Dot 60924 DS1

Download as pdf or txt
Download as pdf or txt
You are on page 1of 97

JOINT TRANSPORTATION

RESEARCH PROGRAM
INDIANA DEPARTMENT OF TRANSPORTATION
AND PURDUE UNIVERSITY

CPT-Based Geotechnical Design Manual,


Volume 2: CPT-Based Design
of Foundations—Methods

Venkata A. Sakleshpur, Monica Prezzi,


Rodrigo Salgado, Mir Zaheer
SPR-4108 • Report Number: FHWA/IN/JTRP-2021/23 • DOI: 10.5703/1288284317347
RECOMMENDED CITATION
Sakleshpur, V. A., Prezzi, M., Salgado, R., & Zaheer, M. (2021). CPT-based geotechnical design manual, Volume 2: CPT-
based design of foundations—Methods (Joint Transportation Research Program Publication No. FHWA/IN/JTRP-
2021/23). West Lafayette, IN: Purdue University. https://doi.org/10.5703/1288284317347

AUTHORS
Venkata A. Sakleshpur Rodrigo Salgado, PhD
Graduate Research Assistant Charles Pankow Professor of Civil Engineering
Lyles School of Civil Engineering Lyles School of Civil Engineering
Purdue University Purdue University

Monica Prezzi, PhD Mir Zaheer, PE


Professor of Civil Engineering Geotechnical Design Engineer
Lyles School of Civil Engineering Indiana Department of Transportation
Purdue University
(765) 494-5034
[email protected]
Corresponding Author

JOINT TRANSPORTATION RESEARCH PROGRAM


The Joint Transportation Research Program serves as a vehicle for INDOT collaboration with higher education in-
stitutions and industry in Indiana to facilitate innovation that results in continuous improvement in the planning,
design, construction, operation, management and economic efficiency of the Indiana transportation infrastructure.
https://engineering.purdue.edu/JTRP/index_html

Published reports of the Joint Transportation Research Program are available at http://docs.lib.purdue.edu/jtrp/.

NOTICE
The contents of this report reflect the views of the authors, who are responsible for the facts and the accuracy of the
data presented herein. The contents do not necessarily reflect the official views and policies of the Indiana Depart-
ment of Transportation or the Federal Highway Administration. The report does not constitute a standard, specifica-
tion or regulation.

ACKNOWLEDGEMENTS
This research was funded with the support provided by the Indiana Department of Transportation (INDOT) through
the Joint Transportation Research Program (JTRP) at Purdue University. The authors would like to thank the agency
for the support. The authors are very grateful for the support received from the project administrator, Peter Becker,
the business owner, Athar Khan, and the study advisory committee, composed of Samy Noureldin and Jose Ortiz,
throughout the duration of the project and for their valuable comments and suggestions. The authors are very grate-
ful to Barry Partridge and Darcy Bullock for their valuable support throughout the project. Special thanks are due
to Alebachew Tilahun, Jonathan Paauwe, and Nayyar Zia Siddiki for sharing the soil investigation data for sites in
Indiana, and to Kamran Ghani, Min Sang Lee, and Victoria Leffel for their comments. The authors would also like to
thank Daniel Alzamora and Derrick Dasenbrock from the Federal Highway Administration (FHWA) for their detailed
comments and suggestions.
TECHNICAL REPORT DOCUMENTATION PAGE

1. Report No. 2. Government Accession No. 3. Recipient’s Catalog No.


FHWA/IN/JTRP-2021/23
4. Title and Subtitle 5. Report Date
CPT-Based Geotechnical Design Manual, Volume 2: CPT-Based Design of June 2021
Foundations—Methods 6. Performing Organization Code

7. Author(s) 8. Performing Organization Report No.


Venkata A. Sakleshpur, Monica Prezzi, Rodrigo Salgado, and Mir Zaheer FHWA/IN/JTRP-2021/23
9. Performing Organization Name and Address 10. Work Unit No.
Joint Transportation Research Program
Hall for Discovery and Learning Research (DLR), Suite 204 11. Contract or Grant No.
207 S. Martin Jischke Drive
SPR-4108
West Lafayette, IN 47907
12. Sponsoring Agency Name and Address 13. Type of Report and Period Covered
Indiana Department of Transportation (SPR) Final Report
State Office Building 14. Sponsoring Agency Code
100 North Senate Avenue
Indianapolis, IN 46204
15. Supplementary Notes
Conducted in cooperation with the U.S. Department of Transportation, Federal Highway Administration.
16. Abstract
This manual provides guidance on how to use the cone penetration test (CPT) for site investigation and foundation design.
The manual has been organized into three volumes. Volume 1 covers the execution of CPT-based site investigations and presents
a comprehensive literature review of CPT-based soil behavior type (SBT) charts and estimation of soil variables from CPT results.
Volume 2 covers the methods and equations needed for CPT data interpretation and foundation design in different soil types, while
Volume 3 includes several example problems (based on instrumented case histories) with detailed, step-by-step calculations to
demonstrate the application of the design methods. The methods included in the manual are current, reliable, and demonstrably the
best available for Indiana geology based on extensive CPT research carried out during the past two decades. The design of shallow
and pile foundations in the manual is based on the load and resistance factor design (LRFD) framework. The manual also indicates
areas of low reliability and limited knowledge, which can be used as indicators for future research.
17. Key Words 18. Distribution Statement
cone penetration test, soil behavior type, shallow foundation, pile No restrictions. This document is available through the
foundation, load and resistance factor design National Technical Information Service, Springfield, VA
22161.
19. Security Classif. (of this report) 20. Security Classif. (of this page) 21. No. of Pages 22. Price
Unclassified Unclassified 97 including
appendices
Form DOT F 1700.7 (8-72) Reproduction of completed page authorized
EXECUTIVE SUMMARY Chapter 3 includes example problems for the estimation of limit
unit shaft resistance and ultimate unit base resistance of
displacement, non-displacement, and partial displacement piles
using CPT data obtained from three sites in Indiana. The
predicted foundation load capacities and settlements were found
Introduction to be in agreement with the measured load test data reported for
these sites.
This manual provides guidance on how to use the cone
penetration test (CPT) for site investigation and foundation
design. The manual has been organized into three volumes. Findings
Volume I covers the execution of CPT-based site investigations,
Not applicable.
a comprehensive literature review of CPT-based soil behavior type
(SBT) charts, and several correlations for the estimation of a soil
variable of interest from CPT results. The volume has been Implementation
organized into two chapters. Chapter 1 details the components of
a CPT system, types of CPT equipment, testing procedures and The CPT-Based Geotechnical Design Manual can be used to
precautions, maintenance of CPT equipment, and planning and train new employees and to facilitate interaction between INDOT
execution of a CPT-based site investigation. Chapter 2 presents engineers, industry, and consultants. Specific implementation
a compilation of correlations for the estimation of a soil variable items for each volume are listed below.
of interest from CPT data, and also presents a comprehensive
review of the chronological development of the SBT classification Volume I
systems that have advanced during the past 55 years of CPT A spreadsheet for the estimation of fundamental soil variables
history. from CPT results was developed. INDOT engineers can use the
Volume II covers the methods and equations needed for CPT spreadsheet on a routine basis to interpret CPT data, generate an
data interpretation and foundation design in different soil types. SBT profile, and obtain the depth profile of a soil property of
The volume has been organized into four chapters. Chapter 1 interest.
provides an introduction to the manual. Chapter 2 presents an
overview of Indiana geology, the typical CPT and soil profiles
found in Indiana, and the influence of these profiles on CPT-based
Volumes II and III
Spreadsheets for the estimation of optimal spacing between
site variability assessment. Chapter 3 details the methods for
estimation of limit bearing capacity and settlement of shallow CPT soundings and CPT-based design of shallow and pile
foundations from CPT data. Chapter 4 describes the methods for foundations were developed. INDOT engineers can use the
estimation of limit unit shaft resistance and ultimate unit base spreadsheets on a routine basis for the design of transportation
resistance of displacement, non-displacement, and partial dis- infrastructure projects in Indiana.
placement piles and pile groups from CPT data. The design of A relationship between cone resistance qc, corrected SPT blow
both shallow and pile foundations is based on the load and count N60, and mean particle size D50 was developed using
resistance factor design (LRFD) framework. data reported by Robertson et al. (1983) and data obtained from
Volume III contains several example problems (based on case 15 sites in Indiana. The relationship can be used to obtain an
histories) with detailed, step-by-step calculations to demonstrate estimate of qc for use in a CPT-based foundation design method
the application of the CPT-based foundation design methods when only SPT blow counts are available for a site.
covered in Volume II. The volume has been organized into three A relationship between critical-state friction angle c, mean
chapters. Chapter 1 includes example problems for the estimation particle size D50, coefficient of uniformity CU, and particle
of optimal spacing between CPT soundings performed in line and roundness R was developed using test data reported for 23 clean
distributed in two dimensions using CPT data obtained from the silica sands in the literature. In the absence of direct shear or
Sagamore Parkway Bridge construction site in Lafayette, Indiana. triaxial compression test results, the relationship can be used to
Chapter 2 contains example problems for the estimation of limit obtain an estimate of c for poorly-graded, clean silica sands with
bearing capacity and settlement of shallow foundations using CPT D50, CU, and R values ranging from 0.15–2.68 mm (0.006–0.105
data reported in literature for sites in the US, UK, and Australia. in.), 1.2–3.1, and 0.3–0.8, respectively.

0
CONTENTS

1. INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Aim of the Manual. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Organization of the Manual . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2. CONSIDERATION OF INDIANA GEOLOGY ON CPT-BASED SITE INVESTIGATIONS . . . . . . . 4
2.1 Overview of Indiana Geology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2 CPT, SPT, and Soil Profiles in Indiana . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.3 Correlation Between CPT Cone Resistance and SPT Blow Count. . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.4 CPT-Based Site Variability Assessment. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.5 Optimal Spacing Between CPT Soundings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.6 Chapter Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3. CPT-BASED DESIGN OF SHALLOW FOUNDATIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.1 Calculation Procedure for Footing Settlement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.2 Calculation Procedure for Limit Bearing Capacity of Footings . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.3 Load and Resistance Factor Design Procedure for Footings. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.4 Chapter Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4. CPT-BASED DESIGN OF PILE FOUNDATIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.1 Calculation Procedure for Limit Shaft Capacity of Single Piles. . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.2 Calculation Procedure for Ultimate Base Capacity of Single Piles . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.3 Load and Resistance Factor Design Procedure for Single Piles . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.4 Load and Resistance Factor Design Procedure for Pile Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.5 Chapter Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
APPENDICES
Appendix A. Critical-State Friction Angle of Sand . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
Appendix B. OCR and K0 of Soil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
Appendix C. Iterative Scheme for Footing Settlement in Sand . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
Appendix D. Penetration Rate Effect on Cone Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
Appendix E. Residual-State Friction Angle of Clay . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
LIST OF TABLES

Table Page

Table 2.1 Geographic information of the CPT locations in Indiana 9

Table 2.2 Soil behavior types associated with the modified Tumay (1985) chart 10

Table 2.3 Soil behavior types associated with the modified Robertson (1990) chart 10

Table 2.4 Vertical variability index, horizontal variability index, and site variability rating for the sites analyzed 23

Table 2.5 Method for estimation of optimal spacing between CPT soundings 26

Table 3.1 wmax/amax values for shallow foundations in sand and clay 27

Table 3.2 Values of a for estimation of primary consolidation settlement of footings in clay 31

Table 3.3 Values of N as a function of sample size n 33

Table 3.4 Values of C1 and C2 to use in Eq. 3.51 as a function of B/L 35

Table 3.5 Resistance factors for footings (D/B # 1) in sand and clay 36

Table 3.6 Resistance factors for footings in sand and clay 36

Table 3.7 Methods for estimation of footing settlement in sand 37

Table 3.8 Methods for estimation of bearing capacity of footings in sand 39

Table 3.9 Methods for estimation of footing settlement in clay 41

Table 3.10 Methods for estimation of bearing capacity of footings in clay 42

Table 4.1 Expressions for Asi for different pile cross-sections 46

Table 4.2 Expressions for Ab for different pile cross-sections 46

Table 4.3 PPDM resistance factors for drilled shafts and CEP piles in sand and clay 48

Table 4.4 ICPDM resistance factors for driven piles in sand and clay 49

Table 4.5 Resistance factors for drilled shafts in sand and clay 49

Table 4.6 Resistance factors for driven piles in sand and clay 49

Table 4.7 Shaft and base efficiencies for a large (464) drilled shaft group in sand for scc 5 2B 51

Table 4.8 Efficiencies for small and large drilled shaft groups in sand 51

Table 4.9 Efficiencies for small and large driven pile groups in sand 51

Table 4.10 Shaft and base efficiencies for a large (464) drilled shaft group in NC clay for scc 5 2B 52

Table 4.11 Efficiencies for small and large drilled shaft and driven pile groups in clay 52

Table 4.12 PPDM equations for the unit shaft and base resistances for nondisplacement piles
(drilled shafts) in sand and clay 54

Table 4.13 MnDOT equations (Modified UniCone method) for the unit shaft and base resistances
for nondisplacement piles (drilled shafts) in sand and clay 54

Table 4.14 PPDM equations for the unit shaft and base resistances for displacement piles driven in sand 55

Table 4.15 ICPDM equations for the unit shaft and base resistances for displacement piles driven in sand) 56

Table 4.16 UWAPDM equations for the unit shaft and base resistances for displacement piles driven in sand 57

Table 4.17 AASHTO equations for the unit shaft and base resistances for displacement piles driven in sand 57

Table 4.18 MnDOT equations (Modified UniCone method) for the unit shaft and base resistances for
displacement piles driven in sand 58

Table 4.19 UPDM equations for the unit shaft and base resistances for displacement piles driven in sand 59

Table 4.20 PPDM equations for the unit shaft and base resistances for displacement piles driven in clay 59

Table 4.21 ICPDM equations for the unit shaft and base resistances for displacement piles driven in clay 60

Table 4.22 UWAPDM equations for the unit shaft and base resistances for displacement piles driven in clay 61
Table 4.23 AASHTO equations for the unit shaft and base resistances for displacement piles driven in clay 61

Table 4.24 MnDOT equations (Modified UniCone method) for the unit shaft and base resistances for displacement piles driven in clay 62

Table 4.25 NDOT equations for the unit shaft and base resistances for displacement piles driven in clay 62
LIST OF FIGURES

Figure Page

Figure 1.1 Schematic of SPT in progress 1

Figure 1.2 Comparison of NSPT values obtained: (a) by different crews using the same SPT equipment
(adapted from Mayne & Harris, 1993), (b) using safety and auto hammers (adapted from Finno, 1989), and
(c) using safety and donut hammers (adapted from Robertson et al., 1983) 2

Figure 1.3 Overview of the cone penetration test 2

Figure 1.4 Typical CPT log 3

Figure 1.5 Results obtained from a SCPTu sounding performed at the Golden Ears Bridge site in Vancouver, Canada 3

Figure 2.1 Bedrock geologic map of Indiana 5

Figure 2.2 Physiographic divisions of southern Indiana 6

Figure 2.3 Sinkhole in Mississippian carbonate rock of Mitchell Plateau in Lawrence County, Indiana 6

Figure 2.4 Close-up view of a sinkhole near Salem Bypass in Washington County, Indiana 6

Figure 2.5 Map showing the tectonic features in Indiana 7

Figure 2.6 Surficial geologic map of Indiana 7

Figure 2.7 Map of southern Indiana showing the distribution of loess deposits (. 1.5 m (5 ft) in thickness) 8

Figure 2.8 Pedological map of Indiana showing the CPT locations 8

Figure 2.9 Modified Tumay (1985) SBT chart 9

Figure 2.10 Modified Robertson (1990) SBT chart 10

Figure 2.11 In situ test profiles for location A in Lake County: (a) CPT-1 profile (qc, fs, FR) and SBT interpreted from modified
Tumay (1985) chart, and (b) N60 profile, (qc/pA)/N60 profile, and soil profile from SPT boring TB-2 12

Figure 2.12 In situ test profiles for location D in Newton County: (a) CPT-2 profile (qc, fs, FR) and SBT interpreted from modified
Tumay (1985) chart, and (b) N60 profile, (qc/pA)/N60 profile, and soil profile from SPT boring TB-1 13

Figure 2.13 CPT-2 profile (qc, fs, FR) and SBT interpreted from modified Tumay (1985) chart for location C in LaPorte County 14

Figure 2.14 In situ test profiles for location E in Tippecanoe County: (a) CPT-3 profile (qc, fs, FR) and SBT interpreted from modified
Tumay (1985) chart, and (b) N60 profile, (qc/pA)/N60 profile and soil profile from SPT boring Pier-7 15

Figure 2.15 Comparison of SBT profiles obtained from sounding CPT-5 at location E in Tippecanoe County using:
(a) modified Tumay (1985) chart (zone numbers listed in Table 2.2), and (b) modified Robertson (1990) chart
(zone numbers listed in Table 2.3) 16

Figure 2.16 CPT-4 profile (qc, fs, FR) and SBT interpreted from modified Tumay (1985) chart for location B in Steuben County 16

Figure 2.17 In situ test profiles for location F in Clinton County: (a) CPT-7 profile (qc, fs, FR) and SBT interpreted from modified
Tumay (1985) chart, and (b) N60 profile, (qc/pA)/N60 profile, and soil profile from boring SPT-8 17

Figure 2.18 In situ test profiles for location G in Madison County: (a) CPT RB-2 profile (qc, fs, FR) and SBT interpreted from
modified Tumay (1985) chart, and (b) N60 profile, (qc/pA)/N60 profile and soil profile from SPT boring TB-2 18

Figure 2.19 In situ test profiles for location H in Decatur County: (a) CPT-1 profile (qc, fs, FR) and SBT interpreted from modified Tumay
(1985) chart, and (b) N60 profile, (qc/pA)/N60 profile and soil profile from SPT boring TB-1 19

Figure 2.20 In situ test profiles for location I in Knox County: (a) CPT-1 profile (qc, fs, FR) and SBT interpreted from modified Tumay (1985)
chart, and (b) N60 profile, (qc/pA)/N60 profile, and soil profile from SPT boring TB-2 20

Figure 2.21 CPT-28 profile (qc, fs, FR) and SBT interpreted from modified Tumay (1985) chart for location J in Vanderburgh County 21

Figure 2.22 Correlation between CPT cone resistance and SPT blow count 21

Figure 2.23 Site variability rating chart 22

Figure 2.24 Site variability ratings for the sites analyzed 24

Figure 2.25 Optimal spacing between CPT soundings performed in line 24

Figure 3.1 Strain influence factor Iz versus depth zf below the footing base 28
Figure 3.2 Influence factor Iq as a function of qb,net =su and H/B for (a) strip footings, (b) square footings, and
(c) rectangular (L/B 5 2) footings 31

Figure 3.3 Examples of two CPT logs in clay and three CPT logs in sand with mean trendlines and range lines 33

Figure 3.4 F versus rB/su0 for a rough footing base in clay 35

Figure 4.1 CPT-based discretization of soil profile for shaft resistance calculation and averaging of
cone resistance for base resistance calculation 44

Figure 4.2 Critical-state friction angle ratio c/c versus mean particle size D50 for silica sands tested against smooth,
lightly rusted, and rusted steel surfaces (Han et al., 2018, 2019a). Interpolation can be used for 1.5 , CU , 2 44

Figure 4.3 Layout of (a) small (163) pile group and (b) large (464) pile group 50

Figure 4.4 Schematic of a 364 pile group with parameters Lg, Bg, and L in (a) plan view and (b) 3D view 52

Figure 4.5 Dutch technique for estimation of qcb 58


1. INTRODUCTION
1.1 Background

Site investigation is an important component of


every infrastructure project and plays a vital role in
project planning, design, and construction. It is akin to
diagnosing patients in medicine because a project site’s
pathology (i.e., the origin, type, spatial distribution,
and properties of soil and rock layers) is evaluated for
engineering purposes (Madhav & Abhishek, 2016,
2017). The main goals of a geotechnical site investiga-
tion are to: (1) identify soil and rock stratigraphy, (2)
establish groundwater level conditions, and (3) estimate
geotechnical design parameters (e.g., strength and
stiffness). Although site investigations involve both soil
and rock characterization, this manual focuses solely on
soil investigations performed using the cone penetration
test (CPT).
Over the past two to three decades, in situ tests have
gained favor over laboratory tests because: (1) in situ
tests are generally faster to perform than laboratory
tests, and (2) laboratory test results are affected by
sample disturbance and represent the properties of only
a few points within a stratum. In contrast, in situ tests,
particularly the CPT, significantly increase the volume
of material investigated at a site and produce more
reliable and repeatable data, thus resulting in substan-
tial cost and time savings.
Among available in situ tests, the standard penetra-
tion test (SPT) and the cone penetration test (CPT) are
the most commonly used tests in practice. The SPT is a Figure 1.1 Schematic of SPT in progress (Salgado, 2008;
crude test that involves driving a standard split-spoon USACE, 2001).
sampler into the ground a distance of 450 mm (18 in.)
from multiple blows using a 630 N (140 lb) hammer penetrometer having a conical tip with 60u apex angle
dropped from a height of 760 mm (30 in.) (Figure 1.1). vertically into the ground at a standard rate of 20 mm/s
The number of blows required for the last 300 mm (12 (0.8 in./s) (ASTM, 2012) (Figure 1.3). The penetrom-
in.) of penetration of the sampler, after an initial seating eter is connected to the lowest rod among a string of
drive of 150 mm (6 in.), is recorded as the raw SPT blow rods pushed down from a truck-mounted, crawler-
count NSPT for the tested depth. mounted, or trailer-mounted rig. The cone penetrom-
The SPT blow count is affected by energy inefficien- eter was originally used to measure only the tip or cone
cies in the drop hammer system and other factors, such resistance qc, defined as the vertical force acting on the
as the effects of the operator, rod length, sampler type, tip of the penetrometer divided by the base area of the
and borehole diameter (Ireland et al., 1970). Although tip. The base area of the cone tip is equal to 1,000 mm2
corrections have been proposed to normalize the NSPT (1.55 in.2) for typical penetrometers that are in
value with respect to these factors (Anderson et al., compliance with ASTM (2012), although penetrometer
2004; Kulhawy & Mayne, 1990; Skempton, 1986), the sizes in practice can vary greatly.
reliability of the SPT remains quite low as test results Over the years, different sensors have been incorpo-
are likely to vary between different crews operating rated into the cone to measure sleeve resistance fs, shear
the same equipment (Look, 2016; Look et al., 2015) wave velocity Vs, pore water pressure u, and other
(Figure 1.2). Consequently, the CPT is gradually parameters (Campanella & Weemees, 1990; Mayne &
replacing the SPT as the preferred in situ test for site Campanella, 2005; Mitchell, 1988; Robertson et al.,
investigations. The greater availability of powerful CPT 1986). The CPT data is generally recorded at 1-to-5-cm
rigs has made it easier for engineers to require that (0.4-to-2-in.) intervals of cone penetration (ASTM,
CPTs be performed as part of site investigations. 2012); however, the data can also be recorded at every
Another reason for the increasing reliance on the 0.2 cm (0.08 in.) of cone penetration depending on the
CPT is the development of sophisticated and reliable level of sophistication of the penetrometer and the data
foundation design methods based on CPT data. acquisition system (Salgado et al., 2015). The data is
The CPT is a quasi-static test and is often used as a directly logged to a field computer in real-time and can be
complement to conventional rotary drilling and sam- used to estimate geostratigraphy, soil types, water table
pling methods. The test is performed by pushing a elevation, and geotechnical design parameters of interest.

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 1


Figure 1.2 Comparison of NSPT values obtained: (a) by different crews using the same SPT equipment (adapted from Mayne &
Harris, 1993), (b) using safety and auto hammers (adapted from Finno, 1989), and (c) using safety and donut hammers (adapted
from Robertson et al., 1983).

Figure 1.3 Overview of the cone penetration test (after ASTM, 2012).

Figure 1.4 shows a typical CPT log, which always resistance was originally thought of as being useful
contains the cone resistance qc and sleeve resistance fs for estimating pile shaft resistance; however, by means
plotted as a function of depth; it may contain more of the friction ratio fs/qc, it has more often been used
information if additional measurements are made. as an indicator of the type of soil through which the
Sleeve friction or sleeve resistance fs is defined as the cone is advanced (Lunne et al., 1997). In general, a
ratio of the shear force acting along the surface of the combination of low qc values and high friction ratio fs/
cylindrical friction sleeve located above the cone tip to qc suggests a clayey soil, whereas for sandy soils, qc
the circumferential area of the sleeve. The circumfer- tends to be high and fs/qc low (Salgado, 2008). Volume I
ential area of the sleeve is equal to 15,000 mm2 (23.25 reviews the charts available in the literature for
in.2) in the standard cone (ASTM, 2012). Sleeve estimating soil behavior type (SBT) from CPT results.

2 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


The seismic piezocone penetration test (SCPTu), a test data obtained at 1-to-5-cm (0.4-to-2-in.) depth
newer version of the CPT, is a hybrid geotechnical- intervals (Campanella et al., 1986; Mayne, 2007; Mayne
geophysical in situ test that provides downhole geophy- & Campanella, 2005; Robertson et al., 1986). Figure 1.5
sical measurements of shear wave velocity Vs at 1-m- shows the results obtained from a SCPTu sounding
depth intervals in addition to the regular penetration performed up to a depth of 95 m (312 ft) at the Golden
Ears Bridge site in Vancouver, Canada. Such high-
quality subsurface data can be efficiently used to
delineate the geostratigraphy of a site and obtain the
required geotechnical parameters for use in foundation
design.
In its simplest application, the CPT offers a quick,
expedient, and economical way to characterize the
ground conditions at a site. According to Mayne
(2007), a 10-m (30-ft)-deep CPT sounding can be
completed in about 15–20 minutes, whereas a conven-
tional soil boring takes about 3–6 times longer to
complete. Since soil samples are not collected and spoils
are not generated during testing, the CPT is less
disruptive from an environmental standpoint and thus
advantageous when investigating environmentally sen-
sitive areas and potentially contaminated sites where
the risk of exposure to hazardous material is high
(Campanella & Weemees, 1990; Fukue et al., 2001;
McKnight et al., 2015; Mondelli et al., 2010; Walker
et al., 2009). The CPT can be performed in most soil
types, ranging from soft-to-stiff clays and loose-to-
dense sands, and silts, but can be difficult to perform in
terrain containing gravels, cobbles, boulders, or other
such obstacles to penetration (Han et al., 2019a,b).
Nonetheless, the almost continuous CPT data permit
clear delineations of soil strata including the thickness
and lateral extent of each layer. In addition, the
penetration process is amenable to theoretical model-
ing, even if the level of sophistication of the required
analyses is such that it remains a topic of advanced
research. The penetration resistance can be either
Figure 1.4 Typical CPT log (Salgado, 2008). correlated with other geotechnical parameters or used

Figure 1.5 Results obtained from a SCPTu sounding performed at the Golden Ears Bridge site in Vancouver, Canada (adapted
from Niazi et al., 2010).

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 3


directly in design; however, its use in design and ultimate unit base resistance of displacement, nondis-
interpretation remains a research need. placement, and partial displacement piles and pile
Soil properties used in geotechnical design are often groups from CPT data.
estimated from a limited number of in situ or laboratory Volume III contains several example problems
tests (due to project budget and time constraints) and (based on instrumented case histories) with detailed,
are thus subject to uncertainty, raising the question as step-by-step calculations to demonstrate the application
to how accurately the soil properties derived from these of some CPT-based foundation design methods covered
tests represent those of the entire site (Madhira & in Volume II.
Sakleshpur, 2018, 2019). Although this uncertainty
cannot be eliminated, it can be addressed by quantify- 2. CONSIDERATION OF INDIANA GEOLOGY ON
ing the variability within individual soundings and of CPT-BASED SITE INVESTIGATIONS
clusters of soundings at a site. Because the CPT is a
more reliable tool than the SPT, it can be used for both 2.1 Overview of Indiana Geology
site variability assessment (Salgado et al., 2015, 2019)
and load and resistance factor design (LRFD) of 2.1.1 Bedrock Geology
foundations (Basu & Salgado, 2012; Han et al., 2015).
Indiana’s bedrock geology has three important
aspects—the first being the topography of the bedrock
1.2 Aim of the Manual
surface. The bedrock of Indiana has undergone erosion
There is a myriad of CPT correlations and CPT- since about 300 million years ago, but it was only
based design protocols in the literature; these correla- during the Ice Age that unconsolidated sediments were
tions and protocols appear in software, producing deposited over the bedrock due to glacial advances and
interpretation results that may be confounding. This retreats across the state. The Ice Age, also known as
leads to confusion among consultants as to which Pleistocene, is a geologic time period that began about
method(s) to use for estimation of soil variables and two million years ago and ended 10,000 years ago;
design of geotechnical structures based on CPT results. during this period, the Earth’s higher and mid-latitude
This manual does not aim to be an exhaustive review of zones experienced extensive glaciation by large, con-
all that can be done with the CPT or of all the possible tinental-scale ice sheets (Wilson, 2008). Thus, the
ways in which CPT results can be used in geotechnical bedrock surface is usually not visible in Indiana because
engineering. The purpose of this manual, written in nearly two-thirds of the state is covered by glacial
concise, objective language, is to provide guidance on material. According to the Indiana Geological and
how to use the CPT specifically for site investigation Water Survey (IGWS), Indiana’s bedrock is exposed
and foundation design. The primary focus of the only in the south-central part of the state, which is
manual is on methods that are current, reliable, and unglaciated, and in localized areas along the Wabash
demonstrably the best available for Indiana geology River—the highest points of the bedrock surface are in
based on extensive CPT research carried out during the Randolph and Wayne counties, while the lowest points
past two decades. The manual also indicates areas of are along the Wabash and Ohio Rivers in Posey and
low reliability and limited knowledge, which can be Vanderburgh counties.
used as indicators for future research. The types of rocks and their spatial distribution form
the second aspect of Indiana’s bedrock geology. Fig-
1.3 Organization of the Manual ure 2.1 shows the bedrock geologic map of Indiana,
which consists of five bedrock units: Pennsylvanian,
The manual has been organized into three volumes. Mississippian, Devonian, Silurian, and Ordovician
Volume I contains two chapters—Chapter 1 details the units. Each unit or formation is tens to hundreds of
components of a CPT system, types of CPT equipment, feet thick and consists primarily of sedimentary rocks,
testing procedures and precautions, maintenance of such as limestone, dolomite, shale, sandstone, and
CPT equipment, and planning and execution of a CPT- siltstone. Each of these sedimentary rocks weathers at a
based site investigation. Chapter 2 presents a compre- different rate and produces unique weathering bypro-
hensive literature review of (a) estimation of soil ducts. For instance, carbonaceous rocks, such as lime-
variables from CPT results and (b) soil behavior type stone and dolomite, dissolve slowly in acid rain and
(SBT) charts. snow to produce sinkholes, caves, and other features
Volume II contains four chapters—Chapter 1 collectively known as karst (West, 2010; White, 1988).
provides an introduction to the manual. Chapter 2 Such soluble rocks having karst or the potential to
presents an overview of Indiana geology, the typical develop karst features account for about 18% of the
CPT and soil profiles found in Indiana, and the land area of the United States (Weary & Doctor, 2014).
influence of these profiles on CPT-based site variability Figure 2.2 shows the karst regions in southern
assessment. Chapter 3 details the methods for estima- Indiana, which include the Mitchell and Muscatatuck
tion of limit bearing capacity and settlement of shallow Plateaus, the Crawford and Norman Uplands, and
foundations from CPT data. Chapter 4 describes the the Charlestown Hills area. The Mitchell Plateau in
methods for estimation of limit unit shaft resistance and south-central Indiana is a karst plateau developed on

4 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


Figure 2.1 Bedrock geologic map of Indiana (Source: IGWS, n.d.).

Mississippian carbonates and extends from the eastern Figure 2.5 shows the tectonic features of Indiana.
part of Owen County down south to the Ohio River in The Kankakee Arch and the Cincinnati Arch constitute
Harrison County and then into Kentucky (Florea et al., a broad anticline, which extends from the northwestern
2018; Gray, 2000; Malott, 1922). The Crawford Upland to the southeastern part of the state (Rupp, 1991).
lies to the west of the Mitchell Plateau and is This anticline is intersected by two faults: the Royal
characterized by ridges and valleys developed on shale, Center Fault and the Fortville Fault. Apart from these
sandstone, and carbonate strata of Mississippian age two faults, there is the Mt. Carmel Fault (in the
(Florea et al., 2018). Karst features have also been Leesville anticline) that extends from Morgan County
detected along the western margin of the Norman south through Monroe and Lawrence counties into
Upland to the east of the Mitchell Plateau as well as in Washington County, and finally, a concentrated region
carbonate strata of Silurian and Devonian age in the of faults in the southwestern part of the state called the
Muscatatuck Plateau and the Charlestown Hills area in Wabash Fault Valley System (Ault & Sullivan, 1982;
southeastern Indiana (Gray, 2000) (Figure 2.2). Karst Hildenbrand & Ravat, 1997; René & Stanonis, 1995;
presents difficulties and challenges to geotechnical Woolery et al., 2018). In general, Indiana is tectonically
engineers due to the presence of underground cavities quiet with practically insignificant movement of the
that may collapse, forming sinkholes. Figure 2.3 and bedrock (Rupp, 1991).
Figure 2.4 show photographs of sinkholes in Lawrence
County and near the Salem Bypass in Washington 2.1.2 Surficial Geology
County, respectively, in Indiana.
The third aspect of Indiana’s bedrock geology is the Figure 2.6 shows the surficial geologic map of
presence of bends and faults in the stratigraphic units. Indiana, which can be broadly divided into four regions

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 5


Figure 2.2 Physiographic divisions of southern Indiana (adapted from Gray, 2000).

Figure 2.3 Sinkhole in Mississippian carbonate rock of


Mitchell Plateau in Lawrence County, Indiana (Frushour,
Figure 2.4 Close-up view of a sinkhole near Salem Bypass in
n.d., as cited in Hasenmueller & Packman, n.d.).
Washington County, Indiana (T. Colglazier & N. Z. Siddiki,
personal communication, November 14, 2017).
(from north to south) based on the type of deposit
encountered. Firstly, large deposits of dune sand, et al., 2018; Wayne & Thornbury, 1951). These glacial
or sand dunes, exist in northern Indiana, particularly till plains are partly bisected by end moraines, which are
along the Lake Michigan shoreline and along the eastern long, arcuate ridges of till, in northeastern Indiana
margins of the Wabash and White Rivers (Argyilan (Brown, 2016; Kassab et al., 2017; Wayne, 1965).
et al., 2018; Cressey, 1928; Hill, 1974; Kilibarda & Finally, thick loess deposits, which contribute to soil
Blockland, 2011; Kilibarda & Shillinglaw, 2014). fertility, lie east of the Wabash and White Rivers and
Secondly, outwash, which is a sorted and stratified south of the Wisconsin glacial boundary, as shown in
mixture of sand and gravel particles transported and Figure 2.7 (Hall & Anderson, 2000; Kim & Kang, 2013;
deposited by glacial meltwater, exists in northern Shaw, 1915).
Indiana and along major river valleys, such as the Eel, Loess is an unstratified, aeolian sediment that
Kankakee, Whitewater, Wabash, White, and Ohio consists mostly of silt with small fractions of clay
Rivers (Logan et al., 1922). Thirdly, glacial till, which (smectite) and fine sand (quartz/feldspar) along with
is an unsorted, unstratified and heterogeneous mixture light carbonate cementation (calcite # 30%) at inter-
of clay-to-boulder size particles deposited by ice, forms particle contacts (Mitchell & Soga, 2005). Loess deposits
flat to hummocky plains in central Indiana (Colgan are typically characterized by low water content (<
et al., 2003; Fleming et al., 1993; Gooding, 1973; Loope 10%), low density (< 1.2 g/cm3 or 74.9 lb/ft3), and loose

6 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


metastable fabric (void ratio 5 0.67–1.50)—they are
strong and incompressible when dry, as evidenced by
several stable vertical cliffs found around the world,
but are collapsible either with saturation alone or with
saturation and loading (Krinitzsky & Turnbull, 1967;
Mitchell & Soga, 2005; Rutledge et al., 1996).
In addition to the aforementioned soil types, organic
soils, such as peat (with organic content . 30%), are
commonly found in the Northern Lake Moraine
Physiographic Region in northern Indiana and occa-
sionally in central Indiana as well (Wilcox et al., 1986;
Wilcox & Simonin, 1988).

2.2 CPT, SPT, and Soil Profiles in Indiana

One of the primary applications of the cone


penetration test is stratigraphic profiling. Figure 2.8
shows the distribution of different soil types in Indiana
and 10 select locations where CPTs were performed by
the Indiana Department of Transportation (INDOT),
the United States Geological Survey (USGS), and
Purdue University. Table 2.1 summarizes the geogra-
Figure 2.5 Map showing the tectonic features in Indiana phic details of the CPT locations marked in Figure 2.8.
(Rupp, 1991). The locations were selected from different parts of the

Figure 2.6 Surficial geologic map of Indiana (after Gray, 2000).

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 7


Figure 2.7 Map of southern Indiana showing the distribution of loess deposits (. 1.5 m (5 ft) in thickness) (Source: Gray, n.d.).

Figure 2.8 Pedological map of Indiana showing the CPT locations.

