Download as pdf or txt
Download as pdf or txt
You are on page 1of 65

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/257564364

Cone Penetration Test Based Direct Methods for Evaluating Static Axial Capacity
of Single Piles

Article  in  Geotechnical and Geological Engineering · August 2013


DOI: 10.1007/s10706-013-9662-2

CITATIONS READS
36 2,556

2 authors:

Fawad Niazi Paul W. Mayne


Purdue University Fort Wayne Georgia Institute of Technology
35 PUBLICATIONS   157 CITATIONS    242 PUBLICATIONS   4,912 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Evaluating CPTu results in sensitive clays View project

Identifying organic clays by CPTu View project

All content following this page was uploaded by Fawad Niazi on 21 December 2014.

The user has requested enhancement of the downloaded file.


Cone Penetration Test Based Direct Methods for Evaluating Static Axial Capacity of Single
Piles: A State-of-the-Art Review

Fawad S. Niazi1 and Paul W. Mayne2


1
Instructor and Graduate Research Assistant, Geosystems Engineering Division, School of Civil & Environmental
Engineering, Georgia Institute of Technology, Atlanta, GA, USA 30332-0355: fniazi6@gatech.edu
2
Professor, Geosystems Engineering Division, School of Civil & Environmental Engineering, Georgia Institute of
Technology, Atlanta, GA, USA 30332-0355: paul.mayne@gatech.edu

Abstract: The direct cone penetration test (CPT) based pile design methods use the measured penetrometer readings

by scaling relationships or algorithms in a single-step process to enable the assessment of pile capacity components

of shaft and base resistance (f p and q b , respectively) for evaluation of full-size pilings. This paper presents a state-of-

the-art review of published works that focus on direct CPT evaluation of static axial pile capacity. The review is

presented in a chronological order to explicate the evolution over the past six decades of an in-situ test based

solution for this soil-structure interaction problem. The objective of this study is an attempt to assemble maximum

published methods proposed as a result of past investigations in one resource to afford researchers and practitioners

with convenient access to the respective design equations and charts. In addition to an all-inclusive summary table

and the design charts, a compilation of significant findings and discussions thereof are presented. Furthermore,

potential future research directions are indicated, with special emphasis on the optimal use of the modern multi-

channel hybrid geophysical-geotechnical seismic CPT to evaluate the complete axial pile load-displacement

response.

Keywords: Pile foundation; Cone penetration test; Piezocone penetration test; Pile load test; Axial pile capacity
1 Introduction

The static axial capacity (Q t ) of pile foundations is calculated from the sum of the shaft capacity (Q s ) and base

capacity (Q b ):

Q t = Q s + Q b = Σ f pi A si + q b A b (1)

where f pi = unit shaft resistance of the ith soil layer through which the pile shaft is embedded; A si = shaft area

providing frictional resistance with the adjacent soil in the ith layer against axial displacement; q b is the unit end

bearing resistance; A b is the pile base area.

The methods for the assessment of pile capacity in terms of its components of shaft and base resistance (f p and

q b , respectively) have followed a constant evolution over the past many decades. Starting from some basic

formulations where f p was related solely to the shear strength of the soil, alternative approaches such as those

incorporating the influence of lateral effective stress coefficients (K) and pile-soil interface friction (δ) were

advanced. Subsequently, other variables which influence the magnitude of f p were studied, including: stress history,

pile length, pile slenderness ratio, soil sensitivity, plasticity of clayey soils, relative density of sandy soils, effective

stress strength of soil, progressive failure mechanism, plugging effect in open-ended pipe piles, soil compressibility

characteristics, pile material, and installation methods. Similarly, the q b component of pile capacity depends

primarily on the conditions around the pile base, including: pile tip configuration [e.g. tip shape, open-ended (OE)

vs. close-ended (CE) etc.], installation method, strength and stiffness properties of the geomaterial encountered at

the pile base, rate of loading, and drainage characteristics. Different researchers have proposed various bearing

capacity theories, and modifications thereof, to evaluate q b of pile foundations, e.g. limit plasticity, elasto-plasticity,

cavity expansion, strength dilatancy, and particle breakage. This continual process of evolution has led to a variety

of sophisticated formulae.

Despite significant contributions being made to the literature in terms of different design methods, many of the

approaches have a number of inherent drawbacks. Soil behavior is, of course, governed by a series of complex

stress-strain changes that occur during installation and subsequent loading. Owing to the difficulties and the

uncertainties in assessing the pile capacity on the basis of the soil strength-deformation characteristics, the most

frequently followed design practice is to refer to the formulae correlating directly the pile capacity components of f p

and q b to the results of the prevalent in-situ tests. Within the domain of these in-situ methods, the cone penetrometer

test (CPT) is one of the most frequently used investigation tool for pile bearing capacity evaluations. Ever since the
first use of CPT in geotechnical investigations, research efforts have advanced the very elementary idea of

considering it as mini-pile foundation. This has resulted in plethora of correlative relationships being developed

between the CPT readings [measured tip stress (q c ) or more proper total tip stress (q t ), sleeve friction (f s ), and

shoulder pore water pressure (u 2 )] and the pile capacity components of f p and q b . Such correlations, although

empirical, have been worked out on the basis of load test results from both instrumented and un-instrumented full

scale piles and are able to accommodate many important variables.

2 Pile Capacity Evaluation from CPT

As commonly reported (e.g. Ardalan et al. 2009; Cai et al. 2009; 2012; Mayne 2007), there are two main approaches

to accomplish axial pile capacity analysis from CPT data: (a) "rational (or indirect) methods" and (b) "direct

methods." The rational methods require a two-step approach. As a first step, CPT data are used to provide

assessments of stress history [preconsolidation stress (σ p ') , and overconsolidation ratio (OCR)], in-situ radial stress

coefficient (K o ) i.e., the ratio of in-situ radial to vertical effective stress (σ ro '/σ vo '), radial stress coefficient at failure

(K f ) i.e., the ratio of radial to vertical effective stress at failure (σ rf '/σ vo '), undrained shear strength (s u ), relative

density (D r ), effective stress strength (φ'), soil total unit weight (γ t ), fundamental soil stiffness [intial shear modulus

(G max ), or initial Young’s Modulus (E max )], interface friction between soil and pile material (δ), and bearing capacity

coefficients (N c , N q ). The pertinent relationships may be found in the related literature, some of which have

previously been summarized by Niazi and Mayne (2010), and Niazi et el. (2010a). Utilizing these input values of

geoparameters, the second step enables the assessment of f p and q b components of pile capacity within a selected

analytical framework.

The pile f p can be evaluated using either total stress analysis (α-method) for clays that relates f p to s u via an

adhesion factor (α), or the effective stress analysis (β-method) for both sands and clays that relates overburden stress

via empirical parameter β. The simplest fundamental formulations of the two approaches for f p are shown below:

α-Method: f p = α ∙s u (2)

β-Method: f p ≈ σ rf '∙tanδ ≈ K f ∙σ vo '∙tanδ ≈ K o ∙σ vo '∙tanδ ≈ β ∙σ vo ' (assuming K o ≈ K f )


(3)

For a pile foundation, an important factor of relevance to q b is the likely strain compatibility differences occurring

between the unmatched mobilization of side resistance and end bearing components during pile loading. For

undrained loading (primarily in clays and cohesive silts) beneath the base, the q b can fully mobilize within tolerable
limits of vertical displacements, usually taken as w/d = 0.1, where w = pile settlement, and d = base diameter. In the

case of drained loading (primarily sands and granular materials), however, it is impractical to assume that full

mobilization of the end bearing resistance occurs for the range of tolerable settlements. Consequently, to achieve a

settlement ratio corresponding to w/d = 0.10, it is customary to use an operational value of q b that is reduced from

the theoretical value [q b = 0.1q b(theory) ]. In sands as well as slow loading (or long-term analysis) in clays and silts,

usually drained conditions are assumed (use bearing capacity coefficient for overburden, N q corresponding to

effective stress analysis). In clays, silts, and soils with low permeability (assuming φ' = 0 for fast loading), usually

undrained conditions are evaluated (use bearing capacity term for cohesion, N c corresponding to total stress

analysis). Thus, the evaluation of q b is performed using relevant coefficients to relate q b either to s u or σ vo ' (q b ≈

N c ∙s u for undrained loading, and q b ≈ 0.1N q ∙σ vo ' for drained loading). Here, N c ≈ 9 for deep foundations, while N q is

function of φ'.

Doherty and Gavin (2011), Jamiolkowski (2003), Patrizi and Burland (2001), and Karlsrud (2012) have

effectively reviewed significant contributions in this regard including works by API (1969; 1975; 1976; 1987;

1993), Chen and Kulhawy (1994), Drewry et al. (1977), Karlsrud et al. (1993; 2005), Kolk and Van der Vende

(1996), Kraft et al. (1981), McClelland (1974), Miller and Luttenegger (1997), Randolph (1983), Randolph and

Murphy (1985), Semple and Rigden (1984), Skempton (1959), and Vijayvergiya and Focht (1972). For the scope of

this paper, no further discussion has been included on the indirect methods. However, summary lists of the various

factors considered by different researchers in their respective studies to make improvements in the predictive

reliability of total stress approach and effective stress approach are presented in Tables 1 and 2, respectively.

In contrast with the rational methods, the direct CPT methods use the measured penetrometer readings by

scaling relationships or algorithms in a single-step process to obtain f p and q b for evaluation of full-size pilings. Fig.

1 presents various paths to evaluate the two components of q b and f p from CPT readings.

The remaining parts of this paper present the basis for development, chronological evolution during the past

over 60 years, and the design formulations and charts for the direct CPT-based methods for axial pile capacity

evaluations. Over 75 different technical papers, reports and PhD research studies were reviewed in this process in an

attempt to present all-inclusive and updated information on the subject. It may be noted that there are still many

reports and manuals (not cited herein) that include design recommendations on the topic, based on various

formulations presented in this paper. Summary information on the scope, effort and significant findings of each

method are included, followed by general discussions and recommendations for future research directions.
3 Direct CPT-Based Methods

In one viewpoint, the cone penetrometer can be considered as a mini-pile foundation, whereby the measured tip

stress and sleeve resistance correspond to the pile end bearing and the component of side friction (Mayne 2007). As

noted by Ardalan et al. (2009) and Eslami and Fellenius (1997), the mean effective stress, compressibility and

rigidity of the surrounding soil medium affect the pile and the cone in a similar manner; thus, eliminating the need to

supplement the field data with laboratory testing and to calculate intermediate values for use in the CPT methods for

pile capacity evaluations. This concept has led to the development of many direct CPT methods, whereby the

measured readings are simply scaled up empirically for the evaluation of full-scale piling. The initial formulations of

such methods was based solely on the measured tip resistance (q c ) derived from mechanical cone penetrometers.

Subsequently, with the introduction of the electrical cone penetrometer, the additional channels measuring sleeve

friction (f s ) and porewater pressures (u 1 and u 2 ), the correction of porewater pressures to the measured tip resistance

to yield corrected tip resistance (q t ), and continuously increasing database of CPT and pile load tests has resulted in

a variety of direct CPT based pile evaluations. Some of these methods provide design equations for either f p or q b ,

while others account for both the components.

In addition to the purely empirical CPT-based direct methods, f p and q b can also be estimated using the

applicable semi-empirical direct methods which require measurements or estimation of additional parameters along

with the penetrometer readings of CPT. Within the scope of this study, the semi-empirical direct methods have also

been grouped with the direct methods.

4 Chronology of Evolution

4.1 Early Work

The early efforts to use CPT data for pile design were directed toward estimating the maximum pile embedment

depth (Begemann 1969). Accordingly, the maximum driving depth for a concrete pile was broadly taken as the

depth where the total cone resistance (point resistance plus total sleeve friction) reached 45 to 50 kN. Based on a

comparative field investigation on q b of a 0.426 m jacked square concrete pile and average q c of six CPT soundings

in medium dense sand, Plantema (1948) concluded that q b ≈ q c . Huizinga (1951) reported on the early efforts in

Holland to estimate q b from CPT data. Hence, q b was computed using average q c in the zone of soil that would be

influenced by the pile base to account for varying q c values; although no specific averaging procedure was detailed.

Huizinga noted from his data that, on the average, measured values of total shaft friction (Q s ) were about twice the
value computed by assuming equal unit friction on the rods and pile. During that period the only measure of side

friction which could be obtained was total friction on the entire length of push-rods.

Meyerhof (1951) concluded that for deep foundations bearing in sand, CPT logs provide better estimates of pile

capacity directly compared to use of theoretical bearing capacity equations. However, he highlighted the possible

anomaly due to scale effect on CPT q c vs. pile q b occurring from the difference in cone and pile diameters,

especially for dense sand layers. Van der Veen and Boersma (1957) explored the scale effect on CPT q c to pile q b

relationship. From the results of load tests on 15 concrete piles and CPT soundings, they concluded that pile q b can

be reasonably evaluated using an average of q c over a depth interval from 3.75d above the pile base to 1.0d below,

where d is pile diameter. In a similar study on 88 pile load tests, Menzenbach (1961) identified the type and strength

of soil as additional factors to be considered while relating pile q b to CPT q c .

4.2 Bogdanovic Method

Bogdanovic (1961) used CPT data to analyze the results of load tests on driven and jacked concrete piles in soft clay

overlying dense gravelly sand in Yugoslavia. Bogdanovic showed that the pile f p may be conveniently calculated as

CPT f s times the ratio of penetrometer to pile circumference, while pile q b estimates are obtained by averaging CPT

q c values from an envelope drawn through the minimum q c values rather than by averaging the actual values over a

depth interval from 4.0d above the pile base to 2.0d below the base. Hence, the concept of minimum path rule was

introduced in the averaging procedure.

4.3 Meyerhof Method

Meyerhof (1956), recognizing the empirical correlations between standard penetration test (SPT) and static CPT,

performed an analysis of the measured results in pile loading tests and CPT readings. He developed a method to

estimate the bearing capacity components (f p and q b ) of driven displacement piles and piles with small displacement

(e.g. steel H piles) in cohesionless soils from q c and f s .

Meyerhof (1976) documented the effects of pile embedment in the different bearing strata (Fig. 2), length of

pile, and the method of pile installation (e.g. driving, drilling etc.). Accordingly, appropriate allowances were

introduced in the design equations. Meyerhof (1983) further expanded the database of CPT and load tests on driven

and drilled shafts to present a refined version of design equations and charts, where the effects of pile base diameter

were also integrated (Fig. 3).


4.4 De Beer’s Contribution

De Beer (1963) provided significant insight into the scale effect problem on the pile q b evaluation. He considered

penetrometers of different sizes and showed that at some depth all penetrometers attain the same q c values. The

effect of this phenomenon on q b of a pile driven into a dense layer underlying a weaker layer can be seen by

referring to Fig. 4. If the layers were penetrated by a cone penetrometer of small diameter (3.57 cm), the depth of

embedment effect would be minimal and the q c would ideally follow the profile of dashed curve. However, if a

large diameter pile were pushed into the second layer, the q b would not equal that of the negligible diameter probe

until the pile had reached a critical depth (L c ) corresponding to the deep foundation condition. De Beer showed that

it is reasonable to assume that the pile base resistance q b curve varies linearly over L c (solid curve); thus, the pile

resistance at any intermediate depth could be determined if the small diameter q c curve and L c were known. He

further showed that L c is a function of the foundation size and sand density.

4.5 Vesić and Kerisel Method

Vesić (1964) and Kerisel (1964) performed experiments on piles and penetrometers (45 to 320 mm diameter) in

sands placed at different densities in rigid wall laboratory test chambers. Both the sets of experiments confirmed De

Beer's theory concerning the depth of embedment effect. According to Vesić's experiments the values for L c vary

from about 6d for loose sand to about 25d for dense sand, whereas, Kerisel's experiments indicate L c values of about

5d for loose sand and approximately 20d for dense sand.

