You are on page 1of 15

Experimental Study of Shape and Depth Factors and

Deformations of Footings in Sand


Firas H. Janabi, S.M.ASCE 1; Rameez A. Raja, S.M.ASCE 2; Venkata A. Sakleshpur, S.M.ASCE 3;
Monica Prezzi, A.M.ASCE 4; and Rodrigo Salgado, F.ASCE 5
Downloaded from ascelibrary.org by Purdue University Libraries on 11/27/22. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Bearing capacity calculation is an important part of shallow foundation design. The expressions for the shape and depth factors
available in the literature for bearing capacity calculation are mostly empirical and are based on results obtained using limit analysis or the
method of characteristics assuming a soil that is perfectly plastic following an associated flow rule. This paper presents the results of an
experimental program in which load tests were performed on model strip and square footings in silica sand prepared inside a half-cylindrical
calibration chamber with a transparent visualization window. The results obtained from the model footing load tests show a significant
dependence of footing penetration resistance on embedment depth. The load test results were subsequently used to determine experimentally
the shape and depth factors for model strip and square footings in sand. To obtain the displacement and strain fields in the sand domain, the
digital image correlation (DIC) technique was used to analyze the digital images collected at different stages during loading of the model
footing. The DIC results provide insights into the magnitude and extent of the vertical and horizontal displacement and maximum shear strain
contours below and around the footing base during penetration. DOI: 10.1061/JGGEFK.GTENG-10874. © 2022 American Society of Civil
Engineers.
Author keywords: Model footing; Bearing capacity; Embedment depth; Shape and depth factors; Digital image correlation (DIC); Sand.

Introduction where N q and N γ = bearing capacity factors; sq and sγ = shape fac-


tors; dq and dγ = depth factors; B = footing width; γ = represen-
The boundary-value problem of footing penetration in soil involves tative soil unit weight that considers the effect of the groundwater
the determination of the displacement and stress fields in the soil table elevation with respect to the footing base; and q0 = surcharge
domain. The bearing capacity equation (Brinch Hansen 1970;
(vertical effective stress) at the footing base level.
Meyerhof 1951, 1963; Terzaghi 1943) is one of the tools that
This equation has been routinely used in foundation engineering
geotechnical engineers typically use to estimate the limit unit
practice for decades. The bearing capacity factor N q for a frictional,
bearing capacity qbL (resistance to plunging) of footings in sand
weightless soil has an exact solution that depends on the friction
(Sakleshpur et al. 2021a, b). Fig. 1 shows the classical failure
angle ϕ of the soil (Lyamin et al. 2007; Reissner 1924; Smith
(collapse) mechanism for a footing with a level base embedded in
2005). The expression for the bearing capacity factor N γ , based on
a uniform sand deposit. The bearing capacity equation for a footing
work done by Martin (2005) using the method of characteristics,
embedded in uncemented sand (c ¼ 0) and subjected to an axial,
is accurate for soils that are assumed to be perfectly plastic and
compressive load has two terms (Salgado 2022)
follow an associated flow rule (in a material following a model with
the Mohr–Coulomb yield criterion, dilatancy angle ψ equal to ϕ),
1 although it is well known that, for real sands, ψ is considerably less
qbL ¼ q0 N q sq dq þ γBN γ sγ dγ ð1Þ
2 than ϕ (Loukidis and Salgado 2011; Salgado 2020; Tatsuoka 1987).
However, limited guidance on the estimation of a representative
1 friction angle for calculation of bearing capacity factors N q and N γ
Ph.D. Candidate, Lyles School of Civil Engineering, Purdue Univ., West
Lafayette, IN 47907; Assistant Lecturer, Dept. of Civil Engineering, College is available in the literature. The difficulty in estimating a repre-
of Engineering, Univ. of Babylon, Babylon, Iraq (corresponding author). sentative value of ϕ for use in the bearing capacity equation is that
ORCID: https://orcid.org/0000-0003-0230-7025. Email: fhashimj@purdue there is not a single value of ϕ to use: each soil element along the
.edu; fhashem77@yahoo.com slip surface is subjected to a different stress state and loading path.
2
Ph.D. Candidate, Lyles School of Civil Engineering, Purdue Univ.,
For instance, a soil element along the slip surface in the active zone
West Lafayette, IN 47907. ORCID: https://orcid.org/0000-0001-7463
-9120. Email: raja4@purdue.edu (Fig. 1) would be subjected to a high mean effective stress and may
3
Ph.D. Candidate, Lyles School of Civil Engineering, Purdue Univ., experience a loading path that approximates triaxial compression,
West Lafayette, IN 47907. Email: vsaklesh@purdue.edu whereas a soil element along the slip surface in the passive zone
4
Professor of Civil Engineering, Lyles School of Civil Engineering, would be subjected to a comparatively lower mean effective stress
Purdue Univ., West Lafayette, IN 47907. Email: mprezzi@ecn.purdue.edu and may experience a loading path that approximates triaxial ex-
5
Charles Pankow Professor in Civil Engineering, Lyles School of Civil tension. In addition, the use of a friction angle obtained based on
Engineering, Purdue Univ., West Lafayette, IN 47907. Email: rodrigo@ the results of analyses that do not account for flow rule nonasso-
ecn.purdue.edu
ciativity may lead to significant error in the calculation of N q and
Note. This manuscript was submitted on March 12, 2022; approved on
September 27, 2022; published online on November 24, 2022. Discussion N γ (Loukidis and Salgado 2009a). The use of a nonassociated flow
period open until April 24, 2023; separate discussions must be submitted rule (ϕ > ψ) is important for accurate estimation of bearing capac-
for individual papers. This paper is part of the Journal of Geotechnical and ity of footings in sand (Salgado 2020; Yin et al. 2001). There has
Geoenvironmental Engineering, © ASCE, ISSN 1090-0241. long been interest in using the classical theories of plasticity also

© ASCE 04022128-1 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2023, 149(2): 04022128


in sand. In addition, we used the digital image correlation (DIC)
technique on the digital images collected during different stages of
model footing loading to obtain the displacement and strain fields
in the sand domain.

Materials and Methods

Fig. 1. Schematic of classical bearing capacity failure mechanism for a Calibration Chamber and Loading Equipment
footing embedded in soil (QL is the limit load).
A DIC calibration chamber (outer diameter = 1,620 mm and
height = 1,200 mm) was used to perform displacement-controlled
load tests on model footings in the Bowen Laboratory at Purdue
Downloaded from ascelibrary.org by Purdue University Libraries on 11/27/22. Copyright ASCE. For personal use only; all rights reserved.

