Anda di halaman 1dari 20

Geomechanics and Geoengineering

ISSN: 1748-6025 (Print) 1748-6033 (Online) Journal homepage: http://www.tandfonline.com/loi/tgeo20

The Second James K. Mitchell Lecture Undisturbed


sand strength from seismic cone tests

P. W. Mayne

To cite this article: P. W. Mayne (2006) The Second James K. Mitchell Lecture Undisturbed
sand strength from seismic cone tests, Geomechanics and Geoengineering, 1:4, 239-257, DOI:
10.1080/17486020601035657

To link to this article: http://dx.doi.org/10.1080/17486020601035657

Published online: 29 Jan 2007.

Submit your article to this journal

Article views: 1536

View related articles

Citing articles: 11 View citing articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=tgeo20

Download by: [65.209.12.68] Date: 22 March 2016, At: 12:29


Geomechanics and Geoengineering: An International Journal
Vol. 1, No. 4, December 2006, 239--257

The Second James K. Mitchell Lecture


Undisturbed sand strength from seismic cone tests
P. W. MAYNE*

Geosystems Engineering, Georgia Institute of Technology, School of Civil and Environmental Engineering, 790 Atlantic Drive,
Mason Building, Atlanta, GA 30032-0355, USA

(Received 30 July 2006; in final form 26 September 2006)

During routine site investigations, high-quality sampling and laboratory testing of sands are not feasible because of inevitable sample disturbance
effects and budgetary constraints. Herein, a select database is compiled for calibration of cone penetration test (CPT) interpretative methods, primarily
from undisturbed frozen sand samples at 15 sites in Japan, Canada, Italy, Norway, and China. The database is used to evaluate the peak secant friction
angle P ¢ in terms of the normalized cone tip resistance ðQ  qt =vo ¢Þ using two analytical approaches: (i) spherical cavity expansion theory with an
operational rigidity index IRR, and (ii) a limit plasticity formulation with appropriate angle of plastification . Backfigured values of IRR and  are
found correlated to the normalized small-strain stiffness G0 =vo ¢. Thus, with measurements taken of both the initial shear modulus and shear strength,
Downloaded by [65.209.12.68] at 12:29 22 March 2016

seismic cone tests offer the opportunity to derive the entire stress--strain--strength response of sands at all depths.

Keywords: Sand; Seismic cone tests; Stress--strain--strength response; Cone penetration test; Calibration; Frozen sand

1. Introduction Reconstituted sand samples have been used in laboratory test


programmes to investigate the stress--strain--strength behaviour of
For several decades, geotechnical site characterization of clean clean sands at different relative densities DR, effective confining
sands has been hampered by the difficulties associated with stresses, drainage conditions (dry, saturated, undrained, drained,
providing the true reference benchmark values of strength and partially saturated), and loading paths (e.g. triaxial compression,
stiffness needed to represent their in situ stress--strain response. extension, plane strain, simple shear, direct shear), as well as other
Undisturbed samples of sand have been difficult or impossible facets (e.g. Lade and Bopp 2005). In many cases, the laboratory test
to acquire in the field because of severe disturbance effects results on reconstituted sands can be interpreted to provide the
caused by insertion of thin-walled tubes, tube bending or buck- stiffness in terms of an equivalent elastic modulus E¢ and strength
ling, alterations caused during removal of the tube from the in terms of a secant friction angle ¢ by adoption of a simple Mohr--
borehole, changes in sand density and/or fabric during trans- Coulomb strength envelope. Additional engineering parameters
portation from field to the laboratory, external vibrations, sam- of interest might include the constrained modulus D¢ = 1/mv,
ple extrusion for trimming or specimen mounting, saturation the dilatancy angle c¢, Poisson’s ratio n¢, the effective cohesion
procedures, and other influences. intercept c¢, the small-strain shear modulus G0 ¼ Gmax ¼ T Vs2 ,
Much of today’s current practice has centred on the use of the threshold strain gth, the damping ratio D, the modulus reduction
reconstituted sand specimens, often prepared by the methods of curves G/G0, the coefficient of secondary compression Ca", the
compaction, air pluviation, moist tamping, slurrying, water sedi- permeability k, the lateral stress coefficient K0, the overconsolida-
mentation, and/or vibration to achieve a desired density. Of tion ratio (OCR), and/or constitutive soil modelling constants
curiosity is the need to check these assumptions and verify that related to strength (M), compressibility (l, ), state (cs), hardening
realistic results are being obtained with respect to true undis- (h), and/or other terms (Been et al. 1987).
turbed sand samples and their behaviour. It is only recently (i.e. Results from laboratory testing on reconstituted sands have
since 2000) that it has been possible to collect a sufficient provided valuable insight into the important factors that control
database on high-quality undisturbed sand samples to allow the effective friction angle, stress--strain response, volume
this level of cross-linking of laboratory and in situ testing on change behaviour under drained loading, and porewater
clean sands. The notions are depicted in figure 1 which shows the response for undrained loading. For instance, the increase of
idealized approach (Class A) and current practice (Class B). strength and stiffness with increasing relative density is well
recognized, as shown by drained triaxial data on quartz sand in
figure 2. Sand strength can be attributed to a basic friction angle
*Corresponding author. Email: paul.mayne@ce.gatech.edu (termed the critical state friction angle cs ¢ ) which is dependent

Geomechanics and Geoengineering: An International Journal


ISSN 1748-6025 print=ISSN 1748-6033 online  2006 Taylor & Francis
http:==www.tandf.co.uk=journals
DOI: 10.1080=17486020601035657
240 P. W. Mayne

Interpreted Soil Engineering Parameters: Preparation Methods


Air
φ′, ψ′, c ′, E′, ν ′, G0, Ψs, λ, κ, Γ, K0, OCR Pluviation

Moist
CLASS A METHOD Triaxial Testing METHOD B Tamped
Undisturbed Young
Frozen Artificial Water
Specimens Sedimentation
Specimens
Slurry
In-Situ e0
In-place DR Estimated
Vibration
Fabric NC
OC e0
Coolant Compaction
Anisotropy DR
System γd
Structure
Age wn
SPT SCPTu

1-d
Ground Disturbed Soil
Freezing Vs Augering
Downloaded by [65.209.12.68] at 12:29 22 March 2016

fs Rotary Drill
Liquid ub Tube Sample
Estimated
Nitrogen Drive Sample
qt Relative
(-20º) N60 Density, DR
Coring

Figure 1. Calibration of in situ penetration tests to laboratory triaxial results from undisturbed frozen samples versus reconstituted sand specimens.

CIDC Triaxial Tests


6
(σ 1 ' /σ 3 ' ) − 1
sinφ ' = φP' = peak friction angle
(σ 1 ' /σ 3 ' ) + 1
5 41.5º
41.0º
Stress Ratio, σ1'/σ3'

38.8º

4 36.7º
35.2º
33.7º φcv'
3
DR = 82.6 %
DR = 74.5 %

2 DR = 59.3 %
DR = 38.5 %
DR = 22.3 %

1
0 5 10 15 20 25 30 35

Axial Strain, εa (%)


Figure 2. Peak and critical state friction angles from drained triaxial tests on quartz sands at different relative densities (data from Koerner 1970).

on the mineralogy, particle shape, size, roundness, and other arrangement, as well as mineralogy and fitting parameters (e.g.
index parameters (Koerner 1970, Cho et al. 2006). The mea- Bolton 1986).
sured peak sand strength p ¢ can be considered as the sum of From a practical standpoint, the use of in situ testing has been
this minimum value cs ¢ and an additional component attribu- widely adopted in geotechnical practice to ascertain the
ted to volume increase tendencies or dilatancy effects (Rowe mechanical properties of sandy soils. For routine use, in situ
1969). These factors can be conveniently expressed in terms of tests for these purposes include the following (e.g. Mitchell and
empirical expressions that account for mean effective confining Gardner 1975, Mitchell 1988, Yu 2004): standard penetration
stress level 0 ¢ and the current relative density DR or packing test (SPT), cone penetration test (CPT), pressuremeter test
The Second James K. Mitchell Lecture 241