4.6 Begemann Method

Begemann (1963) noted from the concurrent work by the Dutch engineers that q b for a pile with its base at the

interface between two soil layers is approximately equal to the average q c of the two layers, concluding that soil

above and below the pile base contributed almost equally to the pile q b . He assumed that the failure at the pile base

followed the classic logarithmic spiral shape, although no clear evidence was given to support this assertion for the

case of a deep foundation. He suggested that it is usually adequate to average q c of the upper layer to a distance 8d

above the pile base and that of the lower layer over 3.5d below the base. For the lower layer, additional weight

should be given to the sounding value at a depth of 3.5d to account for the assumed logarithmic shape where a

significant portion of rupture surface runs near horizontally. Nottingham (1975) reported that this method was

subsequently further refined to account for the presence of thin layers or zones of weak soil which could affect the

shape and position of the rupture surface at the pile base (Fig. 5).
Begemann (1965; 1969) developed design curves for predicting uplift pile friction capacity based on the results

of a large number of uplift load tests and CPT soundings from his adhesion jacket cone developed in 1953. The

driven piles database in these studies consisted of tapered wooden piles, pointed base prefabricated concrete piles

with 45o point, pointless base prefabricated concrete piles, OE steel pipe piles, and I-beam piles. The design curves

along with an example application are shown in Fig. 6. Begemann (1969) also reported that some tests indicate

negligible difference between friction in uplift and compression.

4.7 Aoki and Velloso Method

Aoki and Velloso (1975) presented a method to estimate the bearing capacity of piles from the results of dynamic

penetration tests. Making use of correlations, they obtained the values of q c and f s that would otherwise be measured

in a static CPT. By introducing certain empirically derived factors for different pile and soil types from load test

results, design equations were proposed to estimate the bearing capacity of single piles.

4.8 Nottingham and Schmertmann Method

Apparently, until around mid-70's, no noteworthy work was documented pertaining to the use of CPT data for

predicting pile capacity in clays. This may be attributed to the fact that CPT methods were being developed in

regions where the geology is such that piles are driven through soft clays and organic soils into underlying sands.

Nottingham (1975) is noted to be the first to explored different combinations of soil types to develop design

equations applicable to a wide variety of situations. He also studied the applicability of both mechanical and

electrical friction sleeve penetrometers for predicting the load capacity of displacement piles. His study included 108

load tests on large-scale model piles. Three types of model piles were used: a 102 mm square precast concrete pile; a

102 mm diameter pipe pile; and a step-taper pile with section diameters ranging from 74 to 115 mm. Data from load

tests on 15 full-scale piles were used to verify the equations developed from the model pile tests. He concluded that

the Begemann's procedure was a valid design tool for investigating pile q b in both granular as well as cohesive soils

using either mechanical or electrical penetrometer data. However, he introduced a reduction factor for mechanical

penetrometer used in clays to account for possible increase in q c resulting from friction on the tip mantle. Further, he

introduced new design equations and charts (Figs 7 and 8) for different pile and penetrometer types for both sand

and clay. This method has been commonly referred to as Schmertmann method after some refinement work by

Schmertmann (1978), where limiting q b values were imposed for different soil types and some non-displacement

piles were also included.


4.9 Penpile Method

Clisby et al. (1978) proposed simple design equations, frequently recognized as Penpile method, for the Mississippi

Department of Transportation. The method is applicable to both sand and clay and provides estimates of pile f p as

well as q b . They recommended that the pile q b should be based on the average q c of three CPT q c readings taken

close to the pile base. Later reliability studies on driven and jacked piles (e.g. Abu-Farsakh and Titi 2004; Cai et al.

2009), however, indicated that this method underestimates the pile capacity significantly.

4.10 European Method

De Ruiter and Beringen (1979) proposed another CPT-based method founded on the experience gained on pile

foundations for large offshore structures in northern part of North Sea, with great water depth and adverse sea

conditions. This method, also known as European method or Dutch method, uses different procedures for clay and

sand. In clay, the undrained shear strength (s u ) for each soil layer is first evaluated from q c using cone bearing factor

(N kt ); pile q b is then calculated using bearing capacity factor (N c = 9), while pile f p uses adhesion factor (α)

depending upon the stress history of clay. In sand, pile q b is calculated similar to Schmertmann method, while pile f p

is a function of both f s and q c of respective layers, and the direction of loading (compression vs. tension). In this

method upper limits of 15 MPa and 120 kPa are imposed on q b and f s , respectively.

4.11 Philipponnat Method

Philipponnat (1980) attempted to develop a simple direct CPT-based pile design method by taking account for large

variety of piles and soils. His method was calibrated using a database of 44 different types of instrumented piles,

load tested in different soil types. Appropriate coefficients were introduced in the design equations for estimating

pile f p to account for the pile and soil types. Additionally limiting values of f p were also imposed based on the pile

type. As for pile q b , Philipponnat recommended equal weights for the average q c values of the soil layers 3d above

and below the pile base. However, he suggested removal of extreme values (spikes) from the CPT q c profiles as

these are not expected to have significant impact on q b of piles much larger in size than the penetrometers.

4.12 LCPC Method

Bustamante and Gianeselli (1982) proposed a method based on the analysis of 197 full-scale static load tests

(including compression and uplift) on 96 pile foundations from 48 sites in soils made up of such varied materials as

clay, silt, sand, gravel, weathered rock, mud, peat, weathered chalk, and marl. The foundations taken into account

include both displacement and non-displacement piles. This method is commonly known as the Laboratoire Central

des Ponts et Chaussées (LCPC) method or French method. The approach offers versatility in the variety and types of
different deep foundation systems and geomaterials that can be accommodated. The LCPC method relies solely on

q c readings for evaluating f p along the pile shaft and for q b beneath the pile base. Bustamante and Frank (1997)

presented a simplified version of the LCPC method. Poulos (1989) offered summary graphical approach for pile f p

by LCPC method for clays and sands (Fig. 9). For estimating pile q b , a new approach was introduced in the

averaging procedure to obtain characteristic q c beneath the pile base (Fig. 10). This method also imposes different

upper limits for f p based on pile and soil typology as well as installation methods.

4.13 Cone-m Method

Tumay and Fakhroo (1982) proposed simplified design equations for estimating pile capacity in clayey soils from

electrical CPT. This method, also known as Cone-m method, was presented as a research report to the Louisiana

Department of Transportation. Both f s and q c readings are utilized in this method. Pile f p is calculated using an

adhesion factor and average f s within the layer along the pile; a reduced limiting maximum value of f p appropriate

for the clayey soils is also imposed. The adhesion factor is further calculated from the average f s . Pile q b is

estimated using a procedure similar to Nottingham (1975) and Schmertmann (1978) method.

4.14 Price and Wardle Method

Another contribution in the direction of CPT-pile direct correlations is the method by Price and Wardle (1982),

meant for predicting pile q b and f p from CPT q c and f s , respectively, based on load tests on driven and jacked piles,

and drilled shafts in stiff London clay. The empirically derived respective correlation constants depend upon the pile

typology.

4.15 Gwizdala Method

Gwizdala (1984) studied the behavior of large-diameter drilled shafts in cohesionless soils, and proposed a method

for calculating pile capacity from CPT data. He used 20 load tests data on drilled shafts ranging in diameter from

0.915 m to 2.14 m and in length from 6.1 m to 16.5 m to verify the accuracy of his method. The database included

eight drilled shafts with enlarged base. This method relies on q c readings alone. The variable end bearing and shaft

factors depend upon the soil type and q c readings. The effects of pile diameter as well as enlarged base were also

incorporated into the design equations.

4.16 Kulhawy and Phoon Method

The α-method presented by Kulhawy and Phoon (1993) for drilled shafts in clays may be considered analogous to

the approach used in the Dutch method. Accordingly, the undrained shear strength (s u ) is estimated for each soil
layer adjacent to the pile from the CPT q c , q c(net) = q c – σ vo to account for the vertical overburden stress (σ vo ), and

the cone bearing factor (N kt ). For pile f p , N kt may be taken as 10 for undrained shear strength in isotropically

consolidated triaxial compression mode (s uCIUC ), while for the base, N kt = 15 may be adopted.

4.17 UIUC Method

Alsamman (1995) at University of Illinios at Urbana-Champaign (UIUC) used a database of 95 full-scale drilled

shaft load tests, 48 of which were tested in cohesionless soils, 16 in cohesive soils, and 31 in mixed soils. The

influence of geometry (slenderness ratio, diameter, length, enlarged base), pile installation methodology, type of

bearing layer, and cone type (mechanical vs. electrical) on predicted capacity were assessed. Based on the

correlation of normalized q c of the two cone types presented by Kulhawy and Mayne (1990), separate design curves

(Figs 11 and 12) were given for both pile f p and q b in cohesive and cohesionless soils for mechanical as well as

electrical cone penetrometers. Here the pile bearing capacity of cohesive soils is given as a function of net cone

resistance [q c(net) = q c – σ vo ], while that of cohesionless soils as a function of CPT q c . The influence zone for

calculating q b is simply taken as one diameter beneath the pile base.

4.18 NGI-BRE Method

The Norwegian Geotechnical Institute, Oslo, and Building Research Establishment, London (NGI-BRE) jointly

developed an empirical approach for piles, where Almeida et al. (1996) used data from 43 load tests on OE as well

as CE driven and jacked piles (mostly tested in uplift) at eight clay sites in Norway, UK and USA. For the first time,

pore water pressure measurements of piezocone penetration test (CPTu) were utilized in terms of corrected cone tip

resistance (q t ). The effect of vertical overburden stress was also taken into account by using q t(net) = q t – σ vo .

Following the procedures suggested by Semple and Rigden (1984) and Randolph and Murphy (1985), a reduction in

the pile shaft coefficient (k 1 ) was also recommended for pile slenderness ratio [length/diameter (L/d)] > 60. The

design equations of this method were later updated by Powell et al. (2001) using a database of 63 pile load tests with

eight additional sites from UK, France and Denmark. No specific discussion on the zone of influence near the pile

base was included in this method.

4.19 Politecnico di Torino Method

A direct CPT method for drilled shafts in sands was developed by the Politecnico di Torino, Italy. For clean

quartzitic uncemented normally consolidated (NC) sands, the f p and q b of drilled shafts may be estimated from the

CPT q c (Fioravante 1994; Fioravante et al. 1995). The relationship for q b was developed based on centrifuge model

tests on medium dense (relative density, D r = 48%) to very dense (relative density, D r = 90%) Toyoura siliceous
sand. In sands, for full mobilization of end bearing capacity, very large displacements are required, and thus for

allowable settlements only a fraction of the full resistance is available. To account for this phenomenon, the base

capacity was associated with the relative displacements [pile top displacement to diameter ratio (w t /d)] of 0.1 and

0.05. Using numerical analyses, Lee and Salgado (1999) also developed an improved relationship between

normalized base resistance (q b /q c ) and relative base settlements (w b /d) ranging between 0 and 0.2. They also studied

the effects of sand relative density (D r ) and at-rest lateral earth pressure coefficient (K o ). Neglecting the minor

effects of D r , an average relationship was developed.

4.20 Unicone Method

A generalized direct CPTu method for soil profiles ranging from soft to stiff clay, medium to dense sand, and

mixtures of clay, silt, and sand was proposed by Eslami and Fellenius (1997) based on 142 load tests on a large

variety of piles from 53 sites in 13 countries. Commonly called Unicone method, this is known to be the first and, so

far, the only method that utilizes all three piezocone readings (q t , f s , and u 2 ). This method developed a new soil

profiling chart based on CPTu data: effective cone resistance (q E = q t – u 2 ) and sleeve friction (f s ) (Fig. 13). The

pile f p of each layer is calculated using side correlation coefficients (C se ) based on the soil zones of the profiling

chart and q E of the layer. Similarly the pile q b is calculated using toe correlation coefficient (C te ) based on pile toe

diameter and the geometric mean of q E in the influence zone near the pile toe. The influence zone is dictated by the

pile embedment through upper and lower soil layers.

4.21 KTRI Method

Takesue et al. (1998) found a bilinear relationship between normalized shaft resistance (f p /f s ) and the measured CPT

excess pore water pressures (∆u 2 = u 2 – u o ) using load test data on 6 piles from 4 sites in Japan (Fig. 14). Also

known as Kajima Technical Research Institute (KTRI) method, this approach was developed on bored and driven

steel pipe piles (length: 8.2 to 62.4 m and diameter: 0.8 to 2.5 m), and applies to clays, sands and mixed soil

conditions. As such, the method has shown to work in geomaterials showing negative pore water pressures (Mayne

2007).

4.22 UWA-99 Method

De Nicola and Randolph (1999) described a series of geotechnical centrifuge tests performed on model piles driven

into fine-grained silica flour of varying densities at the University of Western Australia. The fully instrumented OE,

CE, and sleeved model piles were driven and loaded statically to investigate their behavior in homogeneous sand.

Cone penetration tests were also conducted within the sample using a prototype cone. The end-bearing resistance
normalized by q c was plotted against the normalized base displacements (w b /d), leading to the observation that q b /q c

is essentially independent of the sand density, but tends to reduce with increasing overburden stress (σ vo '). On the

basis of these investigations, design equations and curves were developed for q b /q c at w b /d of 0.1 as a function of

σ vo ' and different diameter-to-wall thickness ratios (d/t) for OE piles (Fig. 15). This method utilizes effective stress

approach (β-method) as a means for evaluating fp, and no direct use of CPT reading is involved.

4.23 TCD Method

Lehane and Gavin (2001) presented the results from an experimental program carried out at Trinity College Dublin

(TCD), in which instrumented 40 and 114 mm diameter OE model piles were jacked into loose dry sand in a large

testing chamber. CPT q c profiles were also obtained in the sand deposits. The investigations indicated that the pile

q b may be separated into two components: plug resistance (q plug ), and annulus resistance (q ann ), and separately

expressed as simple functions of the q c and the incremental filling ratio (IFR) prior to loading. The q plug was shown

to be related directly to the CPT q c value and IFR measured prior to the loading, whereas the q ann was seen to be

equivalent to q c (Fig. 16).

Gavin and Lehane (2003) studied the factors affecting the f p of 114 mm diameter OE pipe piles jacked in loose

sand in a similar set of testing chamber experiments at TCD. This test series was designed to investigate the effects

of in situ stress level, pile end condition, directional mode of loading (compression vs. tension or uplift), and degree

of plugging on the development of pile f p . The results indicated that the maximum local f p that can develop at a

given location on an OE pipe pile may be expressed as a function of IFR of the soil plug during installation, the CPT

q c value, pile wall thickness (t), and relative position from pile base (h/r*), where h is the height above the pile base,

and r* is the modified radius of OE pile. The experimental results allowed a new expression to be developed for the

plug resistance during pile installation, and this, used in conjunction with the effective stress approach (β-method),

provide a means of estimating the f p of OE piles from CPT q c . As a result of the findings from these experiments,

and combining the previous results by De Nicola and Randolph (1999) and Lehane and Gavin (2001), modified

design equations were proposed for q b0.1 at w b /d of 0.1. No specific recommendations were given for the zone of

influence beneath the pile base. It should be noted that the design equations in TCD method can be utilized if the

IFR is measured during pile installation.

