 1.75)5.48 + 0.75] - [(2.41)(5.48) + 0.75] (6.48)
A more general analysis, valid for rectangles of any Nm = 5.86 (checks well with an interpolated value from
shape, resting on an upper, stronger layer having strength Fig. 3.28)
parameters c 1 , ¢ 1 and underlain by a lower layer of Ultimate bearing pressure (Eq. 3.36):
strength parameters c 2 , ¢ 2 (Fig. 3.27b) yields, under the
assumption that the slip surfaces are vertical (Vesic, 1970): q0 = (0.22)(5.86) + (0.463) = 1.29 + 0.46 = 1.75 ton/ft 2

qo = [q~ +(1/K)c 1 cot¢il exp {2[1 +(B/L)] Ktan¢ 1 (H/B)} EXAMPLE 3.10: Solve the problem described in Exam-
ple 3.9 under assumption that the undrained shear strength
- (1/K) c 1 cot ¢ 1 (3.40) of the upper layer is 0.53 ton/ft 2 while the strength of the
In this expression K = ( 1 - sin 2 ¢ 1 )/( 1 + sin 2 ¢ 1 ), while lower layer is 0.22 ton/ft 2 .
q" is the bearing capacity of a fictitious footing of the same Shear strength ratio: K = 0.22/0.53 = 0.415
size and shape as the actual footing, but resting on the top Punching index (same as in Example 9) (3 = 1.75
of layer 2 (generally to be evaluated from strength param- Modified bearing capacity factor (Eq. 3.38):
eters c 2 , ¢ 2 and other characteristics of the second layer). Nm = (1/1.75) + (0.415)(5.48) = 0.57 + 2.27 = 2.84
If the upper layer is cohesionless (c 1 = O) with 25° ~
¢ 1 ~ 50° the above expression is reduced to - Ultimate bearing pressure (Eq. 3.36)

q0 = q~ exp {0.67 [ 1 + (B/L )] (H/B)} (3.41) q0 = (0.53)(2.84) + (0.463)(1.00) = I.SO+ 0.46


= 1.96 ton/ft 2
This expression can be used to find a simple expression for
critical depth of the upper layer, beyond which the bearing The small difference in the bearing capacities between
capacity will be little affected by the presence of the l~wer the case analyzed in Example 9 and the present case is un-
soft layer. Denoting by q~ the bearing capacity of the derstandable if one considers the fact that the thickness of
upper layer in infinite mass, one finds: the upper layer is only 6 feet, or less than one fourth of the
foundation width.
3 In (q~/q~)
(H/B)crit = 2(1 +B/L)] (3.42) EXAMPLE 3.11: For the footing from Example 3.2 find
the ultimate bearing capacity if the bearing stratum of me-
A study of this expression, valid if c 1 = 0, reveals that the dium dense sand is underlain by stiff clay (undrained shear
critical depth of a strip foundation should be twice that of strength 0.53 ton/ft 2 ), starting at elevation - 30 feet.
a square foundation, under otherwise identical conditions. The ultimate bearing capacity of sand in infinite mass,
It can also be shown that the critical depth increases from Example 3 .2 is q~ = 35 .1 ton/ft 2 •
roughly in proportion to the angle of shearing resistance The ultimate bearing capacity of a fictitious footing rest-
¢ 1 of the upper layer and the water content w of the lower ing on the lower, clay layer is (see Example 1, for some
layer (if the latter happens to be saturated clay). Most sig- figures):
nificantly, however, the critical depth also varies with the
size of the footing, at least if the lower layer is saturated
q~ = (0.53)(5.14)(1.067) + [(8)(105) + (22)(43))/(2000)
clay. In conditions of Tcheng's tests, for example, the crit- = 2.90 + 0.89 = 3.79 ton/ft 2
ical depth ratio for a two-foot-wide footing should be 6.8,
and for a 20-foot-wide footing 10.3, as compared with 3.2 The critical depth of the upper layer is (Eq. 3.42):
predicted by Eq. 3.42 and 3.5 observed for 2-inch-wide Hcrit = [(3)(28) ln (35.1/3.79)] /{2 [1 + (1/3)]}
models.
A final remark will be made regarding footings resting on = 70 ft> 20 ft
a thin stratum of rock underlain by softer strata. In addi-
Consequently, the bearing capacity of the footing will
tion to examining the possibility of punching shear failure,
be affected by the presence of the stiff clay layer. Its mag-
similar to that occurring in soils, one should also check that
nitude, computed from Eq. 3.41 is
the footing, as designed, does not induce failure of rock in
tension at the bottom of the upper stratum. Such an analy- q0 = (3.79) exp {0.67 [l + (1/3)] (20/28)} = 7.15 ton/ft 2
sis can be made using the concepts developed by Wester-
gaard (1926) for analysis of concrete pavements. For selec-
tion of k-values for the lower stratum, in view of limited 3.15 EFFECT OF RATE OF LOADING
depth of that stratum, see Vesic and Saxena (1970). An
alternate approach is to use the layered solid theory of Bur- All the analyses of bearing capacity presented in preceding
mister (1943 ). paragraphs are conceived for static loading conditions. It is
Bearing Capacity of Shallow Foundations 143
60 ,--~~~~~~~~~~~~~~~~~~~~-,