(PMT), and flat dilatometer test (DMT). So that the in situ test penetration resistances and the relative mapping of ground
readings can be related to the sand stiffness and strength para- variability across a given site. The frozen samples can also be
meters, either the field test measurements are correlated to used to quantify the relative differences between undisturbed
standard laboratory tests on disturbed bulk sand samples taken and reconstituted specimens in the laboratory.
from the field which are reconstituted to the estimated in-place In this paper, an elite database is compiled on undisturbed sand
densities, or, alternatively, large-scale calibration chamber tests samples from 13 different sites that have been carefully tested
with diameters and heights varying from 0.7 to 1.5m have been under triaxial compression conditions in the laboratory. These
prepared to allow insertion of the in situ probes. sands are relatively clean to slightly silty sands (SP to SP--SM)
Of course, it is assumed that the in situ test results can be with 0--15% fines content (mean PF, 4.6%) from five countries.
reliably interpreted to give the actual in-place unit weight (gd), At each site, companion series of SPT and CPTU (piezocone)
relative density (DR), or void ratio (e0) that establish the density were available, and profiles of shear wave velocities (Vs) were
state for the reconstituted samples. In fact, it has been known also obtained, except at one location. Thus the compilation
for quite some time that sand fabric and placement method are affords an opportunity to cross-check and investigate in situ test
as important, or perhaps more significant, than merely the same relationships for sand strength and stiffness which include the
density in establishing the response of sands to loading (e.g. inherent fabric, density, age, and anisotropy of these materials.
Mulilis et al. 1977, Vaid and Sivathayalan 2000). For example,
Figure 3 shows representative results from a set of consolidated
triaxial tests on a silty sand (tailings) having 28% fines content 2. Undisturbed sand database
(FC) reported by Hoeg, et al. (2000). Here, tests showed dra-
Downloaded by [65.209.12.68] at 12:29 22 March 2016

matically different stress--strain responses depending on The database was formulated from a total of 15 effective stress
whether specimens were undisturbed versus artificially pre- friction angles from triaxial compression testing of high-quality
pared at same void ratio (e0 = 0.71) using either air pluviation undisturbed samples of sands. The sands are from four sites in
or by slurry placement. These results were from undrained Japan, six sites in Canada, and one site each in Norway, China,
shear tests and the porewater pressure responses also showed and Italy. Table 1 lists the specific site locations and their
remarkably different behaviour. affiliations. The samples from Japan, Canada, and Italy were
Recent developments in special sampling techniques have obtained using the expensive freezing procedures, while the
now permitted the use of one-dimensional controlled freezing samples from Norway were taken with tubes (frozen after-
of the ground in order to obtain true undisturbed samples of wards, in some cases) and a special Bezier tube sample used
sands (e.g. Hoffman et al. 1996, Mimura 2003). Once recov- for the site in Western Kowloon, China.
ered, the samples are transported to the laboratory and mounted At all sites, field testing included soil test borings with SPT N
in triaxial devices, carefully thawed, consolidated under ambi- values that were adjusted to 60% energy efficiency (N60). At each
ent stress states, and then tested under their true void ratio, site, seismic piezocone (SCPTU) soundings provided continuous
particle packing, and fabric. These test results can be used to readings of cone tip resistance qt, sleeve friction fs, and porewater
provide true reference values of E¢ and ¢ for benchmarking pressures ub measured at the shoulder, with downhole

600
Shear Stress, τ = 0.5(σ1 - σ3)

500

400 e0 (consolidated) = 0.71


σvc' = 500 kPa; σhc' = 250 kPa

300

Undisturbed
200
Pluviated

100 Slurried

0
0 5 10 15 20 25

Axial Strain, εa (%)


Figure 3. Comparisons of undisturbed versus reconstituted specimens from triaxial results on silty sands (tailings) (Hoeg et al. 2000).
242 P. W. Mayne

Table 1. Elite database sites and sources on sands with undisturbed sampling

Sand Location Reference Sand type Sampling method

W. Kowloon China Lee et al. 1999 Hydraulic fill Mazier tube


Yodo Japan Mimura 2003 Natural alluvial Frozen
Yodo Japan Mimura 2003 Natural alluvial Frozen
Yodo Japan Mimura 2003 Natural alluvial Frozen
Natori Japan Matsuo and Tsutsumi 1998 Natural alluvial Frozen
Tone Japan Matsuo and Tsutsumi 1998 Natural alluvial Frozen
Edo Japan Yamashita et al. 2003 Natural alluvial Frozen
Mildred Lake Canada Robertson et al 2000a,b Hydraulic fill Frozen
Massey Canada Robertson et al 2000a,b Natural alluvial Frozen
Kidd Canada Robertson et al 2000a,b Natural alluvial Frozen
J-pit Canada Robertson et al 2000a,b Hydraulic fill Frozen
LL dam Canada Robertson et al 2000a,b Tailings Frozen
Highmont Canada Robertson et al 2000a,b Tailings Frozen
Holmen Norway Lunne et al. 2003 Natural alluvial Tube/frozen
Gioia Tauro Italy Ghionna and Porcino 2006 Natural coarse sand Frozen

shear wave velocities Vs obtained at 1m intervals, except at the additional supplementary information taken from Matsuo and
Kowloon site where only CPTU results were available. It should be Tsutsumi (1998) and Yamashita et al. (2003). The in-place void
Downloaded by [65.209.12.68] at 12:29 22 March 2016

noted that geotechnical site characterization by SCPTu soundings ratios e0 and related void indices emax and emin are given in table 3,
is the most efficient and expedient means to collect multichannelled together with the corresponding relative densities for the sampled
data on many facets of soil behaviour (Mayne and Campanella zones. Isotropically consolidated drained triaxial compression
2005). Table 2 provides a summary of index parameters of the tests (CIDC) and cyclic undrained triaxial tests (CTX) with pore-
sands, including the focal depth of interest, the specific gravity water pressure measurements were performed on undisturbed
of solids Gs, the fines content (FC , 0.075mm), the unifo- samples of these sands. Each friction angle reported represents
rmity coefficient (UC), and the mean grain size D50. the mean secant value from three specimens obtained by the
As indicated in the tables, the undisturbed samples of sands CIDC series, except the Yodo sands individually listed at depths
included natural deposits (generally alluvial), hydraulically of 8.1 and 12.6m where only one CIDC specimen was provided.
placed fills, and tailings from mining operations. Full details of Since drained triaxial tests were performed, the dilation angles
these materials are given in the cited reference sources, as dis- and other parameters (e.g. E50) on these sands are also available
cussed subsequently. In all cases, the samples were situated below for consideration. Series on specimens reconstituted by air pluvia-
the ambient groundwater levels. Therefore all are considered fully tion were also conducted, but are not considered herein.
saturated, excepting the Mildred Lake and LL Dam sites which The single-valued profiles of SPT-N and CPT readings pre-
apparently were less than 100% saturated (Wride et al. 2000). sented by Mimura (2003) are tabulated here. For three sites
(Natori, Edo, and Tone), several sets of three or four N profiles
2.1 Japan and seven to nine CPT profiles were summarized by Matsuo and
Tsutsumi (1998). As the Japanese SPT practice with the auto-
The data from the four natural Japanese sands (Yodo, Natori, matic hammer or tonbi (trip monkey) provides an energy effi-
Edo, and Tone Rivers) were reported by Mimura (2003), with ciency of about 78%, the N values were accordingly corrected to

Table 2. Index parameters and depth ranges of undisturbed sands for database

Sand Depth (m) GWL (m) Mean z (m) Solids Gs Fines (%) UC = D60/D10 D50 (mm)

W. Kowloon 11.8 9 11.8 2.63 1 6.4 0.72


Yodo 8.1 2.1 8.1 2.64 1.9 3.1 0.32
Yodo 10.9 2.1 10.9 2.63 0.27 3.3 0.82
Yodo 12.7 2.1 12.7 2.63 2.1 3 0.62
Natori 8.3 2.1 8.25 2.65 0.23 2 0.22
Tone 7.3 1.4 7.3 2.68 3.78 2 0.18
Edo 3.8 2.1 3.85 2.68 0.42 2.2 0.29
Mildred Lake 27--37 21 32 2.66 10 2.22 0.16
Massey 8--13 1.5 10.5 2.68 ,5 1.57 0.2
Kidd 12--17 1.5 14.5 2.72 ,5 1.78 0.2
J-pit 3--7 0.5 5 2.62 15 2.5 0.17
LL dam 6--10 2.1 8 2.66 8 2.78 0.2
Highmont 8--12 4 10 2.66 10 4 0.25
Holmen 5--15 1 10 2.71 2 3 0.55
Gioia Tauro 1.3--4.2 1.5 3 2.69 0.66 2.1 2
The Second James K. Mitchell Lecture 243

Table 3. In-place void ratio, void limits, and in situ test types for undisturbed sands

Void ratio parameters

Sand Location e0 emax emin DR (%) vo¢ (bar) In situ test methods

Kowloon China 0.492 0.634 0.328 46.5 1.80 SPT CPTu


Yodo Japan 0.820 1.054 0.665 60.2 1.02 SPT SCPTu
Yodo Japan 0.720 0.883 0.569 51.9 1.23 SPT SCPTu
Yodo Japan 0.790 0.921 0.567 37.0 1.43 SPT SCPTu
Natori Japan 0.857 1.167 0.765 77.2 0.87 SPT SCPTu
Tone Japan 0.947 1.330 0.775 69.1 0.84 SPT SCPTu
Edo Japan 1.043 1.227 0.812 44.3 0.51 SPT SCPTu
Mildred Lake Canada 0.768 0.958 0.522 43.6 5.16 SPT SCPTu
Massey Canada 0.970 1.100 0.700 32.5 1.20 SPT SCPTu
Kidd Canada 0.981 1.100 0.700 29.8 1.60 SPT SCPTu
J-pit Canada 0.762 0.986 0.461 42.7 0.55 SPT SCPTu
LL dam Canada 0.849 1.055 0.544 40.3 1.00 SPT SCPTu
Highmont Canada 0.825 1.015 0.507 37.4 1.38 SPT SCPTu
Holmen Norway 0.724 0.840 0.460 30.5 1.10 SPT SCPTu
Gioia Tauro Italy 0.589 0.690 0.450 42.0 0.42 SPT SCPTu