for nonassociative materials. For example, Davis (1968) proposed a University (Fig. 2). The chamber consists of a rigid, half-
redefinition of the friction angle based on the soil’s friction and cylindrical, steel tank with a 620 × 540 mm observation window
dilatancy angles so that these theories could be used consisting of a 102-mm-thick (4 in:) acrylic plexiglass panel (poly-
  methyl methacrylate); the plexiglass panel is shielded by a normal
 −1 cos ψ sin ϕ glass panel on the side of the sand sample. Target dots with known
ϕ ¼ tan ð2Þ
1 − sin ψ sin ϕ coordinates were printed on the inner face of the transparent normal
glass panel to convert the size of the image in pixels to the size of
Shape factors introduce the effect of footing geometry for foot- the object in units of length. The loading frame, which consists of a
ings with finite dimensions (e.g., square, circular, and rectangular steel structure and a hydraulic actuator, can apply a monotonic load
footings), whereas depth factors account for the shear resistance at a rate of 0.1–10 mm=s. The actuator is controlled by an axis
mobilized along the portion of the slip surface above the base of motion controller (model RMC75E) that has dual loop position–
an embedded footing. Several researchers have proposed equations force algorithms to measure the actuator displacement and the cor-
to determine the shape and depth factors either empirically (Brinch responding load during the model footing load tests. Further details
Hansen 1970; Meyerhof 1963; Vesic 1973), by using finite element of the chamber components and testing methodology can be found
(FE) analysis (Antão et al. 2012; Loukidis and Salgado 2009a, in Arshad et al. (2014).
2011; Lyamin et al. 2007; Zhu and Michalowski 2005), by using Given the symmetry of the boundary-value problem, two alu-
the method of characteristics (Zhu et al. 2001), or as a result of minum model footings were used in this study: a half-square foot-
centrifuge tests (Okamura et al. 2002). A summary of the most ing with a width and thickness of 90 mm, and a half-strip footing
commonly used equations available in the literature for the shape with a width and thickness of 50 mm and a length of about 270 mm.
and depth factors can be found in Lyamin et al. (2007) and The coupling between the loading frame and the model footings
Salgado (2022). consists of a ball joint hinge that eliminates any minor bending mo-
In this paper, we attempt to answer two questions: (1) What ment that may occur during loading. A 50-kN tension–compression
should the values of the shape and depth factors be in order to cor- load cell was attached to the piston rod of the actuator to measure
rectly estimate the limit unit bearing capacity of strip and square the force applied during loading of the model footings. To mini-
footings embedded in sand? (2) How do we go about determining mize friction along the interface between the normal glass panel
the values of these factors experimentally? To answer these ques- and the model footing, synthetic grease was used to lubricate the
tions, we performed load tests on model strip and square footings contact surfaces of the square and strip model footings. Further-
placed on the surface and embedded at different depths in Ohio more, for the square footing, a 1-mm groove was drilled into the
Gold Frac sand samples prepared inside a calibration chamber. The load-transfer column to decrease the friction between the glass
results obtained from the model footing load tests were used to panel and the load-transfer column assembly. The effect of a rough
determine experimentally the shape and depth factors for footings interface on the footing load-settlement response was simulated by

Fig. 2. Model footing load test in the DIC calibration chamber using Ohio Gold Frac sand.

© ASCE 04022128-2 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2023, 149(2): 04022128


gluing 180-grit sandpaper to the base and sides of the model foot- displacement and strain fields in the sand domain were obtained
ing, except the side in contact with the glass panel. using an optical measurement technique called digital image
correlation (Doreau-Malioche et al. 2019; Galvis-Castro et al.
2019a, b; Ganju et al. 2020, 2021; Salgado 2013; Tehrani et al.
Test Sand 2016, 2018; Tovar-Valencia et al. 2021, 2018a, b). This technique
Ohio Gold Frac (OGF) sand, which is a silica sand (SiO2 ¼ uses a correlation algorithm to locate and track the same pattern
99.7%), was used in the experiments performed in the DIC calibra- of grey level intensity from an undeformed (reference) image to
tion chamber. It consists of subangular-to-subrounded particles a distorted image (Sutton et al. 2009). The image analysis was
(Han et al. 2018) and is classified as poorly-graded (SP) according performed using the commercial DIC software VIC-2D (Correlated
to the Unified Soil Classification System [ASTM D2487 (ASTM Solutions 2009). Further details on the camera calibration, the
2020)]. The mean particle size D50 , coefficient of uniformity CU, image texture (speckle pattern), and the precision and accuracy of
and coefficient of curvature CC of the sand are 0.6, 1.37, and 1.0, the DIC measurements can be found in Arshad (2014) and Janabi
respectively. The maximum void ratio [ASTM D4254 (ASTM et al. (2022).
Downloaded from ascelibrary.org by Purdue University Libraries on 11/27/22. Copyright ASCE. For personal use only; all rights reserved.

2016a)], minimum void ratio [ASTM D4253 (ASTM 2016b)],


and critical-state friction angle ϕc of the sand in triaxial compres-
sion are 0.81, 0.59, and 31.7°, respectively. Test Protocol and Program
OGF sand samples were prepared inside the calibration chamber
Image Acquisition System and Image Analysis using the air pluviation method (Arshad 2014; Lee et al. 2011).
Details of the sample preparation method for the surface footing
Two complementary metal-oxide-semiconductor (CMOS) cameras (D ¼ 0) and verification of sample density can be found in Janabi
with 5-megapixel resolution were used to collect synchronized et al. (2022). For the embedded footing (D > 0), the sand sample
digital images of the sand domain at a rate of 2 frames per second. was prepared in two stages. In stage 1, the sand was pluviated up
The observation window was illuminated by four light-emitting to the desired embedment depth of the model footing. After the
diode (LED) lamps during loading of the model footing. The installation of the model footing against the plexiglass panel, in
stage 2, the sample preparation advanced up to the desired elevation
of the sand surface above the footing base.
After sample preparation and footing installation, the model
footing was connected to the piston of the hydraulic actuator. The
cameras were then aligned with the observation window of the cal-
ibration chamber. The model footing was monotonically jacked
into the sand at a rate of 0.1 mm=s. The vertical displacement w of
the model footing and the corresponding axial load Q were ex-
tracted from the motion controller data. Since the embedment depth
D of a footing is typically less than the width B of the footing (Chai
and Kutter 2000; Salgado 2022), the model footing load tests were
performed for D=B values ranging from 0 to 1.
Fig. 3 shows a schematic of an axially loaded model footing
embedded in OGF sand prepared inside the DIC chamber. A total
of 14 model footing load tests were performed in uniform dense
and medium-dense OGF sand. Table 1 summarizes the test condi-
Fig. 3. Schematic of model footing embedded in OGF sand pluviated
tions, the footing embedment depth, and the relative density of the
inside the chamber.
sand sample. The test ID used for the sand samples consists of the

Table 1. Model footing load test series in uniform OGF sand


Footing Footing width, Relative embedment Sand sample Relative density, Bearing capacity
Test # Test ID shape B (mm) depth, D=B density DR (%) (kPa)a
1 D-SQ-0B Square 90 0 Dense 91 200
2 D-SQ-0.25B Square 90 0.25 Dense 90 300
3 D-SQ-0.50B Square 90 0.5 Dense 90 405
4 D-SQ-0.75B Square 90 0.75 Dense 91 517
5 D-SQ-1.0B Square 90 1 Dense 91 590
6 MD-SQ-0B Square 90 0 Medium dense 48 48
7 MD-SQ-0.50B Square 90 0.5 Medium dense 50 86
8 MD-SQ-1.0B Square 90 1 Medium dense 50 141
9 D-SQ-0B* Square 50 0 Dense 91 120
10 D-ST-0B Strip 50 0 Dense 90 171
11 D-ST-0.25B Strip 50 0.25 Dense 90 226
12 D-ST-0.50B Strip 50 0.5 Dense 91 266
13 D-ST-0.75B Strip 50 0.75 Dense 91 310
14 D-ST-1.0B Strip 50 1 Dense 90 338
Note: D = dense sand; MD = medium-dense sand; SQ = square footing; and ST = strip footing.
a
For dense sand, the bearing capacity of the footing corresponds to the peak unit load qbL of the footing load-settlement curve, whereas for medium-dense
sand, the bearing capacity of the footing corresponds to the unit load qb;ult defined at 10% relative settlement (w=B ¼ 0.1).

© ASCE 04022128-3 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2023, 149(2): 04022128


Downloaded from ascelibrary.org by Purdue University Libraries on 11/27/22. Copyright ASCE. For personal use only; all rights reserved.