assumed, tacitly, that the footing load Q (Fig. 3. la) is in- 1.60
creased gradually until failure at a loading rate slow enough 1.50
~
that no viscous or inertia effects are felt. This assumption
"' .S
applies to conditions of most ordinary footings, which carry
a certain dead load and are presumed to fail by a single ap-
fo 1.40
1.30 <.'":':
"
plication of excessive static live load. The rate of applica- "
·~ 1.20 ~
tion of these loads affects, under these conditions, the bear-
l 40 I.IO fl"
en
ing capacity only to the extent that it may be related to the ,.;
Data corrected 1.00
rate of drainage of excess pore-water pressures created in ::E to e1=0.800
the supporting soil by the application of the loads. It is 0.90
30~~~~:--~~-'---~~--'-~~~..J......~~--'-~~---l
understood that the selection of shear strength parameters 1000 100 10 0.1 0.01 0.001
c and </J to be introduced in the analysis will be made so as Tinll' t<' l'lt strain, s1..·c
to take care of that effect (see Example 3 .1 ). Fig. 3.31 Effect of rate of strain on undrained strength of a satu-
However, some footings, such as those supporting missile- rated, normally consolidated, fat clay. (After Whitman, 1970.)
launching or blast resistant structures, are subjected to high
live loads of very short duration. The high rates of strain
associated with these impulsive loads may induce viscous ever, what to expect in the case of loose, submerged sand,
and inertia effects in the soil mass. The related phenomena because of transient liquefaction effects.
have been the object of extensive study, mostly by load tests 3) Footings on compacted clay all show a considerable
on model footings on sand and clay (cf., for example, Jack- increase in bearing capacity as the rate of loading changes
son and Hadala, 1964; Richart, 1965; Vesic, Banks, and from static to impact loading conditions. There exists no
Woodard, 1965; Poplin, 1965; Whitman, 1970). The find- direct information about the behavior in the intermediate
ings of these studies can be summarized as follows: range of loading rates. However, on the basis of strength
1) As the rate of loading is increased from about 10-4 tests on clay samples at variable loading rates, we can expect
in/sec (static loading conditions) to about 10 in/sec (impact that the bearing capacity of footings on clay, contrary to
loading conditions) the mode of failure of model footings sand, will increase with increased loading rate (Fig. 3.31 ).
on both dense sand and compacted clay changes from gen- This conclusion is supported by the finding that a good pre-
eral shear to punching shear. This change is explained by diction of not only ultimate bearing capacity but also of
the fact that the inertia effects in the soil mass have a sim- load-displacement behavior of small footings subjected to
ilar effect as overburden pressure (Heller, 1964). transient loads can be obtained by multiplying the stresses
2) In the mentioned range of loading rates, from 10-4 to (or loads) corresponding to any particular displacement by
10 in/sec, footings on dense sand show a slight drop of bear- appropriate strain-rate factors (Jackson and Hadala, 1965).
ing capacity with increased loading rate, followed by a The latter is defined as the ratio of undrained soil strength
steady, slow increase, which is extended all the way into the at a specified strain rate to the undrained soil strength at the
impact velocities range (Fig. 3.30). This trend in variation standard laboratory strain rate.
of bearing capacity is analogous to the trend in variation of In summary, it appears that the conventional, static
shear strength of dry sands observed by Whitman and Healy analyses of bearing capacity can be used for footings sub-
(1962). From the practical point of view, this means that jected to moderately rapid loadings, if the strength param-
the static bearing capacity analyses may be applicable also eters c and </J introduced in the analysis are modified for
in the case of footings subjected to moderately rapid loads, strain rate effects.
provided that the strength parameters are determined by The footings subjected to impact and vibratory loads still
tests at appropriate loading rates. In absence of equigment require a dynamic approach for analysis. Details about
for transient tests, a reduction of ¢-angle of up to 2 may analysis of such footings can be found in Richart ( 1965), as
be in order for dense sand. It is highly questionable, how- well as in Chapters 24 and 25 of this book.