N60 following the recommendations of Skempton (1986). Shear Many sets of CPTs, CPTUs, and SCPTUs have been con-
Downloaded by [65.209.12.68] at 12:29 22 March 2016

wave velocities Vs from the downhole method were obtained ducted at these six sites and the individual records are given by
during SCPTu soundings at these four sites in Japan, as well as in Wride and Robertson (1999). Statistical summary graphs for
the laboratory series (Yamashita et al. 2003). the target zones at each site are available in Wride et al. (2000),
and these have been used to provide mean values for the in situ
tests. For the SPTs, energy measurements were made in the
2.2 Canada field to obtain the N60 values properly. Energy ratios between
50% and 80% were determined during these programmes.
Data on the six Canadian sands were obtained from two natural
deposits of the Fraser River Delta (Massey and Kidd sites), two
tailings basins (Mildred Lake and J-pit), and two tailings dams 2.3 Norway
(LL Dam and Highmont Dam). These sites were investigated for
the large Canadian Liquefaction Experiment (CANLEX) with The information on natural Holmen sand from the Drammen
summary papers given by Robertson et al. (2000a,b) and Wride area of Norway is summarized by Lunne et al. (2003), with
et al. (2000). Many sets of frozen sand samples were taken at supplementary data provided by Gregersen et al. (1973). The
each site for this project and subjected to laboratory testing. A Holmen sand has served as a reference experimental test site
number of series of in situ field tests were also conducted. Thus for the Norwegian Geotechnical Institute (NGI) for five
the reported values of parameters actually represent mean values decades. Here, the very loose clean sands were sampled using
for these sites, as detailed in the cited references. thin-walled tubes; in some cases, the material was frozen within
As the focus of these field experiments was related to soil the tubes after extraction. Four undisturbed specimens were
liquefaction, the majority of the triaxial tests were of both aniso- tested under anisotropically consolidated drained triaxial
tropically and isotropically consolidated undrained compression compression (CADC) conditions, as well as separate triaxial
(CAUC, CIUC) with additional tests performed as triaxial exten- series conducted on pluviated specimens.
sion (CAUE, CIUE) types, all with porewater pressure measure- In-place void ratios and unit weight profiles were obtained in
ments. In addition, limited numbers of specimens were tested the loose Drammen sands using special nuclear density probes
drained (CIDC, CADC) and undrained direct simple shear (NDP) and electrical resistivity probes (ERP). Results from
(DSS). The summary papers concentrate on details related to the CPT, CPTu, and SCPTu soundings at Holmen have been
interpretation of undrained shear strength su in the the context of reported in various papers (Lunne et al. 1986, 2003,
liquefaction susceptibility. From the full comprehensive set of Campanella et al. 1986), as has a representative profile of
data review reports (five volumes) presented by Wride and SPT resistances with depth (Gregersen et al. 1973).
Robertson (1999), the triaxial series also permit evaluations of
the effective stress friction angle ¢ for each of the six sites, as well
as the corresponding value of the critical state friction parameter 2.4 China

The sands at the West Kowloon location in China have been


Mc ¼ ðq=p¢Þf ¼ 6 sin ¢=ð3  sin ¢Þ formed by hydraulic fill placement and were field sampled
using a special Bezier tube (Lee et al. 1999). Direct in-place
measurements of soil density were obtained by both nuclear
for triaxial compression loading. gauge and sand cone methods. Monotonic triaxial tests (CIUC)
244 P. W. Mayne

and cyclic triaxial tests (CTX) with porewater pressure mea- Lake) to as high as 2.0 mm (Gioia Tauro). The clean sands are
surements were performed on an undisturbed sample taken at a from China, Italy, Japan, and Norway, with FC , 4%, while the
depth of about 11.8 m, as well as on specimens reconstituted by dirty sands are from Canada, with 5% , FC ,15%.
moist tamping and air pluviation. In situ testing consisted of Simple statistical analyses gave mean values (and corresponding
borings with SPTs and companion series of piezocone sound- standard deviations) of the index parameters for all sands: specific
ings (CPTU). At the time that this paper was being prepared, no gravity Gs = 2.66  0.03; fines content FC = 4.36  4.49%;
shear wave measurements were available. particle size D50 = 0.35  0.23 mm, and uniformity coefficient
UC = D60/D10 = 2.80  1.19. (NB. Coarse sand from Italy
2.5 Italy with D50 = 2 mm was not included; if it is included,
D50 = 0.46  0.48 mm). It should be noted that it might be better
The Gioia Tauro site is a port located near the toe of Italy. to assume a Gaussian log normal distribution for the particle size
Undisturbed frozen sand specimens were obtained in the shallow and/or uniformity coefficient. Particle shapes vary from sub-
depth range of 1--4 m to procure the natural gravelly coarse sand rounded and subangular to angular. The sands were mostly quart-
(Ghionna and Porcino 2003, 2006). Here, the results of both zitic or siliceous (with major and comparable quantities of quartz
isotropically and anisotropically consolidated cyclic triaxial and feldspar components). however, the Japanese sands also had an
tests (CTX) were used to evaluate the liquefaction potential of appreciable rock mineral component of 9--54% (granitic minerals)
these geomaterials. The final effective stresses from q--p¢ plots and some mica (up to 7.4% for Yodo River sand).
of the CTX series showed Mc = 1.70 in compression The Canadian sands from the Massey, J-pit, Mildred Lake,
(¢ = 41.58) and Me = 1.08 in extension (¢ = 41.28); some and Kidd sites were generally quartzitic (70--90%), with feld-
Downloaded by [65.209.12.68] at 12:29 22 March 2016

limited monotonic tests with confining stresses as high as spar accounting for 5--15%. In contrast, the two tailings dam
c ¢ ¼ 300 kPa showed a basic frictional response at sands (LL Dam and Highmont) contained more unusual miner-
Mc = 1.54 (¢ = 37.88). In situ SPT and SCPTU data from als, including rather high percentage totals of kaolinite, chlor-
the Italian port site are reported in Porcino and Ghionna (2004). ite, smectite, illite, calcite, and mica, totalling some 53% for the
LL Dam and 19% for the Highmont Dam. Specifically, these
two sands were angular in particle shape compared with sub-
3. Index relationships for sands rounded to subangular for the other four Canadian sands. Also,
the sands from LL Dam and Highmont contained 36% and 57%
Of initial interest in the collection of these sands is to confirm quartz, respectively (Wride and Robertson 1999).
that they appear ‘well behaved’ in their index relationships, so The void ratio indices of each sand were determined using
that potential anomalous results can be culled from the database standard laboratory procedures. The maximum void ratio emax
or identified for separate consideration. Considering all 15 sands, ranged from 0.63 to 1.33 (mean 1.02) and the minimum void
they are all clean to slightly silty/clayey and all have ,15% fines ratio emin ranged from 0.33 to 0.81 (mean 0.60). The void ratio
content (FC). By mechanical analyses, they include 10 fine difference (emax -- emin) has been shown to be related to mean
sands, four medium sands, and one coarse sand. The mean particle size D50 (Cubrinovski and Ishihara 1999). Figure 4
particle size D50 ranged from as low as 0.16mm (Mildred illustrates that the 15 datapoints conform to the expected

1.0
Yodo River Natori River
0.9
Tone River Edo River
0.8 Mildred Lake Massey
Kidd J-Pit
0.7 LL-Dam Highmont
emax - emin

Holmen W. Kowloon
0.6 Gioia Tauro C&I 1999
0.5

0.4

0.3
0.2
0.06
0.1
emax − emin = 0.23 +
D50 ( mm )
0.0
0.1 1 10
Mean Particle Size, D50 (mm)
Figure 4. Void range index versus mean grain size for undisturbed sand.
The Second James K. Mitchell Lecture 245

trend borne by prior data. Similarly, in a later study, ðN1 Þ60 11:7
Cubrinovski and Ishihara (2002) observed a set of interrelation- CD * ¼ ¼ ð1Þ
DR 2 ðemax  emin Þ1:7
ships between the extreme void ratios which depends upon
percentage of fines content. Again, the emin versus emax values
from the database of 15 sands can be seen to agree with trends
where ðN1 Þ60 ¼ N60 =ðvo ¢=atm Þ0:5 is the energy-corrected N
for clean (FC , 5%) and dirty (5 , FC , 15%) sands (figure 5).
value normalized to a common reference stress and atm = 1
atm = 1 bar = 100 kPa. Comparison of this approach with
4. Relative densities the available data indicates rather weak agreement (figure 6).
Additional studies of the relative density trends for the
The relative density DR is a common measure of the degree of Canadian sand series by Wride et al. (2000) have suggested
packing in sands. In US practice the term applies strictly to sands that age may be a significant variable.
with FC , 15%, whereas in Japan the relative density can be The results of the CPT can be used similarly in an attempt to
applied to soils with a much higher level of fines (Cubrinovski assess the in-place relative density (e.g. Schmertmann 1978,
and Ishihara 1999). However, in the author’s opinion, DR is a Jamiolkowski et al. 1985). A re-examination of the large sets of
rather weak parameter since it relies on a very small range of void calibration chamber tests by Jamiolkowski et al. (2001) was
ratio extremes (emax -- emin), averaging 0.42 for these sands, and done, with the penetration readings corrected for limited sizes
enlarges this small e to vary from 0 to 100%. of the chambers. It was found that
For the data under consideration, the in-place void ratios e0
have been determined from unit weights on the undisturbed    
Downloaded by [65.209.12.68] at 12:29 22 March 2016