8 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


TABLE 2.1
Geographic information of the CPT locations in Indiana

Notation Soil Type County Approximate Location Details Latitude Longitude

A Dune sand Lake On SR-51/US-6, 900 ft south of I-90 41.5903 -87.2403


B Till in hummocky Steuben On SR-4 over Little Turtle Creek 41.5267 -85.1036
moraine form
C Outwash LaPorte On US-30 over Turf Farm Ditch 41.4058 -86.7389
D Aeolian sand Newton On SR-55 over Gregory Ditch 41.0906 -87.3336
E Outwash Tippecanoe US-52 bridge over Wabash River, Lafayette 40.4511 -86.8929
F Glacial till Clinton 310 ft southwest of INDOT office 40.2777 -86.5342
(1675 IN-28, Frankfort)
G Glacial till Madison On SR-32 over Indian Camp Creek 40.0842 -85.8283
H Glacial till Decatur On US-421, 780 ft southeast of 39.3064 -85.4333
Lost Fork Stream
I Loess with sand Knox On SR-550 over Smalls Creek, 38.7892 -87.4383
1.57 miles west of SR-67
J Lacustrine soil Vanderburgh On W Delaware St, 2.16 miles west of US-41 37.9840 -87.5816

state to demonstrate the effect of Indiana geology on


cone penetration test results.
The raw CPT data collected from each location was
post-processed to obtain profiles of cone resistance qc,
sleeve resistance fs, and friction ratio FR (5 fs/qc). The
USGS and INDOT CPT rigs record data at 5 cm depth
intervals, while the Purdue CPT rig records data at
2 mm depth intervals (Salgado et al., 2015). The cor-
rected, total cone resistance qt was calculated by taking
into account the unbalanced pore water pressure acting
on opposing sides of both the face and joint annulus
of the cone tip (Jamiolkowski et al., 1985; Lunne et al.,
1997; Robertson et al., 1986; Salgado, 2008):

qt ~qc zð1{aÞu2 ðEq: 2:1Þ

where qc 5 measured cone resistance, u2 5 pore water


pressure measured at the shoulder position behind
the cone face, and a 5 cone area ratio (5 0.8 for the
Hogentogler CPT probe (Hogentogler & Co. Inc.,
2004)). According to ASTM D5778 (ASTM, 2012), the
correction of qc to qt is particularly important for CPTs
Figure 2.9 Modified Tumay (1985) SBT chart (Ganju et al.,
in saturated clays, silts, and soils having considerable
2017; Salgado et al., 2019).
amount of fines where substantial pore pressures are
generated during penetration; however, for CPTs in
clean sands, dense to hard geomaterials, and dry soils, modified in order to (a) minimize ambiguities asso-
the correction may be ignored without significant error. ciated with soil behavior types, and (b) make a clearer
It is assumed hereafter that this correction has been distinction between soil intrinsic variables (related
applied whenever it produces nonnegligible changes to closely to soil composition) and soil state variables,
qc, and thus qc will not be distinguished from qt, unless such as relative density, stress state, and fabric.
otherwise stated. Figure 2.9 shows the modified version of the Tumay
A soil profile generation algorithm developed by (1985) SBT chart. In general, a combination of low qc/
Ganju et al. (2017) was used to generate stratigraphic pA (, 10) and high fs/qc values (. 4%) suggests a clayey
profiles from the CPT data obtained at each location. soil, whereas a combination of high qc/pA (. 50) and
The algorithm requires seven input parameters: depth, low fs/qc values (, 2%) suggests a sandy soil; where
corrected cone resistance, sleeve resistance, ground pA 5 reference stress (5 100 kPa or 14.5 psi).
surface elevation, latitude, longitude, and groundwater Table 2.2 summarizes the soil behavior types asso-
table depth. The algorithm was implemented for the ciated with the modified Tumay (1985) chart. Each soil
soil behavior type (SBT) chart proposed originally behavior type that appears in Figure 2.9 is assigned a
by Tumay (1985) and modified subsequently by Ganju zone number. For instance, zones 1 to 7 correspond to
et al. (2017). The original Tumay (1985) chart was clays of different stiffnesses, zone 8 corresponds to

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 9


TABLE 2.2
Soil behavior types associated with the modified Tumay (1985)
chart

Zone Soil Behavior Type

1 Sensitive clay
2 Very soft clay
3 Soft clay
4 Medium stiff clay
5 Stiff clay
6 Very stiff clay
7 Sandy clay or silty clay
8 Clayey silty sand
9 Clayey sand or silt
10 Clayey silt
11 Very dense sand or silty sand
12 Dense sand or silty sand
13 Medium dense sand or silty sand
14 Loose sand or silty sand
15 Very loose sand or silty sand

sands containing fines, and zones 9 and 10 correspond


to clayey sand or silt and clayey silt, respectively. Ganju Figure 2.10 Modified Robertson (1990) SBT chart (Ganju
et al., 2017; Salgado et al., 2019).
et al. (2017) further divided the ‘‘clean sand or silty
sand’’ region of the modified Tumay (1985) chart into
five zones (zones 11 to 15 in Table 2.2) based on the
relative density, which can be estimated from CPT data TABLE 2.3
using the correlation of Salgado and Prezzi (2007). Soil behavior types associated with the modified Robertson (1990)
chart
Apart from the modified Tumay (1985) chart, a
modified version of the Robertson (1990) SBT chart, Zone Soil Behavior Type
which distinguishes clean sand from gravelly sand, was
also used to generate the SBT profile, particularly for 1 Sensitive fine-grained
location E in Tippecanoe County. Figure 2.10 shows 2 Organic clay
3 Clay to silty clay
the modified Robertson (1990) SBT chart according to
4 Clay silt to silty clay
Ganju et al. (2017). The chart uses values of normalized 5 Sand mixtures: silty sand to sandy silt
cone resistance qtn 5 (qt – v0)/v90 and normalized 6 Very dense gravelly sand to sand
friction ratio FRn (%) 5 [fs/(qt – v0)]6100%; where v0 7 Dense gravelly sand to sand
and 9v0 5 in situ vertical total and effective stresses, 8 Medium dense gravelly sand to sand
respectively, at the depth being considered. As the 9 Loose gravelly sand to sand
values of v0 and v90 depend on the unit weights of the 10 Very loose gravelly sand to sand
soil layers at the site and the elevation of the ground- 11 Very dense clean sand to silty sand
water table, the modified Robertson (1990) SBT chart 12 Dense clean sand to silty sand
13 Medium dense clean sand to silty sand
can only be used after the CPT data has been post-
14 Loose clean sand to silty sand
processed. 15 Very loose clean sand to silty sand
Table 2.3 summarizes the soil behavior types asso-
ciated with the modified Robertson (1990) chart.
Similar to the modified Tumay (1985) chart, each soil during cone penetration, on the selected SBT chart.
behavior type that appears in Figure 2.10 is assigned a Secondly, any layer in the initial soil profile with
zone number. Ganju et al., (2017) further divided the thickness less than or equal to 15 cm (5.9 in.) (or 4.2
‘‘gravelly sand to sand’’ region and the ‘‘clean sand to cone diameters) is tagged as a thin layer—a layer in
silty sand’’ region of the modified Robertson (1990) which the CPT probe is unable to develop a cone
chart into five zones each (zones 6 to 10 and 11 to 15 in resistance that is representative of that layer. Finally,
Table 2.3) based on the relative density, which can be the initial soil profile is reanalyzed with the objective of
estimated from CPT data using the correlation of merging the thin layers into the adjacent thick layers to
Salgado and Prezzi (2007). obtain the final soil profile. This is done using three
A total of 23 CPT soundings were analyzed from sequential approaches: (1) the SBT band approach,
locations A–J using the soil profile generation (2) the soil group approach, and (3) the average qc
algorithm developed by Ganju et al. (2017). In this approach, all of which are described in detail by
algorithm, firstly, an initial soil profile is generated by Salgado et al. (2015) and Ganju et al. (2017). The
plotting the qc and FR values, obtained at each depth significance of this methodology is that the final

10 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


generated soil profile will not contain layers thinner chart for both locations A and D agree qualitatively
than 15 cm (5.9 in.). This mitigates the creation of a with the soil profiles obtained from the corresponding
significantly fragmented soil profile littered with SPT boring logs. The soil profiles obtained from the
clusters of layers that are too small to be sensed boring logs are based on laboratory testing of soil
properly by the standard CPT probe. samples collected at depth intervals of 1.5 m (5 ft),
Apart from the CPT, additional independent sam- whereas the SBT profiles generated using the modified
pling may be performed to corroborate the soil profile Tumay (1985) chart are based on nearly continuous
at a site. However, soil behavior types obtained from CPT measurements at 5 cm (2 in.) depth intervals.
SBT charts may not always fully agree with traditional Thus, the SBT profiles contain more soil layers than the
soil classifications based on grain-size distribution and soil profiles obtained from the boring logs because
soil plasticity, such as the Unified Soil Classifica- some of these layers may lie between consecutive SPT
tion System (USCS) (ASTM, 2017) or the American sampling intervals. The qc/pAN60 values for locations
Association of State Highway and Transportation A and D range from about 3 to 8, which is typical for
Officials (AASHTO, 1991), because of the role of soil sandy soils based on their mean particle size D50
fabric and structure (Robertson, 2016). Nonetheless, (Robertson et al., 1983).
a qualitative comparison between the SBT profiles gen-
erated using the selected SBT chart and the soil profiles 2.2.2 Outwash
obtained from in situ boring logs can be instructive.
To complement the CPT profiles obtained at Figure 2.13 shows the CPT profiles (qc, fs, FR) and
locations A–J, the corrected SPT blow count N60 and the SBT profile generated using the modified Tumay
the ratio qc/pAN60 are plotted as a function of depth; (1985) chart for location C in LaPorte County, while
where pA 5 reference stress (5 100 kPa or 14.5 psi). The Figure 2.14 shows the CPT profiles (qc, fs, FR), the
SPTs were performed using an automatic trip hammer SBT profile generated using the modified Tumay (1985)
with an energy ratio of about 80% (Salgado, 2008). As chart, the SPT N60 and qc/pAN60 profiles, and the in situ
both cone resistance and SPT blow count are essentially layer information (with AASHTO group numbers)
penetration resistances, they are closely related. Hence, reported in the boring log for location E in Tippecanoe
plots of qc/pAN60 versus depth may be useful in case a County. Location C lies in the outwash region of
CPT-based design method needs to be used when only northern Indiana, while location E is on the bank of the
SPT blow counts are available for the site. It should be Wabash River near Purdue University. Outwash is a
noted that not all the locations marked in Figure 2.8 mixture of sand, gravel, cobbles, and boulders that are
have SPT borings completed along with CPT sound- transported and deposited by glacial meltwater; it may
ings. Also, it is important to note that the SPT borings also include some modern river alluvium. The SBT
were not carried out at the exact locations of the CPT profile at location C consists of multiple layers of loose-
soundings but were performed within the same project to-very dense sand or silty sand (Figure 2.13).
site. Therefore, the following qc/pAN60 plots for each The soil profile reported in the SPT boring log for
site should be interpreted with caution. location E in Tippecanoe County consists of sandy clay
loam and loose-to-medium dense sandy gravel in the
2.2.1 Dune/Aeolian Sands upper half of the profile and medium dense-to-very
dense sand with gravel, cobbles, and boulders in the
Figure 2.11 and Figure 2.12 show the CPT profiles lower half of the profile (Figure 2.14b). The N60 values
(qc, fs, FR), the SBT profile generated using the range from about 2 to 43, and the qc/pAN60 values range
modified Tumay (1985) chart, the SPT N60 and qc/ from about 3 to as high as 16 due to the presence of
pAN60 profiles, and the in situ layer information (with gravel, cobbles, and boulders in the soil profile. The
AASHTO group numbers) reported in the boring logs modified Tumay (1985) chart includes soil behavior
for location A in Lake County and location D in types ranging from clays to clay-silt-sand mixtures to
Newton County, respectively. Location A is in the dune sands of varying states; however, the chart does not
sand region of northern Indiana, while location D is clearly distinguish sands from sand-gravel mixtures and
slightly further to the south of location A. The gravelly sands. Therefore, a modified version of the
stratigraphic profile obtained from the SPT boring Robertson (1990) SBT chart, which distinguishes clean
log at location A consists of 8 m (26 ft) of medium sand from gravelly sand, was also used to generate the
dense sandy loam followed by 7 m (23 ft) of very loose- SBT profile for location E.
to-medium dense sand and 3 m (10 ft) of dense sandy Figure 2.15 compares the SBT profile generated
loam. On the other hand, the stratigraphic profile from using the modified Robertson (1990) chart with that
the SPT boring log at location D consists of 1.5 m (5 ft) obtained using the modified Tumay (1985) chart for
of very loose-to-loose sand followed by 12.5 m (41 ft) of location E in Tippecanoe County. In order to classify
medium dense sand. the coarse-grained soil layers at the site based on their
The numbers mentioned on the SBT profiles, gen- relative density (using the Salgado and Prezzi (2007)
erated using the modified Tumay (1985) chart, corre- correlation), the saturated unit weight csat, the critical-
spond to the soil zones listed in Table 2.2. The SBT state friction angle c, and the coefficient of lateral
profiles generated using the modified Tumay (1985) earth pressure at-rest K0 of the coarse-grained layers

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 11


Figure 2.11 In situ test profiles for location A in Lake County: (a) CPT-1 profile (qc, fs, FR) and SBT interpreted from modified
Tumay (1985) chart, and (b) N60 profile, (qc/pA)/N60 profile, and soil profile from SPT boring TB-2 (Data source: A. Tilahun,
J. Paauwe, & N. Z. Siddiki, personal communication, December 20, 2017).

were taken as 22.5 kN/m3 (143.2 lb/ft3), 32u, and 0.45, the profile. The mean particle size D50 and gravel
respectively. The SBT profile obtained using the content at the site are in the range of 0.4–4.5 mm
modified Robertson (1990) chart shows layers of very (0.016–0.18 in.) and 5%–50%, respectively (Han et al.,
dense and medium dense gravelly sand to sand, indi- 2019b, 2020). Hence, for sites with high gravel content,
cated by zone numbers 6 and 8, respectively (Table 2.3), the modified Robertson (1990) chart is a better option
between elevations ranging from 149–153 m and 137– for generating SBT profiles from CPT data than the
143 m and a layer of medium dense gravelly sand to modified Tumay (1985) chart. The delineation of gra-
sand at the 128–131 m elevation. In contrast, the SBT velly material in the profile using a CPT-based SBT
profile obtained using the modified Tumay (1985) chart has implications in foundation design because the
chart shows layers of very dense sand or silty sand constitutive response of a sand-gravel mixture is
(indicated by zone number 11) at these elevations and different from that of clean sand, for instance, when
does not capture the presence of gravelly material in subjected to shearing.

12 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


Figure 2.12 In situ test profiles for location D in Newton County: (a) CPT-2 profile (qc, fs, FR) and SBT interpreted from
modified Tumay (1985) chart, and (b) N60 profile, (qc/pA)/N60 profile, and soil profile from SPT boring TB-1 (Data source:
A. Tilahun, J. Paauwe, & N. Z. Siddiki, personal communication, December 20, 2017).

2.2.3 Glacial Till and loam with different percentages of sand, silt, and
clay. The qc/pAN60 values for locations F, G, and H
Figures 2.16, 2.17, 2.18, and 2.19 show the CPT range from 0.5–2.0, 0.5–1.0, and 1.0–3.5, respectively.
profiles (qc, fs, FR), the SBT profiles generated using These ranges are smaller than those reported for the
the modified Tumay (1985) chart, the SPT N60 and qc/ dune/aeolian sand and outwash regions in Sections
pAN60 profiles, and the in situ layer information (with 2.2.1 and 2.2.2, respectively, due to the presence of
USCS/AASHTO group numbers) reported in the smaller particle sizes associated with the soil types
boring logs for locations B, F, G, and H in Steuben, illustrated in Figure 2.16 to Figure 2.19.
Clinton, Madison, and Decatur counties, respectively.
These locations are characterized by glacial till deposits,
2.2.4 Loess with Sand
as shown in Figure 2.8. Location B is in northeastern
Indiana where the till is in a hummocky moraine form, Figure 2.20 shows the CPT profiles (qc, fs, FR), the
locations F and G are in central Indiana where the till is SBT profile generated using the modified Tumay (1985)
mostly in the form of flat plains, and location H is in chart, the SPT N60 and qc/pAN60 profiles, and the in situ
southeastern Indiana where the till is capped by thin layer information (with AASHTO group numbers)
wind-blown silt. The stratigraphic profiles at these obtained from the SPT boring log for location I in
locations consist of layers of sandy silty clay, silty sand, Knox County. This location is in southwestern Indiana,

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 13


Figure 2.13 CPT-2 profile (qc, fs, FR) and SBT interpreted from modified Tumay (1985) chart for location C in LaPorte County
(Data source: A. Tilahun, J. Paauwe, & N. Z. Siddiki, personal communication, December 20, 2017).

which is characterized by wind-blown silt deposits. N60 as a function of mean particle size D50. The chart
The stratigraphic profile obtained from the SPT boring includes data reported by Robertson et al. (1983) and
log consists of 1 m (3 ft) of very loose sand followed by data obtained from 15 sites in Indiana (2 sites each in
3 m (10 ft) of very loose-to-loose loam, 2 m (6.5 ft) of Hamilton, Tippecanoe, Clinton, and Greene counties,
very loose-to-medium dense sandy loam, 8 m (26 ft) of and 1 site each in Jasper, Lake, Newton, Knox, Starke,
soft-to-hard silty loam, and finally unweathered-to- Dubois, and Carroll counties). Starke, Newton, Jasper,
highly-weathered sandstone at a depth of 16.3–21.0 m and Lake counties are located in northern Indiana;
(53–69 ft) below the ground surface. These layers are Hamilton, Tippecanoe, Carroll, and Clinton counties
also captured by the CPT-based SBT profile via zone are in central Indiana; and Greene, Knox, and Dubois
numbers 6–10 (Table 2.2). The N60 values at the site counties are in southern Indiana. The following
range from about 5 to as high as 80, while the qc/pAN60 expression approximates the trend of the 98 data points
values range from 1.0 to 4.5. plotted in Figure 2.22:
 
qc D50 0:25 D50
2.2.5 Lacustrine Soil ~6:95 {0:18 for 0:001ƒ ƒ10 ðEq: 2:2Þ
pA N60 Dref Dref
Figure 2.21 shows the CPT profiles (qc, fs, FR) and
where pA 5 reference stress (5 100 kPa or 14.5 psi),
the SBT profile generated using the modified Tumay
D505 mean particle size, and Dref 5 reference particle
(1985) chart for location J in Vanderburgh County.
size (5 1 mm or 0.0394 in.). The coefficient of determi-
This location is in southern Indiana, near the border
nation R2 and the standard error (SE) of the regres-
with Kentucky, and is characterized by lacustrine soil.
sion are 0.89 and 0.77, respectively. Equation 2.2 may be
Lacustrine soils form under relatively quiet conditions
used to obtain an estimate of qc for use in a CPT-based
at the bottom of lakes and typically consist of silt to
foundation design method when only SPT blow counts
clay-sized particles. The SBT profile generated using the
are available for a site. However, as with any cor-
modified Tumay (1985) chart consists of 4 m (13 ft) of
relation involving the SPT blow count, Eq. 2.2 should
soft-to-very stiff clay and clayey silt underlain by 8 m be used with caution because of the potential error
(26 ft) of medium dense silty sand and 7 m (23 ft) of introduced by the transformation from the SPT blow
sandy clay or silty clay. count (a dynamic resistance) to the CPT cone resistance
(a quasi-static resistance). The qc/pAN60 ratio estimated
2.3 Correlation Between CPT Cone Resistance and SPT using Eq. 2.2 may be decreased by 20%–40%, if needed,
Blow Count to obtain a conservative value of cone resistance.
Equation 2.2 can be further improved as additional
Figure 2.22 shows the correlation between the CPT SPT blow count, cone resistance and D50 data become
cone resistance qc and the corrected SPT blow count available in Indiana.

14 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


Figure 2.14 In situ test profiles for location E in Tippecanoe County: (a) CPT-3 profile (qc, fs, FR) and SBT interpreted from
modified Tumay (1985) chart, and (b) N60 profile, (qc/pA)/N60 profile and soil profile from SPT boring Pier-7 (Data source:
A. Tilahun, J. Paauwe, & N. Z. Siddiki, personal communication, December 20, 2017).

The corrected SPT blow count N60 is expressed as 8


> 0:75 if rod lengthv4 m ð13 ftÞ
(Salgado, 2008): >
>
< 0:85 if 4 m ð13 ftÞƒrod lengthv6 m ð20 ftÞ
Cr ~
N60 ~Ch Cr Cs Cd NSPT ðEq: 2:3Þ > 6 m ð20 ftÞƒrod lengthv10 m ð33ftÞ
> 0:95 if
>
:
1:00 if rod length§10 m ð33 ftÞ ðEq: 2:5Þ
where NSPT 5 measured SPT blow count, Ch 5
hammer correction, Cr 5 rod length correction, Cs 5 
sampler correction, and Cd 5 borehole diameter 1:0 for liner sampler with liner in place
Cs ~
correction: 1:2 for liner sampler without the liner ðEq: 2:6Þ

8
8
> 0:75 for donut hammer ðER~45%Þ < 1:00 for B~65{115 mm ð2:5{4:5 in:Þ
>
>
> Cd ~ 1:05 for B~150 mm ð6:0 in:Þ
< 1:00 for safety hammerðER~60%Þ >
:
Ch ~ 1:15 for B~200 mm ð8:0 in:Þ ðEq: 2:7Þ
>
> 1:20 for pin weight hammer ðER~72%Þ
>
:
1:33 for automatic trip hammer ðER~80%Þ ðEq: 2:4Þ where ER 5 energy ratio, and B 5 borehole diameter.

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 15


Figure 2.15 Comparison of SBT profiles obtained from sounding CPT-5 at location E in Tippecanoe County using: (a) modified
Tumay (1985) chart (zone numbers listed in Table 2.2), and (b) modified Robertson (1990) chart (zone numbers listed in
Table 2.3).

Figure 2.16 CPT-4 profile (qc, fs, FR) and SBT interpreted from modified Tumay (1985) chart for location B in Steuben County
(Data source: A. Tilahun, J. Paauwe, & N. Z. Siddiki, personal communication, December 20, 2017).

16 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


Figure 2.17 In situ test profiles for location F in Clinton County: (a) CPT-7 profile (qc, fs, FR) and SBT interpreted from
modified Tumay (1985) chart, and (b) N60 profile, (qc/pA)/N60 profile, and soil profile from boring SPT-8.

2.4 CPT-Based Site Variability Assessment variability within individual CPT soundings and of the
collection of soundings performed at a site (Cao &
Soil properties used in geotechnical design are often Wang, 2013; Salgado et al., 2015; Xiao et al., 2018). If
estimated from a limited number of in situ or laboratory reasonably quantified, this uncertainty may be used to
tests (due to project budget and time constraints) and select appropriate resistance factors for use in load and
are thus subject to uncertainty, raising the question as resistance factor design (LRFD) of foundations and
to how accurately the soil properties derived from these retaining structures (Foye, 2005; Foye et al., 2006a,b,
tests are representative of the entire site (Phoon & 2009; Kim & Salgado, 2012a,b; Salgado et al., 2011;
Kulhawy, 1999a,b). Although this uncertainty cannot Salgado & Kim, 2014). For sites with high variability,
be eliminated, it can be quantified by analyzing the lower resistance factors could be used to increase the

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 17


Figure 2.18 In situ test profiles for location G in Madison County: (a) CPT RB-2 profile (qc, fs, FR) and SBT interpreted from
modified Tumay (1985) chart, and (b) N60 profile, (qc/pA)/N60 profile and soil profile from SPT boring TB-2 (Data source:
A. Tilahun, J. Paauwe, & N. Z. Siddiki, personal communication, December 20, 2017).

reliability of the foundation design, whereas for sites Salgado et al. (2019) developed the following four-
with low variability, higher resistance factors could be step procedure for CPT-based site variability assess-
used to optimize the construction cost. Based on the ment.
coefficient of variation (COV) of the average strength
parameter (e.g., SPT blow count NSPT) of each soil 1. Generate the SBT profile from the CPT data using an
SBT chart.
layer at a site, Paikowsky (2004) suggested that site
2. Quantify vertical variability via the vertical variability
variability can be classified as low (COV , 25%), index (VVI), which reflects the variability in qc, fs, and
medium (25% # COV # 40%), or high (COV . 40%). soil layering for each CPT sounding.
However, the volume of data available for statistical 3. Quantify horizontal variability via the horizontal varia-
analysis using the SPT is smaller in comparison to the bility index (HVI), which depends on the cross-correla-
CPT, and thus it is better to use a CPT dataset for site tion between cone resistance logs, cone resistance trend
variability assessment. differences, and the spacing between CPT soundings.

18 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


Figure 2.19 In situ test profiles for location H in Decatur County: (a) CPT-1 profile (qc, fs, FR) and SBT interpreted from
modified Tumay (1985) chart, and (b) N60 profile, (qc/pA)/N60 profile and soil profile from SPT boring TB-1 (Data source:
A. Tilahun, J. Paauwe, & N. Z. Siddiki, personal communication, December 20, 2017).

4. Combine both vertical and horizontal variability into an vertical and horizontal directions based on whether the
overall site variability rating (SVR) system. site VVI and HVI values fall in the 0%–33%, 33%–66%,
or 66%–100% range, respectively. Salgado et al. (2015,
Figure 2.23 shows how to categorize a site as being 2019) established a site variability rating, defined in
of low (L), medium (M), or high (H) variability in the terms of a string variable with two characters, each of

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 19


Figure 2.20 In situ test profiles for location I in Knox County: (a) CPT-1 profile (qc, fs, FR) and SBT interpreted from modified
Tumay (1985) chart, and (b) N60 profile, (qc/pA)/N60 profile, and soil profile from SPT boring TB-2 (Data source: A. Tilahun,
J. Paauwe, & N. Z. Siddiki, personal communication, December 20, 2017).

which may take the values, L, M, or H, as shown in the CPT-based site variability assessment algorithm
Figure 2.23. The first letter corresponds to the site VVI, developed by Salgado et al. (2019). The sampling
while the second letter corresponds to the site HVI. For interval for each CPT sounding was at most 5 cm
instance, if the site VVI and HVI values are 47% and (2 in.), and the sounding depths were in the range of 3–
31%, respectively, the site variability rating is ML, 20 m (10–65 ft). Sites A, C, and D have low site VVI
which stands for medium vertical variability and low values because their SBT profiles consist predominantly
horizontal variability. of medium dense-to-very dense sands of similar
Table 2.4 summarizes the computed vertical and behavior. In contrast, the other sites (B and E–J) have
horizontal variability indices for sites in Indiana using medium-to-high site VVI values because their SBT

20 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


Figure 2.21 CPT-28 profile (qc, fs, FR) and SBT interpreted from modified Tumay (1985) chart for location J in Vanderburgh
County (Data source: USGS, n.d.).

variable qc values within the depth of interest between


soundings. The CPT soundings at site G in Madison
County were performed only up to a depth of 3 m
(10 ft) because the project involved the replacement of
an existing structure and widening of the pavement.
Since the HVI value depends on the sounding depth
analyzed, the volume of CPT data obtained from
the shallow, closely-spaced soundings at site G in
Madison County may have been insufficient to render
an HVI value that is representative of the site—this
may have been another reason for the very high HVI
value of 100% obtained for this site. Based on the
procedure outlined previously, each site was assigned a
qualitative site variability rating (SVR), such as LH for
low vertical and high horizontal variability (e.g., site C)
and MH for medium vertical and high horizontal
variability (e.g., sites B, E to G, and J), as shown in
Figure 2.24.

2.5 Optimal Spacing Between CPT Soundings


Figure 2.22 Correlation between CPT cone resistance and The cost of a CPT-based geotechnical site investiga-
SPT blow count. tion is directly proportional to the number of CPT
soundings performed, which in turn depends on site
geology and variability. The cost of a CPT-based site
profiles consist of sandy, silty, and clayey soils with investigation could be reduced by optimizing the spaci-
relatively equal representation; layers of gravelly sand ng between CPT soundings based on the site variability
were also observed for site E in Tippecanoe County. determined from the soundings already performed at
Sites B and G in the glacial till areas of Steuben and the site. Figure 2.25 shows two CPT soundings, X and
Madison counties, respectively, have HVI values of Y, that have already been performed at a site; the
100% due to the presence of soil layers with highly center-to-center spacing between them is sxy.

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 21


1X N
Cxy ~ ðxi {
xÞðyi {
yÞ ðEq: 2:10Þ
N i~1

Cxy
rxy ~ ðEq: 2:11Þ
sx sy

The cross-covariance and cross-correlation coeffi-


cient of a sample dataset can also be calculated using
the functions COVARIANCE.S and CORREL, respec-
tively, in Microsoft Excel. The cross-correlation coeffi-
cient rxy takes values in the –1 to +1 range. A high
cross-correlation coefficient and small qc trend differ-
ence of a CPT pair indicates high correlation and
similarity between the two CPTs, and thus low
variability in the horizontal direction for the site.
 
Step 6: Calculate the average qc difference Dqc,avg 
between CPT soundings X and Y using:
P
N
jxi {yi j
 
Dqc,avg ~ i~1 ðEq: 2:12Þ
N
Figure 2.23 Site variability rating chart (modified from where xi and yi 5 qc values of the ith data point
Salgado et al., 2015).
obtained from CPT soundings X and Y, respectively,
and N 5 number of qc data points contained within the
segment length L.

The optimal spacing (syz)opt between CPT sounding  Step  7: Estimate the maximum credible difference
Dqc,avg  between qc trends for the segment length
max
Y and the next sounding Z can be calculated by
considered using:
following these steps (Ganju et al., 2019; Salgado et al.,
   0:46
2015, 2019): Dqc,avg  L
max
Step 1: Set the analysis (segment) length L as the ~23:86 {4:30
pA LR
minimum of the sounding depths of CPT soundings X
and Y. L
for 1ƒ ƒ30 ðEq: 2:13Þ
Step 2: Determine the number N of cone resistance LR
data points contained within the segment length L.
where L 5 analysis (segment) length, LR 5 reference
Step 3: Calculate the mean cone resistances x  and
length (5 1 m or 3.28 ft), and pA 5 reference stress
y of CPT soundings X and Y, respectively, for the

(5 100 kPa or 14.5 psi). The maximum credible
segment length considered.
difference is determined by considering two idealized
Step 4: Calculate the standard deviations x and
soil profiles, one with a very soft clay layer throughout,
y of the qc values of CPT soundings X and Y,
and the other with sand having 85% relative density
respectively, using: throughout (Salgado et al., 2019).
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Step 8: Calculate the values of functions f0, f1, and f2
u
u 1 X N
using:
s x ~t ðxi { xÞ2 ðEq: 2:8Þ
N{1 i~1 "   #
Dqc,avg 
f0 ~ min   ;1 ðEq: 2:14Þ
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Dqc,avg  max
u
u 1 X N
sy ~ t ðyi { yÞ2 ðEq: 2:9Þ
N{1 i~1 rxy z1
f1 ~ ðEq: 2:15Þ
2
th
where xi and yi 5 qc values of the i data point
obtained from CPT soundings X and Y, respectively.  
sxy
The standard deviation of a sample dataset can also be f2 ~1{ exp {0:25 ðEq: 2:16Þ
LR
calculated using the STDEV function in Microsoft
Excel. where sxy 5 spacing between CPT soundings X and Y,
Step 5: Estimate the cross-covariance Cxy and the and LR 5 reference length (5 1 m or 3.28 ft).
cross-correlation coefficient rxy between CPT sound- Step 9: Estimate the horizontal variability index
ings X and Y using: (HVI) for CPT soundings X and Y using:

22 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


TABLE 2.4
Vertical variability index, horizontal variability index, and site variability rating for the sites analyzed

Number of Sounding Depth


Site Soil Type County Soundings Analyzed (m) Sounding ID VVI (%) Site VVI (%) Site HVI (%) SVR Remarks

A Dune sand Lake 1 19.50 1006751CPT1 30 30 — — Low vertical variability


B Till in hummocky Steuben 4 6.50 0810115RB2 52 53 100 MH Medium vertical variability and high
moraine form 0810115TB1 65 horizontal variability
0810115RB1 28
0810115TB2 67
C Outwash LaPorte 2 10.15 0101453TB1 27 26 79 LH Low vertical variability and high
0101453TB2 25 horizontal variability
D Aeolian sand Newton 1 15.15 1006752CPT2 23 23 — — Low vertical variability
E Outwash Tippecanoe 3 15.55 0400774CPT2 39 40 71 MH Medium vertical variability and high
0400774CPT3 37 horizontal variability
0400774CPT5 43
F Glacial till Clinton 4 3.17 Frankfort02 69 63 96 MH Medium vertical variability and high
Frankfort05 91 horizontal variability
Frankfort06 37
Frankfort07 54
G Glacial till Madison 3 3.00 0101420CPT1 75 57 100 MH Medium vertical variability and high
0101420CPT2 55 horizontal variability
0101420CPT3 40
H Glacial till Decatur 1 8.10 1006241CPT1 69 69 — — High vertical variability

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


I Loess with sand Knox 1 14.40 0800579CPT1 68 68 — — High vertical variability
J Lacustrine soil Vanderburgh 3 19.95 VHC027 43 44 79 MH Medium vertical variability and high
VHC028 45 horizontal variability
VHC033 43

Note: The site VVI is the average of the individual VVIs of all CPT soundings performed at the site. The first letter in SVR corresponds to the site VVI, while the second letter corresponds to
the site HVI. Site HVI values and SVRs could not be assigned for sites where only one CPT sounding was performed.

23
Figure 2.24 Site variability ratings for the sites analyzed.

n n!
Cr ~ ðEq: 2:19Þ
ðn{rÞ!r!

where n Cr 5 number of combinations in which n


objects can be selected r at a time, n 5 number of CPT
soundings already performed at the site, and r 5 2 (for a
pair of CPT soundings). The number of pairs of CPT
soundings available at a site can also be calculated using
Figure 2.25 Optimal spacing between CPT soundings the COMBIN function in Microsoft Excel.
performed in line (modified from Salgado et al., 2015). b. Repeat steps 1 through 9 for all pairs of CPT soundings
performed at the site.
c. Calculate the average of the HVI values for all pairs of
CPT soundings performed at the site.
HVI~1{f2 ½0:8ð1{f0 Þz0:2f1  ðEq: 2:17Þ d. Substitute the average HVI value for the site into Eq.
2.18 to obtain the new spacing for the next CPT
The horizontal variability index ranges from 0 for a sounding. The next CPT sounding will be at a distance
perfectly uniform site to 1 for a highly variable site. no greater than (syz)opt from any sounding already
Step 10: Compute the optimal spacing (syz)opt performed at the site.
between CPT sounding Y and the next sounding Z
using: The procedure for estimation of optimal spacing
between CPT soundings is presented only to provide
  some guidance. The spacing between CPT soundings in
syz opt
~ð1:5{HVIÞsxy ðEq: 2:18Þ the field may be adjusted based on the level of impor-
tance of the structure, knowledge of the site geology,
and soil profile variability.
Equation 2.18 shows that if the value of HVI is
greater than 0.5, the spacing for the next CPT sounding
is decreased, but if the value of HVI is less than 0.5, the 2.6 Chapter Summary
spacing for the next CPT sounding is increased.
Step 11: If the CPT soundings are not performed in In this chapter, an overview of the bedrock and
line but are distributed in two dimensions, execute the surficial geology of Indiana was presented along with
following substeps. the CPT, SPT and soil profiles obtained from ten
different locations across Indiana. About two-thirds of
a. Determine the number of pairs of CPT soundings Indiana is covered by sediments that were transported
performed at the site using: and deposited by glaciers during the Ice Age; the

24 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


bedrock surface is visible only in the south-central developed based on data reported by Robertson et al.
part of the state. The bedrock geology of Indiana (1983) and data obtained from 15 sites in Indiana. The
mainly consists of five bedrock units: Pennsylvanian, correlation may be used to obtain an estimate of qc for
Mississippian, Devonian, Silurian, and Ordovician use in a CPT-based foundation design method when
units, which in turn consist of sedimentary rocks, such only SPT blow counts are available for a site because
as limestone, dolomite, shale, sandstone, and siltstone. CPT-based methods tend to be more reliable. However,
Limestone and dolomite dissolve slowly in water to as with any correlation involving the SPT blow count, it
produce karstic landforms (commonly found in south- should be used with caution because of the potential
ern Indiana) with underground cavities that may error introduced by the transformation from the SPT
collapse, forming sinkholes. The surface geology of blow count (a dynamic resistance) to the CPT cone
Indiana consists of soils transported by wind, water, or resistance (a quasi-static resistance). In such cases when
ice: (a) dune and aeolian sands in northern Indiana, (b) only SPT data is available for the site, it may be
outwash in northern Indiana and along major river preferrable to use SPT-based methods for design
valleys, (c) glacial till in central Indiana, and (d) loess in (though not in clay) instead of CPT-based methods.
southwestern Indiana. A CPT-based site variability assessment methodol-
CPT and SPT data were obtained from 10 select sites ogy developed by Salgado et al. (2019) was applied
across Indiana. The data was analyzed to obtain depth to assess the vertical and horizontal variability of the
profiles of cone resistance qc, sleeve resistance fs, fric- 10 sites in Indiana. The vertical variability of a CPT
tion ratio (FR), corrected SPT blow count N60, and sounding was quantified via the vertical variability
qc/pAN60. The CPT data was post-processed through a index (VVI), which reflects the intra-layer variability,
soil profile generation algorithm developed by Ganju the log variability and the COV of the cone resistance
et al. (2017) to generate SBT profiles for each site using of the sounding. The site VVI was taken as the average
the modified Tumay (1985) SBT chart. According to of the individual VVIs of all CPT soundings performed
this chart, a combination of low qc/pA (, 10) and high at a site. The horizontal variability of a site was
fs/qc values (. 4%) suggests a clayey soil, whereas a quantified via the site horizontal variability index (site
combination of high qc/pA (. 50) and low fs/qc values HVI), which depends on the cross-correlation between
(, 2%) suggests a sandy soil; where pA 5 reference cone resistance logs, cone resistance trend differences,
stress (5 100 kPa or 14.5 psi). For each site, the and the spacing between CPT soundings. The site VVI
CPT-based SBT profiles compared reasonably well and HVI values were combined into an overall site
with the corresponding soil profiles obtained from the variability rating (SVR) system.
SPT boring logs. The SBT profiles account for the A step-by-step procedure for estimation of opti-
presence of thin layers, which are otherwise not mal spacing between CPT soundings was presented
captured by the soil profiles reported in the SPT boring (Table 2.5). However, in order to implement the proce-
logs. This is because the SBT profiles are based on dure, data from at least two CPT soundings are needed
nearly continuous CPT measurements at depth inter- in advance to estimate the optimal spacing of future
vals of 5 cm (2 in.) or less, whereas the soil profiles CPT soundings performed at a site. The procedure
obtained from the SPT boring logs are based on may be further refined through future research, and
laboratory testing of soil samples collected typically at so the use of this procedure in INDOT construction
depth intervals of 1.5 m (5 ft). The modified Tumay projects is optional based on the level of familiarity of
(1985) chart can be used for generating SBT profiles for the engineers with the CPT and the specific site
all soil types in Indiana, except for gravelly materials, investigation goals of the project under consideration.
for which the modified Robertson (1990) chart is more CPT soundings at the desired spacing may be
appropriate. performed based on the level of importance of the
A correlation between cone resistance qc, corrected structure, knowledge of the site geology, and soil profile
SPT blow count N60, and mean particle size D50 was variability.