Fig. 4. Load-settlement response of model footing in dense OGF sand: (a) square footing; and (b) strip footing.

following three parts: (1) the relative density of the sand layer; strip model footings tested in dense and medium-dense OGF sand,
(2) the footing shape; and (3) the footing embedment depth. For respectively, up to relative settlement w=B levels of 0.25–0.30.
example, test ID “D-SQ-1.0B” means that the sand layer is dense, Referring to Fig. 4, for the model square footing, the w=B value
the footing shape is square, and the embedment depth D of the needed to mobilize the limit unit bearing capacity qbL (peak value
footing is equal to its width B. The bearing capacities of the model of qb ) increases from 7% to 13% as D=B increases from 0 to 1,
strip and square footings obtained from the load tests are also in- whereas for the model strip footing, the w=B value needed to mo-
cluded in Table 1. bilize qbL increases from 8% to 16% as D=B increases from 0 to 1.
In contrast, Fig. 5 shows that the load-settlement curves obtained
for the model square footing in medium-dense OGF sand do not
Results and Discussion exhibit a clear peak, regardless of the relative embedment depth
D=B. In addition, Figs. 4 and 5 also show that, irrespective of the
relative density of OGF sand and the shape of the model footing,
Effect of Footing Embedment
the load-settlement response of an embedded footing is stiffer
The effective stress state, fabric, relative density, and intrinsic var- than that of a footing placed on the sand surface. The footing load-
iables of sand control its stress–strain response from small-to-large settlement curves shown in both Figs. 4 and 5 are representative of
strain levels (Salgado et al. 2000). Embedding a footing in sand a series of load tests performed to ensure repeatability of the test
increases its limit unit bearing capacity over that of a surface foot- results. For example, Fig. 6 shows that the load-settlement curves
ing, all other things being equal, because (1) additional work needs obtained for the model square footing on dense and medium-dense
to be done for the sand mass in the passive zone (see Fig. 1) to OGF sand are consistent and repeatable.
move up against the surcharge above the footing base, and (2) addi- To assess the degree of improvement in bearing capacity due to
tional shear resistance is mobilized along the portion of the slip footing embedment, the ratio of the bearing capacity of an em-
surface above the footing base. Figs. 4 and 5 show the effect of bedded footing (D > 0) to that of a surface footing (D ¼ 0) was
embedment depth on the load-settlement response of square and considered. The bearing capacity of the model footing in dense

Fig. 5. Load-settlement response of model square footing in medium- Fig. 6. Example of repeatable test results obtained for model square
dense OGF sand. footing on dense and medium-dense OGF sand.

© ASCE 04022128-4 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2023, 149(2): 04022128


Downloaded from ascelibrary.org by Purdue University Libraries on 11/27/22. Copyright ASCE. For personal use only; all rights reserved.

Fig. 7. Comparison of the values of the bearing capacity ratio obtained in this study with those reported in the literature for (a) square footings in
dense and medium-dense sand; and (b) strip footings in dense sand.

OGF sand corresponds to the peak unit load qbL of the qb versus
w=B curve, whereas the bearing capacity of the footing in medium-
dense OGF sand corresponds to the unit load qb;ult defined at 10%
relative settlement (w=B ¼ 0.1). The results obtained in this study
indicate that, regardless of the sand’s relative density, the improve-
ment in bearing capacity due to the embedment of a model square
footing in OGF sand may be approximated by a linear relationship
[Fig. 7(a)] such that the bearing capacity ratio is equal to 3 for
D=B ¼ 1. In addition, the results obtained from the model square
footing load tests performed in this study are in line with those
obtained from the full-scale square footing load tests performed
by FHWA [as reported by Briaud (2007)] and Lutenegger and
Adams (1998)] [Fig. 7(a)]. The values of the bearing capacity ratio
of Lutenegger and Adams (1998) correspond to medium-dense
sand (DR ¼ 50%), with footing bearing capacity defined as the unit
load at w=B ¼ 0.1. The FHWA footing load-settlement curves
reported by Briaud (2007) for D=B ¼ 0, 0.5, and 1 did not exhibit
Fig. 8. Effect of footing shape and relative embedment depth on the
a distinct peak, and although the footing load-settlement curve for
bearing capacity ratio for model footings in dense OGF sand (lines are
D=B ¼ 0 extended beyond w=B ¼ 0.1, the footings embedded at
best fit to the data points).
D=B ¼ 0.5 and 1 were loaded only to w=B values of 0.059 and
0.074, respectively. Consequently, we determined the value of
the unit load at w=B ¼ 0.1 for the surface footing directly using
Determination of d q
the data reported by Briaud (2007), but for the embedded footings
(D=B ¼ 0.5 and 1), we extrapolated the values of the unit load to Consistent with the classical assumption that dγ ¼ 1, the effect of
w=B ¼ 0.1 to determine the bearing capacity ratio. The values of the shear strength of sand along the portion of the slip surface above
the bearing capacity ratio obtained from the field tests of FHWA the base of an embedded footing is assumed to be fully captured
and Lutenegger and Adams (1998) are of the order of 2.6–2.7 by the depth factor dq. The principle of superposition has been used
for D=B ¼ 1, which is in reasonable agreement with the value to estimate the limit unit bearing capacity of footings in sand for
(¼ 3.0) obtained in this study. Fig. 7(b) shows that the results decades; however, it is incapable of properly capturing the varia-
obtained from the model strip footing load tests performed in this tion of dq with footing embedment depth even in a simple Mohr–
study are in good agreement with those obtained from centrifuge Coulomb material (Lyamin et al. 2007). According to Lyamin et al.
tests performed in dense Toyoura sand (DR ¼ 86%) by Kimura (2007), the contribution of the shear strength of soil located above
et al. (1985). the level of the footing base is lost when the overburden soil is
The bearing capacity of footings embedded in sand also depends replaced by an equivalent surcharge, and, in addition, the lateral
on the footing shape. Fig. 8 shows that, for dense OGF sand, the extent of the collapse mechanism is greater in the presence of soil
increase in bearing capacity due to footing embedment is greater above the footing base when compared with an equivalent surface
for the model square footing than for the model strip footing. For a footing. Furthermore, these authors found that the simulation of
relative embedment depth D=B of 0.5, the bearing capacity ratio overburden soil as an equivalent surcharge produces a different soil
ranges from 1.55 for the model strip footing to 2.0 for the model response below the footing base when compared with soil-on-soil
square footing, whereas for D=B ¼ 1, the bearing capacity ratio interaction. Therefore, we have directly embedded the model strip
ranges from about 2.0 for the model strip footing to 3.0 for the footing in OGF sand to determine experimentally the depth
model square footing. factor dq.

© ASCE 04022128-5 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2023, 149(2): 04022128


 
The work of Lyamin et al. (2007) on limit analysis is based on 1 þ sin ϕ π tan ϕ
the assumption that sand follows an associated flow rule (ϕ ¼ ψ), Nγ ¼ e − 0.6 tanð1.33ϕÞ ð8Þ
1 − sin ϕ
which is not the best approximation to reality for a sand. Loukidis
and Salgado (2009a), based on results obtained from FE simula- Eq. (8) closely approximates the exact values of N γ obtained by
tions of strip and circular footings on sand that follows a nonasso- Martin (2005) for an associated flow rule using the method of
ciated flow rule and the Mohr–Coulomb yield criterion, showed characteristics.
that the bearing capacity factors N q and N γ decrease by 12%– 6. Determine the depth factor dq from the limit unit bearing capac-
42% and 15%–42%, respectively, in comparison with the values ity qbL jD>0 of a strip footing embedded at a given depth using
of N q and N γ obtained for ϕ ¼ ψ; in which the difference increases
with increasing values of ϕ. Therefore, when using a Mohr– qbL jD>0 − qbL jD¼0
Coulomb yield criterion for estimation of the limit unit bearing dq ¼ ð9Þ
q0 N q
capacity of footings in sand, it is important to use a nonassociated
flow rule for sand, with a dilatancy angle ψ that is smaller than the where q0 = surcharge at the footing base level (¼ γD).
Downloaded from ascelibrary.org by Purdue University Libraries on 11/27/22. Copyright ASCE. For personal use only; all rights reserved.