4.24 Fugro-05 Method

Kolk et al. (2005) presented new design criteria for determining the axial capacity of driven pipe piles in silica

sands, mostly applicable for offshore environment. This method, frequently known as Fugro-05 method, was
developed from the results of a database of 45 OE and CE steel piles driven and load tested (24 tension and 21

compression) in well-documented soil conditions including CPT data. The design equations for pile f p were

basically derived from the effective stress approach (β-method). By assuming that the sand-steel interface friction

angle (δ) is relatively constant after the pile installation, and that the changes in radial stress (∆σ rd ') are minimal for

all database piles, basic formulation of β-method was simplified. Based on the Finite Element analysis by De Nicola

and Randolph (1993), a reduction on f p calculations was imposed for layers near the pile base (h/r* < 4). The design

equations for pile f p are based on CPT q c , h/r*, and directional mode of loading (compression vs. tension or uplift);

whereas, for q b0.1 at w b /d of 0.1, the effects of the pile end condition [i.e., area ratio, A r = 1 – (d i /d)2] have been

integrated. Here, the arithmetic average of CPT q c is taken over the influence zone defined as +1.5d at the pile base.

4.25 UCD-05 Method

A research work by Gavin and Lehane (2005) supported by the University College Dublin (UCD) summarized the

updated experimental findings from field load tests on 20 full-scale instrumented piles at 12 different sites with

diameters ranging from 0.1 to 2 m and pile L/d of 15 to 87. The piles were all driven or jacked in predominantly

dense sand. CPT q c data were available at most of the sites. The factors controlling the development of q b of pipe

piles were investigated to provide the framework for an alternative design method which incorporates a direct

measure of soil plug development into the calculation. Instrumentations in all the load tests provided for the

measurement of plug development and annular resistance. It was established that the degree of plugging experienced

by a pipe pile during installation is the most important variable controlling the q b developed during subsequent static

loading. Consequently, an expression was developed for evaluation of q b of pipe piles, written as a function of the

CPT q c value and final filling ratio (FFR). Again, no specific discussion was included for the zone of influence

beneath the pile base.

4.26 ICP-05 Method

Jardine et al. (2005) provided a comprehensive theoretical and empirical basis to present an updated version of

previously known UK Marine Technology Directorate (MTD) method. The MTD method was a treatment for the

evaluation of axial capacity of offshore driven piles, developed at Imperial College London through four

consecutive PhD studies (Jardine 1985; Bond 1989; Lehane 1992; and Chow 1997). This method, now known as

Imperial College Pile (ICP-05) method, takes account of the CPT q c readings, and provides separate design

equations for piles driven in sand and clay. For sand, the design equations apply to the piles loaded for the first time

via slow maintained load (SML) test conducted around 10 days after driving. For pile f p in sand, the basic form of

equation is that of the effective stress approach (β-method), and following additional factors have been incorporated:
change in radial stress during pile loading (∆σ rd '), pile end conditions (OE vs. CE), directional mode of loading

(tension vs. compression), h/r*, radial displacement due to dilation during loading (∆y) based on pile roughness

(R a ), and the operational shear stiffness of soil (G) obtained from q c readings. The interface friction angle, unless

measured, may be defined as a function of mean particle size (d 50 ) (Fig. 17). For pile q b in sand, separate design

equations were given for OE and CE piles. Pile-to-CPT scale effect (d pile /d CPT ), influences of plugging based on

relative density (D r ) further estimated from q c readings, and pile shape were integrated into the design equations.

For piles in clay, no direct use of CPT readings was suggested for estimating f p ; whereas, for q b , separate

expressions were given for OE and CE piles, giving consideration to the rate mode of loading (drained vs.

undrained), and measure of plugging based on d i /d CPT , where d i is the inner pile diameter. The q c,avg for q b is taken

as arithmetic average over the influence zone of +1.5d at the pile base.

4.27 UWA-05 Method

With refined basis of assumptions, and on examination of new extended database, Lehane et al. (2005) at University

of Western Australia (UWA) presented slightly improved formulations for CPT-based axial capacity of typical

(large diameter) offshore piles in sand. They incorporated additional features accepted as having a controlling

influence on OE driven pile capacity [e.g. IFR and FFR as measures of effective shaft area (A sb,eff ) and effective

base area (A rb,eff ), respectively]. In the absence of soil plug measurements, this method also provides a means for

approximating the IFR and FFR. Similarly, in the absence of laboratory measurements of δ, use of Fig. 17 of ICP-05

method was suggested with an upper limit of tanδ = 0.55. This method provides estimates of both f p and q b from

CPT q c , where q c for base is averaged using Dutch averaging technique.

4.28 NGI-05 Method

The Norwegian Geotechnical Institute (NGI) established a comprehensive database of 85 driven piles tested in sand

at 35 different locations, of which 28 with highest quality rating and CPT data were selected. As a result of their

study, Clausen et al. (2005) presented an alternative empirical calculation method, referred to as NGI-05 method.

The formulations proposed by this method differ in format from the other CPT methods proposed during the same

time. CPT q c is included in the expression for pile f p through a correlation between relative density (D r ) and

normalized q c . Friction fatigue is based on a triangular distribution of f p using z/L degradation (Toolan et al. 1990)

instead of the pile diameter. The parameters for describing differences in end condition and pile material were tied to

the characteristics of the database used. The q b to an average q c value in the vicinity of the pile base for both OE and

CE piles is related via D r . The q b expressions for OE piles is further based on the end conditions (plugged vs.

unplugged), and averaged external skin friction.


4.29 Cambridge-05 Method

White and Bolton (2005) studied the causes of low reported values of q b /q c in sand in contrast with continuum

models (e.g. cavity expansion solutions and strain path method) that suggest q b = q c for steady deep penetration.

They re-examined a database of 29 load tests on a variety of CE piles (steel pipe piles, Franki piles with enlarged

base, and precast square, cylindrical and octagonal concrete piles) and CPT q c data, concluding that low q b /q c ,

which forms basis of the apparent scale effect on the diameter, can be attributed to: partial embedment in the

underlying hard layer (Fig. 18) and partial mobilization of q b for a selected settlement (commonly w t /d of 0.1).

Partial embedment of the pile tip into a hard layer underlying weak material was accounted for by weighting q c ,

whereas, partial mobilization was explained by defining failure according to a plunging criterion. They concluded

that any reduction of q c when estimating the q b of CE piles in sand should be linked to the above factors rather than

pile diameter.

4.30 Togliani Method

Formulated on the basis of years of his experience of designing and predicting the pile behavior, Togliani (2008;

2010) also proposed a simplified direct CPT-based method for displacement and non-displacement, cylindrical as

well as tapered piles. In particular, he combined his experience with the LCPC method and certain previous

recommendations for calculating the capacity of tapered piles to provide design equations for both f p and q b . He

incorporated the influence of f s , besides just q c for estimating pile f p ; whereas, q b , in contrast with the

recommendations of White and Bolton (2005), was still related to the pile slenderness ratio in addition to the

average q c between +8d toe and –4d toe .

4.31 German Method

Kempfert and Becker (2010) developed correlations for pile f p and q b from CPT q c and s u based on the load test

database (up to 1000 load tests on precast concrete, cast-in-place concrete, steel pipe, screw cast-in-place, and micro

piles etc.). Their results, presented in the form of empirically derived charts with upper and lower bound estimates of

f p and q b (Figs 19 and 20), have been integrated into the national German recommendations for piles "EA-Pfähle" of

the German Society of Geotechnics. Based on comparative statistical analysis, relevant discussions are included in

their paper to explain different factors influencing the variations of the two components of pile capacity within the

given bounds.
4.32 UCD-11 Method

Igoe et al. (2010; 2011) presented a new expression for estimating the f p of OE piles in sands using experimental

investigations on model pile installed either by jacking or driving into an artificially-created loose sand deposit in

Blessington, Ireland. The tests provided measurements of the soil-plug development and the radial effective stresses

during installation and subsequent loading cycles. They observed that the radial effective stress after equalization

(before loading) (σ rc ') acting on the external shaft of an OE pile depends on: the load cycles, normalized distance

from the pile tip (h/d), and the degree of plugging experienced during installation defined via IFR and the effective

shaft area ratio (A rs,eff ). The basic formulation is the extension of the TCD-05, ICP-05, Fugro-05, and UWA-05

methods. In the new design equation, a lower limit of σ rc ', controlled by CPT q c and a sand density-based scaling

factor for friction fatigue (η), has been imposed.

4.33 Van Dijk and Kolk (V-K) Method

Van Dijk and Kolk (2011) made another attempt to utilize the u 2 readings of CPTu for axial capacity evaluations for

offshore piles in clays. Accordingly, q t(net) values were used in the proposed design equations to estimate both f p and

q b . Thirty-three high quality load tests on OE and CE piles driven and tested in compression as well as tension at

fifteen different clay sites with equally representative CPT data were selected for this study (lengths and diameters

ranging between 3.6 and 71.4 m, and 0.102 and 0.812 m, respectively). The expression for base capacity was

developed by adopting cone factor N kt of 13 and using the relationship: q b = 9∙s u . For pile f p , they studied factors

such as soil plasticity, OCR, pile length, pile slenderness, relative base displacements, and time between pile driving

and testing, concluding that OCR and pile length, but not by loading direction, soil plasticity, pile diameter and pile

base displacement ratio are the factors influencing the shaft capacity.

4.34 SEU Method

By utilizing the formulations and design methodologies of two earlier methods (Unicone and KTRI) with slight

modification for q b estimation, Cai et al. (2011; 2012) at the Southeast University, Nanjing, China observed trends

from the results of 26 load tests on driven, jacked, and cased piles and CPTu data at two sites in Jiangsu Quaternary

clay deposits in eastern China. For shaft capacity, they presented KTRI-type of bi-linear trends between f p /f s and

∆u 2 from the back-analysis load test and CPTu readings (Fig. 21), while for q b , they adopted the approach of

Unicone method. They attempted to give physical meaning to the toe correlation coefficient (C te ) by associating it

with bearing capacity and cone factors N c and N kt , respectively.


4.35 HKU Method

Yu and Yang (2012) recently presented a new method for estimating the base capacity of OE steel pipe piles in sand.

The method, referred to as the HKU method, is based on CPT q c data. In contrast to most of the previous methods, it

takes into consideration the mechanisms of annulus and plug resistance mobilization. The annulus resistance is

linked with pile slenderness (L/d), whereas the plug resistance is related to the plug length ratio (PLR), defined as

H/L, where H is the length of plug measured at the end of pile installation. The cone tip resistance is averaged over a

zone in the vicinity of pile base by taking into account the failure mechanism of piles in sand, the condition of pile

embedment (i.e. full or partial embedment) (Fig. 22), and the effect of soil compressibility.

4.36 Probabilistic Approach by Chen and Zhang (2013)

Chen and Zhang (2013) adopted a probabilistic approach for predicting the impact of spatial correlation between q c

values of different soil layers on the bearing capacity of driven piles in clay. They considered the spatial correlation

via the correlation coefficients [ρ (b)(i) ] and [ρ (i)(j) ] between q cbV (the spatial average of q c over an interval near the

pile base) and q csVi (the spatial average of q c of the ith soil layer), and between q csVi and q csVj (the spatial average of

q c of the jth soil layer), respectively. Adopting the LCPC method and conducting parametric studies they indicated

the importance of such spatial correlation in the probabilistic prediction of the bearing capacity. The proposed

approach was calibrated against 14 pile load tests. At present, the procedure for implementation of this approach is

considered to be fairly intricate for simple frequent application by the practitioners. Therefore, further details have

not been included herein.

The design equations and charts for 35 direct CPT-based methods are presented in Table 3, particularly noting

the newly available approaches that have recently been developed.

5 Discussion

With regard to the selection and implementation of a suitable CPT-based direct method, numerous factors need

particular consideration and reflection. Some of these issues are discussed here.

5.1 Reliance on CPT Readings and Additional Parameters

It may be noticed from Table 3 and Fig. 1 that, in essence, there are two tiers of CPT-based direct methods. In one

set of methods, referred to herein as pure-empirical direct methods, simple direct relationships between CPT

readings and f p and q b components of pile capacity have been suggested based on the field measurements. Examples
include UIUC method, SEU method, German method, KTRI method, Unicone method, Penpile method, and

Politecnico di Torino method. A second type of these CPT methods, termed herein as semi-empirical direct

methods, additional parameters must be evaluated, beyond just reliance on the CPT readings. These additional

parameters, either measured via laboratory or field investigations, or further estimated from CPT-based correlations,

include: pile characteristics, pile base conditions (OE vs. CE), plugging vs. coring, installation methods, pile and soil

types, pile material, pile-soil interface friction, overburden stress, soil shear strength, soil relative density or

plasticity characteristics, soil frictional characteristics.

Some of the methods falling in the category of semi-empirical direct methods take their original concept from

the rational methods, i.e., total stress approach or effective stress approach; however, these methods also

demonstrate direct use of CPT reading in their design equations/charts. The ICP-05 method, Dutch method, UCD-11

method, TCD-03 method, and UWA-05 method are the examples belonging to this category.

From the list of CPT-based direct methods, 26 methods provide estimates for both f p and q b , while the

remaining 9 methods account for either of the two. Fig. 23 provides an overview of the percent reliance of these

methods on the different combinations of CPT readings to estimate f p and q b .

5.2 Correction of Measured Cone Tip Resistance for the Excess Pore Water Pressure

According to the design equations and charts, together with Fig. 23, most of the CPT-based direct methods rely

solely on CPT tip resistance. Only a few draw on multiple readings from the CPTu data. The methods that utilize tip

stress, either solely or in combination with other CPT readings, to estimate f p and/or q b generally do not account for

correction of excess pore water pressures acting on unequal tip area of the cone to obtain corrected tip stress, q t . The

exceptions include Unicone method, KTRI method, SEU method, Dutch method, and the method by Togliani

(2008). The corrected tip resistance q t is obtained from the following expression:

q t = q c + (1 – a n )u 2

where a n = net area ratio of the particular penetrometer determined through calibration in a triaxial chamber, and u 2

is the field excess porewater pressure measured at the shoulder (behind the tip) position. In clean sands and dense

granular soils, it may be reasonable to assume q t ≈ q c because u 2 remains essentially hydrostatic. In soft to stiff

intact clayey and silty soils, however, considerable excess pore water pressures are generated during cone

penetration, warranting significant corrections to the measured q c in order to obtain q t (Mayne 2007).
5.3 Influence Zone for Pile Base Resistance

All the CPT-based direct methods relate unit base resistance to the cone tip resistance (either q c , or more proper q t ).

In doing so, the cone tip resistance data are averaged over a certain depth interval near the pile base. This range of

depth is commonly termed as an "influence zone." Different methods provide different recommendations for the

extent of influence zone above and below the pile toe, which have been detailed in Table 3. This extent, which is

principally taken to account for the rupture path around to the pile toe, has been defined on the basis of different

theories, including punching failure, general shear failure, and other regions. Eslami and Fellenius (1997) pointed

out that there is no specific evidence to support these theories in relation to the deep foundations. Yet, some

recommendations based on limited experimentation and numerical works (e.g. Meyerhof 1976; De Beer 1963;

Altaee et al. 1992; Eslami and Fellenius 1997) are commonly implemented in practice (see the method specific

details in Table 3). Various considerations that have led to these recommendations include: (1) the trend of cone tip

resistance values around the pile toe, (2) extent of soil variability around the pile toe, (3) pile diameter, (4) pile

embedment depth into the bearing soil layer, (5) existence of weak layer beneath the bearing layer, and (6) soil

compressibility.