3.16 CHOICE OF SAFETY FACTOR


600
o\o
"'°' The analyses described in this chapter are all made with the
.:$' 'v'" purpose of assessing the magnitude of ultimate load Q 0 or
500 <l.J(;r<l.J" ,-e,~o~
ultimate pressure q 0 (Eq. 3.1) at which the foundation may
,._
0:,

~
C,
,~ . ,,,,,~
experience a bearing capacity failure. As mentioned in the
introduction, the foundations are designed so as to possess
3
400 """"
c,"'
an adequate safety against this type of failure.
"'" The assessment of adequate safety of a component of a
~
300 structure is, in the modern view, a complex problem of opti-
t
<.'!
mization, which can be properly resolved only with due
b considerations of serviceability and economy of the struc-
·u
"'0. 200 ture, as well as of probability and consequences of failure.
CT
bl)
While the mathematical methods of analysis of this aspect
of the bearing capacity problem are well advanced, (cf., for
'§"
100 example, Freudenthal; 1961, 1968; Wu and Craft, 1967;
"' Circular footings B = 4 in Ang and Amin, 1969) there are apparently no generally ac-
cepted, consistent criteria that can be recommended for use
0
10-4 10-3 10-2 10-1
in engineering design today.
10
A traditional approach to the choice of safety factors F 8 ,
Loading velocity, in/sec
appearing in Eq. 3.2, is outlined in Table 3.8. This approach
Fig. 3.30 Effect of loading velocity on bearing capacity of surface recognizes that the choice of safety factors should depend
footings on dense sand. (After Vesic, Banks, and Woodard, 1965.) on the character and expected life of the structure as well as
144 Foundation Engineering Handbook

TABLE 3.8. MINIMUM SAFETY FACTORS FOR TABLE 3.9. PARTIAL SAFETY FACTORS FOR DESIGN
DESIGN OF SHALLOW FOUNDATIONS. OF SHALLOW FOUNDATIONS. (AFTER
(AFTER VESIC, 1970.) BRINCH HANSEN, 1965.)
Preliminary note: The selection of safety factors for design cannot
be made properly without assessing the degree of reliability of all Load Factors
other parameters that enter into design, such as design loads,
strength and deformation characteristics of the ~oil mass, etc. In Dead load 1.00
view of this, each case is to be considered separately by the designer. Steady water pressure 1.00
The following table may be used as a guide for permanent structures Fluctuating water pressure 1.20 (1.10)
in reasonably homogeneous soil conditions. Live loads (general) 1.50 (1.25)
Wind loads 1.50 (1.25)
Soil Exploration Earth or grain pressure in silos 1.20 (1.10)

Cate- Characteristics Thorough, Strength Factors


gory Typical Structures of the Category Complete Limited
Cohesion c 2.00 (1.80)
Railway bridges Maximum design
Coefficient of internal friction tan </> 1.20 (1.10)
Warehouses load likely to
A Blast furnaces occur often; 3.0 4.0
Hydraulic consequences Remark: The numbers in parentheses refer to temporary structures
Retaining walls of failure or to extraordinary combinations of loadings (such as dead load +
Silos disastrous most unfavorable live load+ most unfavorable wind load).