friction angle ϕ (Frydman and Burd 1997; Loukidis and Salgado For dense OGF sand with DR ≈ 90%, the representative friction
2009a; Salgado 2020; Yin et al. 2001). and dilatancy angles obtained after the aforementioned procedure
To determine experimentally the depth factor dq, model strip are 49.5° and 22.2°, respectively, whereas the nonassociativity fac-
footings were tested in dense OGF sand at different relative em- tor F is equal to 0.88. These values of ϕ and ψ are not the values at a
bedment depths D=B ranging from 0 to 1. The Bolton (1986) point but rather are values representative of the collapse mechanism
framework was used to compute the dilatancy angle ψ, and the pro- that forms below the footing (Fig. 1), such that when they are
cedure of Loukidis and Salgado (2009a) was adopted to estimate plugged into the bearing capacity equation, they produce the value
the bearing capacity factors N q and N γ . The advantage of using the of bearing capacity obtained for the model strip footing tested in
Loukidis and Salgado (2009a) method for estimating the values of dense OGF sand considering flow rule nonassociativity. Fig. 9
N q and N γ is that flow rule nonassociativity is explicitly considered shows that the depth factor dq, calculated using the aforementioned
in the method equations. The following procedure was used to procedure, decreases with increasing values of D=B for the model
determine the depth factor dq for the model strip footing tested in strip footing in dense OGF sand; the trend is consistent with that
this study: reported by Lyamin et al. (2007); however, a direct one-to-one
1. Measure the limit unit bearing capacity qbL jD¼0 for a strip foot- comparison with their results is not valid due to the implicit
ing placed on the sand surface (D ¼ 0). assumption of an associated flow rule made for sand in their limit
2. Back-calculate the bearing capacity factor N γ for a strip footing analysis calculations. The value of dq for the model strip footing in
placed on the sand surface using dense OGF sand decreases from 1.41 for D=B ¼ 0.25 to 1.08 for
D=B ¼ 1. Contrary to the traditional expressions for dq reported in
2qbL jD¼0
Nγ ¼ ð3Þ the literature (Brinch Hansen 1970; Meyerhof 1963; Vesic 1973),
γB the trend in Fig. 9 suggests that the value of dq does not approach 1
as D=B → 0 but rather increases with decreasing values of D=B;
3. Plug the value of N γ obtained from step 2 into the following this disparity, as discussed by Lyamin et al. (2007), is due to the
equation (Loukidis and Salgado 2009a, 2011): inadequacy of the logic of superposition and separation of the dif-
  ferent contributions to bearing capacity.
1 þ sin ϕ Fðϕ;ψÞπ tan ϕ
Nγ ¼ e − 0.6 tanð1.33ϕÞ ð4Þ
1 − sin ϕ
Determination of s γ
where the nonassociativity factor F is given by (Loukidis and The shape factor sγ accounts for the bearing capacity of different
Salgado 2009a) footing shapes other than strip (e.g., square, circular, or rectangular
footing) placed on the sand surface (D ¼ 0)
Fðϕ; ψÞ ¼ 1 − tan ϕ½tanð0.8ðϕ − ψÞÞ2.5 ð5Þ
The dilatancy angle ψ in Eq. (5) is estimated using Bolton’s
equation for plane strain conditions

ψ ¼ 1.25ðϕ − ϕc Þ ð6Þ

where ϕc = critical-state friction angle of the sand.


4. Solve for the representative friction and dilatancy angles using
Eqs. (4)–(6).
5. Calculate the bearing capacity factor N q using (Loukidis and
Salgado 2009a)

1 þ sin ϕ Fðϕ;ψÞπ tan ϕ


Nq ¼ e ð7Þ
1 − sin ϕ

For sands that are modeled as a material following the Mohr–


Coulomb yield criterion with an associated flow rule (ϕ ¼ ψ), the
nonassociativity factor F is equal to 1, and, consequently, Eq. (7)
Fig. 9. Depth factor dq versus relative embedment depth D=B for
reduces to the form obtained by Reissner (1924), whereas Eq. (4)
model strip footing in dense OGF sand.
reduces to the expression given by Loukidis and Salgado (2011):

© ASCE 04022128-6 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2023, 149(2): 04022128


qbL jBL
sγ ¼ ð10Þ
qbL;strip

where qbL jBL = limit unit bearing capacity of a square, circular, or


rectangular footing placed on the sand surface; and qbL;strip = limit
unit bearing capacity of a strip footing placed on the sand surface.
Several studies have shown that the expressions to compute sγ
depend on the footing aspect ratio B=L and the friction angle ϕ or
the relative density DR of the sand (Loukidis and Salgado 2009a;
Lyamin et al. 2007; Okamura et al. 2002; Vesic 1973). For a given
relative density DR, the shape factor sγ for a square or circular foot-
ing is less than 1 (Loukidis and Salgado 2011; Salgado 2022)
 
Downloaded from ascelibrary.org by Purdue University Libraries on 11/27/22. Copyright ASCE. For personal use only; all rights reserved.

DR
sγ ¼ 1 − 0.23 ð11Þ
100

Eq. (11) is a result of rigorous FE analyses of rough circular Fig. 10. Shape factor sq versus relative embedment depth D=B for
footings (B ¼ 1–3 m) on dense sand (DR ¼ 60%–90%, ϕc ¼ 31.6°) model square footing in dense OGF sand.
performed by Loukidis and Salgado (2011) using an advanced,
two-surface-plasticity sand model; the equation may also be used
for square footings (Salgado 2022). The constitutive model is based qbL jD>0 − qbL jD¼0
sq ¼ ð14Þ
on critical-state soil mechanics and accounts for strain-softening, q0 N q dq
both stress-induced and inherent anisotropy, and flow rule nonas-
sociativity (Loukidis and Salgado 2009b).
For dense OGF sand with DR ≈ 90%, the representative friction
To obtain a direct estimate of the shape factor sγ , we performed
and dilatancy angles obtained after the aforementioned procedure
a load test (test ID: D-SQ-0B*) (Table 1) on a 5-cm-wide model
are 47.2° and 30.9°, respectively, whereas the nonassociativity fac-
square footing placed on the surface of dense OGF sand. This size
tor F is equal to 0.97. Fig. 10 shows that the shape factor sq, cal-
(B ¼ 5 cm) is identical to that of the model strip footing tested in
culated using the above procedure, increases with increasing values
this study. Using Eq. (10), the value of sγ determined based on the
of D=B for the model square footing in dense OGF sand. Contrary
results obtained from the load tests performed on the 5-cm-wide
to the assumption that shape factors are independent of footing
model strip and square footings placed on the surface of dense
depth (Brinch Hansen 1970; Meyerhof 1963; Vesic 1973), we
OGF sand (DR ≈ 90%) is equal to 0.70. This value is in reasonable
see that, for a given friction and dilatancy angle, the shape factor
agreement with that obtained from Eq. (11) (¼ 0.79). Although it
sq depends on the relative embedment depth D=B of the footing;
does not fully validate Eq. (11), it does provide some confidence
this finding is consistent with the findings of Lyamin et al. (2007).
that, at least for dense sand, the equation captures the effect of
The value of sq for the model square footing in dense OGF sand
shape sufficiently closely.
increases from 1.10 for D=B ¼ 0.25 to 1.39 for D=B ¼ 1. The
model footing test data and the methodology for estimation of
Determination of s q shape and depth factors presented in this paper may be used to
benchmark or validate numerical simulations of footing penetration
The shape factor sq accounts for the bearing capacity of different in sand.
footing shapes, other than strip, embedded in sand. The following
procedure was used to determine the shape factor sq for the model
square footing tested in this study Scale Effect
1. Measure the limit unit bearing capacity qbL jD¼0 for a square The bearing capacity factor N γ decreases as the footing size B in-
footing placed on the sand surface (D ¼ 0). creases. This observation has been referred to in the literature as the
2. Back-calculate the corresponding factor N γ sγ for a square foot- size (or scale) effect on shallow foundation limit bearing capacity
ing placed on the sand surface using (De Beer 1965; Bolton and Lau 1989; Cerato and Lutenegger 2007;
Herle and Tejchman 1997; Hettler and Gudehus 1988; Kimura
2qbL jD¼0
N γ sγ ¼ ð12Þ et al. 1985; Perkins and Madson 2000; Siddiquee et al. 1999; Ueno
γB et al. 1998, 2001). Two factors may contribute, depending on
circumstances, to the scale effect observed for footings on sand:
3. Divide the value of N γ sγ obtained from step 2 by the value of sγ
(1) pressure-level effect (Loukidis and Salgado 2011); and
estimated using Eq. (11) to obtain the value of N γ .
(2) particle-size effect (Tatsuoka et al. 1991, 1994). The pressure-
4. Plug the value of N γ obtained from step 3 into Eq. (4). The dilat-
level effect is related to the fact that greater values of footing width
ancy angle ψ for triaxial conditions is given by (Bolton 1986;
B correlate with greater values of footing bearing capacity qbL and
Salgado 2022)
average mean effective stress σm0 in the collapse mechanism, and
ψ ¼ 2ðϕ − ϕc Þ ð13Þ consequently, with smaller representative values of ϕ and ψ for use
in the bearing capacity calculation. Thus, the representative values
5. Solve for the representative friction and dilatancy angles using of ϕ and ψ determined from the results of the 1 g model footing
Eqs. (4), (5), and (13). load tests performed in this study are likely to be greater than those
6. Calculate the bearing capacity factor N q [Eq. (7)] using the ϕ mobilized in the collapse mechanism of full-scale footings tested in
and ψ values obtained from step 5. dense OGF sand.
7. Determine the shape factor sq from the limit unit bearing capac- The effect of pressure level on the bearing capacity factor N γ
ity qbL jD>0 of a square footing embedded at a given depth using may be quantified using (Loukidis and Salgado 2011)