5.4 Averaging Technique for Representative Value of Cone Tip Resistance

The averaging technique over the influence zone is another factor to be considered for adopting a representative

value of cone tip stress for use in the design equations for unit base resistance. Arithmetic averaging is the most

common technique implemented in practice. The CPT readings, typically in coarse grained soils, display squiggly

profiles of arbitrary peaks and troughs that may be representative of thin seams of variable soil. However, in relation

to their influence on the pile, having a much larger size, retaining such readings while averaging can possibly result

in non-representative values of unit base resistance. Therefore, such readings are usually filtered out applying a

"minimum path" rule (e.g. Begemann 1963), or by simply removing the peaks and troughs from the records (e.g.

Bustamante and Gianeselli 1982). Eslami and Fellenius (1997) recommend the use of a geometric averaging

technique (i.e., geomean) over arithmetic averaging, since the former automatically results in a filtered

representative value.

5.5 Shaft and Base Areas for Pile Capacity Calculations

As shown earlier, the total axial capacity (Q t ) of deep foundations is calculated from the sum of Q s and Q b , which

further involve calculation of the shaft area providing frictional resistance (A si ) and the pile base area (A b ) providing

the base resistance, respectively. The shaft area of the ith layer is simply given as the product of the pile perimeter

and the layer thickness (∆z i ). For a circular pile, the perimeter is conveniently obtained as π∙d i , where d i is the pile
diameter in the ith layer. For non-circular solid piles, an equivalent pile diameter can be adopted. Therefore, for

square and rectangular piles, equivalent pile diameter, d i = (2b i + 2w i )/π, where b i and w i are the depth and width of

the pile cross-section, respectively, in the ith layer. For piles that are not solid (e.g. OE pipe piles, H shaped piles),

selection of equivalent pile diameter becomes a bit more complicated, as described subsequently.

For OE driven pipe piles, a soil column enters into the pile through the open base during driving and forms a

plug. During the subsequent static loading of the pile, some amount of shaft friction is additionally mobilized along

the interface between the soil column and the inner pile wall. However, this frictional resistance is accounted for by

considering it part of pile base resistance (and so, it is being discussed later). Therefore, the outer diameter of OE

pipe pile is used for calculations of its shaft capacity.

For the case for H piles, the contact area contributing to shaft capacity is also complicated. During driving, soil

enters the space between the flanges of H piles. Seo et al. (2009) gave the following explanation concerning the

shaft area of H piles. "The vibrations during driving cause the soil near the ground surface to detach from the pile.

During static loading of the pile, depending on the soil type and state, the soil in the space between the flanges of the

pile may behave as a plug and therefore become an integral part of the pile. During loading of the pile, the soil plug

may be further compressed or it may slide with respect to the pile. It is difficult to ascertain in practice what occurs

at the time of loading. If the soil in the space between the flanges becomes fully attached to the pile, then the outer

perimeter of the H pile is typically used in shaft capacity calculations. Otherwise, the full steel-soil interface contact

perimeter, which includes not only the outside and inside of the flanges and their tips but also the web, is used in

shaft capacity calculations." Based on load tests on a fully instrumented H pile driven in a multilayer soil, Seo et al.

(2009) concluded that assumption of full soil-pile interface contact perimeter results in over prediction of shaft

capacity. Accordingly, they recommended two options: (1) consider a reduction factor between 0.41 and 0.88

(average 0.65) when assuming full soil-pile interface contact perimeter; or (2) assume outer perimeter in shaft

capacity calculations to get comparable estimates.

The base area (A b ) of a circular solid or CE pipe pile is simply obtained as π∙d2/4, where d = pile base diameter.

For non-circular solid piles of square or rectangular cross-sections, an equivalent base diameter is obtained from: d =

(4 b∙w/π)0.5, where b and w are the depth and width of the pile cross-section at the base, respectively. For OE pipe

piles, following the recommendations by Lehane and Gavin (2001), Gavin and Lehane (2005), and Yu and Yang

(2012), separate contributions from annulus and plug should be considered (also detailed in Table 3):

Q b = (π/4)∙[d2 q plug + (d2 – d i 2) q ann ]

(4)
where q plug = unit plug resistance of pile; q ann = unit annulus resistance of pile; d = pile outer diameter; d i = pile

inner diameter. Here, q plug accounts for the inner shaft resistance (also pointed out earlier) by incorporating the plug

length ratio (PLR = H/L) at the end of pile installation, where H = plug length measured at the end of pile

installation, and L = pile length.

For H piles, on the other hand, the base area should be selected depending upon the soil response in the space

between the flanges in the influence zone of the pile base. If the soil in the space between the flanges becomes fully

attached to the pile, then the gross cross-sectional area (flange width x depth) should be used, otherwise use the

actual cross-sectional area. Since it is difficult to ascertain the actual situation at the time of loading, use of specific

formulation for H piles given in ICP-05 method (see Table 3) is recommended (also proposed by Seo et al. 2009).

5.6 Reference to the Capacity Criteria

The load-displacement (Q – w) curves obtained from axial load tests on pile foundations can exhibit different shapes

and results. Only a single value of "axial load capacity" is selected from this entire curve for design load purposes.

As for all other geotechnical engineering applications, a common practice in the design of pile foundations is to

apply appropriate factors of safety (FS) to this axial load capacity under the allowable stress (ASD) design [also

called working stress (WSD) design], the serviceability limit state (SLS) design, or the load and resistance factors

(LRFD) design. These FS, in part, account for the inaccuracies and uncertainties of the prediction method.

There are at least 45 different criteria defining the axial load capacity on the load-displacement curves (Hirany

and Kulhawy 1988; Dećourt 1999). It has previously been observed that there is no consensus and the capacities

interpreted via different criteria span over a wide range (e.g. Niazi 2011). Some of the CPT-based direct methods do

not explicitly refer to the specific capacity criterion on the basis of which the respective design equations were

formulated, thus adding a certain level of uncertainty. Without reference to any specific criterion, the engineer in-

charge of a piling project using such methods has to considerably rely on experience and subjective engineering

judgment for applying appropriate FS to strike a balance between economical and safe design.

5.7 Versatility of CPT-based Direct Methods

It is clearly evident from the chronological evolution of CPT-based direct methods detailed above that most of the

past research and developments (even in the recent years) have addressed driven pipe piles in sands, mostly for the

offshore environment. In particular, much less efforts have been directed towards drilled shafts in clays, silts and/or

sands, as well as newer composite piles and developments in the deep foundations industry. Furthermore, since the

last work based on the database of full-scale field pile load tests, some new data have been included in the inventory
of load tests, including those using the modern bi-directional Osterberg-cell (O-cell) type of proofing method. There

is a need to incorporate this latest information in the CPT-based pile design methods.

Most existing CPT-based methods focus only on an estimate of "axial load capacity", without recourse to the

behavior in terms of the complete load-displacement-capacity response. In particular, the shear wave velocity (V s )

component of the newer seismic piezocone test (SCPTu) can provide fundamental shear stiffness (G max ) that can be

exploited within an analytical framework towards a versatile application and extension of pile design from "axial

load capacity" to "complete axial load-displacement" response. During the recent years, only piecemeal efforts have

been employed in that direction for individual case studies (e.g. Mayne and Schneider 2001; Mayne and Elhakim

2002; Mayne and Woeller 2008; Mayne and Niazi 2009; Mayne et al. 2010; Niazi and Mayne 2010; Niazi et al.

2010a, b, c; Reuter, 2010; Pando et al. 2004).

6 Summary and Recommendations for Future Research

A concise review of the direct CPT-based methods that focus on axial pile capacity estimation was presented.

Thereupon, an all-encompassing list of methods was organized showing their respective design equations, and

pertinent pile and soil types. Various design graphs/charts as applicable to different methods were also presented.

This will serve as a single convenient resource for the practitioners and researchers working in pile design and

analysis.

The design equations presented in Table 3 and their corresponding charts are based largely on empirical

methods that have evolved through years of experience. Some of these methods are applicable to a wide variety of

geological and geotechnical settings, and pile types, while others concern only to specific limited cases. Randolph

and Wroth (1982) emphasized that such approaches may work well in familiar situations, but there will inevitably be

uncertainty when novel structures or new types of soil deposits are encountered. Eslami and Fellenius (1997) noted

for Unicone method that further correlation experience will result in adjustments of the correlation coefficients used

in their design equations. Doherty and Gavin (2011) also indicated a similar observation on the extension of such

methods to the design situations which are outside the scope of the databases used in their respective calibrations.

From experience of the authors in application of these methods to different case studies (e.g. Niazi and Mayne

2010; Niazi et al. 2010a, b, c) and through extensive review of the literature, it is recognized that the reliability of no

single method may be regarded as ultimate or universal. This argument is also valid from the fact that different

methods were empirically derived using diverse datasets of piles installed in different types of geomaterials. Other
researchers have also attempted to study the reliability of few selected methods (e.g. Abu-Farsakh and Titi 2004;

Ardalan et al. 2009); however, these were also accomplished on specific piles tested in certain particular

geotechnical settings. In view of the authors, the more recent methods that exploit maximum geotechnical

parameters from the CPT, and which were developed from larger and latest database of pile load tests may be

considered for prediction analysis. However, the results must be checked against the estimates of other recent direct

methods as well as the rational methods before finalizing the design.

Based on this review and its ensuing discussion, following recommendations are offered towards future

application and research on the topic:

1. Due caution must be exercised when applying these empirical approaches to field situations of relevance

encountered in design practice. Engineering judgment must remain as the hallmark for interpretation of the data.

2. Evolution of these methods must continuous by supplementing the database with newer piles and latest

geotechnical site investigations using the modern multi-channel hybrid geophysical-geotechnical SCPTu. The

load test database should include any latest static top-down compression type as well as the new O-cell type of

bi-directional full-scale proofing arrangements.

3. Optimal use of all the readings of SCPTu must be exploited for tying with the complete axial pile response.

Thus, it is prudent to develop on earlier methods that utilize all three penetrometer readings (q t , f s , u 2 ) for axial

capacity evaluations (e.g. Unicone method), rather than work on new method based on limited record. The

geophysical component of V s should be combined in an appropriate empirical-analytical framework to derive a

reliable method for predicting complete load-displacement response under axial loading.

4. Any future work on pile-CPT correlations must attempt to include all soil types in general and clays, silts and

mixed soil types, in particular.

5. Such further developments must explicitly indicate the measurement basis for f p and q b from their datasets, and

the average relationships thereof with some of the more frequently used capacity criteria.

Acknowledgments

The authors thank ConeTec Investigations for their continued support of in-situ testing research at Georgia Tech.
References