Highway bridges Maximum design


tors, while the strength parameters c and <p are reduced by
Light industrial load may occur
B and public occasionally, 2.5 3.5
other partial factors. The advantages of this approach,
buildings consequences which is widely used in some countries, become especially
of failure apparent in analysis of footings for earth retaining struc-
serious tures or footings on slopes. A list of recommended partial
factors is given in Table 3.9.
Apartment and Maximum design
C office load unlikely 2.0 3.0
buildings
EXAMPLE 3.12: For the footing discussed in Example 3.1
to occur
find the safe bearing capacity in both assumed conditions.
Remarks:
Use both the traditional approach with a safety factor of
(1) For temporary structures these factors can be reduced to 75 F 8 = 3.0 and the partial safety factor approach. Assume
percent of the above values. However, in no case should the that the footing reaction comes 65 percent from live load
safety factors lower than 2.0 be used. and 35 percent from dead load.
(2) For exceptionally tall buifdings, such as chimneys and towers,
or generally whenever progressive bearing capacity failure may Condition (a)
be feared, these factors should be increased by 20 to 50 percent. Traditional approach: qa = 1.67/3.0 = 0.56 ton/ft 2
(3) The possibility of flooding of foundation soil and/or removal of Partial safety factor approach:
existing overburden by scour or excavation should be given Nominal strength: c = 0.22/2.0 = 0.11 ton/ft 2
adequate consideration. Nominal bearing capacity:
(4) It is advisable to check both the short-term (end-of-construc-
tion) and long-term stability, unless one of the two conditions qn = (0.11)(5.14)(1.067) + (0.463) = 0.60 + 0.46
is clearly less favorable.
(5) It is understood that all foundations will be analyzed also with = 1.06 ton/ft 2
respect to maximum tolerable total and differential settlement.
If settlement governs the design, higher safety factors must be Average load factor: (0.64)(1.50) + (0.36)(1.00) = 1.32
used. Allowable bearing pressure: qa 1.06/1.32 = 0.80 ton/ft 2
Condition (b)
on the consequences of failure. Thus, a lower safety factor The problem will be solved using the values from Example
can be adopted for temporary, limited life structures. Also, 3. 7, which include the compressibility effects.
a higher safety factor is suggested for structures where the Traditional approach: qa = 3.22/3.0 = 1.07 ton/ft 2
maximum design load will regularly occur and where the Partial safety factor approach:
consequences of failure would be disastrous, as compared Nominal strength: Cn = 0.04/2.0 = 0.02 ton/ft 2
with structures where the maximum design load is unlikely tan </Jn = 0.424/ 1.2 = 0.353; </Jn = 19 .5°
to occur, and where the consequences of failure would not Nominal bearing capacity factors:
be as serious. Higher factors are required for structures in
danger of progressive failure, such as exceptionally tall Ne= 14.4;Nq = 6.l;N,, = 5.0
buildings. Finally, this approach also takes into account the
fact that a thorough knowledge of the soil profile, based on Nominal shape factors:
an extensive exploration and testing program, allows the
designer to assess the ultimate loads with a much higher de- te = 1 + (1/3)(0.42) = 1.14;
gree of confidence, thus reducing the probability of failure. tq = 1 +(1/3)(0.35) = 1,12;
In a more modern version of the traditional approach to
evaluation of safety, the uncertainties involved with differ- r,, = 0.87
ent variables such as foundation loads or soil strength are
introduced separately as "partial safety factors" (cf. Brinch The compressibility factors have to be computed using
Hansen, 1965 ). A nominal state of failure is considered, in actual soil parameters. Example 3.7 gives tee= 0.37; tqe =
which the acting loads are multipli,ed by certain partial fac- t,,e = 0.44.
Bearing Capacity of Shallow Foundations 145