© ASCE 04022128-7 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2023, 149(2): 04022128


equation was developed primarily for full-scale strip footings
(B ¼ 1–3 m) placed on the surface of dense Toyoura sand, with
γ 0 B=pA values ranging from 0.2 to 0.6. Both Toyoura sand and
OGF sand are poorly graded silica sands with comparable particle
morphology (consisting of subangular to angular particles). The
value of γ 0 B=pA obtained for the model strip footing placed on
the surface of dense OGF sand is equal to 0.008, which is about
two orders of magnitude smaller than the above range. In addi-
tion, for the model strip footing (B ¼ 5 cm) on dense OGF sand
(DR ¼ 90%), the value of N γ predicted by Eq. (15) is 509, which
is 23% greater than that obtained in this study (¼ 415) using
Eq. (3). Eq. (15) likely overcorrects for the pressure-level effect
observed for the model strip footing tested on dense OGF sand
Downloaded from ascelibrary.org by Purdue University Libraries on 11/27/22. Copyright ASCE. For personal use only; all rights reserved.

because the equation was developed for a range of γ 0 B=pA values


that are much higher than those for the model footing tests.
Fig. 11. Unit load qb versus relative settlement w=B curve for the mod- However, given the radically different scales, the 23% difference
el square footing on dense OGF sand (D ¼ 0) (DIC analysis was car- is surprisingly small.
ried out for loading stages 1–4). Given that the shape factor sγ [Eq. (10)] is determined as the
ratio of the bearing capacity of a square, circular, or rectangular
  0 −0.4 footing placed on the sand surface to that of a strip footing, all other
D γ B
N γ ¼ 2.82 exp 3.64 R ð15Þ things being equal, the pressure-level effect on sγ will be small,
100 pA
perhaps negligibly so. However, this result may not be the case for
where γ 0 = effective unit weight of sand; DR = relative density; the depth factor dq [Eq. (9)] and the shape factor sq [Eq. (14)]; both
B = footing width; and pA = reference stress (¼ 100 kPa). This depend on the bearing capacity factor N q. N q is a function of the

Fig. 12. Cumulative vertical displacement contours (in mm; displacements negative in compression) for model square footing on dense OGF sand
(D ¼ 0) corresponding to loading stages 1–4.

© ASCE 04022128-8 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2023, 149(2): 04022128


Downloaded from ascelibrary.org by Purdue University Libraries on 11/27/22. Copyright ASCE. For personal use only; all rights reserved.

Fig. 13. Cumulative vertical displacement contours (in mm; displacements negative in compression) for model square footing embedded in dense
OGF sand (D ¼ B) corresponding to loading stages 1–4.

representative friction and dilatancy angles of sand in the footing’s a monotonic increase up to a peak value (qbL ) of 200 kPa at
collapse mechanism. w=B ¼ 7%. The unit load near the end of the model footing load
The particle size effect generally refers to the observation that test was almost constant at 167 kPa. Also shown in Fig. 11 are
the value of N γ becomes larger at an increasing rate as the ratio four stages on the qb versus w=B curve where DIC analysis was
B=D50 decreases (Tatsuoka et al. 1991, 1997; Tejchman and carried out to obtain the corresponding displacement and strain
Herle 1999). However, once this ratio exceeds 50 (Tejchman fields in the sand domain. The four stages considered are (1) unit
and Herle 1999; Toyosawa et al. 2013), the particle size effect load corresponding to 0.5qbL , (2) limit unit bearing capacity qbL
on the footing’s bearing capacity is negligible. The ratio of the (corresponding to the peak value of qb ), (3) unit load in the soft-
width B of the model footing to the mean particle size D50 of ening stage at the point where the slope of the qb versus w=B
OGF sand is equal to 83 for the 5-cm-wide strip and square footings curve changes, and (4) unit load at large relative settlement
and 150 for the 9-cm-wide square footing; both these values are (w=B ¼ 0.26).
greater than 50.
Vertical and Horizontal Displacement Fields
Displacement and Strain Fields below the Footing It is useful to study the evolution of the vertical and horizontal
displacements in the sand domain at different stages of footing
Analysis Methodology loading for both the surface (D ¼ 0) and embedded footings
Fig. 11 shows a plot of unit load qb versus relative settlement w=B (D > 0). Figs. 12 and 13 show the vertical displacement contours
for the model square footing tested on the surface (D ¼ 0) of (in units of mm) for the model square footing placed on the
dense OGF sand. The unit load qb for the model footing shows surface and embedded at 1B, respectively, in dense OGF sand

© ASCE 04022128-9 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2023, 149(2): 04022128


Downloaded from ascelibrary.org by Purdue University Libraries on 11/27/22. Copyright ASCE. For personal use only; all rights reserved.

Fig. 14. Cumulative horizontal displacement contours (in mm) for model square footing on dense OGF sand (D ¼ 0) corresponding to loading
stages 1–4.