Abu-Farsakh MY, Titi HH (2004) Assessment of direct cone penetration test methods for predicting the ultimate
capacity of friction driven piles. ASCE J Geotech Geoenviron Eng 130(9):935–944
Almeida MSS, Danziger FAB, Lunne T (1996) Use of the piezocone test to predict the axial capacity of driven and
jacked piles in clay. Can Geotech J 33(1):33–41
Alsamman OM (1995) The use of CPT for calculating axial capacity of drilled shafts. PhD Thesis, UIUC, IL, 299 p
Altaee A, Fellenius BH, Evgin E (1992) Axial load transfer for piles in sand I: Tests on an instrumented precast pile.
Can Geotech J 29(1):11–20
Aoki N, Velloso DA (1975) An approximate method to estimate the bearing capacity of piles. In: Proceedings of 5th
Pan-American conference of soil mechanics and foundation engineering, Buenos Aires, pp 367–376
API (1993) Recommended practice for planning, designing and constructing fixed offshore platforms – working
stress design. API RP2A, 20th ed. American Petroleum Institute, Washington, DC
API (1987) Recommended practice for planning, designing, and constructing fixed offshore platforms. API RP2A,
17th ed. American Petroleum Institute, Washington, DC
API (1976) Recommended practice for planning, designing, and constructing fixed offshore platforms. API RP2A,
7th ed. American Petroleum Institute, Washington, DC
API (1975) Recommended practice for planning, designing, and constructing fixed offshore platforms. API RP2A,
6th ed. American Petroleum Institute, Washington, DC
API (1969) Recommended practice for planning, designing, and constructing fixed offshore platforms. API RP2A,
1st ed. American Petroleum Institute, Washington, DC
Ardalan H, Eslami A, Nariman-Zahed N (2009) Piles shaft capacity from CPT and CPTu data by polynomial neural
networks and genetic algorithms. Computers and Geotechnics 36:616–625
Begemann HKSPh (1969) The Dutch static penetration test with the adhesion jacket cone. Delft Soil Mechanics and
Delft Hydraulics Laboratory, XIII(1): p. 1
Begemann HKSPh (1965) The maximum pulling force on a single tension pile calculated on the basis of results of
the adhesion jacket cone. In: Proceedings of the 6th international conference on soil mechanics and foundation
engineering, Montreal 2: p. 229
Begemann HKSPh (1963) The use of the static penetrometer in Holland. New Zealand Engineering 18(2): p. 41
Bond AJ (l989) Behaviour of displacement piles in overconsolidated clays. PhD Thesis, Imperial College, London
Bogdanovic L (1961) The use of penetration tests for determining the bearing capacity of piles. In: Proceedings of
the 5th international conference on soil mechanics and foundation engineering, Paris 2: p. 17
Burland JP (1993) Closing address. In: Proceedings of recent large-scale fully instrumented pile tests in clay.
Institute of Civil Engineers, London, pp 590–595
Burland JP (1973) Shaft friction of piles in clay – a simple fundamental approach. Ground Eng 6(3):30–42
Bustamante M, Frank R (1997) Design of axially loaded piles—French practice. In: Design of axially loaded piles –
European practice, Proceedings of the ERTC3 seminar, Brussels, Belguim, Balkema, Rotterdam, The
Netherlands: 161–175
Bustamante M, Gianeselli L (1982) Pile bearing capacity predictions by means of static penetrometer CPT. In:
Proceedings of 2nd European symposium on penetration testing (ESOPT-II), Amsterdam, pp 493–500
Cai G, Liu S, Puppala AJ (2012) Reliability assessment of CPTu-based pile capacity predictions in soft clay
deposits. Eng Geol 141–142:84–91
Cai G, Liu S, Puppala AJ (2011) Evaluation of pile bearing capacity from piezocone penetration test data in soft
Jiangsu Quaternary clay deposits. Mar Geores Geotechnol 29:177–201
Cai G, Liu S, Tong L, Du G (2009) Assessment of direct CPT and CPTu methods for predicting the ultimate bearing
capacity of single piles. Eng Geol 104:211–222
Chandler RJ (1968) The shaft friction of piles in cohesive soils in terms of effective stress. Civ Eng and Pub Works
Review, 60(708):48–51
Chen YJ, Kulhawy FH (1994) Case history evaluation of the behavior of drilled shafts under axial and lateral
loading. EPRI TR-104601, Research Project No. 1493–04
Chen J-J, Zhang L (2013) The effect of spatial correlation of cone tip resistance on the bearing capacity of piles. J
Geotech Geoenviron Eng (manuscript accepted, posted ahead of print)
Chow FC (1997) Investigations into displacement pile behaviour for offshore foundations. PhD Thesis, Imperial
College, London
Clausen CJF, Aas PM, Karlsrud K (2005) Bearing capacity of driven piles in sand, the NGI approach. In:
Proceedings of international symposium on frontiers in offshore geomechanics (ISFOG 2005), Perth, Taylor
& Francis, London, pp 677–681
Clisby MB, Scholtes RM, Corey MW, Cole HA, Teng P, Webb JD (1978) An evaluation of pile bearing capacities,
Vol. I, Final Report, Mississippi State Highway Department
De Beer EE (1963) The scale effect in the transposition of the results of deep sounding tests on the bearing capacity
of piles and caisson foundations. Géotechnique 13(1):39–75
De Nicola A, Randolph MF (1999) Centrifuge modelling of pipe piles in sand under axial loads. Géotechnique
49(3):295–318
De Nicola A, Randolph MF (1993) Tensile and compressive shaft capacity of piles in sand. J Geotech Eng Div
ASCE 119(12):1952–1973
De Ruiter J, Beringen FL (1979) Pile foundations for large North Sea structures. Mar Geotechnol 3(3):267–314
Dećourt L (1999) Behavior of foundations under working load conditions. In: Proceedings of the 11th Pan-American
conference on soil mechanics and geotechnical engineering, Foz DoIguassu, Brazil, 4:453–488
Doherty P, Gavin K (2011) The shaft capacity of displacement piles in clay: a state of the art review. Geotech Geol
Eng 29(4):389–410
Drewry JM, Weidler JB, Hoang ST (1977) Predicting axial pile capacities for offshore platforms. Pet Eng: 41–44
Eslami A, Fellenius BH (1997) Pile capacity by direct CPT and CPTu methods applied to 102 case histories. Can
Geotech J 34(6):880–898
Fellenius BH (2002) Excerpt from Chapter 6 of the red book: direct methods for estimating pile capacity, in
Background to UniCone http://www.fellenius.net, accessed May 22, 2010
Fioravante V, Ghionna VN, Jamiolkowski M, Pedroni S (1995) Load carrying capacity of large diameter bored piles
in sand and gravel. In: Proceedings of 10th Asian regional conference on soil mechanics and foundation
engineering, Beijing 2:3–15
Fioravante V (1994) Centrifuge modeling of driven and bored piles in sand subjected to axial load. Workshop on
pile foundations experimental investigation, analysis and design, Napoli, pp 125–163
Flaate K, Selnes P (1977) Side friction of piles in clay. In: Proceedings of the 9th international conference on soil
mechanics and foundation engineering, Tokyo, pp 517–522
Fleming WGK, Weltman AJ, Randolph MF, Elson WK (1992) Piling Engineering. 2nd ed., Blackie/Halsted Press,
John Wiley and Sons, NY, 390 p
Gavin KG, Lehane BM (2005) Estimating the end bearing resistance of pipe piles in sand using the final filling ratio.
In: Proceedings of international symposium on frontiers in offshore geomechanics (ISFOG 2005), Perth,
Taylor and Francis Group, London, pp 717–723
Gavin KG, Lehane BM (2003) The shaft capacity of pipe piles in sand. Can Geotech J 40:36–45
Gwizdala K (1984) Large diameter bored piles in non-cohesive soils. Swedish Geotechnical Institute, Report No. 26,
Linkoping, Sweden
Hirany A, Kulhawy FH (1988) Conduct and interpretation of load tests on drilled shaft foundations. Report Electric
Power Research Institute EL-5915, 1, Palo Alto, CA
Huizinga TK (1951) Application of results of deep penetration tests to foundation piles. In: Proceedings of building
research congress, London, Div. 1, Part 3: 173 p
Igoe DJP, Gavin KG, O'Kelly BC (2011) Shaft capacity of open-ended piles in sand. J Geotech Geoenviron Eng
137(10):903–913
Igoe DJP, Gavin KG, O'Kelly BC(2010) Field tests using an instrumented model pile in sand. In: Proceedings of
physical modelling in geotechnics, Taylor and Francis Group, pp 775–780
Jamiolkowski M (2003) Deep foundations on bored and auger piles. In: Proceedings of the 4th international
geotechnical seminar on deep foundations on bored and auger piles, Ghent, pp 83–100
Jardine RJ, Chow FC, Overy R, Standing J (2005) ICP design methods for driven piles in sands and clays. Thomas
Telford Publishing, London, 105 p
Jardine RJ (1985) Investigations of pile-soil behaviour with special reference to the foundations of offshore
structures. PhD Thesis, Imperial College, London
Karlsrud K (2012) Prediction of load-displacement behavior and capacity of axially-loaded piles in clay based on
analyses and interpretation of pile load test results. PhD Dissertation, Norwegian University of Science and
Technology, Trondheim: 320 p
Karlsrud K, Clausen CJF, Aas PM (2005) Bearing capacity of driven piles in clay, the NGI approach. In:
Proceedings of frontiers in offshore geotechnics (ISFOG 2005), Perth, Taylor and Francis Group, London, pp
775–782
Karlsrud K, Hansen SB, Dyvik R, Kalsnes B (1993). NGI's pile tests at Tilbrook and Pentre – review of testing
procedures and results. Large-Scale Pile Tests in Clay. Thomas Telford, London: 549–583
Kempfert H-G, Becker P (2010) Axial pile resistance of different pile types based on empirical values. In:
Proceedings of Geo-Shanghai 2010 deep foundations and geotechnical in-situ testing (GSP 205), ASCE,
Reston, VA, pp 149–154
Kerisel JL (1965) Vertical and horizontal bearing capacity of deep foundations in clay. Bearing capacity and
settlement of foundations, Duke University, Durham, NC: 45–51
Kerisel JL (1964) Deep foundations basic experimental facts. In: Proceedings of conference on deep foundations,
Mexico City 1, 7 p
Kolk HJ, Baaijens AE, Senders M (2005) Design criteria for pipe piles in silica sands. In: Proceedings of
international symposium on frontiers in offshore geomechanics (ISFOG 2005), Perth, Taylor and Francis
Group, London, pp 711–716
Kolk HJ, van der Velde E (1996) A reliable method to determine the friction capacity of piles driven into clays. In:
Proceedings of 28th annual offshore technology conference, Houston, TX, pp 337–346
Kraft LM, Focht JA, Amerasinghe SF (1981) Friction capacity of piles driven into clay. J Geotech Eng Div
107:1521–1541
Kulhawy FH (2004) On the axial behavior of drilled foundations. GeoSupport 2004: Drilled shafts, micropiling,
deep mixing, remedial methods, and specialty foundation systems (GSP 124), JP Turner and PW Mayne, eds.,
ASCE, Reston, VA, pp 34–51
Kulhawy FH, Phoon K (1993) Drilled shaft side resistance in clay soil to rock. Design and performance of deep
foundations: piles and piers in soil and rock, ASCE (GSP 38), pp 172–183
Kulhawy FH, Mayne PW (1990) Manual on estimating soil properties for foundation design. Report EL-6800,
Research project 1493-6, Electric Power Research Institute, Palo Atlo, CA
Kulhawy FH, Trautmann CH, Beech JF, O'Rourke TD, McGuire W (1983) Transmission line structure foundations
for uplift-compression loading." Report EL-2870, Electric Power Research Institute, Palo Alto, CA, 412 p
Lehane BM, Schneider JA, Xu X (2005) The UWA-05 method for prediction of axial capacity of driven piles in
sand. In: Proceedings of international symposium on frontiers in offshore geomechanics (ISFOG 2005),
Perth, Taylor and Francis, London, pp 683–689
Lehane BM, Gavin KG (2001) Base resistance of jacked pipe piles in sand. J Geotech Geoenviron Eng 127(6):473–
480
Lehane BM (1992) Experimental investigations of pile behaviour using instrumented field piles. PhD Thesis,
lmperial College, London
Lee JH, Salgado R (1999) Determination of pile base resistance in sands. J Geotech Geoenviron Eng 125(8):673–
683
Lunne T, Robertson PK, Powell JJM (1997) Cone penetration testing in geotechnical practice, Blackie
Academic/London, Routledge, NY, 312 p
Mayne PW, Niazi FS, Woeller DJ (2010) Drilled shaft response in Piedmont residuum using elastic continuum
analysis and seismic piezocone tests. In: Proceedings of Geo-Shanghai 2010 deep foundations and
geotechnical ln-situ testing (GSP 205), ASCE, Reston, Virginia, pp 200 – 205
Mayne PW, Niazi FS (2009) Evaluating axial elastic pile response from cone penetration tests. (2009 Michael
O’Neill Lecture), J Deep Found Inst 3(1):3– 2
Mayne PW (2007) Cone penetration testing – a synthesis of highway practice. NCHRP Synthesis 368,
Transportation Research Board, Washington, DC, 117 p
Mayne PW, Woeller DJ (2008) O-cell response using elastic pile and seismic piezocone tests. In: Proceedings of the
2nd British geotechnical association international conference on foundations (ICOF 2008), Dundee, IHS BRE
Press, UK, 1, pp 235–246
Mayne PW, Elhakim A (2002) Axial pile response evaluation by geophysical piezocone tests. In: Proceedings of the
9th international conference on piling and deep foundations, DFI, Nice, Presses de l'ecole nationale des Ponts
et chaussees, pp 543–550
Mayne PW, Schneider JA (2001) Evaluating axial drilled shaft response by seismic cone. Foundation and ground
improvement (GSP No. 113), ASCE, Reston, VA, pp 655–669
McClelland B (1974) Design of deep penetration piles for ocean structures. J Geotech Eng Div 100:709–747
Menzenbach E (1961) The determination of permissible point loads of piles by means of static penetration tests. In:
Proceedings of the 5th international conference on soil mechanics and foundation engineering, Paris 2: 99 p
Meyerhof GG (1983) Scale effects of ultimate pile capacity. J Geotech Eng 109:797–806
Meyerhof GG (1976) Bearing capacity and settlement of pile foundations. J Geotech Eng Div 102:195–228
Meyerhof GG (1956) Penetration tests and bearing capacity of cohesionless soils. ASCE J Soil Mech Found Div 82,
pp 866–1019
Meyerhoff GG (1951) The ultimate bearing capacity of foundations. Géotechnique 2(4): 301 p
Miller GA, Lutenegger AJ (1997) Influence of pile plugging on skin friction in overconsolidated clay. J Geotech
Geoenviron Eng 123: 525–533
Niazi FS (2011) Axial pile displacement evaluations from seismic piezocone data and backanalysis of load tests. In:
Proceedings of the DFI 36th annual conference on deep foundations. Boston, MA: 191 – 200.
Niazi FS, Mayne PW (2010) Evaluation of EURIPIDES pile load tests response from CPT results. International J
Geoeng Case Hist 1(4):367–386 www.geoengineer.org
Niazi FS, Mayne PW, Woeller DJ (2010a) Review of CPT-based methods for response evaluation of driven piles in
dense sands. In: Proceedings of international conference on geotechnical engineering, Pakistan: 259–266
Niazi FS, Mayne PW, Woeller DJ (2010b) Drilled shaft O-Cell response at Golden Ears bridge from seismic cone
tests. The art of foundation engineering practice, ASCE (GSP No. 198), Reston, VA, pp 452–469
Niazi FS, Mayne PW, Woeller DJ (2010c) Case history of axial pile capacity and load-settlement response by
SCPTu. In: Proceedings of the 2nd international symposium on cone penetration testing (CPT’10), Huntington
Beach, CA, 3, pp. 9–16
Nottingham LC (1975) Use of quasi-static friction cone penetrometer data to predict load capacity of displacement
piles. PhD Thesis, University of Florida
Pando MA, Fernandez AL, Filz GM (2004) Pile settlement predictions using theoretical load transfer curves and
seismic CPT data. In: Proceedings of the 2nd international conference on site characterization (ISC-2):
Geotechnical and geophysical site characterization, Millpress, Rotterdam, pp. 1525–1531
Patel DC (1989) A case study of the shaft friction of bored piles in London clay in terms of effective stress. Thesis,
University of London, Imperial College, as referenced in Patrizi and Burland (2001)
Patrizi P, Burland J (2001) Development in the design of driven piles in clay in terms of effective stresses. Rivista
ltaliana di Geotecnica, 35(3):35–49
Peck RB (1958) A study of the comparative behavior of friction piles. Special Report No. 36, Highway Research
Board
Pelletier JH, Doyle EH (1982) Tension capacity in silty clays – Beta pile tests. In: Proceedings of the 2nd
international conference on numerical methods in offshore piling, Austin, pp 1–19
Philipponnat G (1980) Methode pratique de calcul d'un pieu isole a l'aide du penetrometre statique. Revue Française
de Géotechnique, 10:55–64
Plantema G (1948) Results of a special loading test on a reinforced concrete pile, a so-called pile sounding. In:
Proceedings of the 2nd international conference on soil mechanics and foundation engineering, Rotterdam 2,
112 p
Poulos HG (1989) Pile behavior: theory and applications. Rankine Lecture, Géotechnique 39(3):363–415
Powell JJM, Lunne T, Frank R (2001) Semi-empirical design for axial pile capacity in clays. In: Proceedings of 15th
international conference on soil mechanics and geotechnical engineering, (Istanbul), Balkema, Rotterdam,
The Netherlands, pp 991–994
Powell JJM, Quarterman RST (1988) The interpretation of cone penetration tests in clays with particular reference
to rate effects. Penetration Testing 1988, Vol. 2, (Orlando, FL), Balkema, Rotterdam, The Netherlands, pp
903–909
Price G, Wardle IF (1982) A comparison between cone penetration test results and the performance of small
diameter instrumented piles in stiff clay. In: Proceedings of the 2nd European symposium on penetration
testing, Amsterdam, 2, pp 775–780
Randolph MF (1983) Design considerations for offshore piles. In: Proceedings of conference on geotechnical
practice in offshore engineering, Austin, pp 422–439
Randolph MF, Murphy BS (1985) Shaft capacity of driven piles in clay. In: Proceedings of 17th annual offshore
technology conference, Houston, 1:371–378
Randolph MF, Wroth CP (1982) Recent developments in understanding the axial capacity of piles in clay. Ground
Eng 15(7):17–25
Reese LC, O'Neill MW (1988) Drilled shafts: construction practices and design methods, Publication No. FHWA-
HI-88-042, Federal Highway Administration, Office of Implementation, Washington, DC, 564 p
Reuter GR (2010) Pile capacity prediction in Minnesota soils using direct CPT and CPTu methods. In: Proceedings
of the 2nd international symposium on cone penetration testing (CPT’10). Huntington Beach, CA, Omnipress
Salgado R (2006) The role of analysis in non-displacement pile design. Mod Trends Geomech, In: Springer
Proceedings in Physics, 106, pp 521–540
Schmertmann JH (1978) Guidelines for cone penetration test, performance and design. U.S. Department of
Transportation, Washington, DC, Report No. FHWA-TS-78-209, 145 p
Semple RM, Rigden WJ (1984) Shaft capacity of driven pipe piles in clay. In: Proceedings of symposium on
analysis and design of deep foundations, San Francisco, pp 59–79
Seo H, Yildirim IZ, Prezzi M (2009) Assessment of the axial load response of an H pile driven in multilayered soil.
J Geotech Geoenviron Eng 135(12):1789–1804
Skempton AW (1959) Cast in-situ bored piles in London clay. Géotechnique 9(2):91–125
Takesue K, Sasao H, Matsumoto T (1998) Correlation between ultimate pile skin friction and CPT data.
Geotechnical Site Characterization (2), Rotterdam, pp 1177–1182
Togliani G (2010) Pile capacity prediction using CPT – case history. In: Proceedings of the 2nd international
symposium on cone penetration testing, CPT'10, Huntington Beach, CA, Paper No. 3–13. www.cpt10.com,
accessed Nov 28, 2012
Togliani G (2008) Pile capacity prediction for in-situ tests. In: Proceedings of geotechnical and geophysical site
characterization, Taylor and Francis Group, London, pp 1187–1192
Tomlinson MJ (1957) The adhesion of piles driven in clay soils. In: Proceedings of the 4th international conference
on soil mechanics and foundation engineering, London
Toolan FE, Lings ML Mirza UA (1990) An appraisal of API RP2A recommendations for determining skin friction
of piles in sand. In: Proceedings of 22nd offshore technology conference, Houston, pp 33–42
Tumay MT, Fakhroo M (1982) Friction pile capacity prediction in cohesive soils using electric quasi-static
penetration tests. Interim Research Report No. 1, Louisiana Department of Transportation and Development,
Research and Development Section, Baton Rouge, LA, 275 p
Twine D (1987) Shaft friction of bored, cast in-situ piles in stiff, overconsolidated clays in terms of effective stress.
Thesis, University of London, Imperial College, as referenced in Patrizi and Burland (2001)
Van der Veen C, Boersma L (1957) The bearing capacity of a pile predetermined by a cone penetration test. In:
Proceedings of the 4th international conference on soil mechanics and foundation engineering, London 2, 72 p
Van Dijk BFJ, Kolk HJ (2011) CPT-based design method for axial capacity of offshore piles in clays. In:
Proceedings of frontiers in offshore geotechnics II, Taylor and Francis Group, London, pp 555–560
Vesić AS (1964) Investigations of bearing capacity of piles in sand. In: Proceedings of the conference on deep
foundations, Mexico City 1, 197 p
Vijayvergiya VN, Focht JA (1972) A new way to predict the capacity of piles in clay. In: Proceedings of 4th annual
offshore techonology conference, Houston, pp 269–284
White DJ, Bolton MD (2005) Comparing CPT and pile base resistance in sand. Proc Inst Civil Eng Geotech Eng
158(GE1):3–14
Woodward RJ, Lundgren R, Boitano JD (1961) Pile loading tests in stiff clays. In: Proceedings of the 6th
international conference on soil mechanics and foundation engineering, Paris, pp 177–184
Yu F, Yang J (2012) Base capacity of open-ended steel pipe piles in sand. J Geotech Geoenviron Eng 138(9):1116–
1128
Table 1 Factors considered in the total stress approach (α-methods) for estimating pile unit shaft resistance (f p )
Method/Reference Length Stress Ip su σ vo ' Progressive Plugging
effect history failure effect