samples and the corresponding relative densities calculated qt


DR ¼ 100  0:268  ln pffiffiffiffiffiffiffiffi  1:292 : ð2aÞ
from DR = (emax -- e0)/(emax -- emin). In some cases (Norway, vo ¢
China, and Canada) geophysical logging and nuclear isotope
readings were used to confirm unit weights. The range of in- where both cone tip resistance qt and effective overburden
place e0 varied from as low as 0.49 to as high as 1.04 (mean stress vo ¢ are input in units of kPa. If they are expressed in
e0 = 0.83  0.14), corresponding to a relative density range bars, the equation becomes
of 29% , DR , 77% with an overall mean DR = 46%  14%.
In geotechnical practice the relative density of sands is nor- " ! #
mally assessed using in situ penetration tests. Cubrinovski and qt =atm
DR ¼ 100  0:268  ln pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  0:675 : ð2bÞ
Ishihara (1999) have related the stress-normalized N1-value vo ¢=atm
(N value adjusted to an effective overburden stress of one
atmosphere) to the relative density and void ratio potential
(emax -- emin). On conversion to an energy efficiency of 60%, The data are shown to be consistent with the mean trend 
the relationship becomes: two standard deviations presented in the relationships by
Jamiolkowski et al. (2001) in figure 7, and suggest that the

1.0

0.9 Cubrinovski & Ishihara (2002)


for Clean Sands:
0.8 FC ≤ 5%: emax = 0.072 + 1.53 emin
0.7

0.6
emin

0.5

0.4
Data: FC ≤ 5%
0.3
Data: 5 < FC < 15 % Sands with Fines:
0.2
C&I: FC ≤ 5% 5 < FC < 15%:
0.1 C&I: 5 < FC < 15% emax = 0.25 + 1.37 emin
0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5

emax
Figure 5. Interrelationship of void limits for undisturbed sands in database.
246 P. W. Mayne

200
Yodo River Natori River
Tone River Edo River
Mildred Lake Massey

CD* = (N1)60/ (DR)2


Kidd J-Pit
150 LL-Dam Highmont
Holmen W. Kowloon
Gioia Tauro C&I (1999)

100

50

0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

Void Range Potential (emax-emin)


Figure 6. Normalized SPT: relative density format versus void ratio difference for undisturbed sands as suggested by Cubrinovski and Ishihara (1999).

100
Downloaded by [65.209.12.68] at 12:29 22 March 2016

Japan
90
Relative Density, D R (%)

Canada
80 Norway
70 China
Italy
60 JL&M (2001)
50

40

30

20 ⎡ ⎛ qt / σatm ⎞ ⎤
DR = 100 ⋅ ⎢ 0 .268 ⋅ ln ⎜ ⎟ − 0 .675 ⎥
10 ⎢⎣ ⎜ σ '/ σ ⎟ ⎥⎦
⎝ vo atm ⎠
0
10 100 1000
0.5
Normalized Tip Stress, qt1 = qt /(σvo ')
Figure 7. Relative density from CPT measurements for undisturbed sands in terms of suggested relationship by (Jamiolkowski et al. 2001).

sands may be of low compressibility since they fall on the lower Figure 8 shows the sand data taking ageing and overconsolida-
bounds of the lines. tion effects into consideration, but without varying the compres-
An expression for relative density from CPTs that accounts sibility factor (QC = 1). For Holocene soils less than 10 000
for the additional effects of ageing and overconsolidation is of years old, groundwater fluctuations and a general rise of mean
the form (Kulhawy and Mayne 1991) sea level of approximately 40 m worldwide can easily be used to
justify a degree of overconsolidation, as well as OCRs caused by
 0:5
qt1 mechanical erosion, desiccation, wetting/drying, and other factors
DR ¼ 100  ð3Þ (Locat et al. 2003). According to recent work by Cubrinovski and
305  QA  Qc  QOCR
Ishihara (1999), it is also possible that a particle-size scale effect
may exist in the relationship between DR and CPT measurements.
where QA = 1.2 + 0.05 log(t/100) is an ageing correction Notably, there appears to be wide scatter in all three relative
factor, t is time (years), QC is the sand compressibility factor density correlations for the reasons discussed previously.
(QC has values of 0.9, 1, and 1.1 for high, medium, and low
compressibility, respectively), QOCR = OCR0.2 is a factor
accounting for the degree of overconsolidation, and 5. Effective stress friction angle of sands
qt =atm qt
qt1 ¼ ¼ ¼ qt =ðsvo ¢  satm Þn is
ðvo ¢=atm Þn ðvo ¢  atm Þn The peak effective stress friction angle of sands P ¢ is of interest
the stress-normalized cone tip resistance. in the geotechnical design of walls, footing bearing capacity,
The Second James K. Mitchell Lecture 247

100 on measured P ¢ versus measured DR for the undisturbed sand


AGE (years) = 1 100 10
4
} NC database, partly because this approach does not consider
Relative Density D R (%)

80 } OC effective confining stress level.


For triaxial compression loading, a generalized form is given
OCR=4
60
by (Bolton 1986) whereby:

40
 

TX ¢ ¼ cs ¢ þ 3 DR Q  lnðpf ¢=atmÞ  R  cs ¢ ð4Þ

20
where DR is the relative density (decimal value), Q is an
0
empirical term describing soil mineralogy and compressibility,
0 50 100 150 200 250 300 R is an an empirical fitting parameter, and pf ¢ is the mean
Normalized Tip Stress, qt1 principal effective stress at failure. For common quartz sands,
values of cs ¢ ¼ 338, Q = 10, and R = 1 are often adopted
Figure 8. Relative density from CPT results with consideration of age and (Bolton 1986).
OCR. For triaxial stress paths, the estimated value of pf ¢ can be
taken as approximately twice the effective overburden
end-bearing resistance of deep foundations, pile side fric- stress (Kulhawy and Mayne 1990), but the actual magni-
tion, slope stability concerns, and other important civil
Downloaded by [65.209.12.68] at 12:29 22 March 2016

tude depends on the particular problem. Jamiolkowski et al.


engineering works. The peak friction angle can be consid- (2001) suggest detailed estimates of pf ¢ based on the parti-
ered as having a basic friction component cs ¢ due to cular application, specifically to problems of shallow and
mineralogy, particle shape, size, and roundness (Koerner deep foundations. Bolton (1986) suggested that the para-
1970, Santamarina and Cho 2004) and a dilatancy compo- meter Q took values of 10 for quartz, 8 for limestone, and
nent due to packing arrangement, relative density, test 5.5 for chalk, whereas Jamiolkowski et al. (2001) give
mode, and effective confining stress (Rowe 1969, Bolton quite specific values of cs ¢ and Q for individually identi-
1986, Jamiolkowski et al. 2001). An early relationship fied sands having quartz, quartzitic-feldspathic (siliceous),
suggested by Schmertmann (1978) related the peak friction calcareous, and glauconitic mineralogies. Moreover,
angle to the relative density, depending upon the particle Salgado et al. (2000) provide fitting regressions for several
size of the sand. As 10 of the 15 soils are fine sands, figure 9 series of clean to silty quartz--Ottawa sand mixtures to
shows that this provides a rather conservative estimate based obtain specific parameters for all relevant soil constants

46

44
Meas Triaxial φ' (deg.)

42

40

38

36

34
FINE: 0.075 < D50 < 0.42 mm
32
MEDIUM: 0.42 < D50 < 2 mm

30 COARSE: 2 < D50 < 4.75 mm

28
0 10 20 30 40 50 60 70 80 90 100

Relative Density, DR (%)


Figure 9. Peak friction angle in terms of relative density and sand gradation (relationship suggested by Schmertmann (1978)).
248 P. W. Mayne