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 25


TABLE 2.5
Method for estimation of optimal spacing between CPT soundings (Ganju et al., 2019; Salgado et al., 2015, 2019)

Optimal Spacing Between CPT Soundings (syz)opt Notes


 
syz opt ~ð1:5{HVIÞsxy The horizontal variability index (HVI) ranges from 0 for a perfectly
HVI~1{f2 ½0:8ð1{f0 Þz0:2f1  uniform site to 1 for a highly variable site. If HVI is greater than
"   # 0.5, the spacing for the next CPT sounding is decreased, but if HVI is
Dqc,avg 
f0 ~ min   ; 1 less than 0.5, the spacing for the next CPT sounding is increased.
Dqc,avg max
If the CPT soundings are not performed in line but are distributed in two
PN
dimensions, calculate the average of the HVI values for all pairs of
jx {yi j
  i~1 i CPT soundings performed at the site. Substitute the average HVI value
Dqc,avg ~
  N into the equation for (syz)opt to obtain the new spacing for the next
Dqc,avg   0:46
max 5 23.86
L L CPT sounding.
pA ^4:30 for 1ƒ ƒ30  
LR
 
LR The equation for the maximum average qc difference Dqc,avg max was
rxy z1 sxy obtained by considering two idealized soil profiles, one with a very soft
f1 ~ and f2 ~1{exp {0:25
2 LR clay layer throughout, and the other with sand having 85% relative
Cxy 1XN
density throughout.
rxy ~ and Cxy ~ ðxi {
xÞðyi {yÞ
sx sy N i~1 The cross-correlation coefficient rxy takes values in the –1 to +1 range.
A high cross-correlation coefficient and small qc trend difference of
a CPT pair indicates high correlation and similarity between the two
CPTs, and thus low variability in the horizontal direction for the site.

Note: sxy 5 spacing between two CPT soundings, X and Y, that have already been performed at a site, (syz)opt 5  optimal
 spacing between CPT
sounding Y and the next sounding Z that needs to be performed at the site, HVI 5 horizontal variability index, Dqc,avg  5 average qc difference

between CPT soundings X and Y for the segment length considered, N 5 number of qc data points contained within the segment length, Dqc,avg max 5
maximum credible difference between qc trends for the segment length considered, L 5 analysis (segment) length, LR 5 reference length (5 1 m
or 3.28 ft), pA 5 reference stress (5 100 kPa or 14.5 psi), rxy 5 cross-correlation coefficient between CPT soundings X and Y, sx and sy 5 standard
deviations of the qc values of CPT soundings X and Y, respectively, Cxy 5 cross-covariance between CPT soundings X and Y, xi and yi 5 qc values
of the ith data point obtained from CPT soundings X and Y, respectively, and x  and  y 5 mean cone resistances of CPT soundings X and Y,
respectively, for the segment length considered.

3. CPT-BASED DESIGN OF SHALLOW possible, where u2 5 pore water pressure measured at the
FOUNDATIONS shoulder position behind the cone face (refer to Volume I).
c. Obtain the unit weight of the soil in each layer of
Shallow foundations are typically used to support the profile whenever soil samples are recovered during
small-to-medium-sized structures on competent soils the site investigation. In the absence of soil samples,
near the ground surface. The design of a shallow the reader may refer to Section 2.3.3 of Volume I for
foundation involves two key steps: (a) ultimate limit correlations between the unit weight and CPT data. In
general, the saturated unit weight csat of soil typically
state check, and (b) serviceability limit state check.
ranges from 18–21 kN/m3 (115–135 pcf) for sand, 18.5–
Although both bearing capacity and serviceability 22.5 kN/m3 (118–143 pcf) for silty sand, and 15–18 kN/
criteria should be checked properly, only one of the m3 (95–115 pcf) for clay (Salgado, 2008).
two typically controls the design of shallow foundations
depending on the soil type and loading conditions. Step 2: Set the footing shape (e.g., strip, square,
rectangular, or circular), the preliminary geometry
3.1 Calculation Procedure for Footing Settlement (length L and width B) of the footing, and the
embedment depth D of the footing.
The total settlement w of an axially-loaded footing Step 3: Classify the soil in each layer of the pro-
can be calculated from CPT results by following these file below the footing as either ‘‘sand’’ or ‘‘clay.’’ For
steps. mixed or intermediate soils (i.e., soils containing
Step 1: Obtain the site stratigraphy, the groundwater mixtures of sand, silt, and clay), execute the following
table depth, and the unit weight of the soil in each layer substeps.
of the profile.
a. Sand-silt, sand-clay or sand-silt-clay mixtures: Classify
a. Establish the site stratigraphy either from the boring log these soils as ‘‘clay’’ if fines content FC $ 20% and
or by using a CPT-based soil behavior type (SBT) chart plasticity index PI $ 8%, otherwise classify them as
(refer to Section 2.2.3 of Volume I) or both if possible. ‘‘sand’’ (Carraro et al., 2009; Salgado et al., 2000).
b. Obtain the depth zw of the groundwater table from either b. Sands containing gravel: If a site contains sand layers
the boring log or the depth profile of u2 or both if with gravel content greater than 20%, use the lower-

26 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


bound profile of qc, drawn approximately through the N If the mean particle size D50, coefficient of
valleys of the actual qc profile, for estimating footing uniformity CU, and particle roundness R of the
settlement and bearing capacity. sand are known, estimate the critical-state frict-
ion angle using:
Note: In the absence of soil samples, the reader may  
refer to Section 2.2 of Volume I for estimation of soil D50 f
c ð0 Þ~28:3 ðCU Þ2f ðRÞ{3f ðEq: 3:2Þ
behavior type from CPT results. Dref
Step 4: Correct the raw qc data for the pore water
where Dref 5 reference particle size (5 1 mm or
pressure generated during cone penetration using
0.04 in.), and f 5 exponent (5 0.045). Equation
(ASTM, 2012): 3.2 is applicable for poorly-graded, clean silica
sands with D50 5 0.15–2.68 mm (0.006–0.105
qt ~qc zð1{aÞu2 ðEq: 3:1Þ in.), CU 5 1.2–3.1, and R 5 0.3–0.8. The data
used in the development of this equation along
where qt 5 corrected, total cone resistance, qc 5 mea-
with example calculations can be found in Appen-
sured cone resistance, a 5 cone area ratio (< 0.8 for dix A.
typical CPT probes), and u2 5 pore water pressure N If direct shear or triaxial compression test results
measured at the shoulder position behind the cone face. are available, it is recommended that the critical-
The pore water pressure correction to the qc data may state friction angle be determined from such test
be ignored for coarse-grained soils (e.g., sand and gravel) results.
because qt is approximately equal to qc in such soils.
ii. Calculate the gross unit load qb on the footing base
Step 5: Obtain the footing load and maximum
(including the loads from the superstructure, the
tolerable settlement. weight of the foundation, and the weight of the
a. Obtain the unfactored structural load Q that will be backfill when the excavation is backfilled):
applied on the footing from the structural engineer. QzWftg zWfill
b. Set the maximum tolerable angular distortion amax as qb ~ ðEq: 3:3Þ
A
1/500 (Skempton & MacDonald, 1956) or other such
value specified by a geotechnical code. where Q 5 unfactored column (or wall) load on
c. Set the maximum tolerable settlement wmax of the footing the footing, Wftg 5 weight of the footing (5 ccAt),
from Table 3.1 or other such value specified by a geo- cc 5 unit weight of concrete (< 24 kN/m3 or 150
technical code. pcf), A 5 area of the footing base, t 5 thickness of
the footing, Wfill 5 weight of the backfill 5
Step 6: Calculate the total settlement of the footing. max[cfillA(D – t) ; 0], cfill 5 unit weight of the
backfill, and D 5 depth of embedment of the
a. Total settlement of footings in ‘‘sand’’ (Lee et al., 2008; footing. If the footing is not backfilled, Wfill 5 0.
Lee & Salgado, 2002; Schmertmann, 1970; Schmertmann If the thickness of the footing is unknown, an
et al., 1978). Execute the following substeps for footings ‘‘average’’ unit weight cavg may be used for the
in ‘‘sand,’’ otherwise proceed to step 6(b). material above the footing base to calculate the
gross unit load qb:
i. Determine the critical-state friction angle c of
sand through one of the following options. Q Q c zcfill

qb ~ zcavg D~ z c D ðEq: 3:4Þ


A A 2
N Select a c value between 28u and 36u for silica
sand; sands with rounded, smooth particles with
a poorly-graded particle size distribution have iii. Calculate the influence depth zf0 measured from
values near the low end of this range, while sands the footing base using:
with angular, rough particles with a well-graded  
particle size distribution have values near the zf 0 L
~2z0:4 min ; 6 {1 ðEq: 3:5Þ
high end of this range (refer to Appendix A for B B
additional information if needed).
iv. Calculate the depth zfp measured from the footing
base at which the strain influence factor peaks
TABLE 3.1 using:
wmax/amax values for shallow foundations in sand and clay  
(Salgado, 2008; Skempton & MacDonald, 1956) zfp L
~0:5z0:1 min ; 6 {1 ðEq: 3:6Þ
B B
wmax/amax
Soil Type Isolated Foundations Mat Foundations v. Based on the cone resistance profile, divide the soil
layers within the influence depth zf0 below the
Sand 15LR 20LR footing base into sublayers such that the qc values
Clay 25LR 30LR within each sublayer are either approximately
constant or linear with depth so that a representa-
Note: LR 5 reference length (5 1 m or 39.4 in.). Strip footings are
tive cone resistance can be assigned to each
continuous and behave more like mat foundations than isolated
sublayer.
foundations.

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 27


vi. Estimate the strain influence factor Iz for the where qc 5 representative cone resistance of the
sublayer using (Figure 3.1): sublayer, pA 5 reference stress (5 100 kPa or 14.5
8
> zf   psi), 9h0 5 in situ horizontal effective stress at the
>
< Iz0 z z Izp {Iz0 for zf vzfp middle of the sublayer (5 K09v0), and v90 5 in situ
fp
Iz ~ z {z vertical effective stress at the middle of the sublayer
>
> f0 f
Izp for zfp ƒzf ƒzf 0 ðEq: 3:7Þ
: (Terzaghi, 1943):
zf 0 {zfp
0
sv0 ~sv0 {u0 ðEq: 3:11Þ
where zf 5 vertical distance from the footing base
to the middle of the sublayer, Iz0 5 strain influence where v0 5 in situ vertical total stress at the
factor at the footing base level, and Izp 5 peak middle of the sublayer, u0 5 hydrostatic pore water
strain influence factor: pressure at the middle of the sublayer {5 max[cw
  (z – zw) ; 0]}, cw 5 unit weight of water (5 9.81 kN/
L m3 or 62.45 pcf), z 5 depth measured from the
Iz0 ~ min 0:1z0:0111 {1 ; 0:2 ðEq: 3:8Þ
B ground surface to the middle of the sublayer, and
zw 5 depth of the groundwater table.
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi ix. Estimate the elastic modulus E of the sublayer
u
uqb {sv0 zf ~0
0
using:
Izp ~0:5z0:1t 0  ðEq: 3:9Þ
s v0 zf ~zfp  {0:285  0:4  
E w B DR {0:65
 ~l ðEq: 3:12Þ
where sv0 z ~0 5 in situ vertical qc LR LR 100
0
effective stress at
f 0 
the footing base level, and sv0 z ~z 5 in situ verti-
f fp
8
cal effective stress at the depth corresponding to zfp.
< 0:38
> for young NC silica sand
vii. Determine the coefficient of lateral earth pressure l~ 0:53 for aged NC silica sand
at-rest K0 of the sublayer (refer to Appendix B for >
:
0:91 for over OC silica sand ðEq: 3:13Þ
guidance).
viii. Estimate the relative density DR of the sublayer
using (Salgado & Prezzi, 2007): where w 5 initial guess value for footing settle-
   0  ment (5 wmax established in step 5), B 5 width or
qc s
ln {0:4947{0:1041c {0:841 ln h0 diameter of the footing, LR 5 reference length (5
pA pA
DR ð%Þ~  0  1 m or 3.28 ft), DR 5 relative density of the sub-
s
0:0264{0:0002c {0:0047 ln h0 layer (expressed as a percentage), and l 5
pA
parameter that accounts for the effects of aging
ðEq: 3:10Þ and overconsolidation of sand.

Figure 3.1 Strain influence factor Iz versus depth zf below the footing base (after Salgado 2008; Schmertmann et al., 1978).

28 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


x. Compute the total settlement w of the footing iv. Obtain the small-strain shear modulus G0 profile
using: within the influence depth zG 0 below the footing
n   base from the results of seismic cone penetration
0 

X Izi Dzi
w~C1 C2 qb {sv0 zf ~0 ðEq: 3:14Þ tests (SCPTs) using (Salgado, 2008):
i~1
Ei
cm 2
G0 ~ V ðEq: 3:19Þ
where Dz 5 thickness of the sublayer, n 5 number g s
of sublayers within the influence depth zf0 below where cm 5 unit weight of soil (5 csat if the soil is
the footing base, and C1 and C2 5 depth and time saturated), g 5 acceleration due to gravity (5 9.81
factors, respectively: m/s2 or 32.17 ft/s2), and Vs 5 shear wave velocity
0  ! (refer to Section 2.3.4 of Volume I).
sv0 z ~0
C1 ~1{0:5 f
 ðEq: 3:15Þ If SCPT results are unavailable, the small-strain
qb {s 
0
v0 zf ~0 shear modulus may be estimated using the follow-
ing correlation (Viggiani & Atkinson, 1995):
   0 ng
t G0 100sm0 m
C2 ~1z0:2 log ðEq: 3:16Þ ~Cg R0 g ðEq: 3:20Þ
0:1tR pA pA

where tR 5 reference time (5 1 year), and t 5 where Cg, ng, and mg 5 parameters that depend on
service life of the superstructure (in the same unit 9 0 5 in situ mean effective
the plasticity index PI; sm
as tR). stress at the depth being considered; pA 5 reference
xi. Compare the value of w calculated using Eq. 3.14 stress (5 100 kPa or 14.5 psi); and R0 5 mean
with the initial guess value assumed in substep (ix). stress-based overconsolidation ratio:
If the two values match, then report the value of w 0
!
calculated using Eq. 3.14 as the settlement of the pp 1z2K0,NC
R0 ~ 0 ~OCR pffiffiffiffiffiffiffiffiffiffiffi ðEq: 3:21Þ
footing. However, if they do not match, return to p 1z2K0,NC OCR
substep (ix) and use the new value of w obtained
0
from Eq. 3.14 as the initial guess value for the next where pp 5 value of p9 at the intersection of the
iteration (refer to Appendix C for guidance). recompression line with the normal consolidation
line in n–ln p9 space, n 5 specific volume (5 1+e),
b. Total settlement of footings in ‘‘clay.’’ Execute the follow- K0,NC 5 coefficient of lateral earth pressure at-rest
ing substeps for footings in ‘‘clay,’’ otherwise proceed to for normally consolidated soil (< 0.50–0.75 for NC
step 7. clay), and OCR 5 overconsolidation ratio (refer to
Appendix B for guidance).
Immediate settlement of footings in clay (Foye et al., The parameters Cg, ng, and mg can be calculated
2008) using (Foye et al., 2008; Viggiani & Atkinson,
1995):
i. Obtain the depth profile of undrained shear strength
Cg ~37:9 exp ð{0:045 PIÞ for PIw5% ðEq: 3:22Þ
su below the footing base using (Salgado, 2008):

qt {sv0
su ~ ðEq: 3:17Þ ng ~0:109 ln ðPIÞz0:4374 for PIw5% ðEq: 3:23Þ
Nk

mg ~0:0015 PIz0:1863 for PIw5% ðEq: 3:24Þ


where qt 5 corrected, total cone resistance mea-
sured under undrained conditions, v0 5 in situ
vertical total stress at the depth being considered,
The in situ mean effective stress can be calculated
and Nk 5 cone factor (< 9–15 as long as the CPT is
using:
performed at a penetration rate that is sufficiently
high to ensure undrained penetration (refer to 0 1 0 0

Appendix D); soft NC clays tend to have Nk values sm0 ~ sv0 zksh0 ðEq: 3:25Þ
kz1
near the low end of this range, while stiff OC clays
tend to have Nk values near the high end of this where k 5 1 for plane-strain conditions (e.g., strip
range) (Bisht et al., 2021; Mayne & Peuchen, 2018; footings) and 2 for triaxial conditions (e.g., isolated
Salgado, 2008, 2013, 2014; Salgado et al., 2004). footings), sv90 5 in situ vertical effective stress at the
ii. Average the values of su over a vertical distance of depth being considered, s9h0 5 in situ horizontal
B below the footing base to obtain a representative effective stress at the depth being considered (5
undrained shear strength su . K0s9v0), and K0 5 coefficient of lateral earth
iii. Calculate the influence depth zG 0 below the foot- pressure at-rest (refer to Appendix B for guidance).
ing base within which most of the strains develop The plasticity index PI is the difference between the
using: liquid limit LL and the plastic limit PL of the soil
  (PI 5 LL – PL).
zG 0 L v. Calculate a representative small-strain shear modulus
~ min 1z0:111 {1 ; 2 ðEq: 3:18Þ  0 by taking the weighted average of the G0 values
B B G

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 29


within the influence depth zG 0 below the footing iii. Obtain the initial void ratio e0 of the sublayer using
base: the relationship e0 5 wcGs/S; where wc 5 water
Pn
avg content, Gs 5 specific gravity of solids (5 2.60–2.80
G0,i Hi
i~1 for clay), and S 5 degree of saturation (5 1 for
G 0~ ðEq: 3:26Þ
Pn saturated clay). In the absence of soil samples, the
Hi reader may refer to Section 2.3.1 of Volume I for
i~1
additional information on e0.
iv. Estimate the vertical compressive strain Dez of the
avg
where G0,i 5 average small-strain shear modulus of sublayer using:
layer i, Hi 5 thickness of layer i, and n 5 number of
clay layers within the influence depth zG 0 below the 8  0
>
> Cc sv 0 0 0 0
footing base. >
> log 0 if sv0 ~svp and sv §svp ðNC clayÞ
>
>
1ze0 sv0
vi. Using trial footing dimensions, estimate the net unit >
>  0   0 
>
>
load qb,net on the footing base: >
> 1
C log
svp
zC log
sv
> 1ze0
< s s
0 c s0
v0 vp

qb,net ~qb {cm D ðEq: 3:27Þ Dez ~ 0 0 0


>
> if sv0 vsvp ƒsv ðOC then NC clayÞ
>
>  0
>
>
where qb 5 gross unit load on the footing base >
> Cs sv 0 0 0 0
>
> 1ze0 log 0 if sv0 ƒsvp and sv ƒsvp ðOC clayÞ
(including the loads from the superstructure, the >
> sv0
>
:
weight of the foundation, and the weight of the ðEq: 3:31Þ
backfill when the excavation is backfilled; refer to
Eqs. 3.3 and 3.4), and cmD 5 total overburden where 9v0 5 initial (or in situ) vertical effective
stress at the footing base level. stress at the middle of the sublayer before the stress
vii. Obtain the influence factor Iq from Figure 3.2; increment is applied, v9 5 current vertical effective
H 5 thickness of the clay layer below the footing stress at the middle of the sublayer after the stress
base, and B 5 footing width. For circular footings, increment is applied and full primary consolidation
an equivalent footing width may be obtained by has taken place (5 9v0 + Dv), 9vp 5 preconsolida-
equating the cross-sectional area of the footing with tion stress, Cc 5 compression index, and Cs 5
that of an equivalent square. swelling index.
viii. Estimate the representative small-strain Young’s In the absence of laboratory consolidation test
modulus E  0 of clay below the footing base using: results, the compression index Cc may be estimated
using the following approximate correlation (Wroth
E 0
 0 ~2ð1znÞG ðEq: 3:28Þ & Wood, 1978):
1
where n 5 Poisson’s ratio (5 0.5 for undrained Cc & Gs PIð%Þ ðEq: 3:32Þ
200
conditions).
ix. Compute the immediate settlement wi of the footing where PI 5 plasticity index (expressed as a
using: percentage). The swelling index Cs typically ranges
from 0.1Cc to 0.2Cc.
qb,net B v. Compute the 1D consolidation settlement wc1D of
wi ~Iq 0 ðEq: 3:29Þ
E the clay layer below the footing base using:
P
n
wc1D ~ Dez,i Dzi ðEq: 3:33Þ
i~1
Primary consolidation settlement of footings in clay
(Skempton & Bjerrum, 1957) where Dzi 5 thickness of sublayer i, and n 5
number of sublayers.
i. Divide the clay layer below the footing base into n vi. Compute the primary consolidation settlement wc
sublayers of thickness Dz. of the footing using:
ii. Calculate the vertical stress increment Dv at the
middle of each sublayer caused by the applied load wc ~½Azað1{AÞwc1D ðEq: 3:34Þ
Q using the 2-to-1 stress distribution rule:
8
> Q ÐH
>
> Bzz for strip footings Ds3 dz
>
> f
>
>
>
>
> 4Q a~ 0H ðEq: 3:35Þ
<  2 for circular footings Ð
Dsv ~ p Bzzf Ds1 dz
>
> 0
>
> Q
>
>   for rectangular footings
>
> Bzzf Lzzf
>
> where A 5 Skempton’s pore pressure parameter (<
:
ðEq: 3:30Þ 0.5–0.75 for NC clay and 0.3–0.5 for OC clay), D1
5 major principal stress increment, D3 5 minor
where zf 5 vertical distance from the footing base principal stress increment, and H 5 thickness of the
to the middle of the sublayer. Q takes units of load clay layer below the footing base.
per unit length for strip footings and units of load Table 3.2 summarizes the values of a for circular
for all other footings. and strip footings as a function of H/B. For square

30 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


Figure 3.2 Influence factor Iq as a function of qb,net =su and H/B for (a) strip footings, (b) square footings, and (c) rectangular
(L/B 5 2) footings (Foye et al., 2008; Salgado, 2008).

TABLE 3.2
Values of a for estimation of primary consolidation settlement of footings in clay (Skempton & Bjerrum, 1957)

Coefficient a
Normalized Thickness of Clay Layer H/B Circular Footing (B/L 5 1) Strip Footing (B/L 5 0)

0 1.00 1.00
0.25 0.67 0.74
0.5 0.50 0.53
1 0.38 0.37
2 0.30 0.26
4 0.28 0.20
10 0.26 0.14
‘ 0.25 0

footings, the value of a for a circular footing with (i.e., excessive settlement). Repeat step 6 to optimize the
the same cross-sectional area as that of a square design if needed. However, if w . wmax, return to step 6
footing may be used. For rectangular footings with and revise the footing geometry.
0 , B/L , 1, obtain the value of a by interpolation. Step 8: Angular distortion check.
vii. Sum the values of wi and wc to obtain the total settle-
Execute the following substeps for each pair of
ment w of the footing. Note that if significant secon-
dary consolidation is expected at the site, it should
adjacent footings at the site.
be considered together with primary consolidation. a. Compute the angular distortion a for the selected footing
pair using:
Step 7: Total settlement check. Dw
a~ ðEq: 3:36Þ
Compare the estimated total settlement w of the Lcc
footing with the maximum tolerable settlement wmax
selected in step 5. If w # wmax, the footing design is where Dw 5 differential settlement, and Lcc 5 span or
satisfactory with respect to the serviceability limit state center-to-center distance between the two footings.

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 31


b. Compare the estimated angular distortion a for the given i. Using the values of zw, B and D determined from
footing pair with the maximum tolerable angular distor- Section 3.1, calculate the value of the unit weight c
tion amax selected in step 5. If a # amax, the footing design to use in the bearing capacity equation:
is satisfactory with respect to the ultimate/serviceability 8
limit state (i.e., excessive differential settlement). If a . >
> cb if zw vD
>
<  
amax, redo the footing design until the maximum tole- zw {D
c~ cb z ðcm {cb Þ if Dƒzw ƒDzB
rable angular distortion criterion is satisfied. If the criterion >
> B
>
:
cannot be satisfied, consider alternative design solutions, cm if zw wDzB ðEq: 3:39Þ
such as the use of grade beams; combined footings;
replacement of foundation soil with compacted, coarse- where zw 5 depth of the groundwater table, B 5
grained material; geosynthetic-reinforced foundation bed; footing width or diameter, D 5 depth of embed-
mat (or raft) foundations; pile foundations; and piled rafts. ment of the footing, cm 5 moist unit weight of sand,
cb 5 buoyant unit weight of sand (5 csat – cw), csat
5 saturated unit weight of sand, and cw 5 unit
3.2 Calculation Procedure for Limit Bearing Capacity of weight of water (5 9.81 kN/m3 or 62.45 lb/ft3).
Footings ii. Estimate the relative density DR of sand at a depth
of B/2 below the footing base using:
The limit unit bearing capacity qbL of an axially-    0 
qc,CAM s
loaded footing can be calculated from CPT results by ln
pA
{0:4947{0:1041c {0:841 ln h0
pA
following these steps. DR ð%Þ~  0 
sh0
Step 1: Determine the nominal or characteristic cone 0:0264{0:0002c {0:0047 ln
pA
resistance qc,CAM. ðEq: 3:40Þ
a. Combine the cone resistance profiles obtained from all where qc,CAM 5 conservatively assessed mean
CPT soundings performed at the site. Note that, for (CAM) cone resistance at a depth of B/2 below
fine-grained soils (e.g., silts and clays), the cone the footing base (obtained from Eq. 3.37), v90 5 in
resistance should be corrected for pore water pressure situ vertical effective stress at a depth of B/2 below
u2 using Eq. 3.1. the footing base, 9h0 5 in situ horizontal effective
b. Perform a linear regression on the cone resistance data stress at a depth of B/2 below the footing base (5
points to obtain the mean trend of the data with depth K09v0), pA 5 reference stress (5 100 kPa or 14.5
(Figure 3.3). When performing the regression, consider psi), c 5 critical-state friction angle (refer to step
only those data points that follow the general trend of 6(a)(i) of Section 3.1), and K0 5 coefficient of
the qc profile and ignore any outliers or regions that lateral earth pressure at-rest (refer to Appendix B
contain significant scatter in the data. for guidance).
c. Draw lines (parallel to the mean trendline) bounding the iii. Calculate the peak friction angle p of sand using
cone resistance data points, as shown in Figure 3.3. (Bolton, 1986):
d. Determine the relationship of cone resistance with
( " 0
!# )
depth that is exceeded by 80% of the measurements DR 100smp
using (Foye et al., 2006b): p ~c zAy Q{ ln {RQ
100 pA
qc,CAM ðzÞ~Eqc ðzÞ{0:84sqc ðEq: 3:37Þ ðEq: 3:41Þ
where qc,CAM(z) 5 conservatively assessed mean (CAM)
 
cone resistance determined using the 80% exceedance 1 L
criterion (Becker, 1996) (as a function of depth z), Eqc(z) Ay ~ min z8 ; 5 ðEq: 3:42Þ
3 B
5 equation of the mean trendline obtained from the
regression analysis, and qc 5 standard deviation of where pA 5 reference stress (5 100 kPa or 14.5 psi),
cone resistance (Foye et al., 2006a): Q and RQ 5 fitting parameters that depend on the
intrinsic characteristics of sand (Q 5 10 and RQ 5 1
(qc, max {qc,min )sample for clean silica sand), and 9mp 5 representative mean
sqc ~ ðEq: 3:38Þ
Ns effective stress (Loukidis, 2006; Salgado, 2008):
where qc,max 5 value of cone resistance at any depth    
0 cB 0:7 B
z on the upper bound line, qc,min 5 value of cone smp ~20pA 1{0:32 ðEq: 3:43Þ
pA L
resistance on the lower bound line at the same depth
z at which qc,max was computed (see Figure 3.3), and iv. Calculate the shape factors sq and sc using (Lyamin
N 5 number of standard deviations of qc (obtained et al., 2007):
from Table 3.3).
   
 D 0:7{0:01p B 1{0:16ðBÞ
D

sq ~1z 0:098p {1:64
Step 2: Calculate the limit unit bearing capacity of B L
the footing.
ðEq: 3:44Þ
a. Limit unit bearing capacity of footings in ‘‘sand.’’
Execute the following substeps for footings in ‘‘sand,’’  B
sc ~1z 0:0336p {1 ðEq: 3:45Þ
otherwise proceed to step 2(b). L

32 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


Figure 3.3 Examples of two CPT logs in clay and three CPT logs in sand with mean trendlines and range lines (after Foye et al.,
2006a).

TABLE 3.3
Values of N as a function of sample size n (after Tippett, 1925)

n N n N n N

2 1.128379 12 3.258457 100 5.0152


3 1.692569 13 3.335982 200 5.492108
4 2.058751 14 3.406765 300 5.755566
5 2.325929 15 3.471828 400 5.936396
6 2.534413 16 3.531984 500 6.073445
7 2.704357 17 3.587886 600 6.183457
8 2.847201 18 3.640066 700 6.275154
9 2.970027 19 3.688965 800 6.353645
10 3.077506 20 3.734952 900 6.422179
11 3.172874 50 4.498153 1,000 6.482942

Note: n 5 number of cone resistance data points contained within the upper and lower bound lines (see Figure 3.3). For intermediate values of n,
the value of N may be obtained by linear interpolation.

   
For circular footings, the sq and sc equations should Nc ~ Nq {0:6 tan 1:33p ðEq: 3:48Þ
be multiplied by an additional term equal to 1 +
0.0025p and 1 + 0.002p, respectively.
vii. Compute the limit unit bearing capacity qbL of the
v. Estimate the depth factor dq using (Lyamin
footing using (Lyamin et al., 2007):
et al., 2007):
     
  D {0:27 qbL ~ sq dq q0 Nq z0:5 sc dc cBNc ðEq: 3:49Þ
dq ~1z 0:0036p z0:393 ðEq: 3:46Þ
B
where q0 5 surcharge (vertical effective stress)
at the footing base level, and dc 5 depth factor
vi. Calculate the bearing capacity factors Nq and (5 1). For strip footings, the shape factors sq and
Nc using (Loukidis & Salgado, 2011; Reissner, sc are equal to 1. Note that additional factors
1924): would have to be added to the bearing capacity
equation (Eq. 3.49) to account for load inclination,
1z sin p p tan p footing base inclination, and ground inclination, as
Nq ~ e ðEq: 3:47Þ
1{ sin p needed.

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 33


b. Limit unit bearing capacity of footings in ‘‘clay.’’ Execute Step 1: Obtain the nominal dead load DLn and the
the following substeps for footings in ‘‘clay,’’ otherwise nominal live load LLn on the footing from the
proceed to step 3. superstructure design.
i. Determine the undrained shear strength su profile Step 2: Set the load factors for dead load and
below the footing base from CPT results using live load, LFDL and LFLL, as 1.25 and 1.75, respec-
(Foye et al., 2006a,b; Salgado, 2008): tively (AASHTO, 2020). These load factors correspond
to the Strength I limit state (basic load combination
qc,CAM ðzÞ{sv0 ðzÞ relating to the normal vehicular use of the bridge
su ðzÞ~ ðEq: 3:50Þ
Nk without wind), as defined by AASHTO (2020). The
where qc,CAM(z) 5 conservatively assessed mean discussion of other limit states, such as Strength II–V,
(CAM) cone resistance (as a function of depth z) Extreme Event I and II, Service I–IV, and Fatigue I and
corrected for pore water pressure u2, v0(z) 5 in situ II are beyond the scope of the manual—information
vertical total stress (as a function of depth z), and about these limit states can be found in AASHTO
Nk 5 cone factor (< 9–15 as long as the CPT is per- (2020).
formed at a penetration rate that is sufficiently high Step 3: Calculate the nominal resistance Rn of the
to ensure undrained penetration (refer to Appendix footing using:
D); soft NC clays tend to have Nk values near the
low end of this range, while stiff OC clays tend to Rn ~qbL,net A ðEq: 3:54Þ
have Nk values near the high end of this range).
ii. Using Eq. 3.50, determine the strength gradient where qbL,net 5 net limit unit bearing capacity of the
r with depth and the undrained shear strength su0 at footing (5 qbL – q0), qbL 5 limit unit bearing capacity
the footing base level. of the footing (obtained from Section 3.2), q0 5
iii. Determine the correction factor F from Figure 3.4
surcharge at the footing base level, and A 5 area of
based on whether the su profile below the footing
base resembles profile 1 or profile 2. Profile 1
the footing base.
represents an NC clay deposit with su increasing Step 4: Obtain the resistance factor.
linearly with depth from a nonzero value su0 at the Table 3.5 summarizes the resistance factors for load
footing base level. Profile 2 represents an NC clay and resistance factor design of footings using the
deposit below a certain depth, with the footing base bearing capacity equations (Eqs. 3.49 and 3.53)
resting on an OC crust for which su is constant with presented in this chapter, while Table 3.6 summarizes
depth; zf 5 depth measured from the footing base. the resistance factors and footing design methods
iv. Estimate the shape factor ssu and depth factor dsu advocated by AASHTO (2020).
using (Salgado, 2008; Salgado et al., 2004):
Step 5: Verify that the following LRFD inequality is
8 9
> > satisfied (Foye et al., 2006b; Salgado, 2008):
>
> >
>
>
< >
=
B 2:3
ssu ~1zC1 "   # {1:3 ðRF ÞRn §LFDL DLn zLFLL LLn ðEq: 3:55Þ
L>> rB 0:509 >
>
>
> >
>
:exp 0:353 ;
s
u0
rffiffiffiffi If Eq. 3.55 is satisfied, the footing design is
D satisfactory with respect to the ultimate limit state
zC2 ðEq: 3:51Þ
B (i.e., classical bearing capacity failure). Repeat steps 3
rffiffiffiffi to 5 to optimize the design if needed. However, if
D Eq. 3.55 is not satisfied, return to step 3 and revise the
dsu ~1z0:27 ðEq: 3:52Þ
B footing geometry.
Note: The following equation may be used, if needed,
where B 5 footing width, L 5 footing length, and
to obtain an equivalent factor of safety (FS) for
C1 and C2 5 coefficients that depend on the aspect
ratio B/L of the footing (Table 3.4).
the footing design produced using LRFD (Salgado,
v. Compute the limit unit bearing capacity qbL of the 2008):
footing using (Salgado, 2008):

LLn
LFDL zLFLL DL n
rB FS~bR
ðEq: 3:56Þ
qbL ~Fssu dsu 1z su0 Nc zq0 ðEq: 3:53Þ LLn
z1 RF
4su0 Nc DLn

where Nc 5 bearing capacity factor (5 2 + p <


5.14) (Prandtl, 1920, 1921), and q0 5 surcharge
where bR 5 bias factor (5 R/Rn), R 5 mean resistance
(vertical total stress) at the footing base level.
of the footing (calculated from qbL using the mean cone
3.3 Load and Resistance Factor Design Procedure for resistance profile (Figure 3.3)), and Rn 5 nominal
Footings resistance of the footing (calculated from qbL using the
conservatively assessed mean cone resistance qc,CAM
Load and resistance factor design (LRFD) of axially- obtained from Eq. 3.37). To obtain a quick estimate of
loaded footings can be done from CPT results by the equivalent factor of safety, the value of the bias
following these steps. factor bR may be taken as 1.

34 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


Figure 3.4 F versus rB/su0 for a rough footing base in clay.