Nominal bearing capacity: Brinch Hansen, J. (1965), The philosophy of foundation design:
Design criteria, safety factors and settlement limits, in: Bearing
q = (0.02)(14.4)( 1.14 )(0.37) + (0.463)(6.1)(1.12)(0.44) Capacity and Settlement of Foundations, Proceedings of a Sym-
posium held at Duke University, April 5/6, 1965, pp. 9-13.
+ (1/2)( 43)(28)(5.0)(0.87)(0.44 )/(2000) = 0.12 + 1.39 Brinch Hansen, J. (1970), A Revised and Extended Formula for
Bearing Capacity,-Bulletin No. 28, Danish Geotechnical Institute,
+ 0.57 = 2.08 ton/ft 2 Copenhagen, pp. 5-11.
Load factor, same as above, is equal 1.32. Brown, J. D. and Meyerhof, G. G. (1969), Experimental study of
Allowable bearing pressure: Qa = 2.08/1.32 = 1.58 ton/ft 2 bearing capacity in layered clays, Proceedings, Seventh Intern.
Conf Soil Mech. Found. Engrg., Mexico City, Vol. 2, pp. 45-51.
Buisman, A. S. K. (1935), De weerstand van paalpunten in zand,
De Ingenieur 50, pp. Bt. 25-28, 31-35 ..
3.17 BEARING CAPACITY ACCORDING TO BUILDING Buisman, A. S. K. (1940), Grondmechanica, Waltman, Delft, pp.
CODES 190.
Burmister, D. M. (1943), The theory of stresses and displacements in
Most building codes contain some information on bearing layered systems and application to the design of airport runways,
capacity of shallow footings, usually presented in the form Proceedings of the Highway Research Board, 23, pp. 126-148.
De Beer, E. E. (1970), Experimental determination of the shape
of tables relating allowable foundation pressures to certain factors and the bearing capacity factors of sand, Geotechnique
soil types. For example, a code may indicate "safe pres- 20, No. 4, pp. 387-411.
sures" of 3 to 6 ton/ft 2 for dense sand or hard clay and 1.5 Button, S. J. (1953), The bearing capacity of footings on a two-layer
to 3 ton/ft 2 for medium dense sand or stiff clay, etc. When- cohesive subsoil, Proc. Third Intern. Conf Soil Mech. Found.
ever this information is actually based on local experience, Engrg., Zurich, Vol. 1, pp. 332-335.
it should be considered as a helpful indication of pressures Caquot, A. (1934), Equilibre des massifs a frottement interne,
that have been used in a given locality without causing dis- Gauthier-Villars, Paris, pp. 1-91.
tress to the structures built in the past. The limited value of Caquot, A. and Kerisel, J. (1953), Sur le terme de surface dans le
this information will, perhaps, be best understood by point- calcul des fondations en milieu pulverulent, Proceedings, Third
International Conference on Soil Mechanics and Foundation En-
ing out some of the serious shortcomings of the mentioned gineering, Zurich, Vol. I, pp. 336-337.
tables: Cox, A. D., Eason, G., and Hopkins, H. G. (1961), Axially symmet-
1) The character of the bearing stratum is given in de- ric plastic deformations in soils, Philosophical Transactions of the
scriptive terms, often very vague and without specification Royal Society of London, Series A, 254, pp. 1-45.
of physical properties of the soil in qJJestion. De Beer, E. E. (1949), Grondmechanica, Deel II, Funderingen N. V.
2) The underlying strata are assumed to have no effect Standard Boekhandel, Antwerpen, pp. 41-51.
on safe bearing capacity. De Beer, E. E. (1°963), The scale effect in the transposition of the
3) Such important factors as size, shape, and depth of results of deep sounding tests on the ultimate bearing capacity of
piles and caisson foundations, Geotechnique 11, No. 1, pp.
foundation and the position of the water table are normally
39-75.
assumed to have no effect on bearing capacity. De Beer, E. E. (1965a), Bearing capacity and settlement of shallow
4) The type and statical system of the structure sup- foundations on sand, Bearing Capacity and Settlement of Foun-
ported by the foundation, as well as the character of design dations, Proceedings of a Symposium held at Duke University,
loads are equally assumed to be of no effect. pp. 15-34.
It should be obvious th.2t the information of this kind, De Beer, E. E. (1965b), The scale effect on the phenomenon of pro-
however useful, should never be taken as a substitute for a gressive rupture in cohesionless soils, Proceedings, Sixth Inter-
proper engineering analysis of bearing capacity, following national Conference on Soil Mechanics and Foundation Engineer-
procedures similar to those exposed in this chapter. ing, Montreal, Vol. II, pp. 13-17.
De Beer, E. E. (1967), Proefondervindelijlce bijdrage tot de studie
van het gransdraagvermogen van zand onder funderingen op
staal; Bepaling von der vormfactor Sb, Anna/es des Travaux Pub-
REFERENCES lics de Belgique, 68, No. 6, pp. 481-506; 69, No. 1, pp. 41-88;
No. 4, pp. 321-360; No. 5, pp. 395-442; No. 6, pp. 495-522.
Ang, A. H. S., and Amin, M. (1969), Safety factors and probability De Beer, E. E. and Vesic, A. (1958), Etude experimentale de la
in structural design, Journal of the Structural Division ASCE_; 95, capacite portante du sable sous des fondations directes etablies
No. ST7, pp. 1389-1405. ' en surface, Anna/es des Travaux Publics de Belgique 59, No. 3,
Antoine, J., L'Herminier, R. L., and Bachelier, M. (1953), Force pp. 5-58.
portante de pieux de grand diametre fondes sur un bane calcaire De Beer, E. E. and Ladanyi, B. (1961), Etude experimentale de la
de faible epaisseur; Proceedings, Third Intern. Conf Soil Mech. capacite portante ·du sable sous des fondations circulaires etablies
Found. Engrg., Zurich, Vol. 2, p. 3-6. en surface, Proceedings, Fifth Intern. Conf Soil Mech. Found.
Berezantsev, V. G. (1952), Osesimetrichnaia zadacha teorii pred- Engrg., Pari,, Vol. 1, pp. 577-581.
el'nogo ravnovesia sypuchei sredy, Gostekhizdat, Moskva, pp. Desai, C. S. and Reese, L. C. (1970), Analysis of circular footings on
81-120. layered soils, ProceedinfES ASCE, Journal of the Soil Mechanics
a
Biarez, J., Burel, M., and Wack, B. (1961), Contribution l'etude de and Foundations Division 96, SM4, pp. 1289-1310.
la force portante des fondations; Proceedings, Fifth International Darken, W. (1969), Die Einfluss des Aussermittigkeit auf die Grund-
Conference on Soil Mechanics and Foundation Engineering, Paris, bruchlast lotrecht beanspruchter Oberfliichengrundungen auf
Vol. 1, pp. 603-609. nichtbindigen Boden, Mitteilungen aus dem Institut fiir Ver-
Brinch Hansen, J. (1953), Earth Pressure Calculation, Danish Tech- kehrswasserbau, Grundbau und Bodenmechanik, Technische
nical Press, Copenhagen, pp. 1-271. Hochschule Aachen; Heft 44.
Brinch Hansen, J. (1955), Simpel beregning af fundamenters Eason, G. and Shield, R. T. (1960), The plastic indentation of a
baereevne,Jngeni</>ren, 64, No. 4, pp. 95-100. semi-infinite solid by a perfectly rough circular punch, Zeitschrift
Brinch Hansen, J. (1957), General Report, Division Illa, Proceedings, fur Angewandte Mathematik und Physik, Vol. 11, pp. 33-42.
Fourth International Conference on Soil Mechanics and Founda- Freudenthal, A. M. (1961), Safety, reliability and structural design,
tion Engineering, London, Vol. II, pp. 441-447. Journal of the Structural Division ASCE 87, No. ST3, pp. 1-16.
Brinch Hansen, J. (1961), A General Formula for Bearing Capacity, Giraudet, P. (1965), Recherches experimentales sur !es fondations
Bulletin No. 11, Danish Technical Institute, Copenhagen, pp. a
soumises des efforts inclines ou excentres, Annales des Pants et
38-46. Chaussees 135, No. 3, pp. 167-193.
146 Foundation Engineering Handbook