ðcs ¢; Q; RÞ. In that study, the parameters Q and R also given in table 4. The energy-corrected and stress-normal-
appear to vary with the relative density of the sand. The ized SPT N value is given by ðN1 Þ60 ¼ ðN60 Þ=ðvo ¢Þ0:5 ,
best fit for their clean laboratory prepared Ottawa sand where the effective overburden stress is in units of bars
series gave cs ¢ ¼ 29:08; Q ¼ 9:0, and R = 0.49. Similar or atmospheres (1 bar = 100 kPa). Details of this normal-
laboratory studies by Lee et al. (2004) also showed ranges ization procedure are given elsewhere (Skempton 1986,
for these three parameters. Kulhawy and Mayne 1990, Robertson et al. 2000a,b). The
When using this framework in the context of in situ mean (1 SD) for the (N1)60 values in these sands was
testing, the penetration resistance is utilized to evaluate 20.4  14.0.
DR within (4) to obtain peak P ¢. From a reliability view- The cone tip resistance qt is presented in two normalized
point, this is a rather weak approach as the DR correlations formats in table 4, with Q ¼ ðqt  vo Þ=vo ¢ and
from SPT or CPT are not particularly strong (and are qt1 ¼ qt =ðvo ¢Þ0:5 , where qt and vo ¢ are in bars or atmospheres
codependent on additional factors such as age, compressi- (Robertson and Wride 1998). Mean values (1 S.D.) in the
bility, and/or OCR), and the user must also adopt appro- sands are Q = 102.1  81.0 and qt1 = 94.9  58.3. The CPT
priate values of material constants ðcs ¢; Q; and RÞ as well sleeve friction is expressed in two formats for friction ratio:
as assessing the final mean stress pf ¢. Therefore it is more Rf(%) = fs/qt, following Schmertmann (1978), and
satisfying to link penetration data to P ¢ more directly, F(%) = fs/(qt -- vo), as recommended by Wroth (1988).
either through calibrated theoretical formulae or by empiri- Corresponding statistics gave mean Rf = 0.63  0.42% and
cal relationships. F = 0.65  0.43%. The measured excess porewater pressures
were small for all sands, and table 4 presents the normalized
Downloaded by [65.209.12.68] at 12:29 22 March 2016

u=vo ¢ values (-0.13  0.20). The alternative parameter


6. Direct SPT and CPT methods Bq = u/(qt -- vo) was very small with calculated with
mean -0.001  0.005.
Table 4 also includes the material index Ic from the CPT
A number of direct CPT methods have been proposed for
soil behavioural type (SBT), as discussed by Robertson and
evaluating P ¢ from cone tip stress in sands. These include
expressions based on analytical formulations (limit plasticity Wride (1998). The mean value for the sands was
theory, cavity expansion), critical state soil mechanics (state Ic = 1.64  0.37, putting the majority near soil class 6
parameter), numerical simulations (finite elements), and (‘clean sands’) and some in class 5 (‘silty sands’). The
empirical trends from statistical regressions. Reviews of Italian sand was correctly identified as class 7 (‘gravelly
selected methodologies have been given by Mitchell and sand’). For the very low excess porewater pressures, Ic is a
Lunne (1978), Mitchell and Keaveny (1986), Jamiolkowski function of Q and F only. However, the reliability of CPT
and Robertson (1988), Lunne et al. (1994, 1997), (Mayne classification charts based on sleeve friction measurements
et al. 1995), Chen and Juang (1996), Yu and Mitchell (1998), can be somewhat dependent upon the type of cone penetrom-
and Jamiolkowski et al. (2001). eter (tension versus subtraction sleeve), surface smoothness,
A tabulation of the individual peak friction angles and in wear, and other factors (e.g. Lunne et al. 1986), and so SBT
situ penetration test data for the 15 undisturbed sands is should only be used as a guide. The final column includes the

Table 4. Summary of triaxial, SPT, CPT, and Vs data for undisturbed sands

Normalized
CPT resistances and readings
SBTa
Sand Laboratory triaxial type ¢ (deg) SPT (N1)60 Q F fs/qt (%) qt1 u/vo¢ Ic Vs1 (m/s)

Kowloon CIUC 38.1 21 55.4 1.84 1.80 74.5 0.067 2.23 NA


Yodo CIDC 42.4 27 193.7 0.95 0.94 175.9 -0.051 1.63 197.2
Yodo CIDC 38.4 35 102.6 1.03 1.01 101.6 -0.190 1.81 213.1
Yodo CIDC 39.1 31 96.0 0.89 0.88 105.7 -0.186 1.76 195.2
Natori CIDC 40.9 50 225.7 0.30 0.30 207.3 -0.482 1.04 218.1
Tone CIDC 41.7 32 154.5 0.24 0.24 146.7 -0.458 1.07 203.2
Edo CIDC 39.7 22 156.1 0.73 0.72 110.6 -0.082 1.55 163.8
Mildred Lake CIUC, CKoUC 39.6 18 32.0 0.73 0.70 73.8 0.025 1.99 156.4
Massey CIUC, CKoUC 36.7 10 49.4 0.40 0.38 53.4 -0.089 1.63 168.2
Kidd CIUC, CKoUC 37.3 13 52,4 0.37 0.36 68.3 -0.020 1.59 177.4
J-pit CIUC, CKoUC 32.7 3 31.0 0.87 0.75 20.4 0.185 2.08 127.1
LL dam CIUC, CKoUC 39.1 5 38.6 0.41 0.39 39.4 0.011 1.73 153.1
Highmont CIUC, CKoUC 41.5 5 35.2 0.38 0.38 43.9 0.112 1.74 141.3
Holmen CKoUC, CKoDC 33.2 1 29.4 0.42 0.40 27.7 -0.325 1.83 156.6
Gioia Tauro CIDC, CIUC, CTX 41.5b 30 279.2 0.26 0.26 174.3 -0.261 0.93 221.0
a
Soil index Ic from soil behavioural type (SBT) according to Robertson and Wride (1998). For Bq = 0, Ic = [(3.47 -- log Q)2 + (log F + 1.22)2]0.5.
b
Result for final q/p¢ from CTX series; monotonic tests give 37.88.
The Second James K. Mitchell Lecture 249

normalized shear wave velocity Vs1 ¼ Vs =ðvo ¢=atm Þ0:25 Tauro also fits well. However, the data from the six loose
(Andrus et al. 2004, Robertson et al. 2000a,b). Statistics Canadian sands and the very loose Norwegian sand fall
gave mean values Vs1 = 178  30m/s for these sands well above the curve, while the Natori sand falls below.
(n = 14). The SPT may reflect more of an undrained or partly
undrained condition during penetration in loose sands and
silty sands because of its dynamic cyclic application.
6.1 SPT methods

For the SPT, only empirical relationships are available for


evaluating p ¢ (e.g. Stroud 1988, Kulhawy and Mayne 1990). 6.2 CPT methods
The following approximate expression can be generated using
the chart evaluation provided by Schmertmann (1975): The evaluation of the peak friction angle of sands from CPT can
be based on theory, numerical simulations, and/or empirical
 0:34 trends (Yu and Mitchell 1998). Based on bearing capacity
N60
¢  arctan ð5Þ considerations, Meyerhof (1976) suggested that the friction
12:2 þ ðvo ¢=atm Þ
angle of sands was related to the limiting static cone resistance
(qL = qc). As illustrated by figure 11, this appears to be a
conservative approach, in part because the influence of effec-
More recently, an expression was derived by Hatanaka and tive confining stress level is not considered.
Uchida (1996) based on correlations between frozen sand sam- Using bearing capacity theories as a guide, Robertson and
Downloaded by [65.209.12.68] at 12:29 22 March 2016

ples tested in consolidated drained triaxial compression tests Campanella (1983) presented CPT data from five artificial
and SPT field data. If these N values are corrected to an energy sands that had been prepared in flexible-walled calibration
efficiency of 60% (Skempton 1986), the expression is chambers, but not corrected for boundary size effects. The
well-known relationship has been adopted by many in geotech-
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi nical practice (e.g. Lunne et al. 1997) and can be expressed
¢  208 þ 15:4  ðN1 Þ60 ð6Þ conveniently as follows:

The application of (6) to the undisturbed sand database is P ¢ ¼ arctan½0:1 þ 0:38  logðqt =vo ¢Þ: ð7Þ
presented in figure 10, which shows only fair agreement.
As the relationship (6) was derived from Japanese sands, it
shows good agreement with three of the four undisturbed As can be seen in figure 12, this approach appears to
sands from Japan (Yodo, Edo, and Tone), as well as with overestimate friction angles at medium to high normalized
the Chinese sand (Kowloon). The Italian sand from Gioia cone stresses where qt =vo ¢ > 60. At lower values of qt =vo ¢

50

45
Triaxial φ' (deg.)

40

35

Yodo River Natori River


Tone River Edo River
Mildred Lake Massey
30 Kidd J-Pit
LL-Dam Highmont
Holmen W. Kowloon
Gioia Tauro Hatanaka & Uchida (1996)

25
0 10 20 30 40 50 60

Normalized N-Value, (N1)60


Figure 10. Comparison of measured triaxial friction angles with SPT relationship proposed by Hatanaka and Uchida (1996).
250 P. W. Mayne

45 boundary effects caused by the limited sizes of the cham-


bers (Salgado et al. 1998). The resulting expression for
evaluating peak friction angle applies to normalized cone
resistances (qt1 . 20) in quartz sands:
Triaxial φ ' (deg.)