TABLE 3.4 from the mean effective stress, plasticity index, and
Values of C1 and C2 to use in Eq. 3.51 as a function of B/L OCR.
(Salgado, 2008; Salgado et al., 2004) The method for estimation of primary consolidation
settlement of a footing in clay is basically a modifica-
B/L C1 C2
tion of that used to estimate the one-dimensional
1 (circle) 0.163 0.210 consolidation settlement caused by the application of
1 (square) 0.125 0.219 an instantaneous uniform load extending to infinity
0.50 0.156 0.173 horizontally; the modification accounts for the three-
0.33 0.159 0.137 dimensional effects that arise due to the finite size of the
0.25 0.172 0.110 footing. In this method, the main soil variables are
0.20 0.190 0.090
initial void ratio, compression index, swelling index,
and preconsolidation stress. If significant secondary
consolidation is expected at the site, it should be
considered together with primary consolidation.
3.4 Chapter Summary The limit unit bearing capacity of a footing in clay is
calculated assuming that the loads are applied rapidly
In this chapter, detailed, step-by-step procedures for compared to the drainage rate of clay and that the short
computing the total settlement w and limit unit bearing term is the critical loading condition; therefore, loading
capacity qbL of axially-loaded footings from CPT takes place under undrained conditions. In contrast, the
results in sand (silica sand) and clay were presented. limit unit bearing capacity of a footing in sand is
Guidelines for footings installed in mixed or inter- calculated assuming drained conditions. The main soil
mediate soils, such as sand-silt or sand-clay mixtures, variable in the bearing capacity equation is the peak
were provided based on the concept of floating versus friction angle in the case of sand and the undrained
nonfloating soil fabric. shear strength in the case of clay. The undrained shear
Methods for estimation of immediate settlement of strength su can be estimated from CPT results through
footings in sand and clay require a representative value the cone factor Nk, which typically ranges from 9–15
of the elastic modulus of the soil below the footing depending on soil type, stress state and history, and
under drained and undrained conditions, respectively. stress path (e.g., triaxial compression versus direct
For sands, the ratio of the elastic modulus to the simple shear).
cone resistance is a function of footing settlement Load and resistance factor design (LRFD) proce-
level, footing size, and relative density. For clays, the dures for footings in sand and clay were presented. The
elastic modulus is obtained through the small-strain nominal resistance of the footing is calculated through
shear modulus, which can be estimated either from the a nominal value of cone resistance, which is defined as a
shear wave velocity (if SCPT results are available) or conservatively assessed mean (CAM) value that is

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 35


TABLE 3.5 TABLE 3.6
Resistance factors for footings (D/B # 1) in sand and clay Resistance factors for footings in sand and clay (AASHTO, 2020)
(modified from Foye et al., 2006b)
Method/Soil/Condition RF
RF [bT 5 3.0 (pf,T < 10–3)]
Theoretical method (Munfakh et al., 2001) for 0.50
Footing Type Sand Clay
footings in clay
Strip footing 0.25 0.70 Theoretical method (Munfakh et al., 2001) for 0.50
Rectangular footing 0.35 0.75 footings in sand using CPT
Semi-empirical methods (Meyerhof, 1956) for 0.45
Note: bT 5 target reliability index and pf,T 5 target probability of footings in sand and clay
failure (a value of 10–3 means that one in every 1,000 footings would Plate load test 0.55
fail). The resistance factors were developed by Foye et al. (2006b)
using reliability analysis and they correspond to the CPT-based Note: The resistance factors were developed using both reliability
footing design methods covered in this chapter. The RF values for theory and calibration by fitting to working stress design (WSD)
rectangular footings may also be used for square and circular footings. (Allen, 2005). In general, WSD safety factors for footing bearing
capacity range from 2.5 to 3.0, corresponding to a resistance factor
of about 0.55 to 0.45, respectively (AASHTO, 2020). According to
exceeded by 80% of the measured qc data points. The AASHTO (2020), calibration by fitting to WSD controlled the
value of qc,CAM depends on the standard deviation selection of the resistance factor when limited statistical data were
of qc, which is estimated from the range of qc values available.
(i.e., the difference between the maximum and mini-
mum values of qc) contained within the sample dataset.
This difference is related to the number of standard prepared so that the methods can be easily referred to
deviations of qc, which is a function of the sample size. when needed. The design methods covered in this
When using LRFD, it is important to note that the chapter are not mandatory for design in INDOT
resistance factors are always tied to the specific design contracts, and other CPT-based methods, some of
methods and equations for which they were developed. which are summarized in Table 3.7 to Table 3.10, may
Finally, summary tables for the CPT-based footing be used as deemed appropriate for the site and loading
design methods covered in this chapter have been conditions under consideration.

36 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


TABLE 3.7
Methods for estimation of footing settlement in sand

Reference Total Settlement w Notes

n
 
Lee & Salgado (2002) 0 

P Izi Dzi Schmertmann’s method was modified by Lee and Salgado (2002) and Lee
w~C1 C2 qb {sv0 z ~0
Lee et al. (2008) f Ei et al. (2008) based on results obtained from nonlinear finite element
i~1
0  !
Schmertmann (1970)   analyses and cavity expansion analyses (using the program CONPOINT)
sv0 zf ~0 t
Schmertmann et al. (1978) C1 ~1{0:5 0
 ; C2 ~1z0:2 log for isolated and strip footings (B 5 1–3 m) on silica sand (DR 5 30%–90%).
qb {sv0 z ~0 0:1tR
f It accounts for the effects of aging and overconsolidation of sand on the
8 zf  
>
< Iz0 z Izp {Iz0 for zf vzfp estimation of a representative elastic modulus (from cone resistance) within
z
Iz ~ zf 0 {zfp f
the zone of influence of the footing.
>
: Izp for zfp ƒzf ƒzf 0 The method captures the nonlinearity of the footing load-settlement curve
zf 0 {zfp
zf 0

L

zfp

L
 caused by the degradation of the elastic modulus of sand with increasing
~2z0:4 min ; 6 {1 ; ~0:5z0:1 min ; 6 {1 footing settlement level.
B B B B
 
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u 0
 ffi It can be used to calculate from the cone resistance either the load on a given
L u
t
qb {sv0 z ~0
f footing or the area of the footing for a given load corresponding to a user-
Iz0 ~ min 0:1z0:0111 {1 ; 0:2 ; Izp ~0:5z0:1 0

B s  v0 zf ~zfp defined tolerable settlement.
 {0:285  0:4   The value of l is equal to 0.38 for young NC silica sand, 0.53 for aged NC
E w B DR {0:65
~l silica sand, and 0.91 for OC silica sand.
qc LR LR 100

n
 
AASHTO (2020) 0 

P Izi Dzi The equation, originally proposed by Schmertmann et al. (1978), has been
~0
w~C1 C2 qb {sv0 z
Schmertmann et al. (1978) f rewritten by AASHTO (2020) in a way that requires specific units for
i~1 144XEi
E~0:028qc certain variables: z in ft; qc in ksi; and qb and 9v0 in ksf.
X51.25 for L/B 5 1, 1.75 for L/B $ 10, and a linearly-interpolated The parameters zf0, Iz0, zfp, and Iz can be determined from the strain influence
value for L/B between 1 and 10. The equations for C1, C2, and diagrams provided in either Schmertmann et al. (1978), Salgado (2008), or
Izp are the same as in the method above. AASHTO (2020).

Mayne et al. (2012) "  0:345 #2 The method was developed by fitting an equation to a database of footing
1 qb L
Mayne & Dasenbrock (2018) w~B load test results (122 footings on noncalcareous sands) after normalizing
hs qt,net B
MnDOT (Dagger et al., 2018) the unit load and settlement of the footing with respect to cone resistance
hs50.58 for clean sand and qt,net 5 qt – v0; where qt 5 qc + (1 – a)u2 and footing size, respectively.
qt,net is an average net cone resistance measured over a vertical distance The equation is applicable for L/B 5 1–23, D/B 5 0–2.2, and qc 5 0.9–21.6

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


of 1.5B below the footing base. For clean sand, the correction of MPa.
qc to qt is negligible and since overburden stresses are small, particularly (L/B)0.345 is an influence factor for rectangular footings based on the elasticity
for shallow foundations, qt,net < qt < qc. theory solution by Giroud (1968).
For silts that experience drained loading with no excess pore water pressures
being developed, hs 5 1.12. For mixed or intermediate soils, hs can be
determined from the soil behavior type index Ic, as illustrated in Dagger et al.
(2018).

(Continued)

37
38
TABLE 3.7
(Continued)

Reference Total Settlement w Notes


Gavin et al. (2009) w~wi zwc ; where The method approximates the shape of the footing load-settlement curve by
qb B an initial linear component (with no modulus degradation) followed by a
wi ~
k nonlinear (parabolic) component up to wi/B 5 10%.
for the linear stage (0 # wi/B # wy/B), and The value of the exponent n was determined by equating the value of qb
2 31n
obtained from the equation for the nonlinear stage (at wi/B 5 10%) with a
6 qb 7 value of 0.2qcb.
wi ~B4
1{n 5
w The effect of creep is modeled using:
k By
   
for the nonlinear stage (wy/B , wi/B , 10% and wy/B $ 0.03%) wc t qb 2
~m ln ; where m~0:02
2 3 B tR qb,ult
6 qcb 7 wc 5 creep component of settlement
1:61{ ln4 wy
5
k m 5 creep coefficient
4E0
n~ wB
and k~ t 5 time elapsed since the application of the load increment
y pð1{n0 2 Þ
2:30z ln tR 5 reference time corresponding to the onset of creep settlement (in the
B
qcb is the average cone resistance over the depth of influence zf0 below the same unit as t)
footing base, which is given by: qb,ult 5 value of qb at wi/B 5 10% (< 0.2qcb)
 0:75
zf 0 B
~
LR LR

Lehane (2019) qb 5 0.05qcb for a short-term relative settlement w/B of 1% The equations are based on centrifuge and field load test results of footings in
Liu & Lehane (2021) qb 5 0.04qcb for a long-term relative settlement w/B of 1% sand.
qcb is the average cone resistance over the depth of influence zf0 below the Short-term and long-term (creep) settlement refer to the settlement observed
footing base, which is given by: zf0/LR 5 (B/LR)0.7 about 1 day and 30 years, respectively, after the application of the load to
the footing.
0 
Note: C1 and C2 5 depth and time factors, respectively; qb 5 unit load on the footing base; sv0 z ~0 5 in situ vertical effective stress at the footing base level; Iz 5 strain influence factor; Dz 5
f
thickness of sublayer; E 5 elastic modulus; n 5 number of sublayers; tR 5 reference time; t 5 service life of the superstructure (in the same unit as tR); zf0 5 influence depth measured from the
footing base; zf 5 vertical distance from the footing base to the middle of the sublayer; Iz0 5 strain influence factor at the footing base level; Izp 5 peak strain influence factor; zfp 5 depth
0 
measured from the footing base at which the strain influence factor peaks; sv0 zf ~zf p 5 in situ vertical effective stress at the depth corresponding to zfp; qc 5 cone resistance; a 5 cone area ratio
(< 0.8 for typical CPT probes); u2 5 pore water pressure measured at the shoulder position behind the cone face; LR 5 reference length (5 1 m or 3.28 ft); L 5 footing length; B 5 width or
diameter of the footing; DR 5 relative density (expressed as a percentage); E0 5 small-strain Young’s modulus [5 2G0(1+n0)]; G0 5 small-strain shear modulus; n0 5 small-strain Poisson’s ratio
(5 0.1–0.2); and wy/B 5 normalized yield settlement level (< 0.03%).

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


TABLE 3.8
Methods for estimation of bearing capacity of footings in sand

Reference Bearing Capacity Notes


   
Bolton (1986) qbL ~ sq dq q0 Nq z0:5 sc dc cBNc The equations for the shape factors sq and sc and depth factor dq are based on the
Lyamin et al. (2007) 8
c if z vD results of rigorous lower- and upper-bound limit analyses of circular,
>
< b zw {D
Loukidis & Salgado (2011) w rectangular, and strip footings in sand. For a strip footing (B/L 5 0) placed on
c~ cb z ðcm {cb Þ if Dƒzw ƒDzB
Salgado (2008) >
: B the surface (D/B 5 0) of a sand deposit, the shape and depth factors reduce to
cm if zw wDzB a value of 1.
    D
  D 0:7{0:01p B 1{0:16ðB Þ For circular footings, the sq and sc equations should be multiplied by an
sq ~1z 0:098p {1:64
B L additional term equal to 1 + 0.0025p and 1 + 0.002p, respectively.
 B The equation for the bearing capacity factor Nc fits almost perfectly the exact
sc ~1z 0:0336p {1
L values of Nc obtained by Martin (2005) using the method of characteristics or
 
  D {0:27 slip-line method, even for very low values of friction angle.
dq ~1z 0:0036p z0:393 and dc 5 1
B The parameter A is equal to 3 for triaxial conditions (e.g., square and circular
1z sin p p tan p     footings with L/B 5 1) and 5 for plane-strain conditions (e.g., elongated
Nq ~ e and Nc ~ Nq {0:6 tan 1:33p
1{ sin p rectangular footings with L/B $ 7 and strip footings). For rectangular footings
0
( " !# )  
DR 100smp 1 L
p ~c zAy Q{ ln {RQ with 1 , L/B , 7, A is interpolated between 3 and 5 using Ay ~ z8 .
100 pA 3 B
 0:7   The representative mean effective stress 9mp obtained from the equation is not a
0 cB B
smp ~20pA 1{0:32 mean stress around the footing but rather a value of mean stress that will lead
pA L to the value of limit unit resistance qbL. The equation for 9mp accounts for soil
   0 
qc,CAM s anisotropy, for the difference between the friction and dilatancy angles, and for
ln {0:4947{0:1041c {0:841 ln h0
pA p the evolution of soil properties during loading toward limit bearing capacity
DR ð%Þ~  0  A
sh0 failure.
0:0264{0:0002c {0:0047 ln
pA
 
Meyerhof (1956) qcb B D This is an empirical method with B and D in the units of ft and qcb in ksf.
qbL ðksf Þ~ Cwq zCwc
AASHTO (2020) 40 B AASHTO (2020) provides the values of the water table correction factors Cwq
qcb is an average cone resistance measured over a vertical and Cwc in the form of a table. Using those values, we derived the following
distance of B below the footing base. easy-to-use equations for Cwq and Cwc:
8
h z
i < 0:5 for
w
zw vD  
1 zw {D

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


Cwq ~min 0:5z0:5 ; 1 and Cwc ~
D : min 0:5z ; 1 for zw §D
3 B

Mayne & Woeller (2014) w


0:5 L{0:345 The method was developed by fitting an equation to a database of footing load
Mayne & Dasenbrock (2018) qb,ult ~hs qt,net test results (31 footings in 13 silica sands and 11 footings in 4 silt deposits) after
B max B
MnDOT (Dagger et al., 2018) hs 5 0.58 and (w/B)max 5 12% for clean sand normalizing the unit load and settlement of the footing with respect to cone
qt,net 5 qt – v0; where qt 5 qc + (1 – a)u2 resistance and footing size, respectively.
qt,net is an average net cone resistance measured over a The footing geometries consisted of 29 square, 7 rectangular (nearly square), and
vertical distance of 1.5B below the footing base. For 6 circular footings with B 5 0.5–6.1 m and D 5 0–2.35 m.
clean sand, the correction of qc to qt is negligible and since The sands were of different age (recent, Holocene, Pleistocene) and geologic
overburden stresses are small, particularly for shallow origin (alluvial, marine, glaciofluvial, deltaic, aeolian, and residual) with D50 5
foundations, qt,net < qt < qc. 0.1–0.4 mm and qc 5 0.9–10.7 MPa.
For silts that experience drained loading with no excess pore water pressures
being developed, hs 5 1.12 and (w/B)max 5 10%. For mixed or intermediate
soils, hs can be determined from the soil behavior type index Ic, and (w/B)max
can be interpolated based on the value of hs.

(Continued)

39
40
TABLE 3.8
(Continued)

Reference Bearing Capacity Notes


Lehane (2019) qb,ult ~aqcb Lehane (2013) compiled a database of load test results for 47 footings in sand
a 5 0.16 for 10% relative settlement with B 5 0.25–3 m, D 5 0.1–1.6 m, and qcb 5 3.5–14.5 MPa. The equation for
qcb is the average cone resistance over the depth of influence zf0 qb,ult predicts 80% of the footing load test results to within 25% of the
below the footing base, which is given by: zf0/LR 5 (B/LR)0.7 measured qb,ult values at 10% relative settlement.
Liu and Lehane (2021) proposed the following equation to obtain lower-bound
estimates of qb,ult corresponding to a long-term (creep) w/B ratio of 10%: qb,ult
(long term) 5 0.1qcb

Gavin et al. (2009) qb,ult ~aqcb The value of a was determined based on the observed load-settlement response
a < 0.2 for 10% relative settlement (qb/qcb versus w/B) of model and full-scale, square footings in sand.
qcb is the average cone resistance over the depth of influence zf0
below the footing base, which is given by: zf0/LR 5 (B/LR)0.75
Lee & Salgado (2005) qb,ult ~aqcb The method was developed based on results obtained from nonlinear finite
qcb is an average cone resistance measured over a vertical distance element analyses and cavity expansion analyses (using the program
of B below the footing base. CONPOINT) for circular footings (B 5 1–3 m) on Ottawa sand
The value of a for 20% relative settlement can be obtained from (DR 5 30%–90%).
the table provided by Lee and Salgado (2005) as a function of The equation is also applicable for square footings so long as an equivalent area
DR, K0, and B. is considered. However, for footing shapes other than circular or square,
introduction of shape factors would be required.

Note: qbL 5 limit unit bearing capacity (i.e., the unit load at which the footing plunges into the ground), q0 5 surcharge (vertical effective stress) at the footing base level, c 5 unit weight of soil
below the footing base, B 5 footing width, sq and sc 5 shape factors, dq and dc 5 depth factors, Nq and Nc 5 bearing capacity factors, zw 5 depth of the groundwater table, cm 5 moist unit weight
of soil, cb 5 buoyant unit weight of soil, p 5 peak friction angle, L 5 footing length, D 5 depth of embedment of the footing, c 5 critical-state friction angle, DR 5 relative density, Q and RQ 5
fitting parameters that depend on the intrinsic characteristics of sand (Q 5 10 and RQ 5 1 for clean silica sand), qc,CAM 5 conservatively assessed mean (CAM) cone resistance at a depth of B/2
below the footing base (Eq. 3.37), 9v0 5 in situ vertical effective stress at a depth of B/2 below the footing base, 9h0 5 in situ horizontal effective stress at a depth of B/2 below the footing base (5
K09v0), K0 5 coefficient of lateral earth pressure at-rest, pA 5 reference stress (5 100 kPa or 14.5 psi), qb,ult 5 ultimate unit bearing capacity (mobilized at a given relative settlement w/B), w 5
footing settlement, a 5 cone area ratio (< 0.8 for typical CPT probes), u2 5 pore water pressure measured at the shoulder position behind the cone face, and LR 5 reference length (5 1 m or 3.28
ft).

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


TABLE 3.9
Methods for estimation of footing settlement in clay (Foye et al., 2008; Salgado, 2008; Skempton & Bjerrum, 1957)

Immediate Settlement wi Notes

qb,net B The method was developed based on results obtained from nonlinear finite element analyses of strip,
wi ~Iq
E
0 square, and rectangular footings on clay.
E
 0 ~2ð1znÞG  qb,net
0   The influence factor Iq is determined from Figure 3.2 as a function of H/B, , and footing geometry;
zG 0 L su
~ min 1z0:111 {1 ; 2 where qb,net 5 qb – cmD and su 5 average undrained shear strength over a vertical distance of B below
B B
n the footing base.
P avg
G0,i Hi The undrained shear strength profile can be obtained from CPT results using su 5 [(qt – v0)/Nk]; where
G
 0 ~ i~1
n qt 5 qc + (1 – a)u2 and Nk 5 cone factor (< 9–15; soft NC clays tend to have Nk values near the low end
P
Hi of this range, while stiff OC clays tend to have Nk values near the high end of this range).
i~1
cm 2 The small-strain shear modulus G0 is averaged over the influence depth zG 0 below the footing base to
G0 ~ V (from seismic cone penetration test data), or
g s obtain G 0 ; where Gavg 5 average small-strain shear modulus of layer i, Hi 5 thickness of layer i, and
0,i
 0 ng
G0 100sm0 m n 5 number of clay layers within the influence depth zG 0 below the footing base.
~Cg R0 g
pA pA Parameters Cg, ng, and mg depend on the plasticity index PI (Viggiani & Atkinson, 1995):
0 0
1 0

sm0 ~ sv0 zksh0 Cg ~37:9 expð{0:045PIÞ for PIw5%


kz1 ng ~0:109 lnðPIÞz0:4374 for PIw5%
0
!
pp 1z2K0,NC mg ~0:0015PIz0:1863 for PIw5%
R0 ~ 0 ~OCR p ffiffiffiffiffiffiffiffiffiffiffi
p 1z2K0,NC OCR
Consolidation Settlement wc Notes
n
X The coefficient a is determined from Table 3.2 as a function of H/B and footing geometry.
wc ~½Azað1{AÞwc1D ; where wc1D ~ Dez,i Dzi
The initial void ratio can be obtained using the relation e0 5 wcGs/S; where wc 5 water content, Gs 5
i~1
8  0  specific gravity of solids (5 2.60–2.80 for clay), and S 5 degree of saturation. In the absence of soil
> Cc sv
>
>
> log 0 for NC clay samples, the reader may refer to Section 2.3.1 of Volume I for additional information on e0.
> sv0
0 0 In the absence of laboratory consolidation test results, the compression index may be estimated using the
>
> 1ze0 "
> ! !#
< 1 svp s

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


Dez ~ Cs log 0 zCc log 0 v for OC then NC clay correlation of Wroth and Wood (1978): Cc 5 GsPI(%)/200.
> 1ze0 sv0 svp
>
>
>  0  The swelling index Cs typically ranges from 0.1Cc to 0.2Cc.
>
> Cs s
>
>
: log 0 v for OC clay The vertical stress increment Dv caused by the footing load can be calculated using the 2-to-1 stress
1ze0 sv0 distribution rule.
0 0
sv ~sv0 zDsv

Note: Iq 5 influence factor; qb,net 5 net unit load on the footing base; qb 5 gross unit load on the footing base; H 5 thickness of the clay layer below the footing base; B 5 footing width; L 5
footing length; D 5 depth of embedment of the footing; qt 5 corrected, total cone resistance measured under undrained conditions; qc 5 cone resistance; a 5 cone area ratio (< 0.8 for typical
CPT probes); u2 5 pore water pressure measured at the shoulder position behind the cone face; v0 5 in situ vertical total stress at the depth being considered; E  0 5 representative small-strain
Young’s modulus; n 5 Poisson’s ratio (5 0.5 for undrained conditions); cm 5 unit weight of soil; g 5 acceleration due to gravity (9.81 m/s2 or 32.17 ft/s2); Vs 5 shear wave velocity (refer to Section
2.3.4 of Volume I); pA 5 reference stress (5 100 kPa or 14.5 psi); s9m0 5 in situ mean effective stress at the depth being considered; s9h0 5 in situ horizontal effective stress at the depth being
considered (5 K0s9v0); K0 5 coefficient of lateral earth pressure at-rest; k 5 1 for plane-strain conditions (e.g., strip footings) and 2 for triaxial conditions (e.g., isolated footings); R0 5 mean stress-
based overconsolidation ratio; K0,NC 5 coefficient of lateral earth pressure at-rest for normally consolidated soil (< 0.5–0.75 for NC clay); OCR 5 overconsolidation ratio; wc1D 5 one-
dimensional consolidation settlement; A 5 Skempton’s pore pressure parameter (< 0.5–0.75 for NC clay and 0.3–0.5 for OC clay); s9v0 and s9v 5 initial (in situ) and current vertical effective
stresses, respectively, at the depth being considered; Dsv 5 vertical stress increment; s9vp 5 preconsolidation stress; e0 5 initial void ratio; Dz 5 thickness of the sublayer; n 5 number of sublayers;
and Dez 5 vertical compressive strain.

41
42
TABLE 3.10
Methods for estimation of bearing capacity of footings in clay

Reference Bearing Capacity Notes



Davis & Booker (1973) rB The equation for the shape factor ssu is based on a least-squares fit to the
qbL ~Fssu dsu 1z su0 Nc zq0
Salgado et al. (2004) 4su0 Nc values of ssu obtained from the computer program ABC (Martin, 2004),
8 9
Salgado (2008) >
> >
> which is based on the method of characteristics.
> > rffiffiffiffi
B<
>
2:3
>
= D The equation for the depth factor dsu is based on the results of rigorous lower-
ssu ~1zC1 "  0:509 # {1:3 zC2 and upper-bound limit analyses of footings embedded in clay.
L>> rB >
> B
>exp 0:353
>
:
>
>
; The undrained shear strength profile can be obtained from CPT results using
su0
rffiffiffiffi su 5 [(qc,CAM – v0)/Nk]; where Nk 5 cone factor (< 9–15; soft NC clays
D tend to have Nk values near the low end of this range, while stiff OC clays
dsu ~1z0:27
B tend to have Nk values near the high end of this range).

Mayne & Woeller (2014) w


0:5 L{0:345 The method was developed by fitting an equation to a database of footing
MnDOT (Dagger et al., 2018) qb,ult ~hs qt,net load test results (12 footings in 6 intact clays and 11 footings in 5 fissured
B max B
hs 5 1.47 and (w/B)max 5 7% for fissured clay clays) after normalizing the unit load and settlement of the footing with
hs 5 2.70 and (w/B)max 5 4% for intact clay respect to cone resistance and footing size, respectively. The footing
qt,net 5 qt – v0; where qt 5 qc + (1 – a)u2 geometries consisted of 13 square, 1 rectangular, and 9 circular footings
qt,net is an average net cone resistance measured over a vertical distance of with B 5 0.4–5.0 m.
1.5B below the footing base For mixed or intermediate soils, hs can be determined from the soil behavior
type index Ic, and (w/B)max can be interpolated based on the value of hs.

Lehane (2019) qbL &0:45qcb The method is based on load-settlement data (qb/qcb versus w/B) compiled
qb,all &0:25qcb for 1% relative settlement from undrained footing load tests in 5 clays. qcb is the average cone
resistance (corrected for pore water pressure u2) over a vertical distance of
B below the footing base.

Note: qbL 5 limit unit bearing capacity (i.e., the unit load at which the footing plunges into the ground), F 5 correction factor (Figure 3.4), ssu 5 shape factor, dsu 5 depth factor, qc,CAM 5
conservatively assessed mean (CAM) cone resistance (corrected for pore water pressure u2), r 5 rate of increase of undrained shear strength su with depth, su0 5 undrained shear strength at the
footing base level, Nc 5 bearing capacity factor (5 2 + p < 5.14), q0 5 surcharge (vertical total stress) at the footing base level, B 5 footing width, L 5 footing length, C1 and C2 5 coefficients
that depend on the aspect ratio B/L of the footing (Table 3.4), qb,ult 5 ultimate unit bearing capacity (mobilized at a given relative settlement w/B), qb,all 5 allowable bearing pressure, qt 5
corrected, total cone resistance measured under undrained conditions, qc 5 cone resistance, a 5 cone area ratio (< 0.8 for typical CPT probes), u2 5 pore water pressure measured at the shoulder
position behind the cone face, v0 5 in situ vertical total stress at the depth being considered, and LR 5 reference length (5 1 m or 3.28 ft).

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


4. CPT-BASED DESIGN OF PILE FOUNDATIONS plasticity index PI $ 8%, otherwise classify them as
‘‘sand’’ (Carraro et al., 2009; Salgado et al., 2000).
Piles can be classified into three categories based on b. Sands containing gravel: If a site contains sand layers
the changes caused to the state of in situ soil during with gravel content greater than 20%, use the lower-
their installation: (1) nondisplacement piles (e.g., drilled bound profile of qc, drawn approximately through the
shafts), (2) partial-displacement piles (e.g., H-piles and valleys of the actual qc profile, for estimating the pile
open-ended pipe (OEP) piles), and (3) full-displacement capacity (Ganju et al., 2020; Han et al., 2019b, 2020).
piles (e.g., closed-ended pipe (CEP) piles). A pile derives Note: In the absence of soil samples, the reader may
its load-carrying capacity by two mechanisms: (a) shaft refer to Section 2.2 of Volume I for estimation of soil
resistance, which is the friction or adhesion along the behavior type from CPT results.
pile shaft with the surrounding soil, and (b) base Step 4: Correct the raw qc data for the pore water
resistance, which is the compressive resistance at the pressure generated during cone penetration using
contact of the pile base with the underlying soil. Shaft (ASTM, 2012):
resistance is fully mobilized for small pile head settle-
ments (on the order of 0.25%–1% of the pile diameter),
whereas complete mobilization of pile base resistance qt ~qc zð1{aÞu2 ðEq: 4:1Þ
requires large pile head settlements (on the order of
15%–25% of the pile diameter) (Salgado, 2008). where qt 5 corrected, total cone resistance, qc 5
measured cone resistance, a 5 cone area ratio (< 0.8
4.1 Calculation Procedure for Limit Shaft Capacity of for typical CPT probes), and u2 5 pore water pressure
Single Piles measured at the shoulder position behind the cone face.
The pore water pressure correction to the qc data may
The limit shaft capacity QsL of a single, isolated, be ignored for coarse-grained soils (e.g., sand and
axially-loaded pile can be calculated from CPT results gravel) because qt is approximately equal to qc in such
by following these steps. soils.
Step 1: Obtain the site stratigraphy, the groundwater Step 5: Using the cone resistance values obtained
table depth, and the unit weight of the soil in each layer from step 4, divide the soil layers in contact with the
of the profile. pile shaft into sublayers, as shown in Figure 4.1. The
sublayers should satisfy the following criteria.
a. Establish the site stratigraphy either from the boring log
or by using a CPT-based soil behavior type (SBT) chart a. The cone resistance values within each sublayer should
(refer to Section 2.2.3 of Volume I) or both if possible. be either approximately constant or linear with depth
b. Obtain the depth zw of the groundwater table from either so that a representative cone resistance, indicated by the
the boring log or the depth profile of u2 or both if possible, grey vertical bars in Figure 4.1, can be assigned to each
where u2 5 pore water pressure measured at the shoulder sublayer.
position behind the cone face (refer to Volume I). b. The sublayer should consist of the same soil type, i.e.,
c. Obtain the unit weight of the soil in each layer of either ‘‘sand’’ or ‘‘clay.’’
the profile whenever soil samples are recovered during
the site investigation. In the absence of soil samples, the Step 6: Calculate the in situ vertical effective stress
reader may refer to Section 2.3.3 of Volume I for 9v0 at the middle of each sublayer using (Terzaghi,
correlations between the unit weight and CPT data. In 1943):
general, the saturated unit weight csat of soil typically
0
ranges from 18–21 kN/m3 (115–135 pcf) for sand, 18.5– sv0 ~sv0 {u0 ðEq: 4:2Þ
22.5 kN/m3 (118–143 pcf) for silty sand, and 15–18 kN/
m3 (95–115 pcf) for clay (Salgado, 2008). where sv0 5 in situ vertical total stress at the middle of
the sublayer, u0 5 hydrostatic pore water pressure at
Step 2: Select the pile type and decide the pile length. the middle of the sublayer {5 max[cw(z – zw) ; 0]}, cw 5
a. Set the pile type and the embedment length L of the pile unit weight of water (5 9.81 kN/m3 or 62.45 pcf), z 5
based on the soil profile at the site. depth measured from the ground surface to the middle
b. If a competent bearing layer, such as dense sand, stiff of the sublayer, and zw 5 depth of the groundwater
clay, or rock, exists at a reasonable depth from the table.
ground surface, embed the pile base in the bearing layer Step 7: Calculate the limit unit shaft resistance of pile
to ensure that the contribution of that layer toward the segments in contact with ‘‘sand’’ sublayers. Execute the
base resistance can be realized. following substeps if the sublayer is ‘‘sand,’’ otherwise
proceed to step 8.
Step 3: Classify the soil in each layer that is in contact
with the pile as either ‘‘sand’’ or ‘‘clay.’’ For mixed or a. Calculate the in situ horizontal effective stress h9 0 (5
intermediate soils (i.e., soils containing mixtures of K09v0) at the middle of the sublayer, where K0 5
sand, silt, and clay), execute the following substeps. coefficient of lateral earth pressure at-rest (refer to
Appendix B for guidance).
a. Sand-silt, sand-clay or sand-silt-clay mixtures: Classify b. Determine the critical-state friction angle c of the
these soils as ‘‘clay’’ if fines content FC $ 20% and sublayer through one of the following options.

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 43


Figure 4.1 CPT-based discretization of soil profile for shaft resistance calculation and averaging of cone resistance for base
resistance calculation (after Salgado, 2008).

i. Select a c value between 28u and 36u for silica


sand; sands with rounded, smooth particles with a
poorly-graded particle size distribution have values
near the low end of this range, while sands with
angular, rough particles with a well-graded particle
size distribution have values near the high end of
this range (refer to Appendix A for additional
information if needed).
ii. If the mean particle size D50, coefficient of uni-
formity CU, and particle roundness R of the
sublayer are known, estimate the critical-state
friction angle using:
 f
D50
c ð0 Þ~28:3 ðCU Þ2f ðRÞ{3f ðEq: 4:3Þ
Dref

where Dref 5 reference particle size (5 1 mm or


0.04 in.), and f 5 exponent (5 0.045). Equation 4.3
is applicable for poorly-graded, clean silica sands
with D50 5 0.15–2.68 mm (0.006–0.105 in.), CU 5
1.2–3.1, and R 5 0.3–0.8. The data used in the
development of this equation along with example
calculations can be found in Appendix A.
iii. If direct shear or triaxial compression test results Figure 4.2 Critical-state friction angle ratio dc/c versus
are available, it is recommended that the critical- mean particle size D50 for silica sands tested against smooth,
state friction angle be determined from such test lightly rusted, and rusted steel surfaces (Han et al., 2018,
results. 2019a). Interpolation can be used for 1.5 , CU , 2.

c. Set the critical-state interface friction angle c of the the limit unit shaft resistance qsL of the pile segment in
sublayer. contact with a sand sublayer using:
i. For precast concrete piles, set c/c 5 0.95.  0 0 
qsL ~ Fload src zDsrd tan dc ðEq: 4:4Þ
ii. For cast-in-place concrete piles, set c/c 5 1.00.
iii. For steel piles, set c/c 5 0.80–0.85. If the D50 and where Fload 5 factor that accounts for loading direction
0
CU values of the sand are known, obtain the value (5 0.8 for tension and 1.0 for compression), src 5 local
of c/c from Figure 4.2. radial effective stress acting on the pile segment after
0
installation, and Dsrd 5 increase in local radial effective
d. H-piles in ‘‘sand’’: Following the Imperial College pile stress associated with constrained dilation during pile
design method (ICPDM) (Jardine et al., 2005), compute loading:

44 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


0 2 31{0:38 Step 8: Calculate the limit unit shaft resistance of pile
 0 0:13
B 6 7C segments in contact with ‘‘clay’’ sublayers. Execute the
0 s
src ~0:029qc v0 Bmax6rhffiffiffiffiffiffi ; 87C ðEq: 4:5Þ
pA @ 4 A 5A following substeps if the sublayer is ‘‘clay,’’ otherwise
b
p proceed to step 9.
a. Select a c value between 15u and 30u for clay; high-
0
Dsrd ~2qc ½0:0203z0:00125g plasticity clays with high smectite and clay contents tend
0 1 to have values near the low end of this range, while low-
B Dr C plasticity clays with low smectite and clay contents tend
{1:216|10{6 g2 {1 B ffiC
@rffiffiffiffiffi ðEq: 4:6Þ to have values near the high end of this range (refer to
Ab A Table E.1 of Appendix E). If laboratory shear test
p results (e.g., triaxial compression) are available, it is
recommended that the critical-state friction angle be
qc determined from such test results.
pA b. Select a r,min value between 5u and 15u for clay (refer
g~ sffiffiffiffiffiffiffi ðEq: 4:7Þ
0
sv0 to Appendix E for guidance). If ring shear test results
pA are available, it is recommended that the minimum
residual-state friction angle be determined from such
test results.
where pA 5 reference stress (5 100 kPa or 14.5 psi), h 5 c. CEP piles and drilled shafts in ‘‘clay’’ (PPDM).
vertical distance from the middle of the sublayer to the
pile base, Dr 5 radial displacement of soil during pile i. Determine the undrained shear strength su of the
loading (< 0.02 mm (0.8 mil) for lightly rusted steel piles), sublayer from CPT results using (Salgado, 2008):
and Ab 5 area of the pile base (refer to Table 4.2). qt {sv0
e. Drilled shafts, CEP and OEP piles in ‘‘sand’’: Following su ~ ðEq: 4:11Þ
Nk
the Purdue pile design method (PPDM) (Han et al., 2017,
2019b), compute the limit unit shaft resistance qsL of the where qt 5 corrected, total cone resistance mea-
pile segment in contact with a sand sublayer using: sured under undrained conditions, v0 5 in situ
8 vertical total stress at the middle of the sublayer,
0
>
< Fload Ksv0 tan dc for CEP piles and Nk 5 cone factor (< 9–15 as long as the CPT is
0
qsL ~ Ksv0 tan dc for drilled shafts ðEq: 4:8Þ performed at a penetration rate that is sufficiently
>
: 0 high to ensure undrained penetration (refer to
K ð1{0:66 PLRÞ sv0 tan dc for OEP piles
Appendix D); soft NC clays tend to have Nk values
where Fload 5 factor that accounts for loading direction near the low end of this range, while stiff OC clays
(< 0.5–0.6 for tension (Galvis-Castro et al., 2019) and tend to have Nk values near the high end of this
1.0 for compression), PLR 5 plug length ratio, and K 5 range) (Bisht et al., 2021; Mayne & Peuchen, 2018;
lateral earth pressure coefficient: Salgado, 2008, 2013, 2014; Salgado et al., 2004).
ii. Following the Purdue pile design method (PPDM)
8 2 3
>   (Basu et al., 2009, 2014; Chakraborty et al., 2013),
>
> qc
>
> 6 0:01 7   compute the limit unit shaft resistance qsL of the
> 6 7
>
>
> 0:2z 6 sffiffiffiffiffiffi pAffi
{0:2 7 exp {0:14h pile segment in contact with a clay sublayer using:
>
> 6 0 7 LR
>
> 4 sh0 5
>
<
pA qsL ~asu ðEq: 4:12Þ
K~
>
> for CEP and OEP piles
>
>
>
>   0   8  0 
>
> 0:67K0 DR sv0 > s
>
> p ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi  exp 1:5{0:35 ln >
>
> A1 zð1{A1 Þ exp { v0 ðc {r, min ÞA2 for CEP piles
>
> exp 0:3 K {0:4 100 pA >
> pA
>
> 0 <
:  {0:05   0  
for drilled shafts ðEq: 4:9Þ a~ su s :
>
>
> 0 A1 zð1{A1 Þ exp { v0 ðc {r, min ÞA2
>
> sv0 pA
>
:
where pA 5 reference stress (5 100 kPa or 14.5 psi), h 5 for drilled shafts ðEq: 4:13Þ
vertical distance from the middle of the sublayer to the
pile base, LR 5 reference length (5 1 m or 3.28 ft), and 8 0
8 0
DR 5 relative density (expressed as a percentage): < 0:75 for c  r;min ƒ5
> < 0:75 for c  r;min ƒ5
>
A1 ~ 0:43 for c  r;min §12 and A1 ¼ 0:40 for c  r;min §120
0
>
: >
:
   0  for CEP piles for drilled shafts
qc s
ln {0:4947{0:1041c {0:841 ln h0
pA pA (Eq. 4.14)
DR ð%Þ~  0 
sh0 8  
0:0264{0:0002c {0:0047 ln > su
pA >
> 0:55z0:43 ln 0 for CEP piles
< sv0
A2 ~   ðEq: 4:15Þ
ðEq: 4:10Þ >
> s
> 0:40z0:30 ln 0u for drilled shafts
:
sv0
For OEP piles, the plug length ratio (PLR) used in the
equation for qsL is that measured at the specific depth
where qsL is calculated. If the PLR is not measured, it where pA 5 reference stress (5 100 kPa or 14.5 psi).
can be approximated using the same equation (Eq. 4.29) For 5u , c – r,min , 12u, obtain the value of A1 by
provided for the incremental filling ratio (IFR). interpolation.