corresponding to the four stages described previously. The hori- the relative embedment depth D=B of the footing increases
zontal axis in these figures is the radial or horizontal distance x from 0 to 1.
from the centerline of the model footing to the point of interest in In addition to the vertical displacement, the horizontal dis-
the sand domain, normalized by the width B of the footing; the placement of the sand particles also contributes to the develop-
vertical axis is the depth z from the free surface of the sand to ment of strains in the sand domain. Figs. 14 and 15 show that,
the point of interest in the domain, also normalized by the width for all four stages of model footing loading, the effect of footing
B of the footing. Stage 1 of the unit load-relative settlement embedment on the magnitude of the horizontal displacement of
curve is reached at w=B ¼ 1%, regardless of footing embedment. the sand particles is small. Similar to the vertical displacement
At stage 2 of model footing loading, the unit load on the footing contours, the horizontal displacement contours for stage 4 extend
base increases with embedment depth, and the corresponding rel- down by 1.4B–2.0B below the base of the model square footing
ative settlement increases from 7% for the surface footing to 13% and laterally by more than 2.5B from the footing centerline as
for the embedded footing (D=B ¼ 1). A relative settlement of the relative embedment depth D=B of the footing increases from
10% is needed to attain stage 3 for the surface footing, but 0 to 1.
for an embedded footing with D=B ¼ 1, the relative settlement
needed to attain stage 3 increases to a value of 15%. Stage 4 of Shear Strain Fields
model footing loading is reached at relative settlement levels of Figs. 16 and 17 show the maximum engineering shear strain
26% and 30% for the surface footing and embedded footing contours (in percent) for the model square footing placed on the
(D=B ¼ 1), respectively. DIC analyses of the digital images col- surface and embedded at 1B, respectively, in dense OGF sand at
lected during loading of the model square footing in dense OGF different stages of footing loading. The maximum engineering
sand show that, for stage 4, the vertical displacement contours shear strains were computed using the built-in strain computation
extend down by 1.5B–2.0B below the footing base and laterally function in the VIC-2D software; further details of the proce-
by about 2.5B to more than 2.5B from the footing centerline as dure can be found in Ganju et al. (2021). From Figs. 16 and 17,

© ASCE 04022128-10 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2023, 149(2): 04022128


Downloaded from ascelibrary.org by Purdue University Libraries on 11/27/22. Copyright ASCE. For personal use only; all rights reserved.

Fig. 15. Cumulative horizontal displacement contours (in mm) for model square footing embedded in dense OGF sand (D ¼ B) corresponding to
loading stages 1–4.

we see that the maximum shear strains develop at the edges and footing increases from 0 to 1. Referring to the fourth loading
localize along the sides of the model square footing in stage 1; stage for the model square footing placed on the surface (Fig. 16)
for the subsequent loading stages, the maximum shear strain con- and embedded at 1B (Fig. 17) in dense OGF sand, note that,
tours expand further into the sand domain (both vertically and for comparable levels of maximum shear strain, a larger vol-
laterally). A maximum shear strain of about 4% is sufficient for ume of sand is sheared below and around the embedded footing
the model square footing to reach stage 1, regardless of the em- than the surface footing. This finding implies a higher level of
bedment depth. For the second and third loading stages, the post- energy dissipation, which, in turn, is the source of the greater
processing of the images shows that the maximum shear strain unit load required to attain stage 4 for the model square footing
increases from 44% and 70%, respectively, for the surface foot- embedded at 1B in dense OGF sand over that placed on the sand
ing, to 65% and 85%, respectively, for the embedded footing due surface.
to greater confinement and unit base resistance of the footing
at these loading stages [Fig. 4(a)]. However, irrespective of the
embedment depth, the model square footing tested both on the Conclusions
surface and at D=B ¼ 1 in dense OGF sand exhibits similar
levels of maximum shear strain (>90%) at stage 4 of footing Displacement-controlled penetration tests were performed using
loading. The maximum shear strain contours show that, for all half-square and half-strip model footings in a half-cylindrical
four loading stages, the effective zone of sand that is influenced chamber with imaging capabilities. The model footings were
by the model square footing increases from 1.5B to 2.0B below placed at different embedment depths in dense and medium-dense
the footing base as the relative embedment depth D=B of the Ohio Gold Frac (OGF) sand. For the embedded model footings,

© ASCE 04022128-11 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2023, 149(2): 04022128


Downloaded from ascelibrary.org by Purdue University Libraries on 11/27/22. Copyright ASCE. For personal use only; all rights reserved.

Fig. 16. Cumulative maximum shear strain contours (in percent) for model square footing on dense OGF sand (D ¼ 0) corresponding to loading
stages 1–4.

the sand samples were prepared in two stages using the pluviation square footing is in reasonable agreement with that obtained using
method. Digital images of the sand domain were simultaneously the equation developed by Loukidis and Salgado (2011). The
captured during footing penetration to obtain the corresponding model footing test data and the methodology for estimation of
displacement and strain fields in the sand domain. shape and depth factors presented in this paper serve as a useful
To quantify the degree of improvement in bearing capacity due reference and benchmark for validation of numerical simulations
to footing embedment, the ratio of the bearing capacity of an em- of footing penetration in sand.
bedded footing (D > 0) to that of a surface footing (D ¼ 0) was Digital image correlation (DIC) analysis of the digital images
considered. The results obtained from the model footing load tests collected during loading of the model square footing in dense
in dense OGF sand show that the bearing capacity ratio, for the OGF sand show that, for large relative settlement (w=B ¼ 26%),
model square footing, increases with the relative embedment depth the vertical displacement contours extend down by 1.5B–2.0B
D=B of the footing and reaches a value of 3 for D=B ¼ 1. For the below the footing base and laterally by about 2.5B to more
model strip footing, the bearing capacity ratio is approximately than 2.5B from the footing centerline as D=B increases from 0
equal to 2 for D=B ¼ 1. to 1. The effect of footing embedment on the magnitude of the
Based on the results obtained in this study, a procedure was pro- horizontal displacement of the sand particles was small for the dif-
posed to determine experimentally the shape and depth factors for ferent footing loading stages considered in this study. For stage 4
model strip and square footings embedded in OGF sand. The trends of model footing loading, the variations of the vertical and lateral
of the values of sq and dq with D=B obtained in this study are con- extent of the horizontal displacement contours with D=B were
sistent with those reported by Lyamin et al. (2007) based on finite similar to those reported for the vertical displacement contours.
element limit analysis. Contrary to the assumption that shape fac- Regardless of embedment depth, the maximum shear strain in
tors are independent of footing depth, the shape factor sq for the the sand domain was greater than 90% for stage 4, at which point
model square footing was found to depend on the relative embed- the sand could no longer support the load transferred by the model
ment depth D=B of the footing. The shape factor sγ for the model square footing.

© ASCE 04022128-12 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2023, 149(2): 04022128


Downloaded from ascelibrary.org by Purdue University Libraries on 11/27/22. Copyright ASCE. For personal use only; all rights reserved.

Fig. 17. Cumulative maximum shear strain contours (in percent) for model square footing embedded in dense OGF sand (D ¼ B) corresponding to
loading stages 1–4.

Data Availability Statement ASTM. 2016a. Standard test methods for maximum index density and unit
weight of soils using a vibratory table. ASTM D4253. West
Some or all data, models, or code that support the findings of this Conshohocken, PA: ASTM.
study are available from the corresponding author upon reasonable ASTM. 2016b. Standard test methods for minimum index density and
request. unit weight of soils and calculation of relative density. ASTM D4254.
West Conshohocken, PA: ASTM.
ASTM. 2020. Standard practice for classification of soils for engineering
purposes (Unified Soil Classification System). ASTM D2487. West
References
Conshohocken, PA: ASTM.
Antão, A. N., M. V. da Silva, N. Guerra, and R. Delgado. 2012. “An upper Bolton, M. D. 1986. “The strength and dilatancy of sands.” Géotechnique
bound-based solution for the shape factors of bearing capacity of 36 (1): 65–78. https://doi.org/10.1680/geot.1986.36.1.65.
footings under drained conditions using a parallelized mixed f.e. formu- Bolton, M. D., and C. K. Lau. 1989. “Scale effects in the bearing
lation with quadratic velocity fields.” Comput. Geotech. 41 (Apr): capacity of granular soils.” In Vol. 2 of Proc., 12th Int. Conf. on Soil
23–35. https://doi.org/10.1016/j.compgeo.2011.11.003. Mechanics and Foundation Engineering, 895–898. Rotterdam,
Arshad, M. I. 2014. “Experimental study of the displacements caused by Netherlands: A. A. Balkema.
cone penetration in sand.” Ph.D. dissertation, Lyles School of Civil Briaud, J.-L. 2007. “Spread footings in sand: Load settlement curve
Engineering, Purdue Univ. approach.” J. Geotech. Geoenviron. Eng. 133 (8): 905–920. https://doi
Arshad, M. I., F. S. Tehrani, M. Prezzi, and R. Salgado. 2014. “Experimental .org/10.1061/(ASCE)1090-0241(2007)133:8(905).
study of cone penetration in silica sand using digital image correlation.” Brinch Hansen, J. 1970. “A revised and extended formula for bearing
Géotechnique 64 (7): 551–569. https://doi.org/10.1680/geot.13.P.179. capacity.” Danish Geotech. Inst. Bull. 28: 5–11.