40 " #
ðqt =atm Þ
P ¢ ¼ 17:68 þ 11:0  log : ð8Þ
ðv0 ¢=atm Þ0:5
Yodo River Natori River
35 Tone River Edo River
Mildred Lake Massey The normalization of cone penetration test data has been sug-
Kidd
LL-Dam
J-Pit
Highmont
gested as a means to account for effective stress levels and
Holmen W. Kowloon associated depth increases (e.g. Wroth 1988). Accordingly, a
Gioia Tauro Meyerhof 1976
normalization of tip resistance might take the form
30 Q ¼ ðqt  vo Þ=vo ¢, which has been used in soil behavioural
0 50 100 150 200 250 300 350 type classification charts (Robertson 1990). While this term
Limiting Tip Stress, qt = qL (bars) appears to work well in soft clays and silts, the normalization
is less successful in clean sands and intermixed soils (Olsen and
Figure 11. Comparison of measured triaxial friction angles with CPT rela- Mitchell 1995). As noted by Jamiolkowski and Robertson
tionship proposed by Meyerhof (1976).
(1988), there is little loss in accuracy for CPTs in sands by
Downloaded by [65.209.12.68] at 12:29 22 March 2016

adopting Q  ðqt =vo ¢Þ. Thus it is interesting to note the


revised normalization scheme offered by Robertson and
50 Wride (1998) with
φ'= arctan[ 0.1 + 0.38log(q t /σvo ' )]
Total %: kaolinite, smectite, ðqt  vo Þ=atm ðqt  vo Þ
45 calcite, illite, and chlorite Q* ¼ ¼
ðvo ¢=atm Þn ðvo ¢  atm Þn
Triaxial φ' (deg.)

0
18 0
0 0
where the exponent n requires an iterative approach and can
40 26
5 0
0
vary from n = 0.5 in clean sands to n = 1.0 in clays.
0 Japan Of course, the effective vertical stress is only one of three
10
10 Canada principal stress directions and it has been suggested that a more
35 Norway
7 China
appropriate normalization for cone tip resistance is in terms of
Italy the horizontal effective stress. Houlsby and Hitchman (1988)
5 R&C 1983 evaluated calibration chamber test results on sands and
30 expressed the triaxal peak friction angle in the form
0 50 100 150 200 250 300

Normalized Tip Stress, qt /σvo'


P ¢ ¼ 98 þ 6:25  lnðqt =h0 ¢Þ ð9Þ
Figure 12. Comparison of measured triaxial friction angles with CPT
relationship proposed by Robertson and Campanella (1983).
Using a series of laboratory tests on reconstituted sands to
calibrate a constitutive model, Lee et al. (2004) generated simu-
the friction angles of the three sands are very much under- lated cone tip resistances from a numerical code that were in turn
estimated. Most of the sands are composed of particles that back-related to evaluate the peak friction angle from
are mainly quartzitic or siliceous (quartz and feldspar) or con-
tain other hard rock mineral types. Two sites from Canada (LL
Dam and Highmont Dam) have more complex mineralogies, as P ¢ ¼ 15:68  ðqt =h0 ¢Þ0:171 ð10Þ
noted earlier, with higher percentages of kaolin, illite, chorite,
mica, and smectite. However, there is considerable uncertainty in the field
The type of stress normalization can be shown to have an assessment of the effective horizontal stress ho ¢ ¼ K0 vo ¢
influence on this approach. An alternative scheme is to because K0, which is the coefficient of lateral earth stress,
normalize cone tip stress by the square root of effective can be considered as a function of load history, friction
overburden stress, or use a power term expression, as angle, overconsolidation ratio, and other factors (Mayne
suggested by Jamiolkowski et al. (1985), Olsen and and Kulhawy 1982).
Mitchell (1995), and others. In the assessment of flexible- Self-boring pressuremeter data (SBPMT) were obtained
walled CPT calibration chamber test data by Kulhawy and at the six Canadian sites and generally showed 0.4 , K0
Mayne (1990), the qt data were corrected to account for , 0.7 for these sands. At the Holmen sand site in Norway,
The Second James K. Mitchell Lecture 251

SPBMT indicated K0 < 0.4. However, no SPBMT or field strain gR and the rigidity index is its reciprocal; thus
K0 data were reported for the sands in Italy, Japan, and IR = 1/gR.
China. For the undisturbed sand database, a common Vesić (1977) formulated his equations in terms of mean
assumption by the reference sources was that their sands effective stress o ¢; however, from a practical standpoint, the
had an estimated in-place K0 = 0.5 (generally representa- effective vertical stress can be the operative overburden term.
tive of normally consolidated soils). While this appears For a frictional soil (c¢ = 0), spherical cavity expansion (SCE)
oversimplistic for all sands, this value has been adopted theory gives
here, with figure 13 showing rather weak agreement with
both (9) and (10) in these cases.    
qt 3  ð1 þ sin ¢Þ 4 sin ¢
N ¼ ¼  exp  ¢ tan ¢   IRR ð11Þ
v0 ¢ 3  sin ¢ 2 ð1  sin ¢Þ 3ð1 þ sin ¢Þ

7. Vesić cavity expansion theory where IRR = IR/(1 + IR"vol) is the reduced rigidity index and
"vol is the volumetric strain in the plastic zone. A later devel-
Cavity expansion theory is a well-known solution of the opment by Carter et al. (1986) removed the reliance on rigidity
penetrating probe problem. It has been successfully used index and volumetric strain by derivation using the dilatancy
to represent undrained penetration cases involving driven angle c¢.
pilings (Vesić 1977), pressuremeter tests (Yu 2004), A mapping of the paired sets of measured triaxial friction
piezocone profiling of OCRs in clays (Mayne 1991), and angles with corresponding normalized cone resistances over the
Downloaded by [65.209.12.68] at 12:29 22 March 2016

porewater pressure dissipations (Burns and Mayne 2002), SCE solution is presented in figure 14 for undisturbed sands.
as well as in other applications. By extending the The operational values of the rigidity index IRR are seen to
undrained cavity expansion problem to the drained case range from 7 to 310. In fact, individual back-calculated IRR
and taking volumetric strains into consideration, Vesić values can be determined for each of the 15 undisturbed sand
(1972, 1977) offered a means of describing the pile end- sets. These operational IRR values are listed in table 5 for each
bearing resistance in terms of the soil friction angle P ¢ sand.
and rigidity index IR, where IR = G/ max is the ratio of The SCE solution is presented for use in evaluating deep
the shear modulus to the shear strength. Alternatively, if foundation end-bearing resistance, and therefore the nor-
the soil stress--strain curve is represented by a simple malized tip stress ðQ  qt =vo ¢Þ is expressed as a function
bilinear elastic--plastic relationship, the shear modulus of both P ¢ and IRR. For the interpretation of the CPT, what
intersects the shear strength line at a reference shear is desired is the inverse problem with an expression that

50
Yodo River Natori River
Tone River Edo River
Mildred Lake Massey
Kidd J-Pit
LL-Dam Highmont
Holmen W. Kowloon
Gioia Tauro H&H (1988)
45 Lee et al. (2004)
Triaxial φ' (deg.)

40

35
Note: K0 = 0.5
σho' = K0 σvo'

30
0 100 200 300 400 500

Normalized Tip Stress, qt* = qt/σho'


Figure 13. Comparision of measured triaxial friction angles with CPT relationships proposed by Houlsby and Hitchman (1988) and Lee et al. (2004)
252 P. W. Mayne

Cavity Expansion Theory (Vesic 1972, 1977) Early studies suggested that the operational rigidity index IRR
1000
was inversely related to the CPT friction ratio Rf = fs/qt (%).
Irr = 500 Italy
Irr = 200 Japan
However, a check on two such relationships (Vesić 1977, Baldi
Bearing Factor, Nσ = qt/σ vo'

Irr = 100 China et al. 1981) showed rather poor correlation, as did other hypothe-
Irr = 50 Canada sized links. However, a rather substantial relationship was found
Irr = 20 Norway
Irr = 10 between IRR and the normalized shear modulus G0 =vo ¢ obtained
from shear wave velocity measurements Gmax ¼ G0 ¼ T Vs2 :

100 All sands ðn ¼ 14; r 2 ¼ 0:891Þ :


 
ðG0 =v0 ¢Þ 2:45 ð13aÞ
IRR  :
110

The total mass density rT of the soil can be estimated from Vs


and the depth (Mayne 2001). It was determined that the LL
Dam and Highmont Dam sites would be best separated from the
10 other data because of their more unusual mineralogies. In that
30 35 40 45 case, the regression shown in figure 16 gave
Effective Friction Angle, φ ' (deg)
Most sands ðn ¼ 12; r 2 ¼ 0:927Þ :
Downloaded by [65.209.12.68] at 12:29 22 March 2016