Hansen, B. and Christensen, N. H. (1969), Discussion on theoretical Ohde, J. (1950), Der Eindringungswiderstand von Fundamenten als
bearing capacity of very shallow footings, Proceedings ASCE, Grundlage fiir die Festlegung der zulassigen Baugrundbelastung,
Journal of the Soil Mech. and Found. Division, 95, No. SM-6, pp. Bautechnik 27, No. 8, pp. 272-277.
1568-72. Peck, R. B. (1965), Bearing capacity and settlement: Certainties and
Hartmann, W. (1925), Uber die Integration der Differentialglei- uncertainties, in: Bearing Capacity and Settlement, Proceedings
chungen des ebenen Gleichgewichtszustandes fiir den allgemein- of a Symposium held at Duke University, April 5/6, 1965, pp.
plastischen Korper, Thesis, University of Gottingen. 3-8:
Heller, L. W. (1964), Failure modes of impact-loaded footings on Peck, R. B. and Bryant, F. G. (1953), The bearing capacity failure of
dense sand, Technical Report R-281, U.S. Naval Civil Engineering the Transcona elevator; Geotechnique 3, pp. 201-208.
Laboratory, Port Hueneme, California, pp. 1-31. Poplin, J. K. (1965), Dynamic bearing capacity of soils, Report 2:
Hencky, H. (1923), Uber einige statisch bestimmte Falle des Dynamically loaded small-scale footing tests on dry dense sand,
Gleichgewichts in plastischen Korpern; Zeitschrift angew. Math. Techn. Report 3-599, U.S. Army Engineer Waterways Experi-
und Mech. Vol. 3, pp. 241-246. ment Station, Vicksburg, Miss., 133 pp.
Hvorslev, M. J. (1970), The basic sinkage equations and bearing ca- Prandtl, L. (1921), Uber die Eindringungsfestigkeit plastisher Baus-
pacity theories, Techn. Report M-70-1, U.S. Army Waterways toffe und die Festigkeit von Schneiden, Zeitschrift fur Ange-
Experiment Station, Vicksburg, Miss. wandte Mathematik und Mechanik, 1, No. 1, pp. 15-20.
Ishlinskii, A. I. (1944), Osesimmetrichnaia zadacha teorii plastich- Richart, F. E., Jr. (1965), Dynamically loaded foundations, in:
nosti i p'roba Brinellia, Prikladnaia matematika i mekhanika 8, · Bearing Capacity and Settlement of Foundations, Proceedings of
No. 3, pp. 201-208. a Symposium held at Duke University, April 5/6, 1965, pp.
James, C. H. C., Krizek, R. J., arid Baker, W. H. (1969), Bearing 69-81. -
capacity of purely cohesive soils with a nonhomogeneous Reissner, H. (1924), Zurn Erddruckproblem, Proceedings, First In-
strength distribution, Highway Research Record, No. 282, pp._ ternational Conference on Applied Mechanics, Delft, pp. 295-
48-56. . 311.
Jackson, J. G., Jr., and Hadala, P. F. (1964), Dynamic bearing capac- Schultze, E. (1952), Der Widerstand des Baugrundes gegen schriige
ity of soils, Report 3: The application of similitude to small-scale Sohlpressungen, Die Bautechnik 29, No. 12, pp. 336-342.
footing tests, Technical Report 3-599, U.S. Army Waterways Schultze, E. and Horn, A. (1967), The base friction for horizontally
Experiment Station, Vicksburg, Miss. loaded footings in sand and gravel, Geotechnique 17, No. 4, pp.
Kerisel, J. (1967), Scaling laws in soil mechanics, Proceedings, Third 329-347.
Panamerican Conference on Soil Mechanics and Foundation En- Siva Reddy, A. and Srinivasan, R. J. (1967), Bearing capacity of
gineering, Caracas, Vol. III, pp. 69-92. footings on layered clays, Proceedings ASCE, Journal of the Soil
Little, A. L. (1961), Foundations, E. Arnold, London. Mechanics and Foundations Division 93, No. SM2, pp. 83-99.
Lundgren, H. and Mortensen, K. (1953), Determination by the the- Skempton, A. W. (1951), The bearing capacity of clays, Proceedings,
ory of plasticity of the bearing capacity of continuous footings Building Research Congress, London, pp. 180-189.
on sand, Proceedings, Third International Conference on Soil Skempton, A. W. Yassin, A. A. and Gibson, R. E. (1953), Theorie de
Mechanics and Foundation Engineering, Zurich, Vol. I, pp. 409- la force portante des pieux, Anna/es de l'Institut Technique du
412. Biitiment et des Travaux Publics No. 63-64, pp. 285-290.
Mandel, J. (1965), Interference plastique de semelles filantes; Pro- Sokolovskii, V. V. (1960), Statics of Soil Media, Butterworth, Lon-
ceedings, Sixth Intern. Con[. Soil Mech. Found. Engrg., Montreal, don, pp. 1-237.