Tomlinson (1957) x x x √ x x x

Peck (1958) x x x √ x x x

Skempton (1959) x x x √ x x x

Woodward et al. (1961) x x x √ x x x

Kerisel (1965) x x x √ x x x

API (1969) x √ x x √ x x

McClelland (1974) x √ x √ x x x

Vijayvergiya and Focht (1972) √ x x √ √ x x

Drewry et al. (1977) x x x √ x x x

API (1975, 1976) x √ √ √ √ x x

Kraft et al. (1981) √ x x √ √ √ x

Randolph (1983) √ x x √ x √ x

Semple and Rigden (1984) √ √ x √ √ x x

Randolph and Murphy (1985) x √ x √ √ x x

API (1987) x √ x √ √ x x

API (1993) x √ x √ √ x x

Karlsrud et al. (1993) x x √ √ √ x x

Chen and Kulhawy (1994) x x x √ x x x

Kolk and van der Velde (1996) √ √ x √ √ x x

Miller and Lutenegger (1997) x x x x x x √

Jamiolkowski (2003) x √ x √ √ x x

NGI-05 (Karlsrud et al. 2005) x √ √ √ √ x x

Salgado (2006) x √ x √ x x x
German Method (Kempfert and
x x x √ x x x
Becker 2010)

Karlsrud (2012) x √ √ √ √ x x
Notes: I p = plasticity index; s u = undrained shear strength; σ vo ' = effective overburden stress
Table 2 Factors considered in the effective stress approach (β-method) for estimating pile unit shaft resistance (f p )
Method/Reference σr' δ φ' OCR K σ vo ' L d su Dr St Ip

Chandler (1968); Burland (1973); Pelletier and Doyle (1982) √ √ √ x √ √ x x x x x x

Meyerhof (1976) x √ √ √ x √ x x x x x x

Flaate and Selnes (1977) x x x √ x √ √ x x x x x

Kulhawy et al. (1983) √ √ √ x √ √ x x x x x x

Twine (1987); Patel (1989) x x x x x √ x x x x x x

Reese and O'Neill (1988) x x x x x √ √ x x x x x

Fleming et al. (1992) √ √ x x √ √ x x x x x x

Burland (1993) x x x √ x √ x x √ x x x

de Nicola and Randolph (1999) x x x x x √ x x x √ x x

Patrizi and Burland (2001) x x x √ x √ x x √ x x x

ICP-05 Method (Jardine et al., 2005) √ √ x √ √ x √ √ √ x √ √

Karlsrud (2012) x x x √ x √ x x x x x √
Notes: σ r ' = radial effective stress; δ = soil-pile interface friction angle; φ' = effective friction angle; OCR (= σ p '/σ vo ') is the overconsolidation ratio; Κ (= σ r '/σ v ')
radial effective stress coefficient; σ vo ' = effective vertical stress; L = pile length; d = pile diameter; s u = undrained shear strength of clay; D r = relative density of
sand; S t = clay’s sensitivity; I p = plasticity index
Table 3 Summary of direct CPT-based pile design methods
Method/Reference Design Equations

Pile unit side resistance (f p ) Pile unit end bearing (q b )

Bogdanovic (1961) f p = f s ∙[(π∙d CPT )/(π∙d)] q b = q ca(tip)


(for driven and jacked concrete piles
in dense sand)

Begemann (1963; 1965; 1969) f p = fctn(q c , f s and pile type) q b = (q c1 + q c2 )/2 (also see Fig. 5)
(for driven piles in sandy soils) (Fig. 6) q c1 = [(q 1 + q 2 + q 3 + ... + q n ) + nq n ]/2n

Meyerhof (1956; 1976; 1983) f p = n sf ∙f s , or Driven piles:


(for driven piles and drilled shafts in f p (kPa) = 0.5∙n sq ∙q c (q c in MPa) Short piles in sand (where L < L c ):
sandy soils) q b = q l ∙L/L c < q l
n sf = 1 (driven piles), 0.7 (drilled shafts)
n sq = 1 (driven piels); 0.5 (drilled shafts) Long piles driven through weak strata to bearing
embedment depth, z d (where L > L c ):
Deep firm sand deposit of great thickness, H (H/d > 20):
q b = q l1 + (q l2 – q l1 )∙z d /(10d) < q l2
(also see Figs 2 and 3)
Thin sand bearing layer (H/d < 20) overlying a weak
deposit: q b = q l2 + (q l1 – q l2 )∙H'/(10d) < q l1
(also see Figs 2 and 3)

For piles with d > 0.5m multiply the above values of q b


with reduction factor, R b :
R b = [(d + 0.5)/2d]n < 1
where n = 1 for loose sand, 2 for medium dense sand, and
3 for dense sand

Drilled shafts:
Reduce q b to 30% of that determined form above

Notes: d CPT = penetrometer diameter; d = pile diameter; q ca(tip) = average of q c values from an envelope drawn using minimum path rule over the interval: 4.0d
above the pile base to 2.0d below the base; q c1 = average of q c over 3.5d below the pile base; q c2 = average q c over 8d above the pile base; n sf and n sq are the
reduction factors applied to the unit shaft resistance according to the pile type; L and L c are pile embedment and critical depths, respectively (L c = 10d – 40d); q l
= limiting unit base resistance (= limiting static cone resistance) = average of q c in a zone ranging from 4d above to 1d below pile base; z d = embedment depth in
dense sand layer (in m); q ll and q l2 are the limiting unit base resistances in upper/middle and lower soil strata = average of q c in the respective strata (Fig. 2); H' =
distance between pile base and the surface of underlying weak deposit (Fig. 2).
Table 3 continued
Method/Reference Design Equations

Pile unit side resistance (f p ) Pile unit end bearing (q b )

Aoki and Velloso (1975) f p = q ca(side) α s /F S < 120 kPa q b = q ca(tip) /F b ≤ 15 MPa
(for piles in all soil types)
α s (%) depends on soil type: sand = 1.4; silty sand = Empirical factor F b depends on pile type: drilled shafts =
2.0; sandy silt = 2.2; silty sand with clay or sandy clay 3.5; driven cast-in-situ = 2.5; steel and PCC = 1.75
= 2.4; clay-sand-silt mix = 2.8 – 3.0; clayey silt = 3.4;
silty clay = 4.0; clay = 6

Empirical factor F S depends on pile type: drilled shafts


= 7.0; driven cast-in-situ = 5.0; steel and PCC = 3.5

Nottingham (1975); In clay: f p = α clay f s ≤ 120 kPa, q b = (q c1 + q c2 ) /2 ≤ 15 MPa (in sands) and 10 MPa (in
Schmertmann (1978) α clay = fctn(f s ) = 0.2 – 1.25 (see Fig. 7 for different pile very silty sands) (also see Fig. 6 for q c1 and q c2 )
(for driven concrete, steel and timber types)
piles, and drilled shafts in all soil For mechanical penetrometers in clay, multiply computed
types) In sand: Q s = α sand q c by 0.6 to account possible increase in q c resulting from
α sand = fctn(z/d) (Fig. 8); for mechanical cone α sand ≈ tip mantle
55% of the value obtained from Fig. 8

For drilled shafts, reduce f p calculated from driven pile


of same size by 25%
Penpile Method f p (in MPa) = f s /(1.5 + 14.47f s ) (f s in MPa) q b = 0.25q ca(tip) for pile tip in clay
(Clisby et al. 1978) q b = 0.125q ca(tip) for pile tip in sand
(for piles in all soil types)

Dutch Method In clay: f p = α s u(side) < 120 kPa; α = 1 for NC clay and In clay: q b = N c s u(tip) ≤ 15 MPa, s u(tip) = q ca(tip) /N kt ,
(de Ruiter and Beringen 1979) 0.5 for OC clay; s u(side) = q ca(side) /N kt ; N kt = 15–20 N c = 9; N kt = 15–20; q ca(tip) = (q c1 + q c2 ) /2
(for offshore piles in all soil types) In sand: f p = min[f s , q c(side) /300 for compression, In sand: similar to Nottingham (1975) and Schmertmann
q c(side) /400 for tension, 120 kPa] (1978) method

Notes: q ca = arithmetic average of q c in a specified zone/soil layer that depends on the method; y = depth at which side resistance is calculated; d = pile diameter;
A s = shaft area; z/d = pile embedment depth to diameter ratio; q c1 = minimum of the average q c values of zones ranging from 0.7d to 3.75d below the pile tip and
q c2 = average minimum q c values over 8d above the pile tip (use minimum path rule); NC = normally consolidated; OC = overconsolidated; α = adhesion factor;
N kt = cone factor range between 15 and 20 depending on local experience; N c = bearing capacity factor.
Table 3 continued
Method/Reference Design Equations

Pile unit side resistance (f p ) Pile unit end bearing (q b )

Philipponnat (1980) f p = q ca(side) α s /F S < f p (max) q b = k b q ca(tip)


(for all pile types in all soil types)
α s = 1.25 (driven PCC piles and drilled shaft with k b : depends on soil type = 0.35 for gravel; 0.4 for sand;
casing); 0.85 [drilled shaft (d < 1.5 m)]; 0.75 [drilled 0.45 for silt; and 0.5 for clay
shaft (d > 1.5 m)]; 1.10 [H-piles (circumscribed q ca(tip) = [q ca(A) + q ca(B) ]/2
perimeter)]; 0.6 (driven/jacked steel pipe piles); 0.3
(OE steel pipe pile)

F S = 50 (clay and calcareous clay); 60 (silt, sandy clay


and clayey sand); 100 (loose sand); 150 (medium dense
sand); 200 (dense sand and gravel)

f p (max) = 120 [driven PCC piles, H-piles (circumscribed


perimeter) and drilled shaft with casing]; 100 [drilled
shaft (d < 1.5 m)]; 80 [drilled shaft (d > 1.5 m)]; 50
(driven/jacked steel pipe piles); 25 (OE steel pipe pile)

LCPC or French Method f p = q side /k s < f p(max) q b = k b q eq (tip) [see Fig. 10 for q eq (tip) ]
(Bustamante and Gianeselli 1982;
Bustamante and Frank 1997) k s = 30 – 150 depending on soil type, pile type, k b for non-displacement pile: 0.375 (clay and/or silt), 0.15
(for all pile types in all soil types) and installation procedure (sand and/or gravel), 0.20 (chalk)
[also see Fig. 9 for specific values of k s and f p(max) ] k b for displacement pile: 0.60 (clay and/or silt), 0.375
(sand and/or gravel), 0.40 (chalk)
Cone-m Method f p = m f sa < f p(max) (< 72 kPa) q b = (q c1 + q c2 ) /4 + q a /2 ≤ 15 MPa
(Tumay and Fakhroo 1982)
(for piles in clays) m = 0.5 + 9.5e(–9fsa)

Price and Wardle (1982) fp = ks fs q b = k b q c(tip)


(for driven and jacked piles, and where k s = 0.53 (driven piles), 0.62 (jacked piles), and
drilled shafts in stiff clay) 0.49 (drilled shafts) k b = 0.35 (driven piles) and 0.3 (jacked piles)
Notes: q ca = average q c in specified zone; q ca(A) and q ca(B) = average q c within 3d above and below the pile base, respectively; moreover, after removing extreme
values of q c , if q ca(A) > q ca(B) , take q ca(A) = q ca(B) ; q eq(tip) = equivalent average of q c values of zone ranging from 1.5d below pile tip to 1.5d above pile tip; d = pile
diameter; f sa = F t /L = average layer friction in tons/ft2, F t = total sleeve friction determined for pile penetration length (L) in the layer; q c1 : average of q c values
of zones 4d below tip; q c2 : average of the minimum q c values 4d below the cone tip; q a : average of the minimum q c values 8d above the pile base.
Table 3 continued
Method/Reference Design Equations

Pile unit side resistance (f p ) Pile unit end bearing (q b )

Gwizdala (1984) f p = k s q ca(side) q b = k b q ca(tip) (for d < 5); q b = (5/d)0.5 k b q ca(tip) (for d > 5)
(for drilled shafts in cohesionless
soils) Respective values of k s for q c values of 2.5, 10 and 20 Respective values of k b for q c values of 2.5, 5, 7.5, 10, 15
MPa = gravel: (80, 120, 180); coarse to medium sand: and 20 MPa = gravel: (0.8, 0.65, 0.54, 0.45, 0.35, 0.3);
(100, 150, 230); fine sand, silty sand: (130, 190, 300) coarse to medium sand: (0.7, 0.55, 0.45, 0.36, 0.27, 0.23);
fine sand, silty sand: (0.6, 0.47, 0.39, 0.31, 0.22, 0.18)
Kulhawy and Phoon (1993); f p = 0.5σ atm (s uCIUC /σ atm )0.5 q b = 0.62∙q c(net) ; using N kt = 15
Kulhawy (2004); Lunne et al. (1997) f p = 0.158[q c(net) ∙σ atm ]0.5 using N kt = 10 for CIUC
(for drilled shafts in clays)
Alsamman (1995) Cohesionless soils: f p (kPa) = fctn[q c(side) (MPa)] Cohesionless soils: q b (MPa) = fctn[q c(tip) (MPa)]
(for drilled shafts in cohesionless Cohesive soils: f p (kPa) = fctn[q c-net(side) (MPa)] Cohesive soils: q b (MPa) = fctn[q c-net(tip) (MPa)]
and cohesive soils) (Figs 11 and 12) (Figs 11 and 12)
NGI-BRE Method f p = q t(net) /k 1 q b = q t(net) /k 2
(Almeida et al. 1996; Powell et al.
2001; Powell and Quarterman 1988) k 1 = 10.5 + 13.3log[q t(net) /σ vo '] k 2 = N kt /9; N kt = 15 (soft – firm intact clays), 25 to 35
(for driven and jacked piles in clays) (fissured to hard clays)
Politecnico di Torino Method f p (MPa) = [q c (MPa)/274]0.75 For w t /d = 0.05: q b,0.05 = 0.1q c (very dense sand); 0.14q c
(Fioravante 1994; Fioravante et al. based on average trend (medium dense sand)
1995) (for drilled shafts in sands) For w t /d = 0.10: q b,0.1 = 0.14q c (very dense sand); 0.17q c
(medium dense sand)
Lee and Salgado (1999) This method does not indicate a means for evaluating f p q b ≈ q t /[1.90 + 0.62/(w b /d)]
(for piles in sands) q b,0.1 ≈ 0.123q t (for w b /d = 0.1)
Unicone Method f p = C se ∙q E q b = C te ∙q Eg
(Eslami and Fellenius 1997; qE = qt – u2
Fellenius 2002) C te is generally taken as 1; for pile diameter d > 0.4 m, C te
(for all pile types in all soil types) (see Fig. 13 for C se ) = 1/(3d)
Notes: q ca = zone specific average q c ; σ atm = atmospheric pressure = 100 kPa; s uCIUC = undrained shear strength in isotropically-consolidated triaxial compression
mode; q c-net = q c – σ vo ; q t(net) = q t – σ vo ; σ vo = overburden stress; N kt = cone factor; σ vo ' = effective overburden stress = σ vo – u o ; w t and w b = displacements at
pile head and base; d = pile diameter; C se = side correlation coefficient from soil classification chart (from q t ; f s and u 2 ); C te = toe correlation coefficient; q Eg is
the geometric average of q E values over the influence zone after correction for u 2 and adjustment to σ vo '; influence zone for Gwizdala method from 4d (3d for
enlarged base) below to 1d above the pile toe, for Unicone method from 4d below pile tip to 8d above pile toe if pile is installed from weak soil into dense soil,
and from 4d below pile toe to 2d above pile toe when pile is installed from dense soil into weak soil, while for Alsamman method it is 1d below the pile toe.
Table 3 continued
Method/Reference Design Equations