Figure 14. Peak effective triaxial friction angles of undisturbed sands


mapped onto spherical cavity expansion theory solution of Vesić (1977)  
ðG0 =v0 ¢Þ 2:21 ð13bÞ
IRR 
85
relates P ¢ deterministically to the CPT measurements.
Thus an approximation was sought over the range The LL and HM sands can be represented empirically by a
308  P ¢  508 which relates P ¢ as a function of Q and similar shifted relationship (shown by the broken line) by
IRR, which is given by replacing the term 85 with 134.
Based on understanding of small-strain stiffness obtained
  primarily from laboratory resonant column tests on soils (e.g.
log 0:8  Q  IRR 0:25
¢  138  : ð12Þ Hardin 1978), a likely parameter would have been
3:3 þ logðIRR Þ
G0 =ðvo ¢atm Þ0:5 ; however, the normalization scheme in (13)
gave a much better regression coefficient. Recent work by
Using the measured Q and back-calculated IRR values, figure 15 Jardine (1995), Jefferies and Shuttle (2005), and Jardine et al.
shows that (12) gives a good match to the measured triaxial (2005) representing soil stiffnesses ranging from small to inter-
friction angles ranging from 32.78 to 42.48, with errors in mediate to high strain levels has used a similar normalization
reproduction no more than 0.58 compared with the exact theory. form (i.e. G0 =o ¢) to that in (13) with success.
The approximation is valid for the full ranges 20 , Q ,300 and Therefore the use of seismic cone penetration tests in sands is
7 ,IRR ,250. advantageous in that it provides both the measured tip stress qt

Table 5. Values of operational rigidity index IRR from SCE theory and plasticification angle  from NTH limit plasticity solution for undisturbed sands

Sand Mean z (m) ¢ (deg) CPT Q SCE IRR NTH  (deg) G0a (bar) vo¢ (bar)

Kowloon 11.8 38.1 56.6 25.9 -4.9 630 1.80


Yodo 8.1 42.4 195.5 133 -24.2 630 1.02
Yodo 10.9 38.4 104.4 82 -25.5 800 1.23
Yodo 12.7 39.1 97.9 63 -19.2 740 1.43
Natori 8.25 40.9 227.5 232 -37.7 780 0.87
Tone 7.3 41.7 156.4 99 -20.9 597 0.84
Edo 3.85 39.7 157.4 143 -32.4 340 0.51
Mildred Lake 32 39.55 33.2 7.4 20.7 980 5.16
Massey 10.5 36.74 50.5 26 -7.6 538 1.20
Kidd 14.5 37.3 54.7 27.4 -7.7 702 1.60
J-pit 5 32.74 30.5 18.5 -8.4 194 0.55
LL damb 8 39.13 40.2 11.4 12.2 400 1.00
Highmontb 10 41.45 37.4 7 24.2 395 1.38
Holmen 10 33.2 23.6 10 5.3 358 1.10
Gioia Tauro 3 41.5 280.5 307 -40.9 566 0.42
a
G0 = Gmax = small strain shear modulus from field shear wave velocity.
b
Higher percentages of kaolin, chlorite, smectite, and other minerals.
The Second James K. Mitchell Lecture 253

Spherical Cavity Expansion Approximation undrained penetration by limit plasticity developed by the
45
Norwegian Institute of Technology (NTH) (Janbu and
Regression
Best Fit Line (b = 0): Senneset 1974, Sandven et al. 1988) since it allows the
Meas. = 1.001 Pred. evaluation of effective friction angle for all soil types
n = 15; r2 = 0.991 (clays, silts, sands) by consideration of both tip resistance
Meas Triaxial φ'

40 and measured excess porewater pressures.


In the general case of soil represented by a Mohr--
Coulomb strength criterion (¢, c¢), the normalized cone
China resistance number is given by Nm ¼ ðqt  vo Þ=ðvo ¢ þ a¢Þ,
Japan
35 where the attraction a¢ = c¢cot¢. For the more specific
Canada
Norway
case of drained penetration with c¢ = 0, the cone resistance
Italy number becomes the normalized cone tip resistance defined
1:01 earlier, such that Nm ¼ Q ¼ ðqt  vo Þ=vo ¢  qt =vo ¢. In
30
the classical sense, the end-bearing resistance for sands is
30 35 40 45 represented by qt ¼ Nq vo ¢, where Nq is the bearing factor
φ' from SCE using Q and IRR given by

Figure 15. Comparison of measured triaxial friction angle with forward pre- Nq ¼ qt =vo ¢
diction by approximate expression using normalized CPT Q and operational  
1 ð14Þ
Downloaded by [65.209.12.68] at 12:29 22 March 2016

rigidity index IRR. 2


¼ tan 458 þ ¢  exp½ð  2Þ  tan ¢
2

where  is the angle of plastification. For the case of


1000
general shear and a failure zone that extends below the
tip and up around to the shoulder level of the cone, Nq is
Rigidity Index, IRR

Japan
given for the case where  = 0 (Terzaghi 1943). More
100
Norway flexibility is obtained with a variable size failure zone
Italy
Canada and can be used to account for soil compressibility effects.
LL & HM Therefore Senneset et al. (1989) suggest the range -308 
Trend
Trend LL-HM   +158 for soils.
10 A mapping of the triaxial friction angles and corresponding
qt =vo ¢ measurements on the NTH plasticity solution is shown
in figure 17. The data fall within the expected bounds for .
By iteration, (14) has been fitted to the undisturbed sand data-
1 base to back-calculate specific values of the angle of plastifica-
100 1000 10000 tion , as listed in table 5.
Ratio Gmax/σ vo' In lieu of iteration, an approximation can be made for the
range 308  P ¢  508:
Figure 16. Observed relationship between operational rigidity index IRR
and normalized small-strain modulus from shear wave measurements  
(G0/vo¢). ln½0:94  ðqt =v0 ¢Þ
¢  arctan ð15Þ
4:87 þ 0:035  

and the shear wave velocity Vs, thus allowing evaluations of the
Comparison of the measured ¢ with predicted values
necessary parameters ðqt =vo ¢ and IRR Þ to determine ¢ via
from approximation (15) using measured Q and back-
SCE theory.
fitted  from theory are in good agreement, as shown by
figure 18.
Guidance on the selection of  for different soil conditions
8. NTH limit plasticity theory is given by Senneset et al. (1988, 1989). Sands can be sum-
marized into loose (+108    +208), medium (-58    +
Limit plasticity has long been used to evaluate the end- 58), dense (-108    -208), and very dense (-308). Of
bearing resistance of both shallow and deep foundations course, now the problem is reliance on the rather poor-quality
(e.g. Meyerhof 1951), penetrating wedges (Durgunoglu correlation of DR with field CPT measurements to assess the
and Mitchell 1975a,b), walls and slopes (Atkinson 1981), sand consistency. Therefore a direct relationship for  was
and cones during undrained penetration (e.g. Konrad and sought and found to relate well to the normalized shear mod-
Law 1987) and drained penetration (e.g. Mitchell and ulus G0 =vo ¢, as illustrated in figure 19; it is given by (n = 14,
Keaveny 1986). Of particular interest is the effective stress r2 = 0.843).
254 P. W. Mayne

NTH Limit Plasticity Theory (Senneset et al. 1982; 1989) 50


1000 o o
40 β = 195 - 33.7 ln(G0/σvo')

Backfigured β (deg)
30 2
n = 12 r = 0.924
20
10 Japan
Bearing Factor, Nq = qt/σvo'

0 Canada
100
-10 Norway

-20 Italy

-30
Beta = -40
-30 -40
-20
-10 -50
10 0
+10 0 500 1000 1500
+20 Normalized Modulus, G0/σvo'
+30
-β Japan
Figure 19. Observed relationship between operational angle of plastification 
Canada
+β Norway and normalized small-strain modulus from shear wave measurements (G0/vo¢)
China
Italy
1
30 35 40 45 50 strain gs taken through the origin, or G = /gs. Correspondingly,
Effective Friction Angle, φ' (deg) the slope of deviator stress q = 1 -- 3 versus axial strain "a is
Downloaded by [65.209.12.68] at 12:29 22 March 2016

Figure 17. Mapping of friction angles from undisturbed sands onto NTH used to provide an equivalent elastic secant modulus E¢ = (1 --
limit plasticity solution of (Senneset et al. 1988, 1989) NTH solution. 3)/"a. On the other hand, for numerical simulations involving
stepped finite elements or finite-difference solutions, it is com-
mon to use a tangent value of shear modulus (Gtan = /gs)
NTH Method
45 or the alternative tangent Etan = (1 -- 3)/"a. In these
Regression situations, a bulk modulus K¢ may also be needed for numerical
Best Fit Line: codes (e.g. Jardine et al. 2005). In all cases, it turns out that soil
Measured Triaxial φ'

b=0 behaviour is highly non-linear from initial loading to working


Meas = 0.997 Pred. loads to peak and post-peak states.
40 n = 15 The initial stiffness of soils is represented by the small-strain
r2 = 0.998 shear modulus Gmax or the corresponding elastic Young’s mod-
ulus Emax = 2Gmax(1 + n), where n is Poisson’s ratio taken as
China n¢ = 0.2 for drained conditions and nu = 0.5 for the undrained
Japan case. For convenience, the author has adopted a modified
35 Canada hyperbola (Fahey and Carter 1993) since it can be expediently
Norway used to evaluate a secant shear modulus G from the initial
Italy Gmax value, depending upon the level of mobilized stress
1:1 (/ max = 1/FS), where FS is the factor of safety. The scheme
30 can also be used to obtain tangent moduli, if needed.
30 35 40 45 In the original proposed version, two parameters (f and g)
were used to obtain the secant modulus reduction factors
Approx φ' from Q and β
(Fahey 1998) by one of the following equations:
Figure 18. Comparison of measured triaxial friction angle with forward-pre-
diction by approximate expression using normalized CPT Q and plastification G=Gmax ¼ 1  f ð=max Þg ð17aÞ
angle .