© ASCE 04022128-13 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2023, 149(2): 04022128


Cerato, A. B., and A. J. Lutenegger. 2007. “Scale effects of shallow foun- Lutenegger, A., and M. Adams. 1998. “Bearing capacity of footings on
dation bearing capacity on granular material.” J. Geotech. Geoenviron. compacted sand.” In Proc., 4th Int. Conf. on Case Histories in Geotech-
Eng. 133 (10): 1192–1202. https://doi.org/10.1061/(ASCE)1090-0241 nical Engineering, 1216–1224. Rolla, MO: Univ. of Missouri-Rolla.
(2007)133:10(1192). Lyamin, A. V., R. Salgado, S. W. Sloan, and M. Prezzi. 2007. “Two- and
Chai, J., and B. Kutter. 2000. “Shallow foundations.” In Bridge engineering three-dimensional bearing capacity of footings in sand.” Géotechnique
handbook, edited by W. F. Chen and L. Duan. Boca Raton, FL: CRC 57 (8): 647–662. https://doi.org/10.1680/geot.2007.57.8.647.
Press. Martin, C. M. 2005. “Exact bearing capacity calculations using the method
Correlated Solutions. 2009. VIC-2D reference manual. Columbia, SC: of characteristics.” In Vol. 4 of Proc., 11th Int. Conf. on Computer
Correlated Solutions. Methods and Advances in Geomechanics, Turin, 441–450. Bologna,
Davis, E. H. 1968. “Theories of plasticity and the failure of soil masses.” Italy: Patron Editore.
In Soil mechanics: Selected topics, edited by I. K. Lee, 341–380. Meyerhof, G. G. 1951. “The ultimate bearing capacity of foundations.”
Butterworth, London. Géotechnique 2 (4): 301–332. https://doi.org/10.1680/geot.1951.2.4
De Beer, E. E. 1965. “Bearing capacity and settlement of shallow founda- .301.
tions.” In Proc., Symp. on Bearing Capacity and Settlement of Foun- Meyerhof, G. G. 1963. “Some recent research on the bearing capacity of
Downloaded from ascelibrary.org by Purdue University Libraries on 11/27/22. Copyright ASCE. For personal use only; all rights reserved.

dations, 15–34. Durham, NC: Duke Univ. foundations.” Can. Geotech. J. 1 (1): 16–26. https://doi.org/10.1139/t63
-003.
Doreau-Malioche, J., A. Galvis-Castro, R. Tovar-Valencia, G. Viggiani,
Okamura, M., A. Mihara, J. Takemura, and J. Kuwano. 2002. “Effects of
G. Combe, M. Prezzi, and R. Salgado. 2019. “Characterising processes
footing size and aspect ratio on the bearing capacity of sand subjected to
at sand-pile interface using digital image analysis and X-ray CT.”
eccentric loading.” Soils Found. 42 (4): 43–56. https://doi.org/10.3208
Géotech. Lett. 9 (4): 254–262. https://doi.org/10.1680/jgele.18.00232.
/sandf.42.4_43.
Frydman, S., and H. J. Burd. 1997. “Numerical studies of bearing capacity
Perkins, S. W., and C. R. Madson. 2000. “Bearing capacity of shallow foun-
factor Nγ.” J. Geotech. Geoenviron. Eng. 123 (1): 20–29. https://doi.org dations on sand: A relative density approach.” J. Geotech. Geoenviron.
/10.1061/(ASCE)1090-0241(1997)123:1(20). Eng. 126 (6): 521–530. https://doi.org/10.1061/(ASCE)1090-0241
Galvis-Castro, A. C., R. D. Tovar-Valencia, R. Salgado, and M. Prezzi. (2000)126:6(521).
2019a. “Compressive and tensile shaft resistance of nondisplacement Reissner, H. 1924. “Zum erddruckproblem.” In Proc., 1st Int. Congress of
piles in sand.” J. Geotech. Geoenviron. Eng. 145 (9): 04019041. https:// Applied Mechanics, 295–311. Delft, Netherlands: Waltman.
doi.org/10.1061/(ASCE)GT.1943-5606.0002071. Sakleshpur, V. A., M. Prezzi, R. Salgado, and M. Zaheer. 2021a. CPT-
Galvis-Castro, A. C., R. D. Tovar-Valencia, R. Salgado, and M. Prezzi. based geotechnical design manual, Volume 2: CPT-based design of
2019b. “Effect of loading direction on the shaft resistance of jacked foundations (methods). Joint Transportation Research Program Publica-
piles in dense sand.” Géotechnique 69 (1): 16–28. https://doi.org/10 tion No. FHWA/IN/JTRP-2021/23. West Lafayette, IN: Purdue Univ.
.1680/jgeot.17.P.046. Sakleshpur, V. A., M. Prezzi, R. Salgado, and M. Zaheer. 2021b. CPT-
Ganju, E., A. C. Galvis-Castro, F. Janabi, M. Prezzi, and R. Salgado. 2021. based geotechnical design manual, Volume 3: CPT-based design of
“Displacements, strains, and shear bands in deep and shallow penetra- foundations (example problems). Joint Transportation Research Pro-
tion processes.” J. Geotech. Geoenviron. Eng. 147 (11): 04021135. gram Publication No. FHWA/IN/JTRP-2021/24. West Lafayette, IN:
https://doi.org/10.1061/(ASCE)GT.1943-5606.0002631. Purdue Univ.
Ganju, E., F. Han, M. Prezzi, R. Salgado, and J. S. Pereira. 2020. “Quan- Salgado, R. 2013. “The mechanics of cone penetration: Contributions from
tification of displacement and particle crushing around a penetrometer experimental and theoretical studies.” In Vol. 1 of Proc., Geotechnical
tip.” Geosci. Front. 11 (2): 389–399. https://doi.org/10.1016/j.gsf.2019 and Geophysical Site Characterization 4, edited by R. Q. Coutinho and
.05.007. P. W. Mayne, 131–153. London: Taylor & Francis Group.
Han, F., E. Ganju, R. Salgado, and M. Prezzi. 2018. “Effects of interface Salgado, R. 2020. “Forks in the road: Rethinking modeling decisions
roughness, particle geometry, and gradation on the sand-steel interface that defined the teaching and practice of geotechnical engineering.”
friction angle.” J. Geotech. Geoenviron. Eng. 144 (12): 04018096. In Proc., Int. Conf. on Geotechnical Engineering Education (GEE)
https://doi.org/10.1061/(ASCE)GT.1943-5606.0001990. 2020: Technical Committee (TC306) of ISSMGE, 1–17. Athens,
Herle, I., and J. Tejchman. 1997. “Effect of grain size and pressure level on Greece: National Technical Univ. of Athens.
bearing capacity of footings on sand.” In Deformation and progressive Salgado, R. 2022. The engineering of foundations, slopes and retaining
failure in geomechanics, edited by A. Asaoka, T. Adachi, and F. Oka, structures. 2nd ed. Boca Raton, FL: CRC Press.
781–786. Oxford, UK: Pergamon. Salgado, R., P. Bandini, and A. Karim. 2000. “Shear strength and stiffness
Hettler, A., and G. Gudehus. 1988. “Influence of the foundation width on of silty sand.” J. Geotech. Geoenviron. Eng. 126 (5): 451–462. https://
the bearing capacity factor.” Soils Found. 28 (4): 81–92. https://doi.org doi.org/10.1061/(ASCE)1090-0241(2000)126:5(451).
/10.3208/sandf1972.28.4_81. Siddiquee, M. S. A., T. Tanaka, F. Tatsuoka, K. Tani, and T. Morimoto.
1999. “Numerical simulation of bearing capacity characteristics of strip
Janabi, F. H., V. A. Sakleshpur, M. Prezzi, and R. Salgado. 2022. “Strain
footing on sand.” Soils Found. 39 (4): 93–109. https://doi.org/10.3208
influence diagrams for settlement estimation of square footings on lay-
/sandf.39.4_93.
ered sand.” J. Geotech. Geoenviron. Eng. 148 (5): 04022025. https://doi
Smith, C. C. 2005. “Complete limiting stress solutions for the bearing
.org/10.1061/(ASCE)GT.1943-5606.0002770.
capacity of strip footings on a Mohr–Coulomb soil.” Géotechnique
Kimura, T., O. Kusakabe, and K. Saitoh. 1985. “Geotechnical model tests 55 (8): 607–612. https://doi.org/10.1680/geot.2005.55.8.607.
of bearing capacity problems in a centrifuge.” Géotechnique 35 (1): Sutton, M. A., J.-J. Orteu, and H. W. Schreier. 2009. Image correlation for
33–45. https://doi.org/10.1680/geot.1985.35.1.33. shape, motion and deformation measurements: Basic concepts, theory
Lee, J., M. Prezzi, and R. Salgado. 2011. “Experimental investigation of the and applications. 1st ed. New York: Springer US.
combined load response of model piles driven in sand.” Geotech. Test. Tatsuoka, F. 1987. “Discussion on ‘The strength and dilatancy of sands’ by
J. 34 (6): 1–15. https://doi.org/10.1520/GTJ103269. Bolton, 1986.” Géotechnique 37 (2): 219–226. https://doi.org/10.1680
Loukidis, D., and R. Salgado. 2009a. “Bearing capacity of strip and circu- /geot.1987.37.2.219.
lar footings in sand using finite elements.” Comput. Geotech. 36 (5): Tatsuoka, F., S. Goto, T. Tanaka, K. Tani, and Y. Kimura. 1997. “Particle
871–879. https://doi.org/10.1016/j.compgeo.2009.01.012. size effects on bearing capacity of footing on granular material.”
Loukidis, D., and R. Salgado. 2009b. “Modeling sand response using two- In Deformation and progressive failure in geomechanics, edited by
surface plasticity.” Comput. Geotech. 36 (1–2): 166–186. https://doi.org A. Asaoka, T. Adachi, and F. Oka, 133–138. Oxford, UK: Pergamon.
/10.1016/j.compgeo.2008.02.009. Tatsuoka, F., M. Okahara, T. Tanaka, K. Tani, T. Morimoto, and M. S. A.
Loukidis, D., and R. Salgado. 2011. “Effect of relative density and stress Siddiquee. 1991. “Progressive failure and particle size effect in bearing
level on the bearing capacity of footings on sand.” Géotechnique 61 (2): capacity of a footing on sand.” In Proc., Geotechnical Engineering
107–119. https://doi.org/10.1680/geot.8.P.150.3771. Congress, GSP 27, 788–802. Reston, VA: ASCE.