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 45


d. OEP piles and H-piles in ‘‘clay’’ (ICPDM). residual interface friction angle using (Maksi-
mović, 1989; Salgado, 2008):
i. Obtain the overconsolidation ratio (OCR) of the c {r,min
sublayer (refer to Appendix B and Section 2.3.7 of dr &r ~r;min z 0 ðEq: 4:20Þ
s
Volume I for guidance). If laboratory consolidation 1z 0
smedian
test results (e.g., oedometer test or constant rate of
strain (CRS) test) are available, it is recommended 9 ian is the value of 9 at which the
where med
that the OCR be determined from such test results. friction angle is equal to the average of r,min and
ii. Estimate the sensitivity St of the sublayer using: c (refer to Figure E.1 in Appendix E), and 9, in
su the context of pile shaft resistance calculation, is
St ~ ðEq: 4:16Þ the horizontal effective stress 9h on the pile
sur
operative at the time of shearing:
where su 5 ‘‘undisturbed’’ or in situ undrained shear 0 0
strength of the sublayer (refer to step 8(c)(i)). The sh ~Fload Ksv0 ðEq: 4:21Þ
remolded undrained shear strength sur of the
sublayer may be estimated using the following where Fload 5 0.8 regardless of the loading
approximate correlation (Wroth, 1979): direction, and 9v0 5 initial (in situ) vertical
effective stress at the middle of the sublayer.
sur According to the data compiled by Maksimović
&0:017|102ð1{LIÞ ðEq: 4:17Þ
pA (1989), the value of 9median is in the range of 20–
150 kPa (3–22 psi) depending on the clay type
where pA 5 reference stress (5 100 kPa or 14.5 psi), and mineralogy.
LI 5 liquidity index (5 (wc – PL)/PI), wc 5 water N If results from ring shear interface tests per-
content, PI 5 plasticity index (5 LL – PL), LL 5 formed for the applicable value of normal
liquid limit, and PL 5 plastic limit. In the absence effective stress (Ramsey et al., 1998) are available,
of soil samples, the reader may refer to Sections it is recommended that the residual interface fric-
2.3.10.5 and 2.3.10.6 of Volume I for additional tion angle be determined from such test results.
information on sur and St, respectively.
iii. Estimate the lateral earth pressure coefficient K of v. Following the Imperial College pile design method
the sublayer using (Jardine et al., 2005): (ICPDM) (Jardine et al., 2005), compute the limit
unit shaft resistance qsL of the pile segment in
K~½2:2z0:016OCR{0:87 log St  contact with a clay sublayer using:
 {0:20 0
h qsL ~Fload Ksv0 tan dr ðEq: 4:22Þ
OCR0:42 max ;8 ðEq: 4:18Þ
R
8 rffiffiffiffiffiffi Step 9: Repeat steps 7 and 8 to obtain the limit unit
>
< Ab for H-piles
> shaft resistance qsL for each ‘‘sand’’ and ‘‘clay’’ sublayer
p
R~ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ðEq: 4:19Þ in contact with the pile shaft.
>
>
: R2 {R2 for OEP piles Step 10: Compute the limit shaft capacity QsL of the
o i
pile using:
where h 5 vertical distance from the middle of the P
n
sublayer to the pile base, Ro 5 outer radius of OEP QsL ~ qsLi Asi ðEq: 4:23Þ
pile, Ri 5 inner radius of OEP pile, and Ab 5 area i~1
of the pile base (Table 4.2).
where Asi 5 pile shaft area interfacing with sublayer
iv. Determine the residual interface friction angle r of
the sublayer through one of the following options.
i (Table 4.1), and n 5 number of sublayers in contact
with the pile shaft.
N Using the values of c and r,min obtained from
steps 8(a) and 8(b), respectively, estimate the TABLE 4.2
Expressions for Ab for different pile cross-sections
TABLE 4.1
Expressions for Asi for different pile cross-sections Pile Cross-Section Pile Base Area Ab

Circle (CEP) pB2/4


Pile Cross-Section Pile Shaft Area Asi
Square B2
Circle pBDzi Rectangle BwBl
Square 4BDzi H-section1 2bftf + (2Xp + tw)(d – 2tf)
Rectangle 2(Bw + Bl)Dzi
Note: B 5 pile diameter (or width in the case of a square pile); Bw
H-section 2(bf + d)Dzi
and Bl 5 width and length, respectively, of the cross-section of a
Note: B 5 pile diameter (or width in the case of a square pile); Bw rectangular pile (in plan); bf 5 width of flange; d 5 depth of H-section;
and Bl 5 width and length, respectively, of the cross-section of a tf 5 thickness of flange; and tw 5 thickness of web.
1
rectangular pile (in plan); bf 5 width of flange; d 5 depth of H-section; For H-piles, Xp 5 bf /8 if bf /2 , (d – 2tf) , bf and Xp 5 b2f /[16(d –
and Dzi 5 thickness of sublayer i. 2tf)] if (d – 2tf) $ bf (De Beer et al., 1980; Jardine et al., 2005).

46 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


4.2 Calculation Procedure for Ultimate Base Capacity of underlying weak layer (Xu, 2007); therefore, we can
Single Piles calculate the value of qcb from step 1(a). Note that
piles should be sufficiently embedded in a strong,
The ultimate base capacity Qb,ult of a single, isolated, competent layer, whenever possible, to avoid
axially-loaded pile can be calculated from CPT results serviceability issues.
by following these steps.
Step 2: Calculate the ultimate unit base resistance
Step 1: Estimate the average cone resistance qcb at the
pile base. qb,ult of the pile.
a. For piles bearing in ‘‘sand,’’ calculate the ultimate unit
a. Execute the following substeps, depending on the pile
base resistance qb,ult of the pile using (Han et al., 2017,
design method, to estimate the average cone resistance
2019b; Jardine et al., 2005; Lehane et al., 2005):
qcb at the pile base.
8
>
> qcb for H-piles (ICPDM)
i. For the Purdue pile design method (PPDM), >
>
>
> ð1{0:0058DR Þqcb for CEP piles (PPDM)
calculate the value of qcb by averaging the cone >
>
>
>    0 0:4
< DR 1:83 sh0
resistance over a vertical distance within 1B above qb,ult ~ 62pA 100 for drilled shafts (PPDM)
> pA
and 2B below the pile base. >
>
> h i
>
> min 0:21ðIFRÞ{1:2 qcb ; 0:6qcb for OEP piles (PPDM)
ii. For the Imperial College pile design method >
>
>
>
:
(ICPDM), calculate the value of qcb by averaging ðEq: 4:28Þ
the cone resistance over a vertical distance within
1.5B above and 1.5B below the pile base. "  0:2 #
Note: If the soil within the averaging zone is clay, Bi
use the corrected, total cone resistance qt (Eq. 4.1), IFR& min 1; ðEq: 4:29Þ
1:5LR
instead of qc.
where IFR 5 incremental filling ratio, Bi 5 inner
b. If the pile base is embedded in a competent (strong) but
diameter of OEP pile, and LR 5 reference length (5 1 m
thin layer (e.g., dense sand or stiff clay) below which
or 39.4 in.). Equation 4.29 can be used to estimate the
there happens to be a weak layer (e.g., loose sand or soft
IFR if plug length measurements are unavailable, but if
clay), then execute the following substeps to estimate the
they are available, then average the IFR over the last 3B
average cone resistance qcb at the pile base.
of pile driving. The relative density DR of the bearing
layer can be estimated from CPT results using (Salgado
i. From the cone resistance profile, determine the & Prezzi, 2007):
representative cone resistances, qc,w and qc,s, of the    0 
weak and strong layers, respectively. qcb s
ln {0:4947{0:1041c {0:841 ln h0
ii. Estimate the sensing distance Hs using (Xu, 2007; pA pA
DR ð%Þ~ 0

Xu & Lehane, 2008): s


0:0264{0:0002c {0:0047 ln ph0A
 
Hs qc,w ðEq: 4:30Þ
~1:41{2:52 ln ðEq: 4:24Þ
B qc,s
where 9h0 5 in situ horizontal effective stress (5 K09v0)
The sensing distance is the vertical distance from at a depth of L + (B/2), 9v0 5 in situ vertical effective
the layer interface at which the cone resistance first stress at a depth of L + (B/2), pA 5 reference stress (5
starts changing as the cone moves toward it 100 kPa or 14.5 psi), c 5 critical-state friction angle
(Salgado, 2014; Tehrani et al., 2018). (refer to step 7(b) of Section 4.1), and K0 5 coefficient of
iii. Determine the vertical distance H from the pile base lateral earth pressure at-rest (refer to Appendix B for
to the interface between the strong and weak layers. guidance).
iv. If H # Hs, calculate the value of qcb using the b. For piles bearing in ‘‘clay,’’ calculate the ultimate unit
following equations (Xu & Lehane, 2008): base resistance qb,ult of the pile using (Jardine et al., 2005;
Salgado, 2006, 2008):
 
qcb qc,w qc,w 8
~ z 1{ > qcb for H-piles (ICPDM)
qc,s qc,s qc,s >
>
< 10s for CEP piles (PPDM)
u
    qb,ult ~
H >
> cb qcb for OEP piles (ICPDM)
>
:
exp { exp A1 zA2 ðEq: 4:25Þ
B 9:6su for drilled shafts (PPDM) ðEq: 4:31Þ

where su 5 undrained shear strength of the bearing layer,


  estimated from CPT results using (Salgado, 2008):
qc,w
A1 ~ min {0:22 ln z0:11; 1:5 ðEq: 4:26Þ
qc,s qcb {sv0
  su ~ ðEq: 4:32Þ
qc,w Nk
A2 ~ min {0:11 ln {0:79; {0:2 ðEq: 4:27Þ
qc,s
where sv0 5 in situ vertical total stress at a depth of L +
(B/2), Nk 5 cone factor [< 9–15 as long as the CPT is
However, if H . Hs, the base resistance of the pile performed at a penetration rate that is sufficiently high to
will not be affected much by the presence of the ensure undrained penetration (refer to Appendix D); soft

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 47


NC clays tend to have Nk values near the low end of this Step 3: Obtain the nominal limit shaft capacity QnsL
range, while stiff OC clays tend to have Nk values near and the nominal ultimate base capacity Qnb,ult of the pile
the high end of this range], and cb 5 coefficient (5 0.4 if
Eq. 4.33 is satisfied and 1.0 otherwise):
by following the steps outlined in Sections 4.1 and 4.2,
respectively.
Bi qcb
z0:45 v36 ðEq: 4:33Þ
dc pA Step 4: Obtain the resistance factors.
where Bi 5 inner diameter of OEP pile, dc 5 cone
a. Purdue pile design method (PPDM): Table 4.3 sum-
diameter, and pA 5 reference stress (5 100 kPa or 14.5
marizes the PPDM resistance factors, RFs and RFb, for
psi).
the pile shaft and base resistances, respectively, based
Step 3: Multiply the ultimate unit base resistance on the selected pile type and the predominant soil type
qb,ult obtained from step 2 with the pile base area Ab to at the site. The resistance factors may be adjusted as
deemed necessary for sites with high soil variability in the
obtain the ultimate base capacity Qb,ult of the pile:
vertical and horizontal directions. Further research is
Qb,ult ~qb,ult Ab ðEq: 4:34Þ needed to develop PPDM resistance factors for OEP
piles in sand.
Table 4.2 summarizes the expressions for Ab for b. Imperial College pile design method (ICPDM): Table 4.4
different pile cross-sections. For OEP piles bearing in summarizes the ICPDM resistance factors for driven
sand (PPDM), calculate the value of Ab using the gross piles in sand and clay. The resistance factors may be
adjusted as deemed necessary for sites with high soil
cross-sectional area (pB2/4) of the pile base. For OEP
variability in the vertical and horizontal directions.
piles bearing in clay (ICPDM), calculate the value of c. AASHTO: Table 4.5 and Table 4.6 summarize the
Ab using the gross cross-sectional area (pB2/4) of the resistance factors advocated by AASHTO (2020) for
pile base if Eq. 4.33 is satisfied, otherwise use the drilled shafts and driven piles, respectively, in sand and
annulus area of steel. clay.
Step 4: Compute the ultimate load capacity Qult of
the pile using: Step 5: Verify that the following LRFD inequality is
Qult ~QsL zQb,ult ðEq: 4:35Þ satisfied (Basu & Salgado, 2012; Foye et al., 2009):

where QsL 5 limit shaft capacity of the pile, and RFs QnsL zRFb Qnb,ult §LFDL DLn zLFLL LLn ðEq: 4:36Þ
Qb,ult 5 ultimate base capacity of the pile. The ultimate
If Eq. 4.36 is satisfied, the pile design is satisfactory
pile load capacity Qult obtained from Eq. 4.35
for the selected target probability of failure. Repeat
corresponds to a pile head settlement w equal to 10%
steps 3 to 5 to optimize the design if needed. However,
of the pile diameter B. For piles of noncircular cross-
if Eq. 4.36 is not satisfied, return to step 3 and revise the
section (e.g., H-piles), an equivalent pile diameter may
pile geometry.
be obtained by equating the cross-sectional area of the
pile with that of an equivalent circle. Note: The following equation may be used, if
needed, to obtain a factor of safety (FS) based on the
Working Stress Design (WSD) method (Han et al.,
4.3 Load and Resistance Factor Design Procedure for 2015):
Single Piles

Load and resistance factor design (LRFD) of a TABLE 4.3


single, isolated, axially-loaded pile can be done from PPDM resistance factors for drilled shafts and CEP piles in sand
and clay (modified from Han et al., 2015)
CPT results by following these steps.
Step 1: Obtain the nominal dead load DLn and the pf,T 5 10–4
nominal live load LLn on the foundation from the Predominant Soil Type
Pile Type at the Site RFb RFs
superstructure design.
Step 2: Set the load factors for dead load and live Drilled shaft Sand 0.70 0.65
load, LFDL and LFLL, as 1.25 and 1.75, respectively Drilled shaft Clay 0.65 0.70
(AASHTO, 2020). These load factors correspond to the CEP pile Sand 0.30 0.60
Strength I limit state (basic load combination relating CEP pile Clay 0.65 0.65
to the normal vehicular use of the bridge without wind),
Note: The resistance factors were developed by Han et al. (2015)
as defined by AASHTO (2020). The discussion of other based on results obtained from Monte Carlo simulations. For layered
limit states, such as Strength II–V, Extreme Event I and clay deposits (soft over stiff layers), the values of RFb and RFs should
II, Service I–IV, and Fatigue I and II are beyond the be decreased by 25% and 20%, respectively.
scope of the manual—information about these limit Notation: pf,T 5 target probability of failure (a value of 10–4 means
states can be found in AASHTO (2020). that one in every 10,000 piles would fail).

48 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


TABLE 4.4
ICPDM resistance factors for driven piles in sand and clay (modified from Kim et al., 2011; Kim & Lee, 2012)

bT 5 3.5 (pf,T < 2610–4)


Pile Type Predominant Soil Type at the Site RFb RFs

CEP and OEP pile Sand 0.56 0.45


CEP and OEP pile Clay 0.58 0.58

Note: The resistance factors were developed by Kim et al. (2011) and Kim and Lee (2012) based on results obtained from reliability analyses
performed using the first-order reliability method (FORM). The RF values listed in Table 4.4 are the lowest among the values reported by Kim
DLn Qnb,ult
et al. (2011) and Kim and Lee (2012) for different combinations of and n . These values may also be used for H-piles as the design equations
LLn QsL
are similar to those for CEP and OEP piles.
Notation: bT 5 target reliability index and pf,T 5 target probability of failure (a value of 2610–4 means that one in every 5,000 piles
would fail).

TABLE 4.5
Resistance factors for drilled shafts in sand and clay (AASHTO, 2020)

Resistance Factor
Method/Condition Predominant Soil Type at the Site RFb RFs

a-method (Brown et al., 2010) Clay 0.40 0.45


b-method (Brown et al., 2010) Sand 0.50 0.55
Static load test (compression) Sand/Clay 0.70 0.70

Note: The resistance factors were developed based on statistical analysis of load test data combined with reliability theory (Paikowsky et al.,
2004), fitting to allowable stress design (ASD), or both (Allen, 2005). For piles subjected to uplift (tension), the resistance factor RF is equal to 0.35
for the a-method, 0.45 for the b-method, and 0.60 for pile design based on static load test results.

TABLE 4.6
Resistance factors for driven piles in sand and clay (AASHTO, 2020)

Method/Condition Predominant Soil Type at the Site Resistance Factor RF

CPT method (Nottingham & Schmertmann, 1975) Sand/Clay 0.50


Static load test (compression) Sand/Clay 0.75–0.801

Note: The resistance factors were developed based on statistical analysis of load test results combined with reliability theory (Paikowsky et al.,
2004), fitting to allowable stress design (ASD), or both (Allen, 2005). For piles subjected to uplift (tension), the resistance factor RF is equal to
0.40 for the CPT method and 0.60 for pile design based on static load test results. Since a single value for the resistance factor was provided by
AASHTO (2020), this value may be used for both the shaft and base components (i.e., RF 5 RFs 5 RFb).
1
Additional information can be found in AASHTO (2020), including resistance factors for conditions when dynamic tests are performed on
the piles.

Cn QnsL zQnb,ult necessary to install multiple piles as a group to support


FS~ n
~ ðEq: 4:37Þ the load. Load and resistance factor design (LRFD)
D DLn zLLn
of axially-loaded pile groups can be done from CPT
where Cn 5 nominal capacity, and Dn 5 nominal results by following these steps.
demand. Step 1: Obtain the nominal dead load DLn and the
nominal live load LLn on the foundation from the
superstructure design.
4.4 Load and Resistance Factor Design Procedure for
Pile Groups Step 2: Set the load factors for dead load and live
load, LFDL and LFLL, as 1.25 and 1.75, respectively
When the axial load from the superstructure exceeds (AASHTO, 2020). These load factors correspond to the
the resistance offered by a single pile, as is the case for Strength I limit state (basic load combination relating
foundations of skyscrapers, bridge piers and abutments, to the normal vehicular use of the bridge without wind),
power plants, and offshore oil platforms, it becomes as defined by AASHTO (2020). The discussion of other

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 49


limit states, such as Strength II–V, Extreme Event I and
II, Service I–IV, and Fatigue I and II are beyond the
scope of the manual—information about these limit
states can be found in AASHTO (2020).
Step 3: Obtain the nominal limit shaft capacity QnsL,i
and the nominal ultimate base capacity Qnb,ult,i of a
single pile in the group by following the steps outlined
in Sections 4.1 and 4.2, respectively.
Step 4: Set the pile center-to-center spacing scc and
the configuration (or layout) of the pile group.
Step 5: LRFD of pile groups in ‘‘sand.’’
Execute the following substeps if the pile group is
installed in a soil profile that consists predominantly of
‘‘sand,’’ otherwise proceed to step 6.
a. Determine the average (representative) relative density
of the sand layer(s) crossed by the pile group (using
Eq. 4.10) and the relative density of the bearing layer in
which the pile group is embedded (using Eq. 4.30).
b. For small drilled shaft groups (e.g., 162, 163,
and 262 groups) (Figure 4.3a), the efficiencies gs,i
and gb,i for the shaft and base resistances, respectively,
are equal to 1.0 for a pile head settlement of 30 mm (1.2
in.). For a large, drilled shaft group (e.g., 464 group)
(Figure 4.3b), refer to Table 4.7 for the values of gs,i
and gb,i. Further research is needed to develop rigorous
values of gs,i and gb,i for driven pile groups in sand; in
the meantime, the same values for drilled shaft groups
may also be used for driven pile groups if deemed
appropriate. Alternatively, Table 4.8 and Table 4.9
summarize the efficiencies advocated by AASHTO Figure 4.3 Layout of (a) small (163) pile group and (b) large
(2020) for drilled shaft groups and driven pile groups, (464) pile group (Han et al., 2015, 2019c).
respectively, in sand.
c. Obtain the resistance factors, RFs and RFb, for the pile
shaft and base resistances, respectively, from step 4 of
Section 4.3. Step 6: LRFD of pile groups in ‘‘clay.’’
d. Verify that the following LRFD inequality is satisfied Execute the following substeps if the pile group is
(Han et al., 2015): installed in a soil profile that consists predominantly of
‘‘clay.’’
np np
P P
RFs gs,i QnsL,i zRFb gb,i Qnb,ult,i a. Individual pile failure ultimate limit state.
i~1 i~1
i. For small drilled shaft groups (e.g., 162, 163, and
§LFDL DLn zLFLL LLn ðEq: 4:38Þ 262 groups) (Figure 4.3a), the efficiencies gs,i and
gb,i for the shaft and base resistances, respectively,
where np 5 number of piles in the group. If Eq. 4.38 is are equal to 1.0 for a pile head settlement of 30 mm
satisfied, the pile group design is satisfactory for (1.2 in.). For a large drilled shaft group (e.g., 464
the selected target probability of failure. Repeat steps group) (Figure 4.3b), refer to Table 4.10 for the
3 to 5 to optimize the design if needed. However, if values of gs,i and gb,i. Further research is needed to
Eq. 4.38 is not satisfied, return to step 3 and revise the develop rigorous values of gs,i and gb,i for driven
design. pile groups in clay; in the meantime, the same
Note: The following equation may be used, if needed, to values for drilled shaft groups may also be used for
obtain a factor of safety (FS) based on the Working driven pile groups if deemed appropriate.
Stress Design (WSD) method: Alternatively, Table 4.11 summarizes the efficien-
np np cies advocated by AASHTO (2020) for drilled shaft
P P
gs,i QnsL,i z gb,i Qnb,ult,i groups and driven pile groups in clay.
n
C ii. Obtain the resistance factors, RFs and RFb, for the
FS~ n ~ i~1 i~1
ðEq: 4:39Þ
D DLn zLLn pile shaft and base resistances, respectively, from
step 4 of Section 4.3.
where Cn 5 nominal capacity, and Dn 5 nominal iii. Verify that the following LRFD inequality is
demand. satisfied (Han et al., 2015):

50 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


TABLE 4.7
Shaft and base efficiencies for a large (464) drilled shaft group in sand for scc 5 2B (Han et al., 2015; Han, Salgado, 2019)

Relative Density DR 5 50% Relative Density DR 5 80%


Pile Head Settlement w Efficiency Center Pile Side Pile Corner Pile Center Pile Side Pile Corner Pile

30 mm (1.2 in.) gb,i 1.14 0.90 0.80 0.93 0.78 0.74


gs,i 0.63 1.01 1.06 0.94 1.25 1.01
50 mm (2.0 in.) gb,i 1.28 0.96 0.81 1.16 0.86 0.77
gs,i 0.80 1.19 1.16 1.23 1.51 1.04

Note: The value of 50 mm (2 in.) for the pile head settlement is based on the tolerable settlement criteria for frame structures and bridges.
Settlements beyond 50 mm (2 in.) would lead to serviceability issues, while those approaching 100 mm (4 in.) would lead to structural damage
(Bozozuk, 1978). For intermediate values of w and DR, the values of gs,i and gb,i can be obtained by linear interpolation.
Notation: B 5 pile diameter, scc 5 pile center-to-center spacing, gs,i 5 efficiency for shaft resistance of the ith pile in the group, and
gb,i 5 efficiency for base resistance of the ith pile in the group.

TABLE 4.8
Efficiencies for small and large drilled shaft groups in sand (AASHTO, 2020)

Group Configuration scc Special Conditions gi

Single row (e.g., 162 and 163 groups) 2B — 0.90


$ 3B — 1.00
Multiple row (e.g., 262 and 464 groups) 2.5B — 0.67
3B — 0.80
$ 4B — 1.00
Single and multiple rows $ 2B Pile cap is in firm contact with medium dense or denser 1.00
soil, and no scour is expected below the cap
Single and multiple rows $ 2B Pressure grouting is used along the sides of the pile to 1.00
restore lateral stress losses caused by pile installation,
and the pile base is pressure grouted

Note: For intermediate values of scc, the value of gi can be obtained by linear interpolation. For pile groups bearing on a strong soil layer of
limited thickness overlying a weak deposit, the nominal resistance of the pile group is taken as the lesser of (a) the sum of the individual nominal
resistances of each pile in the group, and (b) the nominal resistance of the pile group against block failure, with consideration to the punching of the
pile group into the underlying weak layer (AASHTO, 2020).
Notation: B 5 pile diameter, scc 5 pile center-to-center spacing, and gi 5 efficiency of the ith pile in the group (5 gs,i 5 gb,i).

TABLE 4.9
Efficiencies for small and large driven pile groups in sand (AASHTO, 2020)

Group configuration scc Condition gi

Single and multiple rows $ 2.5B No weak layer is present below the pile base 1.00

Note: The value of gi is equal to 1 regardless of whether the pile cap is or is not in contact with the ground. For pile groups bearing on a strong
soil layer of limited thickness overlying a weak deposit, the nominal resistance of the pile group is taken as the lesser of (a) the sum of the individual
nominal resistances of each pile in the group, and (b) the nominal resistance of the pile group against block failure, with consideration to the
punching of the pile group into the underlying weak layer (AASHTO, 2020).
Notation: B 5 pile diameter, scc 5 pile center-to-center spacing, and gi 5 efficiency of the ith pile in the group (5 gs,i 5 gb,i).

np np
P P b. Block failure ultimate limit state.
RFs gs,i QsL,i zRFb gb,i Qb,ult,i
i~1 i~1

i. Determine the length Lg and width Bg of the pile


§LFDL DLn zLFLL LLn ðEq: 4:40Þ
group, as shown in Figure 4.4.
ii. Set the limit unit shaft resistance qsL of the pile
where np 5 number of piles in the group. group is equal to su , the average (representative)
Note: Equation 4.39 may be used, if needed, to undrained shear strength along the pile length. The
obtain a factor of safety (FS) based on the undrained shear strength profile along the pile shaft
Working Stress Design (WSD) method. can be obtained from CPT data using Eq. 4.11.

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 51


   
iii. Estimate the limit unit base resistance qbL of the pile RFs 2 Bg zLg LqsL zRFb Bg Lg qbL
group using (Salgado, 2008; Skempton, 1951):
   §LFDL DLn zLFLL LLn ðEq: 4:42Þ
Bg L
qbL ~5su 1z0:2 1z ðEq: 4:41Þ
Lg 12Bg

Note: The following equation may be used, if


where su 5 undrained shear strength at a depth of
needed, to obtain a factor of safety (FS) based on
L + (Bg/3), and L 5 pile embedment length. the Working Stress Design (WSD) method.
iv. Set both the shaft and base resistance factors,
 
RFs and RFb, as equal to 0.60 for driven pile Cn 2 Bg zLg LqsL zBg Lg qbL
groups and 0.55 for drilled shaft groups (AASHTO, FS ~ n ~ ðEq: 4:43Þ
D DLn zLLn
2020).
v. Verify that the following LRFD inequality is where Cn 5 nominal capacity, and Dn 5 nominal
satisfied: demand.

TABLE 4.10
Shaft and base efficiencies for a large (464) drilled shaft group in NC clay for scc 5 2B (Han et al., 2015)

Pile Head Settlement w Efficiency Center Pile Side Pile Corner Pile

30 mm (1.2 in.) gb,i 0.96 1.01 1.00


gs,i 0.38 0.77 0.98
50 mm (2.0 in.) gb,i 1.02 1.06 1.03
gs,i 0.46 0.85 1.03

Note: The value of 50 mm (2 in.) for the pile head settlement is based on the tolerable settlement criteria for frame structures and bridges.
Settlements beyond 50 mm (2 in.) would lead to serviceability issues, while those approaching 100 mm (4 in.) would lead to structural damage
(Bozozuk, 1978). For intermediate values of w, the values of gs,i and gb,i can be obtained by linear interpolation. Further research is needed to
develop rigorous values of gs,i and gb,i for pile groups in OC clay, but until then, the same values for NC clay may also be used for OC clay.
Notation: B 5 pile diameter, scc 5 pile center-to-center spacing, gs,i 5 efficiency for shaft resistance of the ith pile in the group, and
gb,i 5 efficiency for base resistance of the ith pile in the group.

TABLE 4.11
Efficiencies for small and large drilled shaft and driven pile groups in clay (AASHTO, 2020)

Group Configuration scc Condition gi

Single and multiple rows 2.5B Pile cap is not in firm contact with the ground and 0.65
the soil at the ground surface is soft
Single and multiple rows $ 6B Same as above 1.00

Note: For intermediate values of scc, the value of gi can be obtained by linear interpolation. If the pile cap is not in firm contact with the ground
but the soil is stiff, gi 5 1.0. If the pile cap is in firm contact with the ground, gi 5 1.0.
Notation: B 5 pile diameter, scc 5 pile center-to-center spacing, and gi 5 efficiency of the ith pile in the group (5 gs,i 5 gb,i).

Figure 4.4 Schematic of a 364 pile group with parameters Lg, Bg, and L in (a) plan view and (b) 3D view (Salgado, 2008).

52 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


c. If Eqs. 4.40 and 4.42 are satisfied, the pile group design is mechanical response of soil during shearing, such as
satisfactory with respect to the ultimate limit states of relative density and stress state in the case of sands, and
individual pile failure and block failure, respectively. undrained shear strength, critical-state friction angle,
Repeat steps 3, 4 and 6 to optimize the design if needed. and minimum residual-state friction angle in the case of
However, if either Eq. 4.40 or Eq. 4.42 is not satisfied,
clays, improves the capability of a CPT-based design
return to step 3 and revise the design.
method to predict the ultimate load capacity of a pile.
These variables can be determined using the guidance
and relationships provided in the manual. Also, pile
4.5 Chapter Summary design methods that consider the effect of soil plugging
during the installation of open-ended pipe piles are
In this chapter, detailed, step-by-step procedures for expected to provide more realistic estimates of pile
computing the limit shaft capacity QsL and the ultimate capacity than methods that do not consider this effect.
base capacity Qb,ult of a single, isolated, axially-loaded The capacities mobilized by a closed-ended pipe pile
pile from CPT results in sand (silica sand) and clay were and an open-ended pipe pile are different (Han et al.,
presented. The limit unit shaft resistance qsL and the 2019b; Paik et al., 2003), and pile design methods that
ultimate unit base resistance qb,ult of a pile designed do not differentiate between these pile types do not
using the PPDM depend on the critical-state friction consider installation effects on pile capacity.
angle c and relative density DR in the case of sands, Shaft degradation is a process by which the unit shaft
and the undrained shear strength su and friction angles, resistance at a given depth along the pile decreases as
c and r,min, in the case of clays; r,min 5 minimum the pile is driven down further from that depth (Lehane
residual-state friction angle. The undrained shear et al., 1993; Randolph, 2003; Randolph et al., 1994;
strength su can be estimated from CPT results through White & Lehane, 2004). This degradation, however,
the cone factor Nk, which typically ranges from 9–15 is not properly accounted for in the purely direct
depending on soil type, stress state and history, CPT-based pile design methods (i.e., methods that
and stress path (e.g., triaxial compression versus rely only on CPT data to the exclusion of other
direct simple shear). In addition to some of these variables). Furthermore, because of greater varia-
variables, the ICPDM relies on other key parameters, bility in sleeve resistance measurements (among other
such as residual interface friction angle r, sensitivity St, issues), fs is not a reliable parameter for use in
and overconsolidation ratio OCR in the case of clays. foundation design (Schneider et al., 2008), which is
For base resistance calculations, both the PPDM and why the modern pile design methods (e.g., PPDM,
the ICPDM average the cone resistance qc around the ICPDM, UWAPDM, and UPDM) rely instead on the
pile base according to some formula and relate the cone resistance qc, among other variables, and contain a
ultimate unit base resistance qb,ult of the pile to the shaft resistance degradation term in the design equa-
representative (average) cone resistance qcb, which tions. The PPDM, ICPDM, UWAPDM, and UPDM
serves as a proxy for the limit unit base resistance qbL are based on the 10% relative settlement criterion,
of the pile. i.e., the methods predict the ultimate load capacity of
Guidelines for piles installed in mixed or intermedi- the pile corresponding to a pile displacement equal to
ate soils, such as sand-silt-clay mixtures and gravelly 10% of the pile diameter (except for certain cases, such
sand, were provided. In addition, load and resistance as floating piles in soft clay, where the limit load is
factor design (LRFD) procedures for single piles and achieved after relatively small settlements (Basu &
pile groups were presented, and potential areas for Salgado, 2014)).
future research have been indicated. When using A final note is in order regarding the use of the cone
LRFD, it is important to note that the resistance resistance to obtain other soil parameters of interest.
factors are always tied to the specific design methods The cone resistance is a single measurement, but it
and equations for which they were developed. depends on more than one variable. For example, in
Summary tables for the CPT-based pile design simple terms, the cone resistance qc in sand depends on
methods covered in this chapter have been prepared two state variables—relative density DR and in situ
so that the methods can be easily referred to when horizontal effective stress 9h0 [5 f(K0, OCR)]—and one
needed. The design methods covered in this chapter are intrinsic variable—critical-state friction angle c. The
not mandatory for design in INDOT contracts, and cone resistance can be used to estimate DR if the other
other CPT-based methods, some of which are summar- two variables (9h0 and c) are known, but it cannot be
ized in Table 4.12 to Table 4.25, may be used as used to determine all three variables. This needs to be
deemed appropriate for the site and loading conditions kept in mind as engineers may be tempted to obtain the
under consideration. Pile design methods that rely values of more than one variable from qc, which is a
solely on the measured values of cone resistance to the single measurement. Interpreting CPT results can be
exclusion of other information that may be available at likened to solving a system of equations: the number of
design time will not be as accurate as methods that equations must be equal to the number of unknowns to
consider all the available information. In this sense, be determined. If only one measurement is available, we
they are, in fact, less conservative. The inclusion of key cannot determine multiple independent variables from
intrinsic and state variables known to control the that one measurement.

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 53


4.5.1 Design Methods for Nondisplacement Piles (Drilled Shafts) in Sandy and Clayey Soils

TABLE 4.12
PPDM equations for the unit shaft and base resistances for nondisplacement piles (drilled shafts) in sand and clay

Soil Type and References Limit Unit Shaft Resistance qsL Ultimate Unit Base Resistance qb,ult

Sand (Han et al., 2017) 0    0 0:4


qsL ~Ksv0 tan dc DR 1:83 sh0
  0   qb,ult ~62pA
K 0:67 DR s 100 pA
~ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi exp 1:5{0:35 ln v0
K0 exp 0:3 K0 {0:4 100 pA

Clay (Chakraborty et al., qsL ~asu qb,ult59.6su


 {0:05   0  
2013; Salgado, 2006) su s  
a~ 0 A1 zð1{A1 Þ exp { v0 c {r, min )A2
sv0 pA

A1 5 0.75 for c – r,min # 5u, 0.40 for c – r,min $ 12u and a linearly
interpolated value for 5u , c – r,min , 12u
 
su
A2 ~0:4z0:3 ln 0
sv0

Note: The method predicts the ultimate load capacity Qult of the pile corresponding to a pile head settlement w equal to 10% of the pile diameter B.
The equation for the ultimate unit base resistance qb,ult of drilled shafts in sand is applicable for L/B , 50. The method is intended to estimate the
shaft resistance in clay after dissipation of the excess pore water pressure generated during pile installation. The relative density DR and undrained
shear strength su can be estimated from CPT results using the equations provided in the chapter.
Notation: PPDM 5 Purdue pile design method, pA 5 reference stress (5 100 kPa or 14.5 psi), K 5 coefficient of lateral earth pressure,
9v0 5 in situ vertical effective stress at the depth being considered, dc 5 critical-state interface friction angle (which, for drilled shafts, is equal to the
internal critical-state friction angle c of the soil), 9h0 5 in situ horizontal effective stress at the depth being considered (5 K09v0), K0 5 coefficient of
lateral earth pressure at-rest (Appendix B), and r,min 5 minimum residual-state friction angle (Appendix E).