Pile unit side resistance (f p ) Pile unit end bearing (q b )

KTRI Method Estimates f p from measured f s and ∆u 2 This method does not indicate a means for evaluating q b
(Takesue et al. 1998) (Fig. 14)
(for all pile and soil types)
UWA-99 Method This method utilizes effective stress approach (β- CE: q b0.1 = q c ∙λ∙m∙(w b /d)/(w b /d + c) at w b /d = 0.1
(De Nicola and Randolph 1999) method) as a means for evaluating f p . No direct use of where m = 0.7, c = 0.03; λ = 1.75 – 0.5σ vo '/σ atm (for σ vo ' <
(for driven pipe piles in medium to CPT reading is involved. 200kPa), λ = 0.75 (for σ vo ' > 200 kPa)
very dense homogeneous sand) (also see Fig. 15)

OE: see Fig. 15

TCD-01 Method This method does not indicate a means for evaluating f p CE: q b ≈ q c ; Q b = q b A b
(Lehane and Gavin 2001)
(for base capacity of OE and CE OE: q plug = fctn(q c , IFR) and q ann ≈ q c (Fig. 16)
jacked pipe piles installed in loose Q b = q plug A plug + q ann A ann
sands where IFR is measured during
pile installation)

TCD-03 Method f p = f L ∙σ rc '∙tanδ f ' q b0.1 = {λ q c [(1 – 0.9IFR)∙r i 2 + 2rt]}/r2


(Gavin and Lehane 2003; de Nicola f L = 0.8 for piles in tension and 1.0 for piles in
and Randolph 1999) compression; σ rc ' = 0.029q b (σ vo '/σ atm )0.12 ∙ (h/r*)−0.38; For σ vo ' < 200 kPa: λ = 1.23 – 0.35σ vo '/σ atm
(for OE jacked pipe piles installed in where q b = (q plug r i 2 + q ann 2rt)/r2
loose sands where the IFR is For σ vo ' > 200 kPa: λ = 0.53
measured during pile installation) For IFR < 1: q plug = q c (1 – IFR) + IFR∙q plug,min ;
IFR > 1: q plug = q plug,min
where q ann = q c ; q plug,min = 0.1q c

Notes: ∆u 2 = excess pore water pressure = u 2 – u o (measured in kPa); u o = hydrostatic pore water pressure; λ = reduction factor for effective vertical overburden
stress; σ vo ' = σ vo – u o ; w b = displacement at the pile base; d = pile diameter; σ atm = atmospheric pressure = 100 kPa; q ann = unit resistance below the pile annulus;
q plug = plug unit end bearing resistance; IFR = incremental filling ratio = ∆L p /∆z; ∆L p = incremental change of plug length; ∆z = change in penetration depth; Q b
= total base capacity; A plug and A ann are the pile plug and annulus areas, respectively; f L = reduction factor for loading direction; σ rc ' = radial effective stress
following full equalization of pore pressure; δ f ' = interface friction angle at failure; h = height above the pile tip; r* = modified radius term for OE piles = (r2 –
r i 2)0.5; r = external radius; r i = internal radius; t = pile wall thickness; q plug,min = minimum plug base resistance; q b0.1 = unit end bearing at a tip displacement of
10% of the pile diameter.
Table 3 continued
Method/Reference Design Equations

Pile unit side resistance (f p ) Pile unit end bearing (q b )

Fugro-05 Method Compression loading: q b0.1 = 8.5q c,avg ∙(σ atm /q c,avg )0.5A r 0.25 at w b /d of 0.1
(Kolk et al. 2005) h/r* > 4: f p = 0.08q c (σ vo '/σ atm )0.05 ∙(h/r*)−0.9 A r = 1 − (d i /d)2
(for driven piles in silica sands, h/r* < 4: f p = 0.08q c (σ vo '/σ atm )0.05 (4)−0.9(h/4r*)
mostly for offshore piling) Tension loading:
f p = 0.045q c (σ vo '/σ atm )0.15 ∙ [max(h/r*, 4)−0.85]
For noncircular piles, equivalent circular area is used to
assess r*

UCD-05 Method This method does not indicate a means for evaluating f p For CE pile: q b = q c ; Q b = q b A b
(Gavin and Lehane 2005)
(for base resistance of OE and CE For OE pile:
driven and jacked pipe piles in dense q ann = q c and q plug = (0.8 – 0.7 FFR)q c (for FFR < 1)
sands) Q b = q plug A plug + q ann A ann

ICP-05 Method f p = a(σ rc ' + ∆σ rd ')tanδ f ' CE pipe piles: q b0.1 = q c,avg ∙max[1−0.5log(d/d CPT ),0.3]
(Jardine et al. 2005) σ rc ' = 0.029b q c (σ vo '/σ atm )0.13∙[max(h/r*,8)]−0.38
(for driven piles loaded first time via ∆σ rd ' = 2G∆y/r*; ∆y ≈ 2R a ≈ 0.02 mm OE pipe piles: Unplugged: q b0.1 = q c,avg ∙A r ;
SML test conducted around 10 days a = 0.9 (OE piles in tension), 1.0 (all other cases); Plugged: q b0.1 = q c,avg ∙max[0.5 − 0.25∙log(d/d CPT ), 0.15,
after driving in sandy soils) b = 0.8 (tension), 1.0 (compression); δ f ' measured or Ar]
estimated as fctn(d 50 ) (Fig. 17) where A r = 1 – (d i /d)2
G = 185q c q c1N −0.7, or If d i > 2.0(D r − 0.3) or d i > 0.083q c,avg /σ atm d CPT , pile is
G = q c [0.0203 + 0.00125 q c1N – 1.216e–6q c1N 2]–1 unplugged, otherwise plugged (d i in meters);
q c1N = (q c /σ atm )/(σ vo '/σ atm )0.5 D r = 0.4 ln(q c1N /22)
Non-circular piles (use equivalent base area): square/
rectangular: q b0.1 = 0.7q c,avg ; H section: q b0.1 = q c,avg
Notes: h = height above the pile tip; r* = modified radius for OE piles = (r2 – r i 2)0.5; r and r i = external and internal radii; for square, rectangular and H-piles r*
= (A b /π)0.5; A b = base area = w∙b (square/rectangular section), A s + 2X p (D – 2T) (H section); w = width, b = breadth; A s = area of H section, X p = B/8 if B/2 < (D
– 2T) < B, X p = B2/[16(D – 2T)] if (D – 2T) > B, B = flange width, D = total depth, T = flange thickness; σ vo ' = effective overburden stress = σ vo – u o ; σ atm =
atmospheric pressure = 100 kPa; q c,avg = q c averaged + 1.5d over pile tip level; A r = area ratio; d and d i = outer and inner pile diameters; q ann = unit resistance
below the pile annulus; q plug = plug unit end bearing resistance; FFR = final filling ratio; A plug and A ann = pile plug and annulus areas; Q b = total base capacity;
σ rc ' = radial effective stress following pore pressure equalization; ∆σ rd ' = change in radial stress during pile loading; δ f ' = interface friction angle at failure; d 50 =
mean particle size; G = operational shear modulus; ∆y = radial displacement due to dilation during loading; R a = average pile roughness; q b0.1 = unit end bearing
at tip displacement of 10% pile diameter; q c,avg = q c averaged + 1.5d over pile tip level; d CPT = 0.036 m; D r = as decimal, nominal relative density.
Table 3 continued
Method/Reference Design Equations

Pile unit side resistance (f p ) Pile unit end bearing (q b )

ICP-05 Method This method utilizes effective stress approach (β- CE pipe piles:
(Jardine et al. 2005) method) as a means for evaluating f p . No direct use of q b = 0.8q c,avg (undrained loading)
(for driven piles in clays) CPT reading is involved q b = 1.3q c,avg (drained loading)

OE pipe piles:
Unplugged: q b = q c,avg (undrained loading);
q b = 1.6q c,avg (drained loading)
Plugged: q b = 0.4q c,avg (undrained loading);
q b = 0.65q c,avg (drained loading)

If (d i /d CPT + 0.45q c,avg /σ atm ) < 36, pile is plugged


otherwise unplugged during static loading

UWA-05 Method f p =(f t /f c ) [0.03q c A rs,eff 0.3[max(h/d,2)−0.5] + ∆σ rd ']tanδ cv ; q b0.1 = q c,avg ∙(0.15 + 0.45A rb,eff )
(Lehane et al. 2005) f t /f c = 1 (compression), 0.75 (tension)
(for driven piles in sandy soils) A rb,eff = 1 − FFR(d i /d)2;
A rs,eff = 1 − IFR(d i /d) ; where IFR = ∆L p /∆z
2

If IFR is not measured: FFR = IFR averaged over last 3d of the pile penetration,
Average IFR = min{1,[d i (m)/1.5]0.2} where IFR = ∆L p /∆z
∆σ rd ' = 2G∆y /r*; G ≈ 185q c q c1N −0.7, or
G = q c [0.0203 + 0.00125q c1N – 1.216e–6q c1N 2]–1 If IFR not measured: FFR = min{1,[d i (m)/1.5m]0.2}
∆y = 2R a ≈ 0.02 mm; q c1N = (q c /σ atm )/(σ vo '/σ atm )0.5

Notes: q c,avg = q c averaged + 1.5d over pile tip level; d i = inner pile diameter; for non-circular piles equivalent pile diameter is adopted in the manner described
for piles in sand; d CPT = 0.036 m; σ atm = atmospheric pressure = 100 kPa; f t /f c = ratio of tension to compression capacity; A rs,eff = effective shaft area ratio; δ cv =
constant volume friction angle (in absence of laboratory measurement, use Fig. 17 of ICP-05 method with an upper limit of tanδ cv = 0.55); h = height above pile
tip; IFR = incremental filling ratio; ∆L p = incremental change of plug length; ∆z = change in penetration depth; d = pile outer diameter; ∆σ rd ' = change in radial
stress during pile loading; G = operational shear modulus; ∆y = radial displacement due to dilation during loading; R a = average pile roughness; r* = modified
radius for OE piles = (r2 – r i 2)0.5; where r i = internal pile radius = d i /2; q b0.1 = unit end bearing at a tip displacement of 10% of the pile diameter; q c,avg = q c
averaged using Dutch averaging technique; A rb,eff = effective base area ratio; FFR = final filling ratio.
Table 3 continued
Method/Reference Design Equations

Pile unit side resistance (f p ) Pile unit end bearing (q b )

NGI-05 Method f p = (z/L)∙σ atm F Dr F sig F tip F load F mat > = 0.1∙σ vo ' CE: q b0.1 = q c,tip ∙[0.8/(1 + D r 2)]
(Clausen et al. 2005) F Dr = 2.1(D r – 0.1)1.7 OE: q b0.1 = min[q b0.1(plugged) , q b0.1(unplugged) ]
(for driven piles in sandy soils) F sig = (σ vo '/σ atm )0.25 Plugged: q b0.1 = q c,tip ∙[0.7/(1+3D r 2)]
F tip = 1.0 (driven OE), 1.6 (driven CE) Unplugged: q b0.1 = q c,tip ∙A r + 12∙f p,avg ∙L∙(1 − A r )/(πd i )
F load = 1.0 (tension), 1.3 (compression) D r = 0.4ln(q c1N /22)
F mat = 1.0 for steel and 1.2 for concrete q c1N = (q c /σ atm )/(σ vo '/σ atm )0.5
D r = 0.4ln(q c1N /22) A r = 1 − (d i /d)2
q c1N = (q c /σ atm )/(σ vo '/σ atm )0.5
Cambridge-05 Method This method does not indicate a means for evaluating f p q b ≈ 0.9q c,avg
(White and Bolton 2005)
(based on CE pipe piles, Franki piles q c,avg = average q c in zone + 1.5d; for partial embedment in
with enlarged base, and PCC piles, transition zone (–2<z b /d<8):
jacked or driven through upper soft q c,avg = {q c,weak + [(q c,hard – q c,weak )(z b /d + 2)]/10}
into lower hard sand layer) (see Fig. 18 for further description of partial embedment
reduction factor)
Togliani (2008) Q shaft = Q cylinder + Q taper q b = k 3 q c(toe)
(for cylindrical and tapered driven Q cylinder = π∙d mean ∙L∙f cylinder k 3 = χ + [0.01(L pile /d toe )]
piles and drilled shafts in all soil Q taper = π/4∙[d top 2 – d bottom 2]∙f taper χ = 0.2 (driven piles), 0.1 (drilled shafts)
types) f cylinder = k 1 q c 0.5 & f taper = q c (k 2 d mean /d base )
k 1 = β [1.2(0.8 + R f /8)] for R f < 1
k 1 = β [1.1(0.4 + ln(R f )] for R f > 2
k 1 = β [1.2(0.8+R f /8) + 1.1(0.4+Ln(R f )]/2 for 1 < R f <
2
β = 1 (driven displacement piles), 0.6 (non-
displacement driven and CFA), 0.5 (bored)
k 2 = 1.2 (for q c < 3 MPa); 1.0 (for q c > 3 MPa)
German Method Provides upper and lower bound estimates of f p (kPa) Provides upper and lower bound estimates of q b (MPa)
(Kempfert and Becker 2010) based on q c (measured in MPa) based on q c (measured in MPa)
(for piles in sandy soils) (Figs 19a and 20) (Figs 19b)
Notes: z/L = layer depth to pile length ratio; D r = relative density in decimal, may be greater than 1.0; q b0.1 = unit end bearing at a tip displacement of 10% of the
pile diameter; q c,tip = average q c in the zone at the pile base; A r = area ratio at pile base; d i = pile inner diameter; d = pile outer diameter; f p,avg = average external
friction over pile embedment depth L; q c,avg = average q c in zone around the pile base described in Fig. 18 ; d mean = R f = friction ratio = (f s /q c )100; q c(toe) =
average q c between +8d toe and –4d toe ; L pile /d toe = pile slenderness ratio.
Table 3 continued
Method/Reference Design Equations

Pile unit side resistance (f p ) Pile unit end bearing (q b )