© ASCE 04022128-14 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2023, 149(2): 04022128


Tatsuoka, F., M. S. A. Siddiquee, and T. Tanaka. 1994. “Link among J. Geotech. Geoenviron. Eng. 144 (12): 04018092. https://doi.org/10
design, model tests, theories and sand properties in bearing capacity .1061/(ASCE)GT.1943-5606.0001984.
of footing on sand.” In Vol. 1 of Proc., 13th Int. Conf. on Soil Mechanics Toyosawa, Y., K. Itoh, N. Kikkawa, J.-J. Yang, and F. Liu. 2013. “Influence
and Foundation Engineering, 87–88. Rotterdam, Netherlands: A. A. of model footing diameter and embedded depth on particle size effect
Balkema. in centrifugal bearing capacity tests.” Soils Found. 53 (2): 349–356.
Tehrani, F. S., M. I. Arshad, M. Prezzi, and R. Salgado. 2018. “Physical https://doi.org/10.1016/j.sandf.2012.11.027.
modeling of cone penetration in layered sand.” J. Geotech. Geoenviron. Ueno, K., K. Miura, O. Kusakabe, and M. Nishimura. 2001. “Reappraisal
Eng. 144 (1): 04017101. https://doi.org/10.1061/(ASCE)GT.1943-5606 of size effect of bearing capacity from plastic solution.” J. Geotech.
.0001809. Geoenviron. Eng. 127 (3): 275–281. https://doi.org/10.1061/(ASCE)
Tehrani, F. S., F. Han, R. Salgado, M. Prezzi, R. D. Tovar, and A. G. Castro. 1090-0241(2001)127:3(275).
2016. “Effect of surface roughness on the shaft resistance of non- Ueno, K., K. Miura, and Y. Maeda. 1998. “Prediction of ultimate bearing
displacement piles embedded in sand.” Géotechnique 66 (5): 386–400. capacity of surface footings with regard to size effects.” Soils Found.
https://doi.org/10.1680/jgeot.15.P.007. 38 (3): 165–178. https://doi.org/10.3208/sandf.38.3_165.
Tejchman, J., and I. Herle. 1999. “A ‘class A’ prediction of the bearing Vesic, A. S. 1973. “Analysis of ultimate loads of shallow foundations.”
Downloaded from ascelibrary.org by Purdue University Libraries on 11/27/22. Copyright ASCE. For personal use only; all rights reserved.

capacity of plane strain footings on sand.” Soils Found. 39 (5): 47–60. J. Soil Mech. Found. Div. 99 (1): 45–73. https://doi.org/10.1061
https://doi.org/10.3208/sandf.39.5_47. /JSFEAQ.0001846.
Terzaghi, K. 1943. Theoretical soil mechanics. New York: Wiley. Yin, J.-H., Y.-J. Wang, and A. P. S. Selvadurai. 2001. “Influence of
Tovar-Valencia, R. D., A. Galvis-Castro, R. Salgado, and M. Prezzi. 2018a. nonassociativity on the bearing capacity of a strip footing.” J. Geotech.
“Effect of surface roughness on the shaft resistance of displacement Geoenviron. Eng. 127 (11): 985–989. https://doi.org/10.1061/(ASCE)
model piles in sand.” J. Geotech. Geoenviron. Eng. 144 (3): 04017120. 1090-0241(2001)127:11(985).
https://doi.org/10.1061/(ASCE)GT.1943-5606.0001828. Zhu, F., J. I. Clark, and R. Phillips. 2001. “Scale effect of strip and cir-
Tovar-Valencia, R. D., A. Galvis-Castro, R. Salgado, and M. Prezzi. 2021. cular footings resting on dense sand.” J. Geotech. Geoenviron. Eng.
“Effect of base geometry on the resistance of model piles in sand.” 127 (7): 613–621. https://doi.org/10.1061/(ASCE)1090-0241(2001)
J. Geotech. Geoenviron. Eng. 147 (3): 04020180. https://doi.org/10 127:7(613).
.1061/(ASCE)GT.1943-5606.0002472. Zhu, M., and R. Michalowski. 2005. “Shape factors for limit loads on
Tovar-Valencia, R. D., A. C. Galvis-Castro, M. Prezzi, and R. Salgado. square and rectangular footings.” J. Geotech. Geoenviron. Eng. 131 (2):
2018b. “Short-term setup of jacked piles in a calibration chamber.” 223–231. https://doi.org/10.1061/(ASCE)1090-0241(2005)131:2(223).

© ASCE 04022128-15 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., 2023, 149(2): 04022128

You might also like