TABLE 4.13
MnDOT equations (Modified UniCone method) for the unit shaft and base resistances for nondisplacement piles (drilled shafts) in sand
and clay (Dagger et al., 2018)

Soil Type Limit Unit Shaft Resistance qsL Ultimate Unit Base Resistance qb,ult
   
Sand qsL ~qE hpt htc hrate 100:732Ic {3:605 qb,ult ~qcb 100:325Ic {1:218
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Ic is calculated using the same set of equations
Ic ~ ð3:47{ log Qtn Þ2 zð1:22z log Fr Þ2
  n as those in the estimation of qsL.
qt {sv0 pA fs
Qtn ~ 0 and Fr ~ |100%
pA sv0 qt {sv0
 0 
s
n~ min 0:381Ic z0:05 v0 {0:15; 1
pA

Clay Use the same equation as for sand Use the same equation as for sand

Note: The method predicts the maximum load capacity Qmax of the pile (i.e., the maximum load applied on the piles considered in the database).
For most (. 90%) of the pile load tests considered in the database, the value of Qmax was nearly equal to the value of Qult based on the 10% relative
settlement criterion (i.e., the load corresponding to a pile head settlement w equal to 10% of the pile diameter B). The following adjustment was
proposed to estimate Qult from Qmax: Qult 5 0.986Qmax (Niazi & Mayne, 2016).
The value of the exponent n is approximately equal to 1 for clay, 0.75 for silt, and 0.5 for sand. For mixed or intermediate soils, iterative
calculations are needed to determine the value of Ic. For the first iteration, the method recommends the use of n 5 1 to obtain an initial value of
Ic at the depth being considered. In the next iteration, this initial value of Ic is used to update the value of n, which is then used to obtain a new value
of Ic. The process is repeated until the value of Ic converges, which is generally after the third cycle. Additional information on sensitive clays can be
found in Niazi and Mayne (2016).
The representative cone resistance qcb for base resistance calculation is qE averaged over a vertical distance of B below the pile base (Dagger et al.,
2018).
Notation: MnDOT 5 Minnesota Department of Transportation, B 5 pile diameter, qE 5 effective cone resistance (5 qt – u2); qt 5 corrected,
total cone resistance; fs 5 sleeve resistance; u2 5 pore water pressure measured at the shoulder position behind the cone face; Ic 5 soil behavior type
index; Qtn 5 normalized cone resistance; Fr 5 normalized friction ratio; v0 and 9v0 5 in situ vertical total and effective stresses, respectively, at the
depth being considered; pA 5 reference stress (5 100 kPa or 14.5 psi); pt 5 coefficient for pile type (5 0.84 for drilled shafts); tc 5 coefficient for
loading direction (5 0.85 for tension and 1.11 for compression); and rate 5 coefficient for loading procedure (5 1.09 for constant rate of
penetration test and 0.97 for maintained load test).

54 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


4.5.2 Design Methods for Displacement Piles in Sandy Soil

TABLE 4.14
PPDM equations for the unit shaft and base resistances for displacement piles driven in sand (modified from Han et al., 2019b)

Pile Type Limit Unit Shaft Resistance qsL Ultimate Unit Base Resistance qb,ult
0
Closed-ended pipe pile qsL ~Fload Ksv0 tan dc qb,ult5(1–0.0058DR)qcb
 
{ah
K~Kmin zðKmax {Kmin Þ exp
LR
0:01(qc =pA )
Kmax ~ q ffiffiffiffiffiffiffiffiffiffiffiffiffiffi
0
sh0 =pA
Kmin50.2 and a50.14
0
h i
Open-ended pipe pile qsL ~K ð1{0:66PLRÞsv0 tan dc qb,ult ~ min 0:21ðIFRÞ{1:2 qcb ; 0:6qcb
K and Kmax take the same formulae as above, "  0:2 #
with Kmin 5 0.2 and a 5 0.14 Bi
IFR& min 1;
1:5LR

Note: The method predicts the ultimate load capacity Qult of the pile corresponding to a pile head settlement w equal to 10% of the pile diameter B.
The method considers open-ended pipe piles in sand to behave as fully-plugged piles during static loading. Accordingly, the ultimate base capacity
Qb,ult of an open-ended pipe pile is calculated using the gross cross-sectional area (pB2/4) of the pile base. The exponential term in the equation for K
accounts for shaft resistance degradation due to pile driving.
For open-ended pipe piles, the plug length ratio (PLR) used in the equation for qsL is that measured at the specific depth where qsL is calculated.
If the PLR is not measured, it can be approximated using the same equation provided for the IFR. IFR is the incremental filling ratio averaged over
the last 3B of pile driving; if not measured, it can be estimated using the equation provided.
The representative cone resistance qcb for base resistance calculation is qc averaged from 1B above to 2B below the pile base.
Notation: PPDM 5 Purdue pile design method, Fload 5 factor that accounts for loading direction (< 0.5–0.6 for tension and 1.0
for compression), pA 5 reference stress (5 100 kPa or 14.5 psi), LR 5 reference length (5 1 m or 39.4 in.), K 5 coefficient of lateral earth pressure,
9v0 5 in situ vertical effective stress at the depth being considered, dc 5 critical-state interface friction angle (Figure 4.2), h 5 vertical distance from
the pile base to the depth being considered, Bi 5 inner diameter of open-ended pipe pile, 9h0 5 in situ horizontal effective stress at the depth being
considered (5 K09v0), K0 5 coefficient of lateral earth pressure at-rest (Appendix B), qc 5 cone resistance, and DR 5 relative density (estimated
from CPT results using Eq. 4.30).

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 55


TABLE 4.15
ICPDM equations for the unit shaft and base resistances for displacement piles driven in sand (Jardine et al., 2005)

Pile Type Limit Unit Shaft Resistance qsL Ultimate Unit Base Resistance qb,ult
0 0  
Closed-ended qsL ~(Fload src zDsrd ) tan dc B
 0 0:13  {0:38 qb,ult ~ max 0:3; 1{0:5 log qcb
pipe pile dc
0 s h 0 2GDr
src ~0:029qc v0 max ;8 and Dsrd ~
pA R R
{1 qc =pA
G~qc 0:0203z0:00125g{1:216|10{6 g2 and g~ qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
0
sv0 =pA

Open-ended pipe Use the same equations as for closed-ended pipe pile but The pile responds as a plugged pile during static
pile with an equivalent pile radius R given by: loading if:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Bi Bi qcb
R~ Ro2 {R2i v0:02ðDR {30Þ or v0:083
LR dc pA
For piles in tension, the value of qsL is decreased Response as a plugged pile during static loading:
further by 10%.    
B Ri 2
qb,ult ~ max 0:15; 0:5{0:25 log ; 1{ 2 qcb
dc Ro
Qb,ult ~qb,ult pR2o
Response as an unplugged pile during static loading:
 
qann,ult ~qcb and Qb,ult ~qann,ult p Ro2 {R2i

H-pile Use the same equations as for closed-ended pipe pile qb,ult 5 qcb
but with an equivalent pile radius R given by:
rffiffiffiffiffiffi
Ab
R~
p
Ab 5 2bf tf + (2Xp + tw)(d – 2tf)
Xp 5 bf /8 if bf /2 , (d – 2tf) , bf , and
Xp 5 bf 2/[16(d – 2tf)] if (d – 2tf) $ bf

Square or Use the same equations as for closed-ended pipe pile qb,ult 5 0.7qcb
rectangular pile but with an equivalent pile radius R given by:
rffiffiffiffiffiffi
Ab
R~
p
Ab 5 BwBl; where Bw and Bl 5 width and length, respectively,
of the pile cross-section (in plan)

Note: The method predicts the ultimate load capacity Qult of the pile corresponding to a pile head settlement w equal to 10% of the pile diameter
B. In addition, the method is intended to predict the pile capacity measured 10 days after driving for ‘‘virgin’’ piles (i.e., piles that have not been
load-tested). The representative cone resistance qcb for base resistance calculation is qc averaged from 1.5B above to 1.5B below the pile base.
Notation: ICPDM 5 Imperial College pile design method, Fload 5 factor that accounts for loading direction (5 0.8 for tension and 1.0 for
compression), Dr 5 radial displacement of soil during pile loading (< 0.02 mm or 0.8 mil for lightly rusted steel piles), pA 5 reference stress
(5 100 kPa or 14.5 psi), LR 5 reference length (5 1 m or 39.4 in.), 9rc 5 local radial effective stress acting on the pile segment after installation, D9rd
5 increase in local radial effective stress associated with constrained dilation during pile loading, 9v0 5 in situ vertical effective stress at the depth
being considered, dc 5 critical-state interface friction angle, Bi 5 inner diameter of open-ended pipe pile, dc 5 cone diameter, R 5 pile radius, h 5
vertical distance from the pile base to the depth being considered, qc 5 cone resistance, DR 5 relative density, Ro 5 outer radius of open-ended pipe
pile, Ri 5 inner radius of open-ended pipe pile, Ab 5 area of pile base, G 5 shear modulus, bf 5 width of flange, d 5 depth of H-section,
tf 5 thickness of flange, tw 5 thickness of web, and qann,ult 5 ultimate unit annulus resistance.

56 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


TABLE 4.16
UWAPDM equations for the unit shaft and base resistances for displacement piles driven in sand (Lehane et al., 2005)

Pile Type Limit Unit Shaft Resistance qsL Ultimate Unit Base Resistance qb,ult

Closed-ended ft 0 0

qb,ult 5 0.6qcb
qsL ~ s zDsrd tan dc
pipe pile fc rc
  {0:5
0 h
src ~0:03qc max ;2
B
2 3{0:75
0 4GDr G 6 qc =pA 7
Dsrd ~ and ~1854qffiffiffiffiffiffiffiffiffiffiffiffiffiffi5
B qc 0
s =p v0 A

Open-ended ft 0 0

qb,ult5(0.15 + 0.45Arb)qcb
qsL ~ s zDsrd tan dc  2
pipe pile fc rc Bi
  {0:5  2 Arb ~1{FFR
0 h Bi B
src ~0:03qc ðArs Þ0:3 max ;2 and Ars ~1{IFR
B B FFR is the final filling ratio, which is defined
2 3{0:75
as the average incremental filling ratio
0 4GDr G 6 qc =pA 7 measured over the final 3B of pile driving; if
Dsrd ~ and ~1854qffiffiffiffiffiffiffiffiffiffiffiffiffiffi5
B qc 0
s =p not measured, it can be roughly
v0 A
approximated by using the same equation
IFR is the average incremental filling ratio measured over the
for the IFR.
final 20B of pile driving; when plug length measurements
are not available, it can be estimated using:
"  0:2 #
Bi
IFR& min 1;
1:5LR

Note: The method predicts the ultimate load capacity Qult of the pile corresponding to a pile base settlement equal to 10% of the pile diameter
(Lehane et al., 2007; Xu, 2007; Xu et al., 2008). In addition, the method is intended to predict the pile capacity measured 10–20 days after driving.
The method considers open-ended pipe piles in sand to behave as fully-plugged piles during static loading. Accordingly, the ultimate base
capacity Qb,ult of an open-ended pipe pile is calculated using the gross cross-sectional area (pB2/4) of the pile base.
The representative cone resistance qcb for base resistance calculation is qc averaged using the Dutch technique (Figure 4.5): qcb 5 0.5(qc1+qc2),
with qc1 5 0.5(qc1a + qc1b), qc1a 5 average of the qc values over a vertical distance of lB below the pile base, qc1b 5 average of the qc values over a
vertical distance of lB below the pile base following a minimum path rule, and qc2 5 average of the qc values over a vertical distance of 8B above the
pile base following a minimum path rule. The value of qc1 is calculated for different l values ranging from 0.7 to 4.0, and the minimum value of
qc1 obtained is used in the calculation of qcb. Additional information about the computation of qc1 and qc2 can be found in Schmertmann (1978).
For open-ended pipe piles, B is replaced by Beff [5 B(Arb)0.5] in the calculation of qcb.
In the absence of plug length measurements, the value of the IFR may also be estimated using: IFR < tanh[0.3(Bi/dc)0.5] (Lehane, 2019). The FFR
can be roughly approximated by using the same equation for the IFR.
Notation: UWAPDM 5 University of Western Australia pile design method, ft/fc 5 ratio of tension to compression capacity (5 0.75 for tension
and 1.0 for compression), Dr 5 radial displacement of soil during pile loading (< 0.02 mm or 0.8 mil for lightly rusted steel piles), pA 5 reference
stress (5 100 kPa or 14.5 psi), LR 5 reference length (5 1 m or 39.4 in.), 9rc 5 local radial effective stress acting on the pile segment after
installation, D9rd 5 increase in local radial effective stress associated with constrained dilation during pile loading, 9v0 5 in situ vertical effective
stress at the depth being considered, dc 5 critical-state interface friction angle, Ars 5 effective shaft area ratio, Arb 5 effective base area ratio,
Bi 5 inner diameter of open-ended pipe pile, Beff5 effective pile diameter, dc 5 cone diameter, h 5 vertical distance from the pile base to the depth
being considered, qc 5 cone resistance, and G 5 shear modulus.

TABLE 4.17
AASHTO equations for the unit shaft and base resistances for displacement piles driven in sand (AASHTO, 2020; Nottingham &
Schmertmann, 1975)

Pile Type Limit Unit Shaft Resistance qsL Limit Unit Base Resistance qbL
( z

Closed-ended pile 0:125 Ks fs for 0ƒzƒ8B qbL ~qcb


qsL ~ B
Ks fs for 8BƒzƒL

Note: The representative cone resistance qcb for base resistance calculation is qc averaged using the Dutch technique (Figure 4.5):
qcb 5 0.5(qc1+qc2), with qc1 5 0.5(qc1a + qc1b), qc1a 5 average of the qc values over a vertical distance of lB below the pile base, qc1b 5 average of the
qc values over a vertical distance of lB below the pile base following a minimum path rule, and qc2 5 average of the qc values over a vertical distance
of 8B above the pile base following a minimum path rule. The value of qc1 is calculated for different l values ranging from 0.7 to 4.0, and the
minimum value of qc1 obtained is used in the calculation of qcb. Additional information about the computation of qc1 and qc2 can be found in
AASHTO (2020).
Notation: Ks 5 correction factor (estimated from the chart provided by AASHTO (2020) as a function of L/B, penetrometer type (electrical
versus mechanical), and pile material (steel, concrete, or timber)), fs 5 sleeve resistance, L 5 embedded length of the pile, B 5 width or diameter of
the pile, z 5 depth measured from the ground surface, and qc 5 cone resistance.

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 57


Figure 4.5 Dutch technique for estimation of qcb (modified from Schmertmann, 1978).

TABLE 4.18
MnDOT equations (Modified UniCone method) for the unit shaft and base resistances for displacement piles driven in sand (Dagger et al.,
2018)

Pile Type Limit Unit Shaft Resistance qsL Ultimate Unit Base Resistance qb,ult
   
Closed-ended pipe pile qsL ~qE hpt htc hrate 100:732Ic {3:605
qb,ult ~ 100:325Ic {1:218 qcb
Open-ended pipe pile qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Ic is calculated using the same set of equations
Ic ~ ð3:47{ log Qtn Þ2 zð1:22z log Fr Þ2
H-pile   n as those in the estimation of qsL.
qt {sv0 pA fs
Qtn ~ 0 and Fr ~ |100%
pA sv0 qt {sv0
 0 
s
n~ min 0:381Ic z0:05 v0 {0:15; 1
pA

Note: The method predicts the maximum load capacity Qmax of the pile (i.e., the maximum load applied on the piles considered in the database).
For most (. 90%) of the pile load tests considered in the database, the value of Qmax was nearly equal to the value of Qult based on the 10% relative
settlement criterion (i.e., the load corresponding to a pile head settlement w equal to 10% of the pile diameter B). The following adjustment was
proposed to estimate Qult from Qmax: Qult 5 0.986Qmax (Niazi & Mayne, 2016).
The value of the exponent n is approximately equal to 1 for clay, 0.75 for silt, and 0.5 for sand. For mixed or intermediate soils, iterative
calculations are needed to determine the value of Ic. For the first iteration, the method recommends the use of n 5 1 to obtain an initial value of
Ic at the depth being considered. In the next iteration, this initial value of Ic is used to update the value of n, which is then used to obtain a new value
of Ic. The process is repeated until the value of Ic converges, which is generally after the third cycle.
The ultimate base capacity Qb,ult of an open-ended pipe pile is calculated using the gross cross-sectional area (pB2/4) of the pile base. The
representative cone resistance qcb for base resistance calculation is qE averaged over a vertical distance of B below the pile base (Dagger et al., 2018).
Notation: MnDOT 5 Minnesota Department of Transportation, B 5 pile diameter, qE 5 effective cone resistance (5 qt – u2); qt 5 corrected,
total cone resistance; fs 5 sleeve resistance; u2 5 pore water pressure measured at the shoulder position behind the cone face; Ic 5 soil behavior type
index; Qtn 5 normalized cone resistance; Fr 5 normalized friction ratio; v0 and 9v0 5 in situ vertical total and effective stresses, respectively, at the
depth being considered; pA 5 reference stress (5 100 kPa or 14.5 psi); pt 5 coefficient for pile type (5 1.13 for driven piles); tc 5 coefficient for
loading direction (5 0.85 for tension and 1.11 for compression); and rate 5 coefficient for loading procedure (5 1.09 for constant rate of
penetration test and 0.97 for maintained load test).

58 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


TABLE 4.19
UPDM equations for the unit shaft and base resistances for displacement piles driven in sand (Lehane et al., 2020)

Pile Type Limit Unit Shaft Resistance qsL Ultimate Unit Base Resistance qb,ult

Closed-ended ft 0 0

qb,ult ~0:5qcb
qsL ~ s zDsrd tan dc
pipe pile fc rc
  {0:4  {0:33  
0 qc h 0 qc dc
src ~ max ;1 and Dsrd ~0:1qc 0
44 B sv0 B

Open-ended pipe ft 0 0

qb,ult ~ð0:12z0:38Arb Þqcb
qsL ~ s zDsrd tan dc  2
pile fc rc Bi
  {0:4  2 Arb ~1{FFR
0 qc h Bi B
src ~ ðArs Þ0:3 max ;1 and Ars ~1{PLR
44 B B FFR is the final filling ratio, which is defined as the
 {0:33  
0 qc dc average incremental filling ratio measured over
Dsrd ~0:1qc 0 the final 3B of pile driving; if not measured, it
sv0 B
PLR is the plug length ratio; when plug length measurements can be roughly approximated by using the same
are not available, it can be estimated using: equation for the PLR.
"   #
Bi 0:5
PLR& tanh 0:3
dc

Note: The method predicts the ultimate load capacity Qult of the pile corresponding to a pile base settlement equal to 10% of the pile diameter.
In addition, the method is intended to predict the pile capacity measured 14 days after driving.
The method considers open-ended pipe piles in sand to behave as fully-plugged piles during static loading. Accordingly, the ultimate base
capacity Qb,ult of an open-ended pipe pile is calculated using the gross cross-sectional area (pB2/4) of the pile base.
For piles installed in relatively homogeneous sands, the representative cone resistance qcb for base resistance calculation is qc averaged from 1.5B
above to 1.5B below the pile base. For piles installed in highly variable soil profiles (i.e., when qc varies significantly in the vicinity of the pile base),
qcb can be either taken as 1.2qc,Dutch or estimated using the procedure developed by Boulanger and DeJong (2018); qc,Dutch 5 qc averaged using the
Dutch technique (Schmertmann, 1978). For open-ended pipe piles, B is replaced by Beff [5 B(Arb)0.5] in the calculation of qcb.
Notation: UPDM 5 Unified pile design method, ft/fc 5 ratio of tension to compression capacity (5 0.75 for tension and 1.0 for compression),
9rc 5 local radial effective stress acting on the pile segment after installation, D9rd 5 increase in local radial effective stress associated with
constrained dilation during pile loading, 9v0 5 in situ vertical effective stress at the depth being considered, dc 5 critical-state interface friction angle
(5 29u in the absence of laboratory interface shear test results), Ars 5 effective shaft area ratio, Arb 5 effective base area ratio, Bi 5 inner diameter
of open-ended pipe pile, Beff 5 effective pile diameter, dc 5 cone diameter, h 5 vertical distance from the pile base to the depth being considered,
and qc 5 cone resistance.

4.5.3 Design Methods for Displacement Piles in Clayey Soil

TABLE 4.20
PPDM equations for the unit shaft and base resistances for displacement piles driven in clay (Basu et al., 2009; Salgado, 2008)

Pile Type Limit Unit Shaft Resistance qsL Ultimate Unit Base Resistance qb,ult

Closed-ended qsL ~asu 10su for short-term resistance
 0  qb,ult &
pipe pile s 12su for long-term resistance
a~A1 zð1{A1 Þ exp { v0 ðc {r, min ÞA2
pA
for short-term resistance, and
 {0:05 ( "  0  #)
su sv0 A3
a~1:28 0 A1 zð1{A1 Þ exp { (c {r, min )
sv0 pA
for long-term resistance
A1 5 0.75 for c – r,min # 5u, 0.43 for c – r,min $ 12u and a
linearly interpolated value for 5u , c – r,min , 12u
   
su su
A2 ~0:55z0:43 ln 0 and A3 ~0:64z0:40 ln 0
sv0 sv0

Note: The method predicts the ultimate load capacity Qult of the pile corresponding to a pile head settlement w equal to 10% of the pile diameter B.
Short-term resistance refers to the resistance available immediately after pile installation (corresponding to zero dissipation of excess pore water
pressure). Long-term resistance refers to the resistance available after dissipation of the excess pore water pressure generated during pile installation.
Notation: PPDM 5 Purdue pile design method, pA 5 reference stress (5 100 kPa or 14.5 psi), 9v0 5 in situ vertical effective stress at the depth
being considered, c 5 critical-state friction angle, r,min 5 minimum residual-state friction angle (Appendix E), and su 5 undrained shear strength
(estimated from CPT results using the equations provided in the chapter).

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 59


TABLE 4.21
ICPDM equations for the unit shaft and base resistances for displacement piles driven in clay (Jardine et al., 2005)

Pile Type Limit Unit Shaft Resistance qsL Ultimate Unit Base Resistance qb,ult
0 
Closed-ended pipe pile qsL ~Fload Ksv0 tan dr 0:8qcb for undrained loading
 {0:20 qb,ult ~
 h 1:3qcb for drained loading
K~ 2:2z0:016OCR{0:87DIvy OCR0:42 max ;8
R
su
DIvy ~log10 St and St ~
sur
Open-ended pipe pile Use the same equations as for closed-ended pipe pile The pile responds as a plugged pile during static
but with an equivalent pile radius R given by: loading if:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Bi qcb
R~ Ro2 {R2i z0:45 v36
dc pA
Response as a plugged pile during static loading:

0:4qcb for undrained loading
qb,ult ~
0:65qcb for drained loading
Qb,ult ~qb,ult pR2o
Response as an unplugged pile during static loading:

qcb for undrained loading
qann,ult ~
1:6qcb for drained loading
 
Qb,ult ~qann,ult p Ro2 {R2i

H-pile Use the same equations as for closed-ended pipe pile qb,ult5qcb
but with an equivalent pile radius R given by:

rffiffiffiffiffiffi
Ab
R~
p
Ab 5 2bftf + (2Xp + tw)(d – 2tf)
Xp 5 bf/8 if bf/2 , (d – 2tf) , bf, and
Xp 5 bf2/[16(d – 2tf)] if (d – 2tf) $ bf

Square or rectangular Use the same equations as for closed-ended pipe pile qb,ult 5 0.7qcb
pile but with an equivalent pile radius R given by:
rffiffiffiffiffiffi
Ab
R~
p
Ab 5 BwBl; where Bw and Bl 5 width and length, respectively,
of the pile cross-section (in plan)

Note: The method predicts the ultimate load capacity Qult of the pile corresponding to a pile head settlement w equal to 10% of the pile diameter
B. In addition, the method is intended to estimate the shaft resistance after dissipation of the excess pore water pressure generated during pile
installation. The representative cone resistance qcb for base resistance calculation is qt averaged from 1.5B above to 1.5B below the pile base.
The residual interface friction angle dr can be determined from the results of ring shear interface tests performed for the applicable value of
normal effective stress (Ramsey et al., 1998). If such test results are unavailable, it is possible to estimate the value of dr by recognizing that it varies
with the normal effective stress 9 acting on the pile shaft, which, for production piles, is typically rough, so that dr is approximately equal to
r. Note that 9, in the context of pile shaft resistance calculation, is the horizontal effective stress 9h on the pile operative at the time of shearing:
9h 5 FloadK9v0.
Notation: ICPDM 5 Imperial College pile design method, Fload 5 0.8 regardless of the loading direction, pA 5 reference stress (5 100 kPa or
14.5 psi), LR 5 reference length (5 1 m or 39.4 in.), qt 5 corrected, total cone resistance, 9v0 5 in situ vertical effective stress at the depth being
considered, Ab 5 area of pile base, Bi 5 inner diameter of open-ended pipe pile, dc 5 cone diameter, R 5 pile radius, h 5 vertical distance from the
pile base to the depth being considered, OCR 5 overconsolidation ratio, su 5 undrained shear strength, DIvy 5 relative void index at yield in e–log
9v space, St 5 sensitivity, sur 5 remolded undrained shear strength, LI 5 liquidity index [5 (wc – PL)/PI], wc 5 water content, PL 5 plastic limit,
PI 5 plasticity index, Ro 5 outer radius of open-ended pipe pile, Ri 5 inner radius of open-ended pipe pile, bf 5 width of flange, d 5 depth of
H-section, tf 5 thickness of flange, tw 5 thickness of web, and qann,ult 5 ultimate unit annulus resistance.

60 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


TABLE 4.22
UWAPDM equations for the unit shaft and base resistances for displacement piles driven in clay (Lehane, 2019; Lehane et al., 2013)

Pile Type Limit Unit Shaft Resistance qsL Ultimate Unit Base Resistance qb,ult
  {0:2
Closed-ended pipe pile h qb,ult<0.5qcb for undrained loading
0:23qt max ;1
R
qsL ~  0:15 tan dr
qt
0
sv0
or
  {0:2
h
qsL ~0:055qt max ;1
R

Open-ended pipe pile Use the same equations as for closed-ended pipe pile Response as a plugged pile during static loading:
but with an equivalent pile radius R given by: qb,ult<0.5qcb for undrained loading
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
R~ Ro2 {R2i

Note: The method is intended to estimate the shaft resistance after dissipation of the excess pore water pressure generated during pile installation
(Lehane, 2019; Lehane et al., 2017). Two equations were proposed for the limit unit shaft resistance qsL and the second one was reported by Lehane
et al. (2013) to be slightly more reliable than the first. The ultimate base capacity Qb,ult of an open-ended pipe pile is calculated using the gross cross-
sectional area (pB2/4) of the pile base.
The residual interface friction angle dr can be determined from the results of ring shear interface tests performed for the applicable value
of normal effective stress (Ramsey et al., 1998). If such test results are unavailable, it is possible to estimate the value of dr by recognizing that it
varies with the normal effective stress 9 acting on the pile shaft, which, for production piles, is typically rough, so that dr is approximately equal to
r. Note that 9, in the context of pile shaft resistance calculation, is the horizontal effective stress 9h on the pile operative at the time of shearing:
9h 5 0.23qt[max(h/R;1)]–0.2/(qt/9v0)0.15.
Notation: UWAPDM 5 University of Western Australia pile design method, qt 5 corrected, total cone resistance, 9v0 5 in situ vertical effective
stress at the depth being considered, R 5 pile radius, h 5 vertical distance from the pile base to the depth being considered, Ro 5 outer radius of
open-ended pipe pile, and Ri 5 inner radius of open-ended pipe pile.

TABLE 4.23
AASHTO equations for the unit shaft and base resistances for displacement piles driven in clay (AASHTO, 2020; Nottingham &
Schmertmann, 1975)

Pile Type Limit Unit Shaft Resistance qsL Limit Unit Base Resistance qbL
( z

Closed-ended pile 0:125Kc fs for 0ƒzƒ8B qbL ~qcb


qsL ~ B
Kc fs for 8BƒzƒL

Note: The representative cone resistance qcb for base resistance calculation is qt averaged using the Dutch technique (Figure 4.5): qcb 5
0.5(qc1+qc2), with qc1 5 0.5(qc1a + qc1b), qc1a 5 average of the qt values over a vertical distance of lB below the pile base, qc1b 5 average of the
qt values over a vertical distance of lB below the pile base following a minimum path rule, and qc2 5 average of the qt values over a vertical distance
of 8B above the pile base following a minimum path rule. The value of qc1 is calculated for different l values ranging from 0.7 to 4.0, and the
minimum value of qc1 obtained is used in the calculation of qcb. Additional information about the computation of qc1 and qc2 can be found in
AASHTO (2020).
Notation: Kc 5 correction factor [estimated from the chart provided by AASHTO (2020) as a function of fs and pile material (steel, concrete, or
timber)], fs 5 sleeve resistance, L 5 embedded length of the pile, B 5 width or diameter of the pile, z 5 depth measured from the ground surface,
and qt 5 corrected, total cone resistance (Eq. 4.1).

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 61


TABLE 4.24
MnDOT equations (Modified UniCone method) for the unit shaft and base resistances for displacement piles driven in clay (Dagger et al.,
2018)

Pile Type Limit Unit Shaft Resistance qsL Ultimate Unit Base Resistance qb,ult
   
Closed-ended pipe pile qsL ~qE hpt htc hrate 100:732Ic {3:605 qb,ult ~ 100:325Ic {1:218 qcb
Open-ended pipe pile qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Ic is calculated using the same set of
Ic ~ ð3:47{ log Qtn Þ2 zð1:22z log Fr Þ2
H-pile   n equations as those in the estimation of qsL.
qt {sv0 pA fs
Qtn ~ 0 and Fr ~ |100%
pA sv0 qt {sv0
 0 
s
n~ min 0:381Ic z0:05 v0 {0:15; 1
pA

Note: The method predicts the maximum load capacity Qmax of the pile (i.e., the maximum load applied on the piles considered in the database).
For most (. 90%) of the pile load tests considered in the database, the value of Qmax was nearly equal to the value of Qult based on the 10% relative
settlement criterion (i.e., the load corresponding to a pile head settlement w equal to 10% of the pile diameter B). The following adjustment was
proposed to estimate Qult from Qmax: Qult 5 0.986Qmax (Niazi & Mayne, 2016).
The value of the exponent n is approximately equal to 1 for clay, 0.75 for silt, and 0.5 for sand. For mixed or intermediate soils, iterative
calculations are needed to determine the value of Ic. For the first iteration, the method recommends the use of n 5 1 to obtain an initial value of Ic at
the depth being considered. In the next iteration, this initial value of Ic is used to update the value of n, which is then used to obtain a new value of
Ic. The process is repeated until the value of Ic converges, which is generally after the third cycle. Additional information on sensitive clays can be
found in Niazi and Mayne (2016).
The ultimate base capacity Qb,ult of an open-ended pipe pile is calculated using the gross cross-sectional area (pB2/4) of the pile base. The
representative cone resistance qcb for base resistance calculation is qE averaged over a vertical distance of B below the pile base (Dagger et al., 2018).
Notation: MnDOT 5 Minnesota Department of Transportation, B 5 pile diameter, qE 5 effective cone resistance (5 qt – u2); qt 5 corrected,
total cone resistance; fs 5 sleeve resistance; u2 5 pore water pressure measured at the shoulder position behind the cone face; Ic 5 soil behavior type
index; Qtn 5 normalized cone resistance; Fr 5 normalized friction ratio; v0 and 9v0 5 in situ vertical total and effective stresses, respectively, at the
depth being considered; pA 5 reference stress (5 100 kPa or 14.5 psi); pt 5 coefficient for pile type (5 1.13 for driven piles); tc 5 coefficient for
loading direction (5 0.85 for tension and 1.11 for compression); and rate 5 coefficient for loading procedure (5 1.09 for constant rate of
penetration test and 0.97 for maintained load test).

TABLE 4.25
NDOT equations for the unit shaft and base resistances for displacement piles driven in clay (Song et al., 2019)

Pile Type and Reference Unit Shaft Resistance Unit Base Resistance

Closed-ended pipe pile qt qb 5 0.54qcb


qs ~
Precast prestressed concrete pile (modified from 60
de Ruiter and Beringen, 1979)

H-pile (modified from Tumay and Fakhroo, 1982) qs ~ min m fs,avg ; 0:72pA qb 5 min [0.5qcb; 150pA]

 
m ~0:45z8:55 exp {0:09fs,avg
Pn
fsi Dzi
i~1
fs,avg ~ n
P
Dzi
i~1

Note: The method is applicable to fine-grained Nebraska soils and predicts the pile capacity that would be obtained from dynamic load tests
performed using the pile driving analyzer (PDA) at the end of initial driving and post-processed using the signal matching program CAPWAP (Case
Pile Wave Analysis Program).
In the de Ruiter and Beringen (1979) method, the representative cone resistance qcb for base resistance calculation is qt averaged using the Dutch
technique (Figure 4.5). In the Tumay and Fakhroo (1982) method, qcb is calculated in a manner similar to the Dutch technique: qcb 5 0.5(qc1+qc2),
with qc1 5 0.5(qc1a + qc1b), qc1a 5 average of the qt values over a vertical distance of 4B below the pile base, qc1b 5 average of the qt values over a
vertical distance of 4B below the pile base following a minimum path rule, and qc2 5 average of the qt values over a vertical distance of 8B above the
pile base following a minimum path rule.
Notation: NDOT 5 Nebraska Department of Transportation, m* 5 modified friction coefficient, fs,avg 5 weighted-average sleeve resistance,
fsi 5 sleeve resistance of soil layer i, Dzi 5 thickness of soil layer i, n 5 number of soil layers in contact with the pile shaft, pA 5 reference stress
(5 100 kPa or 14.5 psi), B 5 pile diameter, and qt 5 corrected, total cone resistance (Eq. 4.1).

62 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


REFERENCES Lafayette, IN: Purdue University. https://doi.org/10.5703/
1288284314282.
AASHTO. (1991). AASHTO M 145: Standard specification Becker, D. E. (1996). Eighteenth Canadian geotechnical
for classification of soils and soil-aggregate mixtures for colloquium: Limit states design for foundations. Part I.
highway construction purposes. American Association of An overview of the foundation design process. Canadian
State Highway and Transportation Officials. Geotechnical Journal, 33(6), 956–983.
AASHTO. (2020). AASHTO LRFD bridge design specifica- De Beer, E., Lousberg, E., De Jonghe, A., Carpentier, R., &
tions (9th ed.). American Association of State Highway and Wallays, M. (1980). Analysis of the results of loading
Transportation Officials. tests performed on displacement piles of different types and
Akinlotan, O. (2017). Mineralogy and palaeoenvironments: sizes penetrating at a relatively small depth into a very
the Weald Basin (Early Cretaceous), Southeast England. dense sand layer. Recent Developments in the Design and
The Depositional Record, 3(2), 187–200. Construction of Piles (pp. 199–211). The Institution of Civil
Allen, T. M. (2005, February). Development of geotechnical Engineers.
resistance factors and downdrag load factors for LRFD Bellotti, R., Jamiolkowski, M., Presti, D. C. F. Lo, & O’Neill,
foundation strength limit state design (Report No. FHWA- D. A. (1996). Anisotropy of small strain stiffness in Ticino
NHI-05-052). National Highway Institute, Federal High- sand. Géotechnique, 46(1), 115–131.
way Administration. Bishop, A. W., Green, G. E., Garga, V. K., Andresen, A., &
Altuhafi, F. N., Coop, M. R., & Georgiannou, V. N. (2016). Brown, J. D. (1971, December). A new ring shear apparatus
Effect of particle shape on the mechanical behavior of and its application to the measurement of residual strength.
natural sands. Journal of Geotechnical and Geoenviron- Géotechnique, 21(4), 273–328.
mental Engineering, 142(12), 04016071. Bisht, V., Salgado, R., & Prezzi, M. (2021, December 1). The
Altuhafi, F. N., Jardine, R. J., Georgiannou, V. N., & Moinet, material point method for cone penetration in clays.
W. W. (2018, June). Effects of particle breakage and stress Journal of Geotechnical and Geoenvironmental Engineering,
reversal on the behaviour of sand around displacement 147(12), 04021158.
piles. Géotechnique, 68(6), 546–555.
Bolton, M. D. (1986). The strength and dilatancy of sands.
Altuhafi, F., O’Sullivan, C., & Cavarretta, I. (2013). Analysis Géotechnique, 36(1), 65–78.
of an image-based method to quantify the size and shape of
Bonaparte, R. (1982). A time-dependent constitutive model
sand particles. Journal of Geotechnical and Geoenviron-
for cohesive soils [Doctoral dissertation, University of
mental Engineering, 139(8), 1290–1307.
California, Berkeley].
Anderson, J. B., Townsend, F. C., & Horta, E. (2004). A brief
Boulanger, R. W. W., & DeJong, J. T. T. (2018). Inverse
study on the repeatability of in-situ tests at the Florida
filtering procedure to correct cone penetration data for
department of transportation deep foundations research
thin-layer and transition effects (pp. 25–44). In Cone
site in Orlando, Florida. In A. V. de Fonseca, & P. W.
Penetration Testing 2018 (1st ed.). CRC Press.
Mayne (Eds.), Geotechnical and Geophysical Site
Characterization, 2, 1597–1604. Bozozuk, M. (1978). Bridge foundations move. Transportation
Research Record: Journal of the Transportation Research
Argyilan, E. P., Johnston, J. W., Lepper, K., Monaghan, G.
W., & Thompson, T. A. (2018). Lake level, shoreline, and Board, No. 678, 17–21.
dune behavior along the Indiana southern shore of Lake Brooker, E. W., & Ireland, H. O. (1965). Earth pressures at
Michigan. Ancient Oceans, Orogenic Uplifts, and Glacial rest related to stress history. Canadian Geotechnical Journal,
Ice: Geologic Crossroads in America’s Heartland. Geological 2(1), 1–15.
Society of America Field Guide, 51, 181–203. Brown, D. A., Turner, J. P., & Castelli, R. J. (2010). Drilled
ASTM. (2012). ASTM D5778-12: Standard test method for shafts: construction procedures and LRFD design methods
electronic friction cone and piezocone penetration testing of (Report No. FHWA NHI-10-016). Federal Highway
soils. ASTM International. Administration.
ASTM. (2017). ASTM D2487: Standard practice for classifi- Brown, R. (2016). Geomorphons: Landform and property
cation of soils for engineering purposes (unified soil classifi- predictions in a glacial moraine in Indiana landscapes.
cation system). ASTM International. Catena, 142, 66–76.
Ault, C. H., & Sullivan, D. M. (1982, October). Faulting in Campanella, R. G., Robertson, P. K., & Gillespie, D. (1986).
southwest Indiana (U.S. Nuclear Regulatory Commission Seismic cone penetration test. Use of In-Situ Tests in
Report NUREG/CR-2908). Indiana Geological Survey. Geotechnical Engineering, GSP 6, 116–130. American
Basu, D., & Salgado, R. (2012). Load and resistance factor Society of Civil Engineers.
design of drilled shafts in sand. Journal of Geotechnical and Campanella, R. G., & Weemees, I. (1990). Development and
Geoenvironmental Engineering, 138(12), 1455–1469. use of an electrical resistivity cone for groundwater conta-
Basu, D., & Salgado, R. (2014). Closure to ‘‘Load and mination studies. Canadian Geotechnical Journal, 27(5),
resistance factor design of drilled shafts in sand’’ by D. Basu 557–567.
and Rodrigo Salgado. Journal of Geotechnical and Geo- Cao, Z., & Wang, Y. (2013, February). Bayesian approach
environmental Engineering, 140(3), https://doi.org/10.1061/ for probabilistic site characterization using cone penetra-
(ASCE)GT.1943-5606.0001055 tion tests. Journal of Geotechnical and Geoenvironmental
Basu, P., Prezzi, M., Salgado, R., & Chakraborty, T. (2014). Engineering, 139(2), 267–276.
Shaft resistance and setup factors for piles jacked in clay. Carraro, J. A. H., Prezzi, M., & Salgado, R. (2009). Shear
Journal of Geotechnical and Geoenvironmental Engineering, strength and stiffness of sands containing plastic or
140(3), 04013026. nonplastic fines. Journal of Geotechnical and Geoenviron-
Basu, P., Salgado, R., Prezzi, M., & Chakraborty, T. (2009). A mental Engineering, 135(9), 1167–1178.
method for accounting for pile setup and relaxation in pile Chakraborty, T. (2009). Development of a clay constitutive
design and quality assurance (Joint Transportation Research model and its application to pile boundary value problems
Program Publication FHWA/IN/JTRP-2009/24). West [Doctoral dissertation, Purdue University].