UCD-11 Method f p = σ rf '∙tanδ f ', where σ rf ' = σ rc ' + ∆σ rd ' This method does not indicate a means for evaluating q b
(Igoe et al. 2010; 2011) σ rc ' = q c [0.025 – 0.0025(h/d)]A r,eff > σ rc ' ,min
(for shaft capacity of OE piles in A rs,eff = 1 – IFR(d i /d)2; where IFR = ∆L p /∆z
sandy soils) σ rc ' ,min = η q c ; η = 0.003(loose sand), 0.006(dense
sand); ∆σ rd ' = 2G∆y/r*; G = 185q c q c1N −0.7, or
G = q c [0.0203+0.00125 q c1N –1.216e–6 q c1N 2]–1
∆y = 2R a = 0.02 mm; q c1N = (q c /σ atm )/(σ vo '/σ atm )0.5

V-K Method f p = k s(z) ∙q t(net) , z q b = 0.7q t(net),avg


(Van Dijk and Kolk 2011) k s(z) = 0.16(h/uL)–0.3[Q t(z) ]–0.4 < 0.08
(for offshore piles in clays) uL = 1.0 m (= 3.3 feet) q t(net) = q t – σ vo
Q t(z) = q t(net),z /σ vo ' at z; where q t(net) = q t – σ vo

SEU Method (Cai et al. 2011; 2012) Estimates f p from measured f s and ∆u 2 Similar to Unicone Method: q b = C te q E
[for PCC thin-wall and high-strength (Fig. 21) qE = qt – u2
caissons, cement fly- ash grave pile
in soft clays (driven or jacked)]

HKU Method This method does not indicate a means for evaluating f p Q b = (π/4)[d2 q plug + (d2 – d i 2) q ann ]
(Yu and Yang 2012)
(for base capacity of OE steel pipe q plug = 1.063exp(–1.933PLR) q ca ; where PLR = H/L
piles in sandy soils) q ann = [1.063 – 0.045(L/d)] q ca > 0.46q ca
q ca = 0.5(M A + M B ) if M A < M B ; otherwise = M B
(also see Fig. 22 for further description)

Notes: σ rf ' = radial effective stress at failure; σ rc ' = radial effective stress after installation, before loading; ∆σ rd ' = change in radial stress during loading; h =
distance from pile toe; d = outer diameter; A rs,eff = effective shaft area ratio; IFR = incremental filling ratio; ΔL p = change in soil plug length during an increase in
pile penetration Δz; d i = inner diameter; η = scale reduction factor for friction fatigue; r* = modified radius for OE piles = (r2 – r i 2)0.5; r i = internal pile radius; G
= operational shear modulus; ∆y = radial displacement due to dilation during loading; R a = average pile roughness; z = depth below surface; uL = unit length to
render expression dimensionless; q t(net),avg = q t(net) averaged + 1.5d near pile tip; ∆u 2 = excess porewater pressure = u 2 – u o (measured in kPa); C te = factor for pile
unit tip resistance similar to the bearing capacity factor, N c , and empirical cone factor N kt ; Q b = total base capacity; PLR = plug length ratio; H = plug length
measured at the end of pile installation; L = pile length; q ca = average q c in the influence zone at the pile base; M A and M B are the averages of q c trace within the
ranges of A or B determined by the geometric means: (q c1 q c2 …. q ci … q cn )1/n; q ci = ith q c reading recorded over the range of A or B (see Fig. 22 for further
description).
CPT Based Evaluations of
Pile Capacity

Direct Rational (Indirect)


Methods Methods

Total Stress Approach Effective Stress Approach

Unit Shaft Resistance (fp): Unit Shaft Resistance (fp):


α-Methods β-Methods
(applicable to fine grained soils) (applicable to coarse and
fctn(su, σvo', OCR, Ip, L, fine grained soils)
plugging, progressive failure) fctn(σr, δ, φ', OCR, K, σvo', L,
d, su, Dr, St, Ip)

Unit Base Resistance (qb): Unit Base Resistance (qb):


undrained loading drained loading
(applicable to fine grained soils) (applicable to coarse grained soils
fctn(su) and
slow loading in fine grained soils)
fctn(φ', σvo', L, d, Dr)

Pure Empirical Methods: Semi-Empirical Methods:


Enable evaluations of fp and qb directly Enable evaluations of fp and qb
using qc (or qt), and/or fs and/or u2 using qc (or qt), and/or fs
along with additional parameters (σr,
δ, φ', K, σvo', L, d, su, Dr, plugging)

Fig. 1 Alternative paths for CPT-based evaluations of f p and q b components of pile capacity
Weak
Weak Soil
Soil ql1
L
ql1 L Dense
Sand
zd
zd ~10d H < 20d
qb
qb ~10d
H’
H > 20d
Weak
ql2 Soil
ql2 Dense
(a) Sand (b)

Fig. 2 Pile unit base resistance and depth in sand stratum beneath weak soil layer: (a) thick sand stratum; (b) thin
sand stratum overlying weak soil (after Meyerhof 1976)
Pile Unit Base Resistance, qb (MPa) 30

25

20

15

10

0
0 5 10 15 20 25 30
Cone Tip Resistance, qc (MPa)
Fig. 3 Relationship between pile base resistance and CPT q c in sand (after Meyerhof 1983)
Penetration Resistance

Weak Soil
3.57 cm penetrometer qc curve
Depth

Strong Soil
Lc

Large diameter qb curve

Fig. 4 Scale effect diagram (adapted from De Beer 1963)


qc
0 20 40 60 80 100
0
qb = (qc1 + qc2)/2

200
e qc1 = Average qc over a distance of xd
below the pile tip (path a-b-c). Sum qc
d values in both the downward (path a-b)
and upward (path b-c) directions. Use
400 L actual qc values along path a-b and
minimum path rule along path b-c.

8d
Compute qc1 for x-values from 0.7 to 3.75
Depth

and use the minimum qc1 value obtained.

600

c
a
xd

qc2 = Average qc over a distance of 8d


above the pile tip (path c-e). Use the
800 b b minimum path rule as for path b-c in
the qc1 computations.

1000
Fig. 5 Begemann procedure for predicting pile q b (adapted from Begemann 1963)
20

18
Cone Tip Resistance, qc (MPa)

1 2 34
16

14

12

10 B C

8
A
6

2
O
0
0 20 40 60 80 100 120 140 160 180 200 100 80 60 40 20
Sleeve Friction, fs (kPa) Reduction Coefficient (%)

Example Calculations Legend


Plot point A for measured average qc = 6.4 MPa, and fs = 130 kPa 1: Wooden pile
Connect points O and A; extend to point B 2: Pointed base prefab concrete pile
Extend horizontally to point C on curve 2 for prefab pile with 45o point with 45o point
Extend vertically down from point C 3: OE steel pipe pile and I-beam pile
Measure reduction coefficient ≈ 55% 4: Pointless base prefab concrete pile
Estimate fp = 130 kPa x 0.55 = 71.50 kPa

Fig. 6 Begemann's graph for predicting pile f p (adapted from Begemann 1965)
Penetrometer-pile friction ratio, αc 1.4
Steel Piles
1.2
Concrete and Timber Piles
1.0

0.8

0.6

0.4

0.2

0.0
0 50 100 150 200
Penetrometer Sleeve Friction, fs
Fig. 7 Design curves for pile f p in clays (after Nottingham 1975; Schmertmann 1978)
Penetrometer-pile friction ratio, αs 3.0
Steel Pipe Pile
2.5 Timber Pile
Square Concrete Pile

2.0

1.5

1.0

0.5

0.0
0 10 20 30 40
Pile depth to diameter ratio, z/d
Fig. 8 Design curves for pile f p in sand (after Nottingham 1975; Schmertmann 1978)
300
Note: Lower limit applies for unreliable
construction. Upper Limit for very careful
Side Resistance, fp (kPa)

250 construction
l
PILE TYPE = IIIB
200

IIIA IA, IIA


150 upper
lower

100 IIB upper


lower
IB
50

0
0 10 20 30 40

Cone Tip Resistance, qc (MPa)

(a)
180
Note: Lower limit applies for unreliable
160 construction. Upper Limit for very careful
Side Resistance, fp (kPa)

construction
140 l
PILE TYPE = IIIB
120

100
IIIA
80
upper upper
60 IA, IIA IB
lower lower
40 IIB

20

0
0 5 10 15 20

Cone Tip Resistance, qc (MPa)


(b)
Fig. 9 LCPC method for pile f p evaluation from CPT q c in: (a) sands; (b) clays (based on Bustamante and Gianeselli
1982; adapted from Poulos 1989)
Pile qc

q’ca
Step 1: Smoothen the raw qc curve
0.7q’ca 1.3q’ca to climate the local irregularities
Depth
Step 2: From the smoothened curve
of qc, calculate q’ca as mean qc
d between 1.5d above and below the
1.5d pile base

Step 3: From the smoothened curve,


clip peak values > 1.3 times q’ca
beneath the pile base and < 0.7
qeq (tip) times q’ca above the base
1.5d
Step 4: Take average of the qc
values of the new clipped
smoothened curve so obtained to
Smoothened qc curve find qeq (tip)

Fig. 10 Calculation procedure for equivalent q c for LCPC method (adapted from Bustamante and Gianeselli 1982)
160
Cohesionless soils fpmax ≈ 132 kPa

120
fpmax ≈ 96 kPa
fp (kPa)

80
Sand/silty sand - mechanical cone
Sand/silty sand - electrical cone
40
Gravelly sand/gravel - electrical cone
Gravelly sand/gravel - electrical cone
0
0 10 20 30 40
qc(side) (MPa)
4
Cohesionless soils

(qb, crit)max ≈ 2.9 MPa


3
qb (MPa)

2
(w/d)crit = 0.1

1
Mechanical cone
Electrical cone
0
0 10 20 30 40
qc(tip) (MPa)
Fig. 11 Design curves for pile f p and q b in cohesionless soils (after Alsamman 1995)
100
Cohesive soils fpmax ≈ 86 kPa

75
fp (kPa)

50

25
Mechanical cone
Electrical cone
0
0 2 4 6 8 10
qc-net(side) (MPa)
3
Cohesive soils

2
qb (MPa)

1
Mechanical cone
Electrical cone
0
0 2 4 6 8 10
qc-net(tip) (MPa)
Fig. 12 Design curves for pile f p and q b in cohesive soils (after Alsamman 1995)
100
Zone No. Soil Type Cse
1 Soft sensitive clay 0.08 5
2 Soft clay and silt 0.05
3 Stiff clay and silt 0.025
4
qE = qt - u2 (MPa)

4 Silty sandy mix 0.01


10 5 Sand 0.004

1 2

0.1
1 10 100 1000
Sleeve friction, fs (kPa)
Fig. 13 Unicone chart for zone numbers and soil types (after Eslami and Fellenius 1997)
6
Clay fp/fs ≈ ∆u2/200 - 0.5
Pile-CPT Friction Ratio, fp/fs

5 Mix Bored cast in-situ piles


Sand
4
Clay
Mix Driven steel piles
3
Sand

2 fp/fs ≈ ∆u2/1250 + 0.76

0
-300 0 300 600 900 1200
Excess Pore Pressure, ∆u2 (kPa)
Fig. 14 KTRI chart for estimating pile f p from CPTu data (after Takesue et al. 1998)
0.9
Closed ended
0.8 Open: d/t = 10
Normalized end-bearing, qb0.1/qc

Open: d/t = 20
0.7
Open: d/t = 40

0.6 Open: d/t = 60

0.5

0.4

0.3

0.2

d = pile external diameter


0.1
t = wall thickness
0
0 50 100 150 200 250 300
Effective overburden stress, σvo' (kPa)
Fig. 15 Variation of normalized q b with effective overburden stress for w b /d = 0.1 (adapted from de Nicola and
Randolph 1999)
1.4
qplug,f mobilized at plunging failure in load
1.2 tests and during installation jacking stages

0.8
qplug/qc

0.6

qplug,f/qc = 0.9756e-1.64(IFR)
0.4 R² = 0.9941

0.2
During Jacking
Load Tests
0
0 0.5 1 1.5
IFR (for given jacking stage/prior to load test)
(a)
1.4

1.2

0.8
qann,f/qc

0.6
qann,f/qc ≈ 1
0.4

0.2

0
0 0.5 1 1.5
IFR
(b)
Fig. 16 Dependence of: (a) q plug,f /q c and (b) q ann,f /q c on IFR for OE jacked pipe piles in sand (after Lehane and
Gavin 2001)
Interface friction angle at failure, δf' 38
Jardine et al. (1992)
36 Shell UK Ltd database
34 CUR (2001) recommendation

32

30

28

26

24

22

20
0.01 0.1 1 10
Mean Particle Size, d50 (mm)
Fig. 17 lnterface friction angle in sand: trends from direct shear interface tests (after Jardine et al. 2005)
Penetration Resistance, qb, qc

qb, qc, weak

Soft layer

Hard layer
qc, hard
Depth, z

qb qb transition zone
–2 < zb/d < 8
zb

qb, qc, hard

Fig. 18 Partial embedment reduction factor on q b (after White and Bolton 2005)
300
1 Driven precast piles
2 Simplex piles
Pile unit shaft resistance, fp (kPa)

250 3 Atlas piles


4 Fundex piles
5 Bored piles 3
200

150 5 2
1
4
100

50
Upper values (50%-quantile)
Lower values (10%-quantile)
0
0 5 10 15 20 25 30
Cone tip resistance, qc (MPa)
(a)
14
1 Driven precast piles
2 Simplex piles
Pile unit base resistance, qb (MPa)

12 3 Atlas piles
4 Fundex piles
1 2
5 Bored piles
10
4

8
3
6

5
4

2
Upper values (50%-quantile)
Lower values (10%-quantile)
0
0 5 10 15 20 25 30
Cone tip resistance, qc (MPa)
(b)
Fig. 19 Upper and lower empirical values of different piles in coarse grained soils for: (a) f p ; (b) q b (after Kempfert
and Becker 2010)
350
1 Driven precast piles
2 Simplex piles
10
Pile unit shaft resistance, fp (kPa)

300 3 Atlas piles


4 Fundex piles
250
5 Bored piles 9
6 Auger piles
7

200 3
8

150 6
2 5
1
4
100
7 VM piles
50 8 RI piles
9 Micro piles
10 RV piles
0
0 5 10 15 20 25 30
Cone tip resistance, qc (MPa)
Fig. 20 Lower empirical values of f p for different piles in coarse grained soils (after Kempfert and Becker 2010)
10
Pile Types
Cement fly-ash grave pile

8 Prestressed concrete thin-wall caisson


Pile-CPT Friction Ratio, fp/fs

Prestressed concrete high-strength caisson

4
fp/fs ≈ ∆u2/250 + 1.30
fp/fs ≈ ∆u2/125 + 0.8

Marine clay
2
Clay
Silty clay
Silty sand
0
-200 0 200 400 600 800 1000
Excess Pore Pressure, ∆u2 (kPa)
Fig. 21 SEU chart for estimating pile f p from CPTu data (after Cai et al. 2011; 2012)
Influence zone

Fig. 22 Influence zone for averaging cone tip resistance near pile base (after Yu and Yang 2012)
100
Estimates for base resistance, qb
90 93

80
Percent reliance (%)

70 Estimates for shaft resistance, fp


64
60
52
50
40
40

30
24
21
20 16 17

10 7 7

0
CPT tip resistance only Corrected CPT qt instead of measured CPT qc
Multiple CPT readings Additional parameters: measured
Additional parameters: estimated via correlations
Fig. 23 Reliance of different CPT-based direct methods on combinations of CPT readings and additional parameters

View publication stats

You might also like