Vol. II, pp. 127-131. Stuart, J. G. (1962), Interference between foundations with special
Mandel, J. and Saleni;:on, J. (1969), Force portante d'un sol sur une reference to surface footings in sand, Geotechnique 12, No. I,
assise rigide; Proc. Seventh Intern. Con[. Soil Mech. Found. pp. 15-22.
Engrg., Mexico, Vol. 2, pp. 157-164. Suklje, L. (1954 ), Ca pa cite portante des couches cohesives peu per-
Meyerhof, G. G. (1951), The ultimate bearing capacity of founda- meables et d'epaisseur limitee; Proceedings, Yugoslav Society of
tions, Geotechnique 2, pp. 301-332. Soil Mechanics and Foundation Engineering, No. 3, pp. 16-26.
Meyerhof, G. G. (1953), The bearing capacity of foundations under Szechy, K. (1967), Der Einfluss der Sohlflachenform von Streifen-
eccentric and inclined loads, Proceedings, Third International fundamenten auf die Tragfahigkeit und Spannungsausbreitung,
Conference on Soil Mechanics and Foundation Engineering, VD! Zeitschrift 109, No. 8, pp. 339-344.
Ziirich, Vol. I, pp. 440-445. Tcheng, Y. (1957), Fondations superficielles en milieu stratifie, Pro-
Meyerhof, G. G. (1955), Influence of roughness of base and ground- ceedings, Fourth Intern. Conf Soil Mech. Found. Engrg., Lon-
water conditions on the ultimate bearing capacity of founda- don, Vol. 1, pp. 449-452.
tions; Geotechnique 5, No. 3, pp. 227-242. Terzaghi, K. (1943), Theoretical Soil Mechanics, J. Wiley & Sons,
Meyerhof, G. G. (1961), The ultimate bearing capacity of wedge- New York.
shaped foundations, Proceedings, Fifth Intern. Con[. Soil Mech. Terzaghi, K. and Peck, R. B. (1948), Soil Mechanics in Engineering
Found. Engrg., Paris, Vol. 2, pp. 105-109. Practice, New York (J. Wiley), Second edition 1966, 729 pp.
Meyerhof, G. G. (1963), Some recent research on the bearing capac- Tschebotarioff, G. P. (1951), Soil Mechanics, Foundations and Earth
ity of foundations, Canadian Geotechnical Journal l, No. 1, pp. Structures, McGraw-Hill, New York, pp. 655.
16-26. Vesic, A. (1963a), Bearing capacity of deep foundations in sand,
Mizuno, T. (1953), On the bearing power of soil under a uniformly National Academy of Sciences, National Research Council, High-
distributed circular load, Proceedings, Third International Con- way Research Record, 39, pp. 112-153.
ference on Soil Mechanics and Foundation Engineering, Zurich, Vesic, A. (1963b), Theoretical studies of cratering mechanisms af-
Vol. I, pp. 446-449. fecting the stability of cratered slopes, Final Report, Project No.
Muhs, H. (1963), Ueber die zulassige Belastung nicht bindigen A-655, Engineering Experiment Station, Georgia Institute of
Boden, Mitteilungen der Degebo, Heft 16, Berlin, pp. 105-121. Technology, Atlanta, Georgia pp. 1-67.
Muhs, H. and Weiss, K. (1969), Die Grenztragfahigkeit und Schief-
Vesic, A. S. (1964), Model investigations of deep foundations and
stellung ausmittig-lotrecht belasteter Einzelfundamente im Sand
scaling laws, Panel discussion, Session II, Proceedings, North
nach Theorie und Versuch; Mitteilungen der Degebo, No. 22, 84
American Conference on Deep Foundations (Congreso Sobre
pp.
Cimientos Profundos) Mexico City, Vol. II, pp. 525-533.
Muhs, H. and Weiss, K. (1969), The influence of the load inclination
on the bearing capacity of shallow footings, Proceedings, Seventh Vesic, A. S. (1965a), Ultimate loads and settlements of deep founda-
Intern. Con[. Soil Mech. Found. Engrg., Mexico City, Vol. 2, pp. tions in sand, Bearing Capacity and Settlement of Foundations,
187-194. Proceedings of a Symposium held at Duke University, pp. 53-68.
Nadai, A. (1963), Theory of Flow and Fracture of Solids; Vol. II, Vesic, A. S. (1965b), Cratering by explosives as an earth pressure
McGraw-Hill Book Co., p. 470. problem, Proceedings, Sixth International Conference on Soil
Nordlund, R. L. and Deere, D. U. (1970), Collapse of Fargo grain Mechanics and Foundation Engineering, Montreal, Vol. II, pp.
elevator, Proceedings ASCE, Journal of the Soil Mechanics and 427-431.
Foundation Division 96, No. SM2, pp. 585-607; Vesic, A. S. (1967), A study of bearing capacity of deep founda-
Bearing Capacity of Shallow Foundations 147