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 63


Chakraborty, T., Salgado, R., Basu, P., & Prezzi, M. (2013). Foye, K. C., Salgado, R., & Scott, B. (2006a). Assessment of
Shaft resistance of drilled shafts in clay. Journal of variable uncertainties for reliability-based design of foun-
Geotechnical and Geoenvironmental Engineering, 139(4), dations. Journal of Geotechnical and Geoenvironmental
548–563. Engineering, 132(9), 1197–1207.
Cho, G.-C., Dodds, J., & Santamarina, J. C. (2006, May). Foye, K. C., Salgado, R., & Scott, B. (2006b). Resistance
Particle shape effects on packing density, stiffness, and factors for use in shallow foundation LRFD. Journal of
strength: natural and crushed sands. Journal of Geotechnical Geotechnical and Geoenvironmental Engineering, 132(9),
and Geoenvironmental Engineering, 132(5), 591–602. 1208–1218.
Colgan, P. M., Mickelson, D. M., & Cutler, P. M. (2003). Ice- Fukue, M., Minato, T., Matsumoto, M., Horibe, H., & Taya,
marginal terrestrial landsystems: Southern-Laurentide Ice N. (2001). Use of resistivity cone for detecting contami-
Sheet margin. In D. A. Evans (Ed.), Glacial Landsystems nated soil layers. Engineering Geology, 60(1–4), 361–369.
(pp. 111–142). Routledge. Galvis-Castro, A. C., Tovar-Valencia, R. D., Salgado, R., &
Coop, M. R., & Lee, I. K. (1993). The behaviour of granular Prezzi, M. (2019). Effect of loading direction on the shaft
soils at elevated stresses. Predictive Soil Mechanics: resistance of jacked piles in dense sand. Géotechnique, 69(1),
Proceedings of the Wroth Memorial Symposium (pp. 186– 16–28.
198). Thomas Telford. Ganju, E., Han, F., Prezzi, M., & Salgado, R. (2020). Static
Cressey, G. B. (1928). The Indiana sand dunes and shorelines of capacity of closed-ended pipe pile driven in gravelly sand.
the Lake Michigan basin. University of Chicago Press. Journal of Geotechnical and Geoenvironmental Engineering,
Dafalias, Y. F., Manzari, M. T., & Papadimitriou, A. G. 146(4), 04020008.
(2006). SANICLAY: Simple anisotropic clay plasticity Ganju, E., Han, F., Prezzi, M., Salgado, R., & Pereira, J. S.
model. International Journal for Numerical and Analytical (2020). Quantification of displacement and particle crush-
Methods in Geomechanics, 30(12), 1231–1257. ing around a penetrometer tip. Geoscience Frontiers, 11(2),
Dagger, R., Saftner, D., & Mayne, P. W. (2018, November). 389–399.
Cone penetration test design guide for state geotechnical Ganju, E., Prezzi, M., & Salgado, R. (2017). Algorithm for
engineers (Report No. MN/RC 2018-32). University of generation of stratigraphic profiles using cone penetration
Minnesota. test data. Computers and Geotechnics, 90, 73–84.
Davis, E. H., & Booker, J. R. (1973, December). The effect of Ganju, E., Salgado, R., & Prezzi, M. (2019). Site variability
increasing strength with depth on the bearing capacity of characterization using cone penetration test data. Proceed-
clays. Géotechnique, 23(4), 551–563. ings of Geo-Congress 2019—Eighth International Conference
DeJong, J. T., Jaeger, R. A., Boulanger, R. W., Randolph, M. on Case Histories in Geotechnical Engineering (pp. 152–157).
F., & Wahl, D. A. J. (2013). Variable penetration rate cone Gao, Z., Zhao, J., Li, X.-S., & Dafalias, Y. F. (2014).
testing for characterization of intermediate soils. In R. Q. A critical state sand plasticity model accounting for fabric
Coutinho & P. W. Mayne (Eds.), Geotechnical and evolution. International Journal for Numerical and Analyti-
Geophysical Site Characterization 4, 1, 25–42. cal Methods in Geomechanics, 38(4), 370–390.
DeJong, J. T., & Randolph, M. (2012). Influence of partial Gasparre, A. (2005, July). Advanced laboratory characterisa-
consolidation during cone penetration on estimated soil tion of London clay [Doctoral dissertation, Imperial
behavior type and pore pressure dissipation measurements. College]. http://hdl.handle.net/10044/1/45389
Journal of Geotechnical and Geoenvironmental Engineering, Gavin, K., Adekunte, A., & O’Kelly, B. (2009). A field investi-
138(7), 777–788. gation of vertical footing response on sand. Proceedings of
Finno, R. J. (1989). Subsurface conditions and pile installation the ICE–Geotechnical Engineering, 162(5), 257–267.
data: 1989 foundation engineering congress test section. Gens, A. (1982). Stress-strain and strength of a low plasticity
Proceedings of Symposium on Predicted and Observed Axial clay [Doctoral dissertation, Imperial College]. http://hdl.
Behavior of Piles, GSP 23, 1–74. American Society of Civil handle.net/10044/1/8410
Engineers. Giroud, J.-P. (1968). Settlement of a linearly loaded rectan-
Fleming, A. H., Brown, S. E., & Ferguson, V. R. (1993). The gular area. Journal of the Soil Mechanics and Foundations
hydrogeologic framework of Marion County, Indiana. Division, 94(4), 813–831.
Indiana Geological Survey Open-File Report 93. Gooding, A. M. (1973). Characteristics of late Wisconsinan
Florea, L. J., Hasenmueller, N. R., Branam, T. D., Frushour, tills in eastern Indiana (Department of Natural Resources
S. S., & Powell, R. L. (2018). Karst geology and hydro- Geological Survey Bulletin 49). Indiana Geological Survey.
geology of the Mitchell Plateau of south-central Indiana. Gray, H. (n.d.). Loess [Webpage]. Indiana Geological and
Ancient Oceans, Orogenic Uplifts, and Glacial Ice: Geologic Water Survey. https://igws.indiana.edu/surficial/loess
Crossroads in America’s Heartland, Geological Society of Gray, H. H. (2000). Physiographic divisions of Indiana
America Field Guide 51, 95–112. Geological Society of (Indiana Geological Survey Special Report 61). Indiana
America. Geological Survey.
Foye, K. C. (2005). A rational, probabilistic method for the Hall, R. D., & Anderson, A. K. (2000). Comparative soil
development of geotechnical load and resistance factor design development of Quaternary paleosols of the central United
[Doctoral dissertation, Purdue University]. Purdue e-Pubs. States. Palaeogeography, Palaeoclimatology, Palaeoecology,
Foye, K. C., Abou-Jaoude, G., Prezzi, M., & Salgado, R. 158(1–2), 109–145.
(2009). Resistance factors for use in load and resistance Han, F., Ganju, E., Prezzi, M., Salgado, R., & Zaheer, M.
factor design of driven pipe piles in sands. Journal (2020). Axial resistance of open-ended pipe pile driven in
of Geotechnical and Geoenvironmental Engineering, 135(1), gravelly sand. Géotechnique, 70(2), 138–152.
1–13. Han, F., Ganju, E., Salgado, R., & Prezzi, M. (2018). Effects
Foye, K. C., Basu, P., & Prezzi, M. (2008, September). of interface roughness, particle geometry, and gradation on
Immediate settlement of shallow foundations bearing on the sand-steel interface friction angle. Journal of Geotech-
clay. International Journal of Geomechanics, 8(5), 300–310. nical and Geoenvironmental Engineering, 144(12), 04018096.

64 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


Han, F., Ganju, E., Salgado, R., & Prezzi, M. (2019a). Kilibarda, Z., & Shillinglaw, C. (2014). A 70 year history of
Closure to ‘‘Effects of interface roughness, particle geome- coastal dune migration and beach erosion along the
try, and gradation on the sand–steel interface friction southern shore of Lake Michigan. Aeolian Research, 17,
angle’’ by Fei Han, Eshan Ganju, Rodrigo Salgado, and 263–273.
Monica Prezzi. Journal of Geotechnical and Geoenviron- Kim, D., Chung, M., & Kwak, K. (2011). Resistance factor
mental Engineering, 145(11), 07019017. calculations for LRFD of axially loaded driven piles in
Han, F., Ganju, E., Salgado, R., & Prezzi, M. (2019b). sands. KSCE Journal of Civil Engineering, 15(7), 1185–1196.
Comparison of the load response of closed-ended and open- Kim, D., & Kang, S.-S. (2013). Engineering properties of
ended pipe piles driven in gravelly sand. Acta Geotechnica, compacted loesses as construction materials. KSCE Journal
14, 1785–1803. of Civil Engineering, 17(2), 335–341.
Han, F., Lim, J., Salgado, R., Prezzi, M., & Zaheer, M. Kim, D., & Lee, J. (2012). Resistance factor contour plot
(2015). Load and resistance factor design of bridge founda- analyses of load and resistance factor design of axially-
tions accounting for pile group–soil interaction (Joint Trans- loaded driven piles in clays. Computers and Geotechnics, 44,
portation Research Program Publication No. FHWA/IN/ 9–19.
JTRP-2015/24) West Lafayette, IN: Purdue University. Kim, D., & Salgado, R. (2012a). Load and resistance fac-
http://dx.doi.org/10.5703/1288284316009 tors for external stability checks of mechanically stabilized
Han, F., Salgado, R., Prezzi, M., & Lim, J. (2017, March). earth walls. Journal of Geotechnical and Geoenvironmental
Shaft and base resistance of non-displacement piles in sand. Engineering, 138(3), 241–251.
Computers and Geotechnics, 83, 184–197. Kim, D., & Salgado, R. (2012b). Load and resistance factors
Han, F., Salgado, R., Prezzi, M., & Lim, J. (2019). Axial for internal stability checks of mechanically stabilized earth
resistance of non-displacement pile groups in sand. Journal walls. Journal of Geotechnical and Geoenvironmental
of Geotechnical and Geoenvironmental Engineering, 145(7), Engineering, 138(8), 910–921.
04019027. Kim, K. K., Prezzi, M., & Salgado, R. (2006). Interpretation of
Hasenmueller, N. R., & Packman, D. M. (n.d.). Karst features cone penetration tests in cohesive soils (Joint Transportation
in Indiana [Webpage]. Indiana Geological and Water Research Program Publication FHWA/IN/JTRP-2006/22).
Survey. https://igws.indiana.edu/bedrock/karst West Lafayette, IN: Purdue University. https://doi.org/10.
Herle, I., & Gudehus, G. (1999, September). Determination of 5703/1288284313387
parameters of a hypoplastic constitutive model from Kim, K., Prezzi, M., Salgado, R., & Lee, W. (2008, August).
properties of grain assemblies. Mechanics of Cohesive- Effect of penetration rate on cone penetration resistance
Frictional Materials, 4(5), 461–486. in saturated clayey soils. Journal of Geotechnical and
Geoenvironmental Engineering, 134(8), 1142–1153.
Hildenbrand, T. G., & Ravat, D. (1997). Geophysical setting
Kirkgard, M., & Lade, P. (1991). Anisotropy of normally
of the Wabash Valley Fault System. Seismological Research
consolidated San Francisco Bay Mud. Geotechnical Testing
Letters, 68(4), 567–585.
Journal, 14(3), 231–246.
Hill, J. R. (1974). The Indiana dunes–legacy of sand (Indiana
Krinitzsky, E. L., & Turnbull, W. J. (1967). Loess deposits of
Geological Survey Special Report 8). Indiana Geological
Mississippi (Geological Society of America Special Paper
Survey.
No. 94). Geological Society of America.
Hogentogler & Co. Inc. (2004). CPT brochure 2004.
Krumbein, W. C., & Sloss, L. L. (1951). Stratigraphy and
Hryciw, R. D., Zheng, J., & Shetler, K. (2016). Particle sedimentation. W. H. Freeman and Company.
roundness and sphericity from images of assemblies by Kulhawy, F. H., & Mayne, P. W. (1990, August). Manual on
chart estimates and computer methods. Journal of Geotech- estimating soil properties for foundation design (Report
nical and Geoenvironmental Engineering, 142(9), 04016038. EPRI EL-6800). Cornell University Geotechnical Engineer-
IGWS. (n.d.). Bedrock geology of Indiana [Webpage]. Indiana ing Group.
Geological and Water Survey. https://igws.indiana.edu/ Ladd, C. C., & Edgers, L. (1972, July). Consolidated-undrained
bedrock direct-simple shear tests on saturated clay (Research Report
Ireland, H. O., Moretto, O., & Vargas, M. (1970, June). No. R72-82). MIT Department of Civil Engineering.
The dynamic penetration test: A standard that is not Ladd, C. C., Foott, R., Ishihara, K., Schlosser, F., & Poulos,
standardized. Géotechnique, 20(2), 185–192. H. G. (1977). Stress deformation and strength character-
Jamiolkowski, M., Ladd, C. C., Germaine, J. T., & istics. Proceedings of the 9th International Conference on
Lancellotta, R. (1985, August). New developments in field Soil Mechanics and Foundation Engineering, 2, 421–494.
and lab testing of soils. Proceedings of the 11th International Ladd, C. C., & Varallyay, J. (1965, July 1). The influence of
Conference on Soil Mechanics and Foundation Engineering the stress system on the behavior of saturated clays during
(pp. 57–153). A. A. Balkema. undrained shear (Research Report No. R65-11). Depart-
Jardine, R., Chow, F., Overy, R., & Standing, J. (2005). ICP ment of Civil Engineering, MIT.
design methods for driven piles in sands and clays. Thomas Lee, J., Eun, J., Prezzi, M., & Salgado, R. (2008). Strain
Telford Publishing. influence diagrams for settlement estimation of both isola-
Jovičić, V., & Coop, M. R. (1997, June). Stiffness of coarse- ted and multiple footings in sand. Journal of Geotechnical
grained soils at small strains. Géotechnique, 47(3), 545–561. and Geoenvironmental Engineering, 134(4), 417–427.
Kassab, C. M., Brickles, S. L., Licht, K. J., & Monaghan, G. Lee, J., & Salgado, R. (2002). Estimation of footing settlement
W. (2017). Exploring the use of zircon geochronology as an in sand. International Journal of Geomechanics, 2(1), 1–28.
indicator of Laurentide Ice Sheet till provenance, Indiana, Lee, J., & Salgado, R. (2005, April). Estimation of bearing
USA. Quaternary Research, 88(3), 525–536. https://doi.org/ capacity of circular footings on sands based on cone pene-
10.1017/qua.2017.71 tration test. Journal of Geotechnical and Geoenvironmental
Kilibarda, Z., & Blockland, J. (2011). Morphology and origin Engineering, 131(4), 442–452.
of the Fair Oaks Dunes in NW Indiana, USA. Geomor- Lehane, B., Liu, Z., Bittar, E., Nadim, F., Lacasse, S., Jardine,
phology, 125(2), 305–318. R., Carotenuto, P., Rattley, M., Jeanjean, P., Gavin, K.,

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 65


Gilbert, R., Bergan-Haavik, J., & Morgan, N. (2020). In Z. dissertation, Purdue University]. Purdue e-Pubs. https://
Westgate (Ed.), A new ‘unified’ CPT-based axial pile capa- docs.lib.purdue.edu/dissertations/AAI3232127/
city design method for driven piles in sand. Proceedings of Loukidis, D., & Salgado, R. (2009). Modeling sand response
4th International Symposium on Frontiers in Offshore using two-surface plasticity. Computers and Geotechnics,
Geotechnics (ISFOG 2020) (pp. 462–477). 36(1–2), 166–186.
Lehane, B. M. (2013). Relating foundation capacity in sands Loukidis, D., & Salgado, R. (2011). Effect of relative density
to CPT qc. In R. Q. Coutinho & P. W. Mayne (Eds.), and stress level on the bearing capacity of footings on sand.
Proceedings of the Fourth International Conference on Géotechnique, 61(2), 107–119.
Cite Characterization: Geotechnical and Geophysical Site Lunne, T., Robertson, P. K., & Powell, J. J. M. (1997). Cone
Characterization 4, 1, 63–81. Taylor & Francis. Penetration Testing in Geotechnical Practice. Blakie
Lehane, B. M. (2019). CPT-based design of foundations. Academic and Professional.
Australian Geomechanics Journal, 54(4), 23–48. Lupini, J. F. (1980). The residual strength of soils [Doctoral
Lehane, B. M., Jardine, R. J., Bond, A. J., & Frank, R. (1993). dissertation, Imperial College]. http://hdl.handle.net/10044/
Mechanisms of shaft friction in sand from instrumented pile 1/11442
tests. Journal of Geotechnical Engineering, 119(1), 19–35. Lupini, J. F., Skinner, A. E., & Vaughan, P. R. (1981). The
Lehane, B. M., Li, Y., & Williams, R. (2013, February). Shaft drained residual strength of cohesive soils. Géotechnique,
capacity of displacement piles in clay using the cone 31(2), 181–213.
penetration test. Journal of Geotechnical and Geoenviron- Lyamin, A. V., Salgado, R., Sloan, S. W., & Prezzi, M. (2007).
mental Engineering, 139(2), 253–266. Two- and three-dimensional bearing capacity of footings in
Lehane, B. M., Lim, J. K., Carotenuto, P., Nadim, F., sand. Géotechnique, 57(8), 647–662.
Lacasse, S., Jardine, R. J., & van Dijk, B. (2017, Madhav, M. R., & Abhishek, S. V. (2016). Ground versus soil:
September). Characteristics of unified databases for driven A new paradigm in geotechnical engineering education.
piles. Proceedings of the 8th International Conference of Proceedings of the International Conference on Geo-
Offshore Site Investigation and Geotechnics: Smarter Solu- Engineering Education (TC 306)–Shaping the Future of
tions for Offshore Developments, 1, 162–191. Society for
Geotechnical Education (SFGE). https://doi.org/10.20906/
Underwater Technology (SUT).
CPS/SFGE-09-0002
Lehane, B. M., Schneider, J. A., & Xu, X. (2005). The UWA-
Madhav, M. R., & Abhishek, S. V. (2017). A non-mechanistic
05 method for prediction of axial capacity of driven piles in
perspective of geotechnical engineering. Proceedings of
sand. In M. J. Cassidy & S. Gourvenec (Eds.), Proceedings
the International Convention on Civil Engineering (ICCE)
of the International Symposium on Frontiers in Offshore
(pp. 7–18).
Geotechnics (pp. 683–689).
Madhira, M., & Sakleshpur, V. A. (2018). Geotechnics of soft
Lehane, B. M., Schneider, J. A., & Xu, X. (2007). CPT-based
ground. In A. Krishna, A. Dey, & S. Sreedeep (Eds.),
design of displacement piles in siliceous sands. In Y.
Geotechnics for Natural and Engineered Sustainable
Kikuchi, M. Kimura, J. Otani, & Y. Morikawa (Eds.),
Technologies (pp. 27–44). Springer.
Advances in Deep Foundations (pp. 69–86). CRC Press.
Madhira, M., & Sakleshpur, V. A. (2019). Mining geotechni-
Lings, M. L., & Dietz, M. S. (2004, May). An improved direct
shear apparatus for sand. Géotechnique, 54(4), 245–256. cal parameters from the ground. 8th Annual Praphulla
Kumar Lecture (pp. 1–35). Indian Geotechnical Society
Liu, Q., & Lehane, B. M. (2021, May). A centrifuge investi-
(Kochi Chapter).
gation of the relationship between the vertical response of
footings on sand and CPT end resistance. Géotechnique, Maksimović, M. (1989). On the residual shearing strengh of
71(5), 455–465. clays. Géotechnique, 39(2), 347–351.
Logan, W. N., Cumings, E. R., Malott, C. A., Visher, S. S., Malott, C. A. (1922). The physiography of Indiana. In W. N.
Tucker, W. M., & Reeves, J. R. (1922). Handbook of Logan, E. R. Cumings, C. A. Malott, S. S. Visher, W. M.
Indiana Geology (Publication No. 21). Department of Tucker, & J. R. Reeves (Eds.), Handbook of Indiana
Conservation. Geology (Publication No. 21, Part 2, pp. 59–256). Indiana
Look, B. (2016). The SPT N-value errors examined with Department of Conservation.
digital technology. In B. M. Lehane, H. E. Acosta- Martin, C. M. (2004, October). User guide for ABC: Analysis
Martı́nez, & R. Kelley (Eds.), Geotechnical and Geo- of bearing capacity (OUEL Report No. 2261/03). University
physical Site Characterization 5, 1, 333–338. Australian of Oxford.
Geomechanics Society. Martin, C. M. (2005). Exact bearing capacity calculations
Look, B. G., Seidel, J. P., Sivakumar, S. T., & Welikala, D. L. using the method of characteristics. Proceedings of 11th
C. (2015). Real time monitoring of SPT using a PDM International Conference of IACMAG, 4, 441–450.
device–the failings of our standard test revealed. Proceed- Mayne, P. W. (2007). Cone penetration testing: A synthesis of
ings of International Conference on Geotechnical Engineering highway practice (NCHRP Synthesis Report 368). Trans-
(ICGE) (pp. 435–438). portation Research Board.
Loope, H. M., Antinao, J. L., Monaghan, G. W., Autio, R. J., Mayne, P. W., & Campanella, R. G. (2005). Versatile site
Curry, B. B., Grimley, D. A., Huot, S., Lowell, T. V., & characterization by seismic piezocone. Proceedings of 16th
Nash, T. A. (2018). At the edge of the Laurentide ice sheet: International Conference on Soil Mechanics and Geotech-
Stratigraphy and chronology of glacial deposits in central nical Engineering (pp. 721–724). International Society for
Indiana. In L. J. Florea (Ed.), Ancient Oceans, Orogenic Soil Mechanics and Geotechnical Engineering.
Uplifts, and Glacial Ice: Geologic Crossroads in America’s Mayne, P. W., & Dasenbrock, D. (2018). Direct CPT method
Heartland, Geological Society of America Field Guide, 51, for 130 footings on sands. In X. Zhang, P. J. Cosentino, &
245–258. M. H. Hussein (Eds.), Proceedings of IFCEE 2018: Inno-
Loukidis, D. (2006). Advanced constitutive modeling of sands vations in Geotechnical Engineering, GSP 299, 135–146.
and applications to foundation engineering [Doctoral American Society of Civil Engineers.

66 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


Mayne, P. W., & Harris, D. E. (1993). Axial load-displacement resistance factor design (LRFD) for deep foundations
behavior of drilled shaft foundations in Piedmont residuum (NCHRP Report 507). Transportation Research Board.
(Report No. 41-30-2175). Federal Highway Administration. Parry, R. H. G. (1960, December). Triaxial compression and
Mayne, P. W., & Peuchen, J. (2018). Evaluation of CPTU Nkt extension tests on remoulded saturated clay. Géotechnique,
cone factor for undrained strength of clays. In M. A. Hicks, 10(4), 166–180.
F. Pisanò, & J. Peuchen (Eds.), Proceedings of the 4th Phoon, K.-K., & Kulhawy, F. H. (1999a). Characterization of
International Symposium on Cone Penetration Testing (pp. geotechnical variability. Canadian Geotechnical Journal,
423–429). CRC Press. 36(4), 612–624.
Mayne, P. W., Uzielli, M., & Illingworth, F. (2012). Shallow Phoon, K.-K., & Kulhawy, F. H. (1999b). Evaluation of
footing response on sands using a direct method based on geotechnical property variability. Canadian Geotechnical
cone penetration tests. Proceedings of GeoCongress 2012 Journal, 36(4), 625–639.
(pp. 664–679). American Society for Civil Engineers. Powers, M. C. (1953, June). A new roundness scale for
Mayne, P. W., & Woeller, D. J. (2014). Generalized direct sedimentary particles. Journal of Sedimentary Research,
CPT method for evaluating footing deformation response 23(2), 117–119.
and capacity on sands, silts, and clays. Proceedings of Geo- Prandtl, L. (1920). Über die Härte Plastischer Körper.
Congress 2014, GSP 234, 1983–1997. Nachrichten von der Gesellschaft der Wissenschaften zu
McKnight, T., Cho, Y. M., Townsend, T. G., & Choate, A. Göttingen, Mathematisch-Physikalische Klasse, 12, 74–85.
(2015). Cone penetration testing for characterizing land- Prandtl, L. (1921). Eindringungsfestigkeit und Festigkeit von
filled municipal solid waste. Journal of Geotechnical and Schneiden. Zeitschrift für Angewandte Mathematik und
Geoenvironmental Engineering, 141(3), 06014018. Mechanik, 1(1), 15–20.
Meehan, C. L. (2006). An experimental study of the dynamic Rahman, S., Salgado, R., Prezzi, M., & Becker, P. J. (2020).
behavior of slickensided surfaces [Doctoral dissertation, Improvement of stiffness and strength of backfill soils through
Virginia Polytechnic Institute and State University]. optimization of compaction procedures and specifications
VTechWorks. http://hdl.handle.net/10919/26074 (Joint Transportation Research Program Publication No.
Meyerhof, G. G. (1956). Penetration tests and bearing capa- FHWA/IN/JTRP-2020/16). West Lafayette, IN: Purdue
city of cohesionless soils. Journal of the Soil Mechanics and University. https://doi.org/10.5703/1288284317134
Foundations Division, 82(1), 1–19. Ramsey, N., Jardine, R., Lehane, B., & Ridley, A. (1998).
Mitchell, J. K. (1988, March). New developments in penetra- A review of soil-steel interface testing with the ring shear
tion tests and equipment. In J. de Ruiter (Ed.), Proceedings apparatus. Offshore Site Investigation and Foundation
of the 1st International Symposium on Penetration Testing, Behaviour: New Frontiers–Proceedings of an International
Conference, 1, 237–258. Society for Underwater Tech-
245–261. A. A. Balkema.
nology (SUT).
Mitchell, J. K., & Soga, K. (2005). Fundamentals of soil
Randolph, M. F. (2003, December). Science and empiricism in
behavior (3rd ed.). John Wiley & Sons.
pile foundation design. Géotechnique, 53(10), 847–875.
Mondelli, G., Giacheti, H., & Howie, J. (2010). Interpretation
Randolph, M. F., Dolwin, J., & Beck, R. (1994). Design of
of resistivity piezocone tests in a contaminated municipal
driven piles in sand. Géotechnique, 44(3), 427–448.
solid waste disposal site. Geotechnical Testing Journal,
Reissner, H. (1924). Zum erddruckproblem. Proceedings of
33(2), 123–136.
the 1st International Congress of Applied Mechanics (pp.
Munfakh, G., Arman, A., Collin, J. G., Hung, J. C. J., &
295–311).
Brouillette, R. P. (2001). Shallow foundations reference
René, R. M., & Stanonis, F. L. (1995). Reflection seismic
manual (Publication No. FHWA-NHI-01-023). Federal
profiling of the Wabash Valley fault system in the Illinois
Highway Administration.
Basin (U.S. Geological Survey Professional Paper 1538-O).
Murthy, T. G., Loukidis, D., Carraro, J. A. H., Prezzi, M., & In K. M. Shedlock & A. C. Johnston (Eds.), Investigations
Salgado, R. (2007). Undrained monotonic response of clean of the New Madrid Seismic Zone.
and silty sands. Géotechnique, 57(3), 273–288. Riemer, M. F., Seed, R. B., Nicholson, P. G., & Jong, H.-L.
Niazi, F. S., & Mayne, P. W. (2016). CPTu-based enhanced (1990). Steady state testing of loose sands: Limiting mini-
UniCone method for pile capacity. Engineering Geology, mum density. Journal of Geotechnical Engineering, 116(2),
212, 21–34. 332–337.
Niazi, F. S., Mayne, P. W., & Woeller, D. J. (2010). Drilled Robertson, P. K. (1990). Soil classification using the cone
shaft O-cell response at Golden Ears Bridge from seismic penetration test. Canadian Geotechnical Journal, 27(1), 151–
piezocone tests. The Art of Foundation Engineering Practice 158.
Congress (pp. 452–469). http://dx.doi.org/10.1061/41093 Robertson, P. K. (2016). Cone penetration test (CPT)-based
(372)22 soil behaviour type (SBT) classification system—an update.
Nishimura, S. (2005). Laboratory study on anisotropy of natural Canadian Geotechnical Journal, 53(12), 1910–1927.
London clay [Doctoral dissertation, Imperial College]. Robertson, P. K., Campanella, R. G., Gillespie, D., & Greig,
Nottingham, L. C., & Schmertmann, J. H. (1975). An investi- J. (1986, June). Use of piezometer cone data. Use of In-Situ
gation of pile capacity design procedures (Engineering and Testing in Geotechnical Engineering, GSP 6, 1263–1280.
Industrial Experiment Station Final Report D629). American Society of Civil Engineers.
University of Florida Department of Civil Engineering. Robertson, P. K., Campanella, R. G., & Wightman, A.
Paik, K., Salgado, R., Lee, J., & Kim, B. (2003, April). (1983). SPT-CPT correlations. Journal of Geotechnical
Behavior of open- and closed-ended piles driven into sands. Engineering, 109(11), 1449–1459.
Journal of Geotechnical and Geoenvironmental Engineering, Rodrı́guez, J. M., Johansson, J. M. A., & Edeskär, T. (2012).
129(4), 296–306. Particle shape determination by two-dimensional image
Paikowsky, S. G., Birgisson, B., McVay, M., Nguyen, T., analysis in geotechnical engineering. Proceedings of Nordic
Kuo, C., Baecher, G., Ayyub, B., Stenersen, K., O’Malley, Conference on Soil Mechanics and Geotechnics (pp. 207–
K., Chernauskas, L., & O’Neill, M. (2004). Load and 218). Danish Geotechnical Society.

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 67


de Ruiter, J., & Beringen, F. L. (1979). Pile foundations for Shaw, E. W. (1915). On the origin of the Loess of
large North Sea structures. Marine Geotechnology, 3(3), southwestern Indiana. Science, 41(1046), 104–108.
267–314. Skempton, A. W. (1951). The bearing capacity of clays.
Rupp, J. A. (1991). Structure and isopach maps of the Building Research Congress, 1, 180–189.
Paleozoic rocks of Indiana (Indiana Geological Survey Skempton, A. W. (1985, March). Residual strength of clays in
Special Report 48). Indiana Geological Survey. landslides, folded strata and the laboratory. Géotechnique,
Rutledge, E. M., Guccione, M. J., Markewich, H. W., 35(1), 3–18.
Wysocki, D. A., & Ward, L. B. (1996, December). Loess Skempton, A. W. (1986). Standard penetration test proce-
stratigraphy of the Lower Mississippi Valley. Engineering dures and the effects in sands of overburden pressure,
Geology, 45(1–4), 167–183. relative density, particle size, ageing and overconsolidation.
Salgado, R. (2006). The role of analysis in non-displacement Géotechnique, 36(3), 425–447.
pile design. In W. Wu & H. S. Yu (Eds.), Modern Trends in Skempton, A. W., & Bjerrum, L. (1957). A contribution to the
Geomechanics, 106, 521–540. Springer. settlement analysis of foundations on clay. Géotechnique,
Salgado, R. (2008). The engineering of foundations (1st ed.). 7(4), 168–178.
McGraw-Hill. Skempton, A. W., & MacDonald, D. H. (1956). The allowable
Salgado, R. (2013). The mechanics of cone penetration: settlements of buildings. Proceedings of the Institution of
contributions from experimental and theoretical studies. Civil Engineers, 5(6), 727–768.
R. Q. Coutinho & P. W. Mayne (Eds.), Geotechnical and Sladen, J. A., D’Hollander, R. D., & Krahn, J. (1985). The
Geophysical Site Characterization 4, 1, 131–153. CRC Press. liquefaction of sands, a collapse surface approach.
Salgado, R. (2014). Experimental research on cone penetra- Canadian Geotechnical Journal, 22(4), 564–578.
tion resistance. Geo-Congress 2014 Keynote Lecture, GSP Song, C. R., Kim, S., Bekele, B., Zhang, J., & Silvey, A.
235, 140–163. American Society of Civil Engineers. (2019). CPT-based pile design (Report No. SPR-P1 M076).
Salgado, R., Bandini, P., & Karim, A. (2000). Shear strength Nebraska Department of Transportation.
and stiffness of silty sand. Journal of Geotechnical and Sukumaran, B., & Ashmawy, A. K. (2001). Quantitative
Geoenvironmental Engineering, 126(5), 451–462. characterisation of the geometry of discret particles.
Géotechnique, 51(7), 619–627.
Salgado, R., Ganju, E., & Prezzi, M. (2019). Site variability
Tehrani, F. S., Arshad, M. I., Prezzi, M., & Salgado, R.
analysis using cone penetration test data. Computers and
(2018). Physical modeling of cone penetration in layered
Geotechnics, 105, 37–50.
sand. Journal of Geotechnical and Geoenvironmental
Salgado, R., & Kim, D. (2014). Reliability analysis of load
Engineering, 144(1), 04017101.
and resistance factor design of slopes. Journal of Geo-
Terzaghi, K. (1943). Theoretical soil mechanics. John Wiley &
technical and Geoenvironmental Engineering, 140(1), 57–73.
Sons.
Salgado, R., Lyamin, A. V., Sloan, S. W., & Yu, H. S. (2004).
Terzaghi, K., Peck, R. B., & Mesri, G. (1996, February). Soil
Two-and three-dimensional bearing capacity of founda-
mechanics in engineering practice (3rd ed.). Wiley.
tions in clay. Géotechnique, 54(5), 297–306.
Thurairajah, A. (1962). Some shear properties of kaolin and of
Salgado, R., & Prezzi, M. (2007). Computation of cavity sand [Doctoral dissertation, University of Cambridge].
expansion pressure and penetration resistance in sands. Tippett, L. H. C. (1925). On the extreme individuals and the
International Journal of Geomechanics, 7(4), 251–265. range of samples taken from a normal population.
Salgado, R., & Prezzi, M. (2014, September–December). Biometrika, 17(3/4), 364–387. Oxford University Press.
Penetration rate effects on cone resistance: Insights from Tsomokos, A., & Georgiannou, V. N. (2010). Effect of grain
calibration chamber and field testing. Soils and Rocks, shape and angularity on the undrained response of fine
37(3), 233–242. sands. Canadian Geotechnical Journal, 47(5), 539–551.
Salgado, R., Prezzi, M., & Ganju, E. (2015). Assessment of Tumay, M. (1985). Field calibration of electric cone penetrom-
site variability from analysis of cone penetration test data eters in soft soils (Publication No. FHWA/LA/LSU-GE-85/
(Joint Transportation Research Program Publication No. 02). Federal Highway Administration.
FHWA/IN/JTRP-2015/04). West Lafayette, IN: Purdue Tumay, M. T., & Fakhroo, M. (1982). Friction pile capacity
University. https://doi.org/10.5703/1288284315523 prediction in cohesive soils using electric quasi-static penetra-
Salgado, R., Woo, S. I., & Kim, D. (2011). Development of tion tests (Interim Research Report No. 1). Louisiana
load and resistance factor design for ultimate and service- Department of Transportation and Development, Research
ability limit states of transportation structure foundations and Development.
(Joint Transportation Research Program Publication No. USACE. (2001, January 1). Engineer manual: Engineering and
FHWA/IN/JTRP-2011/03). West Lafayette, IN: Purdue design – geotechnical investigations (Publication No. EM
University. https://doi.org/10.5703/1288284314618 1110-1-1804). U.S. Army Corps of Engineers.
Schmertmann, J. H. (1970). Static cone to compute static USGS. (n.d.). United states geological survey – earthquake
settlement over sand. Journal of the Soil Mechanics and hazards program [Data set]. https://earthquake.usgs.gov/
Foundations Division, ASCE, 96(3), 1011–1043. research/cpt/data/txt/VHC028.txt
Schmertmann, J. H. (1978, July). Guidelines for cone penetra- Uthayakumar, M., & Vaid, Y. P. (1998). Static liquefaction of
tion test (performance and design) (Report No. FHWA-TS- sands under multiaxial loading. Canadian Geotechnical
78-209). Federal Highway Administration. Journal, 35(2), 273–283.
Schmertmann, J. H., Hartmann, J. P., & Brown, P. R. (1978). Verdugo, R., & Ishihara, K. (1996). The steady state of sandy
Improved strain influence factor diagrams. Journal of the soils. Soils and Foundations, 36(2), 81–91.
Geotechnical Engineering Division, 104(8), 1131–1135. Viggiani, G., & Atkinson, J. H. (1995). Stiffness of fine-
Schneider, J. A., Xu, X., & Lehane, B. M. (2008). Database grained soil at very small strains. Géotechnique, 45(2), 249–
assessment of CPT-based design methods for axial capacity 265.
of driven piles in siliceous sands. Journal of Geotechnical Wadell, H. (1932). Volume, shape, and roundness of rock
and Geoenvironmental Engineering, 134(9), 1227–1244. particles. The Journal of Geology, 40(5), 443–451.

68 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


Wadell, H. (1933). Sphericity and roundness of rock particles. Wroth, C. P. (1979). Correlations of some engineer-
The Journal of Geology, 41(3), 310–331. ing properties of soils. Proceedings of the 2nd Interna-
Walker, G., Jefferies, N., Lennard, M., & Lightfoot, J. (2009). tional Conference on Behaviour of Offshore Structures, 1,
Cone penetration testing of radiologically contaminated 121–132.
burial trenches. Proceedings of the 12th International Wroth, C. P. (1984). The interpretation of in situ soil tests.
Conference on Environmental Remediation and Radioactive Géotechnique, 34(4), 449–489.
Waste Management, 475–486. Wroth, C. P., & Wood, D. M. (1978). The correlation of index
Wayne, W. J. (1965). The Crawfordsville and Knightstown properties with some basic engineering properties of soils.
moraines in Indiana (Indiana Geological Survey Report of Canadian Geotechnical Journal, 15(2), 137–145.
Progress 28). Indiana Geological Survey. Xiao, T., Li, D.-Q., Cao, Z.-J., & Zhang, L.-M. (2018). CPT-
Wayne, W. J., & Thornbury, W. D. (1951). Glacial geology of based probabilistic characterization of three-dimensional
Wabash County, Indiana. (Indiana Geological Survey spatial variability using MLE. Journal of Geotechnical and
Bulletin 5). Indiana Geological Survey. Geoenvironmental Engineering, 144(5), 04018023.
Weary, D. J., & Doctor, D. H. (2014). Karst in the United Xu, X. (2007). Investigation of the end bearing performance of
States: A digital map compilation and database (U.S. displacement piles in sand [Doctoral dissertation, The
Geological Survey Open-File Report 2014-1156). https:// University of Western Australia].
pubs.er.usgs.gov/publication/ofr20141156#:,:text5https% Xu, X., & Lehane, B. M. (2008, April). Pile and penetrometer
3A//doi.org/10.3133/ofr20141156 end bearing resistance in two-layered soil profiles. Géotech-
West, T. R. (2010). Geology applied to engineering. Waveland nique, 58(3), 187–197.
Press. Xu, X., Schneider, J. A., & Lehane, B. M. (2008). Cone
White, D. J., & Lehane, B. M. (2004). Friction fatigue on penetration test (CPT) methods for end-bearing assessment
displacement piles in sand. Géotechnique, 54(10), 645–658. of open- and closed-ended driven piles in siliceous sand.
White, W. B. (1988, May). Geomorphology and hydrology of Canadian Geotechnical Journal, 45(8), 1130–1141.
karst terrains. Oxford University Press. Yang, J., & Wei, L. M. (2012). Collapse of loose sand with the
Wilcox, D. A., Shedlock, R. J., & Hendrickson, W. H. (1986). addition of fines: the role of particle shape. Géotechnique,
Hydrology, water chemistry and ecological relations in the 62(12), 1111–1125.
raised mound of Cowles Bog. Journal of Ecology, 74(4), Yang, Z. X., Jardine, R. J., Zhu, B. T., Foray, P., & Tsuha, C.
1103–1117. H. C. (2010). Sand grain crushing and interface shearing
Wilcox, D. A., & Simonin, H. A. (1988, June). The during displacement pile installation in sand. Géotechnique,
stratigraphy and development of a floating peatland, 60(6), 469–482.
Pinhook Bog, Indiana. Wetlands, 8(1), 75–91. Zheng, J., & Hryciw, R. D. (2015). Traditional soil particle
Wilson, J. (2008). How the Ice Age shaped Indiana (2nd ed.). sphericity, roundness and surface roughness by computa-
Wilstar Media. tional geometry. Géotechnique, 65(6), 494–506.
Woolery, E. W., Whitt, J. W., Van Arsdale, R. B., & Zheng, J., & Hryciw, R. D. (2016). Index void ratios of sands
Almayahi, A. (2018). Geophysical and geological evidence from their intrinsic properties. Journal of Geotechnical and
for quaternary displacement on the caborn fault, Wabash Geoenvironmental Engineering, 142(12), 06016019.
Valley fault system, southwestern Indiana. Seismological
Research Letters, 89(6), 2473–2480.

Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23 69


APPENDICES

Appendix A. Critical-State Friction Angle of Sand

Appendix B. OCR and K0 of Soil

Appendix C. Iterative Scheme for Footing Settlement in Sand

Appendix D. Penetration Rate Effect on Cone Resistance

Appendix E. Residual-State Friction Angle of Clay

70 Joint Transportation Research Program Technical Report FHWA/IN/JTRP-2021/23


APPENDIX A. CRITICAL-STATE FRICTION ANGLE OF SAND

The critical-state friction angle ϕc is simply the friction angle that a given soil has at critical
state. It is independent of soil state (i.e., relative density and confining stress) but depends on
particle size (e.g., D50), morphology (e.g., roundness R and sphericity S), mineralogy (e.g., silicates
versus carbonates), and gradation (e.g., coefficient of uniformity CU) (Han et al., 2018; Salgado,
2008). The value of ϕc for a silica sand typically ranges from 28°–36°; sands with rounded, smooth
particles with a poorly-graded particle size distribution have values near the low end of this range,
while sands with angular, rough particles with a well-graded particle size distribution have values
near the high end of this range (Salgado, 2008). In contrast, the value of ϕc for a carbonate sand
typically ranges from 37°–44° (Altuhafi et al., 2016; Coop & Lee, 1993; Salgado, 2008).

A.1 Roundness
Roundness is a measure of sharpness of the particle corners (Figure A.1). It is defined as
the ratio of the average radius of curvature of the corners of a 2D projection of the particle to the
radius rins of the largest inscribed circle for the same projection (Wadell, 1932):
N
ri
N
R i1
(Eq. A.1)
rins
where ri = radius of curvature of corner i of the particle, and N = number of particle corners. Table
A.1 summarizes the different roundness classes proposed by Powers (1953).

Figure A.1 Definition of roundness for a 2D projected outline of a particle (Hryciw et al., 2016;
Wadell, 1932).

A-1
Table A.1 Classification of particles based on roundness (Powers, 1953)
Roundness Class Roundness Interval Mean Roundness1
Very angular 0.12–0.17 0.14
Angular 0.17–0.25 0.21
Subangular 0.25–0.35 0.30
Subrounded 0.35–0.49 0.41
Rounded 0.49–0.70 0.59
Well-rounded 0.70–1.00 0.84
1
Geometric mean

A.2 Sphericity
Sphericity is a measure of the extent to which a particle resembles the shape of a sphere.
Particle sphericity has been defined in several ways in the literature (Mitchell & Soga, 2005;
Rodríguez et al., 2012); three widely used definitions are detailed below.

1. Diameter sphericity SD: It is defined as the ratio of the diameter Dc of a circle having the same
area as the projected 2D area of the particle to the diameter Dcir of the smallest circle
circumscribed about the 2D projection of the particle (Wadell, 1933):
Dc
SD  (Eq. A.2)
Dcir

2. Width-to-length ratio sphericity SWL: It is defined as the ratio of the width d2 to the length d1
of the particle (Zheng & Hryciw, 2015):
d2
SWL  (Eq. A.3)
d1
The length d1 and width d2 of the particle are defined as the largest and smallest dimensions,
respectively, of a rectangle enclosing the particle; the selected rectangle is the one with the
largest possible dimension circumscribing the particle. The reciprocal of the width-to-length
ratio sphericity is commonly referred to as the elongation ratio.

3. Perimeter sphericity SP: It is defined as the ratio of the perimeter Pc of a circle having the same
area as the projected 2D area A of the particle to the projected perimeter P of the particle
(Altuhafi et al., 2013):
Pc 2  A
SP   (Eq. A.4)
P P
Figure A.2 illustrates the definitions of diameter sphericity SD and width-to-length ratio
sphericity SWL. Figure A.3 shows a chart developed by Krumbein and Sloss (1951) with 20
reference particle silhouettes having roundness and sphericity values ranging from 0.1–0.9 and
0.3–0.9, respectively, in increments of 0.2. If access to digital, computer-based tools, such as
ImageJ and MATLAB, is limited, the chart can be used to estimate particle roundness and
sphericity by comparing the shapes of individual particles viewed under a microscope with the

A-2
reference particle silhouettes given in the chart. The sphericity obtained from the Krumbein and
Sloss (1951) chart is the width-to-length ratio sphericity SWL (Zheng & Hryciw, 2015).

(a) (b)

Figure A.2 Illustrations of (a) diameter Dcir of the smallest circle circumscribed about the 2D
projection of the particle, and (b) length d1 and width d2 of the particle.

Figure A.3 Chart for estimating roundness and sphericity (Krumbein & Sloss, 1951).

A.3 Silica Sand Database


Table A.2 summarizes the intrinsic parameters of 23 clean silica sands reported in the
literature. The parameters include mean particle size D50, coefficient of uniformity CU, roundness
R, sphericity S, minimum void ratio emin, maximum void ratio emax, and critical-state friction angle

A-3
ϕc in triaxial compression. All the sands are poorly-graded, except FS Ohio SW, which is classified
as well-graded according to the Unified Soil Classification System (USCS) (ASTM, 2012). The
number designations for some of the uniform sands (e.g., Ottawa 20–30) listed in Table A.2
indicate the sieve numbers between which the sand particles were retained. The D50, CU, and R
values for the sands are in the range of 0.15–2.68 mm (0.006–0.105 in.), 1.2–7.9, and 0.3–0.8,
respectively. Although different researchers have defined particle sphericity in different ways for
the sands listed in Table A.2, the S values were found to lie within a relatively narrow range of
0.65–0.90 regardless of the definition used. Zheng and Hryciw (2016) also found the S values to
lie within a similar range for the sands considered in their database. They reasoned that sand
particles are usually bulky in nature and that slender, elongated sand particles are rarely found in
practice because such particles are susceptible to breakage.

A-4
Table A.2 Intrinsic parameters of 23 clean silica sands reported in the literature
Gradation Morphology Packing Strength
Sand D50 (mm) CU R S emin emax ϕc (°) Reference
FS Ohio 6–10 2.68 1.31 0.43 0.86 0.66 0.92 34.6 Han et al. (2018)
FS Ohio 10–16 1.59 1.30 0.44 0.83 0.65 0.92 33.7 Han et al. (2018)
FS Ohio 16–20 1.01 1.25 0.40 0.78 0.66 0.97 32.9 Han et al. (2018)
FS Ohio 20–40 0.63 1.42 0.39 0.82 0.62 0.91 31.8 Han et al. (2018)
FS Ohio 50–100 0.23 1.56 0.35 0.82 0.63 0.93 31.7 Han et al. (2018)
FS Ohio Coarse 1.50 2.00 — — 0.45 0.72 33.6 Han et al. (2018)
FS Ohio Fine 0.35 2.00 — — 0.48 0.72 33.4 Han et al. (2018)
FS Ohio SW 1.04 7.90 — — 0.37 0.65 33.21 Han et al. (2018)
Fontainebleau NE34 0.21 1.53 0.45 0.752 0.51 0.90 30.0 Altuhafi et al. (2018); Yang et al. (2010);
Zheng & Hryciw (2016)
Fraser River 0.30 2.40 0.43 0.83 0.68 1.00 33.0 Gao et al. (2014); Sukumaran & Ashmawy (2001);
Uthayakumar & Vaid (1998)
Ham River 0.30 1.59 0.45 0.652 0.59 0.92 32.0 Coop & Lee (1993); Jovičić & Coop (1997);
Zheng & Hryciw (2016)
Lausitz 0.25 3.09 0.51 — 0.44 0.85 32.2 Herle & Gudehus (1999); Zheng & Hryciw (2016)
Leighton Buzzard 0.78 1.27 0.75 0.802 0.51 0.80 30.0 Lings & Dietz (2004); Thurairajah (1962);
Zheng & Hryciw (2016)
Longstone 0.15 1.43 0.30 0.652 0.61 1.00 32.5 Tsomokos & Georgiannou (2010);
Zheng & Hryciw (2016)
M31 0.28 1.54 0.62 0.702 0.53 0.87 30.2 Tsomokos & Georgiannou (2010);
Zheng & Hryciw (2016)
Monterey No. 0 0.38 1.58 — 0.893 0.53 0.86 32.8 Altuhafi et al. (2013); Riemer et al. (1990)
Ohio Gold Frac 0.62 1.60 0.43 0.83 0.58 0.87 32.5 Ganju et al. (2020); Han et al. (2018)
Ottawa Graded 0.31 1.89 0.804 0.904 0.49 0.76 29.5 Carraro et al. (2009)
Ottawa 20–30 0.72 1.18 0.72 0.88 0.50 0.74 29.2 Han et al. (2018)
Q-Rok4 0.63 1.50 0.40 0.73 0.70 1.03 33.0 Unpublished research
Sacramento River 0.30 1.80 — 0.883 0.53 0.87 33.2 Altuhafi et al. (2013); Riemer et al. (1990)
Ticino 0.58 1.50 0.40 0.802 0.57 0.93 33.0 Altuhafi et al. (2016); Bellotti et al. (1996);
Cho et al. (2006);
Toyoura 0.17 1.70 0.35 0.652 0.60 0.98 31.6 Loukidis & Salgado (2009); Verdugo & Ishihara
(1996); Zheng & Hryciw (2016)
Note: D50 = mean particle size, CU = coefficient of uniformity (= D60/D10), emin = minimum void ratio, emax = maximum void ratio, R = roundness, S = diameter sphericity SD
(unless otherwise indicated), and ϕc = critical-state friction angle in triaxial compression (unless otherwise indicated).
The properties of INDOT No. 4 sand, which is a backfill material typically used for retaining wall construction in Indiana, are: D50 = 0.85 mm, CU = 4.58, R = 0.72, SWL =
0.73, emin = 0.29, emax = 0.54, and ϕc = 38.0° in direct shear (Rahman et al., 2020).
1
Obtained from direct shear test results.
2
Width-to-length ratio sphericity SWL (Mitchell & Soga, 2005; Zheng & Hryciw, 2015).
3
Perimeter sphericity SP (Altuhafi et al., 2013).
4
Unpublished research.

A-5
A.4 Simple Correlation
In the absence of direct shear (DS) or triaxial compression (TXC) test results, a simple
approach to critical-state friction angle estimation is to use an equation of the form:
C2
D 
 CU   R 
C3 C4
c  C1  50  (Eq. A.5)
 Dref 
where Dref = reference particle size (= 1 mm or 0.04 in.); and C1, C2, C3, and C4 = regression
coefficients. The values of C1, C2, C3, and C4 were obtained by performing a least squares
regression in Microsoft Excel. The following equation was found to fit the ϕc values reported in
Table A.2 quite well:

D 
c    28.3 50   CU   R
2 3
(Eq. A.6)
 Dref 
where ϕc = critical-state friction angle in triaxial compression, and ζ = exponent (= 0.045). The
adjusted coefficient of determination r2, mean absolute error, and mean absolute percentage error
are 0.89, 0.4°, and 1.3%, respectively. The adjusted r2 is a modified version of r2 that has been
adjusted for the number of independent variables considered in the model. Equation A.6 is
applicable for poorly-graded, clean silica sands with D50 = 0.15–2.68 mm (0.006–0.105 in.), CU =
1.2–3.1, and R = 0.3–0.8; however, it should be used with caution for (a) well-graded sands with
CU ≥ 6, (b) sands with D50, CU and R values that lie outside these ranges, and (c) sands with plastic
or non-plastic fines greater than 5%. Equation A.6 could be further improved through future
research.
Figure A.4 compares the critical-state friction angle predicted using Eq. A.6 with that
obtained from TXC test results for the poorly-graded, clean silica sands listed in Table A.2. The
differences between the predicted and measured ϕc values are within 1°. The value of ϕc predicted
using Eq. A.6 may be decreased by a degree or two, if needed, to obtain a conservative estimate
for use in foundation design. However, we re-emphasize that laboratory direct shear or triaxial
compression test results provide the best means for estimating the critical-state friction angle of
sands, particularly those that contain plastic or non-plastic fines greater than 5% (Carraro et al.,
2009; Murthy et al., 2007).

A-6
40 FS Ohio 6–10
 FS Ohio 10–16
 D50 
c     28.3    CU   R 
2 3
FS Ohio 16–20
 Dref  FS Ohio 20–40
Predicted critical-state friction angle (o)

FS Ohio 50–100
D50 = 0.152.68 mm (0.0060.105 in.) Fontainebleau NE34
Dref = 1 mm (0.04 in.) Fraser River
Ham River
CU = 1.23.1
35 Lausitz
R = 0.30.8 Leighton Buzzard
 = 0.045 Longstone
M31
Ohio Gold Frac
Ottawa Graded
Ottawa 20–30
Q-Rok
Ticino
30 Toyoura

+1o
1o

25
25 30 35 40
Critical-state friction angle obtained from TXC tests (o)
Figure A.4 Comparison of critical-state friction angles obtained from Eq. A.6 and TXC tests on
poorly-graded, clean silica sands.

To evaluate the performance of Eq. A.6 in an unbiased manner, a blind test was performed
on two additional, poorly-graded, clean silica sands—Nerlerk sand and Fujian sand; these sands
were not used in the development of Eq. A.6. The properties of Nerlerk sand are: D50 = 0.23 mm
(0.009 in.), CU = 1.56, R = 0.43, SWL = 0.75, emin = 0.66, emax = 0.89, and ϕc = 30° in triaxial
compression (Sladen et al., 1985); the values of R and SWL are based on Krumbein and Sloss (1951).
The properties of Fujian sand are: D50 = 0.40 mm (0.016 in.), CU = 1.53, R = 0.55, and ϕc = 30.8°
in triaxial compression (Yang & Wei, 2012). The critical-state friction angle of Nerlerk sand and
Fujian sand obtained from Eq. A.6 is shown below.

Nerlerk Sand
 0.045
D   0.009 
c  28.3 50   CU   R  28.3   1.56   0.43
2 3 0.09 0.135
  30.9
 Dref   0.04 
Fujian Sand
 0.045
D   0.016 
c  28.3 50   CU   R  28.3   1.53   0.55
2 3 0.09 0.135
  30.6
 Dref   0.04 

A-7
The difference between the predicted and measured ϕc value is equal to 0.9° for Nerlerk sand and
0.2° for Fujian sand.

A.5 Procedure for Estimation of ϕc from Intrinsic Soil Variables


In the absence of direct shear or triaxial compression test results, the critical-state friction
angle ϕc of a poorly-graded, clean silica sand may be estimated from intrinsic soil variables by
following these steps.
1. Perform a sieve analysis test and obtain the particle-size distribution curve.
2. Determine the mean particle size D50 and the coefficient of uniformity CU (= D60/D10) from
the particle-size distribution curve.
3. Determine the dominant particle size of the sand (i.e., the sieve size with the maximum
percentage by mass of particles retained on the sieve).
4. Select a reasonable number of random particles (say 25 particles) from those retained on
the sieve identified in step 3 and place them in an orderly fashion on a flat surface (e.g.,
glass slide). The number of random particles may be increased or decreased depending on
how variable the morphology is from one particle to the next.
5. Execute one of the following methods, based on the desired level of sophistication, to
determine particle roundness and sphericity.
Method 1 (Visual)
a. Observe the particles through a microscope.
b. Compare the observed shapes of the particles against the reference particle
silhouettes given in the chart by Krumbein and Sloss (1951) (Figure A.3).
c. Determine the roundness R and sphericity S of each particle and average the
values for all the particles selected.
Method 2 (Computational)
a. Observe the particles through a microscope and obtain high-resolution images of
the particles using a digital camera attached to the microscope.
b. Analyze the particle images using the software ImageJ
(https://imagej.nih.gov/ij/download.html) or the MATLAB code developed by
Zheng and Hryciw (2015)
(https://www.mathworks.com/matlabcentral/fileexchange/60651-particle-
roundness-and-sphericity-computation).
c. Determine the roundness R and sphericity S of each particle and average the
values for all the particles selected.
6. Estimate the critical-state friction angle ϕc of the sand using Eq. A.6.

A-8
APPENDIX B. OCR AND K0 OF SOIL

B.1 Overconsolidation Ratio (OCR)


Laboratory consolidation tests, such as the oedometer test or the constant rate of strain
(CRS) test, provide the best means of determining the OCR of clays. In addition, the OCR may be
known from the site history (e.g., if soil was previously removed or structures were demolished at
the site), or it may be deduced from geologic considerations or from in situ testing observations.
A preliminary estimate of the OCR of clay can be obtained from CPT results using the following
approximate correlation (Ladd et al., 1977; Salgado, 2008; Wroth, 1984):
  s    1.25   q N 1.25 
 u v 0 OC   
OCR  max   ; 1  max 
tn k
 ; 1 (Eq. B.1)
  su  v0  NC     su  v0  NC  

where (su/σ′v0)OC = normalized undrained shear strength of an OC clay; (su/σ′v0)NC = normalized


undrained shear strength of the same clay when normally consolidated (≈ 0.2–0.3 for most clays);
qtn = normalized cone resistance (= (qt – σv0)/σ′v0); qt = corrected, total cone resistance measured
under undrained conditions (Eq. 2.1); σv0 and σ′v0 = in situ vertical total and effective stresses,
respectively, at the depth being considered; and Nk = cone factor (≈ 9–15 as long as the CPT is
performed at a penetration rate that is sufficiently high to ensure undrained penetration (refer to
Appendix D); soft NC clays tend to have Nk values near the low end of this range, while stiff OC
clays tend to have Nk values near the high end of this range) (Bisht et al., 2021; Mayne & Peuchen,
2018; Salgado, 2008, 2013, 2014; Salgado et al., 2004). An average Nk value of 12 may be used in
Eq. B.1 to obtain a preliminary estimate of the OCR.
The normalized undrained shear strength (su/σ′v0)NC of an NC clay can be estimated using
(Wroth, 1984):
1.7 sin c
 3  sin  for CIUC conditions
 su   c
   
  NC  sin c  a 2 1 
  v0
for CK 0 UC conditions (Eq. B.2)
 2a  2 
  

3  sin c
a (Eq. B.3)
2  3 2sin c 
where CIUC = isotropically-consolidated undrained triaxial compression, CK 0UC = K0-
consolidated undrained triaxial compression, Λ = plastic volumetric strain ratio (≈ 0.8), and ϕc =
critical-state friction angle (≈ 15°–30° for most clays; high-plasticity clays with high smectite and
clay contents tend to have values near the low end of this range, while low-plasticity clays with
low smectite and clay contents tend to have values near the high end of this range (refer to Table
E.1 of Appendix E)). An alternative expression that provides conservative estimates of (su/σ′v0)NC
for both CIUC and CK0UC test conditions is (su/σ′v0)NC = ϕc/100 (Kulhawy & Mayne, 1990;
Salgado, 2008).
The OCR (= σ′vp/σ′v) of sand may be evaluated based on the geologic history of the site,
where σ′vp = preconsolidation stress, which is the maximum vertical effective stress ever

B-1
experienced by the soil, and σ′v = current vertical effective stress. The reader may also refer to
Section 2.3.7 of Volume I for additional information on the OCR.

B.2 Coefficient of Lateral Earth Pressure At-Rest K0


The coefficient of lateral earth pressure at-rest K0 of soil can be determined using (Brooker
& Ireland, 1965):
K 0  K 0,NC OCR (Eq. B.4)
where K0,NC = value of K0 if the soil is normally consolidated (= 0.40–0.50 for NC sand, with dense
sands tending to have lower values and loose sands having higher values, and 0.50–0.75 for NC
clay) (Salgado, 2008; Salgado & Prezzi, 2007), and OCR = overconsolidation ratio, which is equal
to 1 for NC soil and greater than 1 for OC soil. The reader may also refer to Section 2.3.9 of
Volume I for additional information on K0.

B-2
APPENDIX C. ITERATIVE SCHEME FOR FOOTING SETTLEMENT IN
SAND

Because the representative elastic modulus of each sublayer is a function of footing


settlement, an iterative scheme is needed if we wish to generate a load-settlement curve for a given
footing geometry. Figure C.1 shows the iterative scheme proposed to achieve this objective. An
initial guess value for w (= wmax established in step 5(c) of Section 3.1) is first chosen, and the
representative elastic modulus of each sublayer is then calculated using Eq. 3.12. Next, the footing
settlement computed using Eq. 3.14 is compared with the initial guess value. If the convergence
criterion of 10–5 is satisfied, the value of w obtained from Eq. 3.14 is reported as the footing
settlement corresponding to the load acting on the footing. However, if the convergence criterion
is not satisfied, the footing settlement obtained from Eq. 3.14 is used as the initial guess value for
w in the next iteration. A convergence criterion of 10–5 was found to be adequate with respect to
accuracy and computational time, and convergence was typically achieved within a few iterations.
The iterative scheme can be used to obtain the load-settlement curve of the footing up to a footing
settlement w equal to 10% of the footing size B; however, it should not be used to estimate the
limit unit bearing capacity qbL of the footing (i.e., the unit load on the footing base that causes the
footing to plunge into the ground). The iterations can be performed in Microsoft Excel either by
going to File → Options → Formulas and selecting Enable iterative calculation in the Calculation
options tab or by using the Solver tool. Note that parameters DR, E/qc, and Iz should be calculated
for each sublayer within the influence depth zf0 below the footing base.

C-1
Start

Obtain qc profile from CPT sounding

Input L, B, D, K0, γm, ϕc, Q, wmax

Calculate zf0, zfp, Iz0, Izp, Iz using Eqs. (3.5)–(3.9)

Calculate DR using Eq. (3.10)

Choose initial guess for w [= wold = wmax (say 25 mm or 1 in.)]

Calculate E/qc using Eq. (3.12)

Calculate w (= wnew) using Eq. (3.14)

Check if
Input No wold  wnew
wold = wnew  10 5
wold

Yes

No Check if
Modify L or B wnew ≤ wmax

Yes
Footing design is satisfactory with
respect to the serviceability limit state

End

Figure C.1 Iterative scheme for estimation of footing settlement in sand using CPT results.

C-2
APPENDIX D. PENETRATION RATE EFFECT ON CONE RESISTANCE

Cone penetration at the standard rate of 2 cm/s (0.8 in./s) is fully drained for clean sand
and fully undrained for pure clay. However, for soils containing mixtures of sand, silt, and clay,
cone penetration at the standard rate of 2 cm/s (0.8 in./s) may take place under partially drained
conditions depending on the ratios of these three broad particle size groups and the fabric of the
soil. According to Kim et al. (2008, 2006), the undrained cone resistance is expected to be
measured in CPTs performed with the standard cone (dc = 35.7 mm or 1.4 in.) at the standard rate
(υ = 2 cm/s or 0.8 in./s) in soils having coefficient of consolidation cv values less than roughly
10–4 m2/s (0.15 in.2/s). However, if the cv value of the soil is greater than about 10–4 m2/s (0.15
in.2/s), the CPT sounding should be performed at a faster rate so that the normalized penetration
rate V (= υdc/cv) is greater than 10 (Salgado & Prezzi, 2014). This approach would be the easiest
way to ensure that cone penetration in mixed or intermediate soils takes place under undrained
conditions. However, as this is still a topic of ongoing research, the implementation of this
approach is optional and not mandatory in INDOT construction projects. The alternative would be
to attempt to interpret the results of a CPT sounding actually performed under partial drainage
conditions; however, there are no reliable methods for doing that at the present time. The
coefficient of consolidation can be determined from the results of laboratory consolidation tests or
CPT pore pressure dissipation tests (DeJong & Randolph, 2012), as discussed in Sections 1.3.6
and 2.3.14 of Volume I. Dissipation tests are valuable in clayey soils and they should be done
whenever engineers judge that the value of the information obtained from the test justifies the
expense for the site being investigated.
Volume I of the manual includes a synthesis of the work done by researchers on the aspect
of penetration rate vis-à-vis the drainage conditions. The methodology proposed by DeJong et al.
(2013) to address partial drainage conditions during cone penetration in intermediate soils is
provided in Section 1.3.7 of Volume I. However, this methodology has not been standardized or
formally adopted in practice.

D-1
APPENDIX E. RESIDUAL-STATE FRICTION ANGLE OF CLAY

The residual shear strength τr of clay is the product of the normal effective stress σʹ on the
shearing plane and the tangent of the residual-state friction angle ϕr, which in turn depends on the
value of σʹ, the clay mineralogy, the clay fraction (CF), and the magnitude and rate of shear
displacement. According to Skempton (1985), the shear displacements needed for an intact clay
with CF > 30% and σʹ < 600 kPa to attain residual-state friction angles of ϕr and ϕr + 1° range from
100–500 mm (4–20 in.) and 30–200 mm (1.2–8.0 in.), respectively. Based on the clay fraction of
the soil, different residual-state shearing mechanisms are possible, resulting in different values of
ϕr (Lupini, 1980; Lupini et al., 1981). Based on Skempton's observations on the variation of ϕr
with the clay fraction of sand-bentonite mixtures tested in ring shear, Salgado (2006) proposed the
following equation for ϕr of clay-silt-sand mixtures as a function of the clay fraction at a given
stress level:
 c,mix  r pure clay 
r  r pure clay    52%  CF  %  (Eq. E.1)
 27% 
 
where ϕc,mix = critical-state friction angle of the clay-silt-sand mixture, and  r pure clay = residual-
state friction angle of the clay fraction of the mixture. For CF ≤ 25%, the bulky sand/silt particles
are likely to control the behavior of the mixture and thus ϕr = ϕc,mix, whereas for CF ≥ 52%, the
platy/tube-like/needle-like clay particles are likely to control the behavior of the mixture and thus
ϕr =  r pure clay ≈ 5°, 10°, and 15° for montmorillonite, illite, and kaolinite clay minerals,
respectively (Skempton, 1985). For intermediate values of CF between 25% and 52%, ϕr lies
between ϕc,mix and  r pure clay .
Besides the clay fraction and mineralogy, the residual-state friction angle ϕr also depends
on the magnitude of the normal effective stress σʹ acting on the shearing plane; ϕr decreases
nonlinearly with increasing σʹ (Figure E.1) because a larger normal stress forces greater
realignment of clay particles in the direction of shearing. Soils with high clay fraction (CF ≥ 52%)
and high smectite content, such as London clay, exhibit a significant drop in ϕr with increasing σʹ,
while soils with low clay fraction (CF ≤ 25%) and low smectite content may not exhibit any
residual behavior. Following the work by Maksimović (1989), ϕr can be expressed in terms of σʹ
using (Salgado, 2006):
 
r  r ,min  c r ,min (Eq. E.2)

1

 median
where σ′ = normal effective stress on the plane of shearing, ϕr,min = minimum residual-state friction
angle (attained at large normal effective stress), ϕc = critical-state friction angle, and σʹmedian = value
of σʹ at which the friction angle is equal to the average of ϕr,min and ϕc. At very large stresses, ϕr
reaches an absolute minimum, denoted by ϕr,min. For σ′ on the shearing plane approaching zero, ϕr
approaches the critical-state friction angle ϕc due to the negligible reorientation of the clay particles
in the absence of a normal stress forcing this reorientation to happen.

E-1
Figure E.1 Residual-state friction angle ϕr versus normal effective stress σʹ on the shearing plane
(Salgado, 2006).

Table E.1 summarizes the values of ϕc and ϕr,min of some well-known soils in the literature,
such as Lower Cromer till, Boston blue clay, San Francisco bay mud, London clay, and Weald
clay as a function of their CF and PI values. Although Lower Cromer till is a glacial till composed
of sand (> 50%), clay (= 14%–20%), and almost no silt (Gens, 1982), it has been considered in the
literature to behave like a “clay” but with no residual behavior. Boston blue clay is a low-plasticity,
insensitive, marine clay, composed of illite and quartz (Terzaghi et al., 1996), and does not exhibit
any residual behavior (Ladd & Edgers, 1972). San Francisco bay mud is a highly-plastic silt
containing a large amount of clay-size particles (montmorillonite and illite), organic substances,
shell fragments, and traces of sand (Bonaparte, 1982). London clay is composed of illite, kaolinite,
montmorillonite, and quartz (Gasparre, 2005); both San Francisco bay mud and London clay
exhibit residual strength with sustained shearing beyond the critical state. Figure E.2 illustrates the
fit of Eq. E.2 to ring shear test data for Weald clay. The fit was done by first estimating the value
of ϕc in triaxial compression (Parry, 1960) and then finding the values of σ'median and ϕr,min that
minimize the sum of least squares.

E-2
Table E.1 Critical-state and residual-state strength data for clayey soils reported in the literature

Soil Mineralogy CF (%) PI (%) A ϕc (°) ϕr,min (°) Reference


Boston Blue Clay Illite, quartz 35 13.1 0.37 32.41 — Ladd & Varallyay (1965)

London Clay Kaolinite, illite, 53–62 42–45 0.73–0.79 21.3 9.42 Bishop et al. (1971); Gasparre
montmorillonite, quartz (2005); Nishimura (2005)

Lower Cromer Till Illite, calcite, quartz 14–20 10–12 0.60–0.71 30.0 — Dafalias et al. (2006); Gens
(1982); Lupini et al. (1981)

San Francisco Bay Mud Illite, montmorillonite 47 47 1.00 28.91 16.2 Kirkgard & Lade (1991);
Meehan (2006)

Weald Clay Illite, kaolinite, illite- 52 33 0.63 20.9 8.33 Akinlotan (2017); Bishop et
montmorillonite, al. (1971); Parry (1960)
vermiculite
Note: CF = clay fraction, PI = plasticity index, A = activity (= PI/CF), ϕc = critical-state friction angle in triaxial compression, and ϕr,min = minimum residual-state
friction angle in ring shear.
1
Extrapolated value corresponding to 30% axial strain (Chakraborty, 2009).
2
Value corresponds to blue London clay at Wraysbury (CF = 57%, PI = 43%, A = 0.75). For brown London clay at Walthamstow (CF = 53%, PI = 42%, A =
0.79), ϕr,min = 7.5° (Bishop et al., 1971).
3
Obtained from the fit of Eq. E.2 to ring shear test data reported by Bishop et al. (1971).

E-3
Effective normal stress ' (ksf)
0 4 8 12 16 20
25 25
Test data (Bishop et al. 1971)
Fit using Eq. (E.2)

20 c = 20.9o 20
Residual-state friction angle r (o)

Residual-state friction angle r (o)


r,min = 8.3 o

'median = 51 kPa (1 ksf)

15 15

10 10

5 5

0 0
0 200 400 600 800 1000
Effective normal stress ' (kPa)
Figure E.2 Fit of Eq. E.2 to ring shear test data for Weald clay.

E-4
About the Joint Transportation Research Program (JTRP)
On March 11, 1937, the Indiana Legislature passed an act which authorized the Indiana State
Highway Commission to cooperate with and assist Purdue University in developing the best
methods of improving and maintaining the highways of the state and the respective counties
thereof. That collaborative effort was called the Joint Highway Research Project (JHRP). In 1997
the collaborative venture was renamed as the Joint Transportation Research Program (JTRP)
to reflect the state and national efforts to integrate the management and operation of various
transportation modes.

The first studies of JHRP were concerned with Test Road No. 1 — evaluation of the weathering
characteristics of stabilized materials. After World War II, the JHRP program grew substantially
and was regularly producing technical reports. Over 1,600 technical reports are now available,
published as part of the JHRP and subsequently JTRP collaborative venture between Purdue
University and what is now the Indiana Department of Transportation.

Free online access to all reports is provided through a unique collaboration between JTRP and
Purdue Libraries. These are available at http://docs.lib.purdue.edu/jtrp.

Further information about JTRP and its current research program is available at
http://www.purdue.edu/jtrp.

About This Report


An open access version of this publication is available online. See the URL in the citation below.

Sakleshpur, V. A., Prezzi, M., Salgado, R., & Zaheer, M. (2021). CPT-based geotechnical design man-
ual, Volume 2: CPT-based design of foundations—Methods (Joint Transportation Research Pro-
gram Publication No. FHWA/IN/JTRP-2021/23). West Lafayette, IN: Purdue University. https://
doi.org/10.5703/1288284317347

You might also like