tions, Final Report, Project B-189, Georgia Institute of Technol- of AASHO road test rigid pavements, Report 97, National Coop-
ogy, Atlanta, Georgia, pp. xvi and 264. erative Highway Research Program, National Academy of Sci-
vesic, A. S. (1969), Effects of scale and compressibility on bearing ence, Washington, D.C., 35 pp. •
capacity of surface foundations, Proceedings, Seventh Interna- West, J. M. and Stuart, J. G. (1965), Oblique loading resulting from
tional Conference on Soil Mechanics and Foundation Engineer- interference between surface footings on sand, Proceedings, Sixth
ing, Mexico City, 1969, Vol. III, pp. 270-272. International Conference on Soil Mechanics and Foundation En-
Vesic, A. S. (1970), Research on Bearing Capacity of Soils (unpub- gineering, Montreal, Vol. II, pp. 214-217.
lished). Westergaard, H. M. (1926), Stresses in concrete pavements computed
Vesic, A. S. and Johnson, W. H. (1963), Model studies of beams by theoretical analysis;Public Roads 7,No. 2, pp. 25-35.
resting on a silt subgrade, Proceedings ASCE, Journal of the Soil White, L. S. (1953), Transcona elevator failure: Eyewitness account,
Mechanics and Foundations Division 89, No. SMl, pp. 1-31. Geotechnique 3, No. 5, pp. 209-214.
vesic, A. S., Wilson, W. E., Clough, G. W. and Tai, T. L. (1965), Whitman, R. V. and Healy, K. A. (1962), Shear strength of sands
Engineering properties of nuclear craters, Technical Report No. during rapid loadings, Proceedings ASCE, Journal of the Soil
3-699, Report 2, U.S. Army Engineer Waterways Experiment Mechanics and Foundations Division 88, No. SM2, pp. 99-132.
Station, Vicksburg, Miss., pp. 1-123. Whitman, R. V. (1970), The response of soils to dynamic loadings,
Vesic, A. S., Banks, D. C., and Woodard, J. M. (1965), An experi- Contract Report No. 3-26, U.S. Army Waterways Experiment
mental study of dynamic bearing capacity of footings on sand, Station, Vicksburg, Miss., 200 + 16 pp.
Proceedings, Sixth International Conference on Soil Mechanics Wu, T. H. and Kraft, L. (1967), The probability of foundation
and Foundation Engineering, Montreal, Vol. II, pp. 209-213. safety, Journal of the Soil Mechanics Division, ASCE 93, No.
vesic, A. S. and Saxena, S. K. (1970), Analysis of structural behavior SM5, pp. 213-231.

-·----·---------------------------

You might also like