E=Emax ¼ 1  f ðq=qmax Þg : ð17bÞ


  2228  37:68  lnðGmax =vo ¢Þ ð16Þ

9. Stress--strain strength of sands In the above, no changes in n¢ are made during loading;
however, these could be implemented as needed (Fahey and
The representation of the soil stiffness and strength can be Carter 1993). Poisson’s ratio data for six sands presented
accomplished using simple linear elastic--plastic forms, non- by Lehane and Cosgrove (2000) show relatively constant n¢
linear algorithms, and/or constitutive soil models. The many during loading up to about 0.1% strains, but subsequently
methods available have been reviewed by Duncan (1994) and indicating increases in n¢. Expressions (17) have been used
more recently by (Lade 2005). effectively to represent stress--strain--strength curves in dif-
For analytical solutions, a secant value of shear modulus G is ferent soil types, including clays (Mayne 2001, Elhakim
normally obtained from the slope of shear stress  versus shear and Mayne 2003), silts (Mayne et al. 1999), and clean to
The Second James K. Mitchell Lecture 255

Cone Tip Stress, qt (MPa) Sleeve Friction, fs (kPa) Porewater, ub (kPa) Shear Modulus, Gmax (MPa)
0 10 20 30 0 20 40 60 80 100 -100 0 100 200 300 0 50 100 150
0 0 0 0

1 1 1 1
Edo
2 2 2 River 2
Sand
Dept h ( met er s)
3 3 3 3
Frozen G0 = 34 MPa
4 4 4 4
Sample
5 5 5 5

6 6 6 6

7 7 7 Natori 7
Frozen River
8 8 8 8
Sand
Sample G0 = 78 MPa
9 9 9 9

10 10 10 10

Figure 20. Seismic piezocone results for Edo sand from 0 to 5.2m and Natori sand from 5.6 to 10m (data from Mimura 2003).

silty sands (Fahey 1998, Lehane and Fahey 2002, Lee et al. CIDC Drained Triaxial Tests (Mimura, 2003)
450
2004). For uncemented soils that are not highly structured,
Natori River Sand (8.2 m)
values of f = 1 and g = 0.3  0.1 have been suggested as 400
Downloaded by [65.209.12.68] at 12:29 22 March 2016

Deviatoric stress, q (kPa)


initial ‘guestimates’ (Mayne 2005). 350
The drained strength of sands depends on the particular mode g = 0.4
300
of loading application. For direct simple shear (DSS), the shear
250 0.3
strength can be stated simply as
Edo River Sand (3.8 m)
200
max ¼ vo ¢ tan ¢ g = 0.4
150

100 0.3
while for isotropic triaxial drained compression tests (CIDC) E/Emax = 1 - (q/qmax)g
50
the peak deviator stress is given by
0
qmax ¼ ð1  3 Þmax ¼ 2vo ¢ sin ¢=ð1  sin ¢Þ 0 1 2 3 4
Axial Strain (%)

Figure 21. Measured stress--strain curves from frozen samples at Natori and Edo
For torsional shear tests (TS) and anisotropically consolidated sites with simplified curves derived from SCPTu data (data from Mimura 2003).
triaxial tests (CADC), consideration of the initial K0 state can
be made by stress path analyses. undrained loading, variable stress paths, volumetric strain pre-
In situ SCPTU data from two of the Japanese sand sites are dictions, and cyclic response, as well as representing post-peak
presented in figure 20, with the upper portion of the Edo River site softening. Consequently, future research should be directed
shown for depths of 0--5.5m and the lower portion of the Natori towards the calibration of laboratory and in situ tests within
River sand shown for depths of 6--10m. At the corresponding the framework of constitutive soil models as these will allow
depths of the undisturbed frozen sand samples taken for each site, versatility. The initiation of this database on undisturbed sands
the cone tip stress reading has been used to assess the friction will help to begin the important calibration of parameters
angle ¢ and triaxial strength qmax. The shear wave measurement needed for these constitutive soil models.
is used to determine the initial stiffness (G0 = Gmax). Calculated
stress--strain curves for these two specimens obtained using the
aforementioned algorithms are shown in figure 21. It can be seen 10. Conclusions
that g = 0.3--0.4 gives a good fit for both the 3.8 m sample at Edo
River and 8.2 m sample at Natori River. A special database on 15 undisturbed sands has been compiled.
Although the approach described above has been shown to be It includes in-place void ratio and triaxial strength determina-
simple and reasonable for two sands, it cannot effectively tions, primarily from special frozen specimens. Results of in
address the infinite number of possible stress paths and loading situ penetration tests (SPT and SCPTU) were also available for
conditions needed for a true realistic numerical modelling, cross-correlation of laboratory and field results.
because of its empirical basis. Nevertheless, the use of a con-
stitutive soil model, such as NorSand (Jefferies 1993, Jefferies 1. The data were used to calibrate the spherical cavity
and Shuttle 2005) or MIT-S-1 (Pestana and Whittle 1999), expansion solution of (Vesić 1972, 1977) for evaluating
would be beneficial because these models are capable of hand- ¢ from the normalized cone resistance qt =vo ¢ and the
ling many more facets of soil behaviour, including drained and operational rigidity index IRR. The parameter IRR
256 P. W. Mayne

trended well with normalized small-strain shear mod- Measurement of Soil Properties, Vol. I, pp. 172--189, 1975b (ASCE,
Reston, VA).
ulus G0 =o ¢. Elhakim, A. F. and Mayne, P. W.,. Derived stress--strain--strength of clays from
2. The dataset was also used to calibrate the limit plasticity seismic cone tests. In Deformation Characteristics of Geomaterials, Vol.
solution (Janbu and Senneset 1974) for determination of 1, pp. 81--87, 2003 (Swets & Zeitlinger, Lisse).
Fahey, M., Deformation and in-situ stress measurement. In Geotechnical Site
¢ as a function of qt =vo ¢ and the angle of plastification Characterization, Vol. 1, pp. 49--68, 1998 (Balkema, Rotterdam).
 which describes the size of the failure zone near the Fahey, M. and Carter, J. P., A finite element study of the pressuremeter in sand
tip. The angle  was shown to correlate with G0 =vo ¢. using a nonlinear elastic plastic model. Can. Geotech. J., 30(2), 348--362.
Ghionna, V. N. and Porcino, D., Undrained monotonic and cyclic behavior of a
3. Using a simple algorithm, complete stress--strain-- coarse sand from undisturbed and reconstituted samples. In Deformation
strength curves can be represented with G0 providing Characteristics of Geomaterials, Vol. 1, pp. 527--534, 2003 (Balkema,
the initial stiffness and  max obtained from the Rotterdam).
Ghionna, V. N. and Porcino, D., Liquefaction resistance of undisturbed and
interpreted ¢. reconstituted samples of a natural coarse sand from undrained cyclic
4. Seismic cone tests in sands provide an optimal method of triaxial tests. J. Geotech. Geoenviron. Eng., 2006, 132(2), 194--202.
capturing four independent readings with depth in the same Gregersen, O. S., Aas, G., and DiBiagio, E., Load tests on friction piles in sand.
In Proceedings of the 8th International Conference on Soil Mechanics and
sounding: qt, fs, ub, and Vs. Foundation Engineering, 1973, Vol. 2, pp. 109--118.
Hardin, B. O., The nature of stress--strain behavior of soils. In Earthquake
Engineering and Soil Dynamics, Vol. 1, pp. 3--90, 1978 (ASCE, Reston, VA).
Acknowledgements Hatanaka, M. and Uchida, A., Empirical correlation between penetration resis-
tance and effective friction angle of sandy soil. Soils Found., 1996, 36(4),
The author thanks the National Science Foundation (Award 1--9.
Hoeg, K, Dyvik, R., and Sandbkken, G., Strength of undisturbed versus
CMS-0338445) and the Mid-America Earthquake Center
Downloaded by [65.209.12.68] at 12:29 22 March 2016

reconstituted silt and silty sand specimens. J. Geotech. Geoenviron.


(Award EEC-9701785) for their support. Any opinions, find- Eng., 2000, 126(7), 606--617.
ings, and conclusions or recommendations expressed here are Hoffman, B. A., Sego, D. C., Robertson, P. K., and Fourie, A. B., Thawing
protocol of undisturbed samples of loose sand obtained by in-situ ground
those of the author and do not necessarily reflect those of the freezing. In Proceedings of the 49th Canadian Geotechnical Conference
NSF or MAE. (Frontiers of Geotechnology), 1996, Vol. 1, pp. 71--80.
Houlsby, G. T. and Hitchman, R., Calibration chamber tests of a cone pressure-
meter in sand. Geotechnique, 1988, 38(1), 39--44.
Jamiolkowski, M. and Robertson, P. K., Future trends for penetration testing. In
References Penetration Testing in the UK, pp. 321--342, 1988 (Thomas Telford,
London).
Andrus, R. D., Piratheepan, P., Ellis, B. S. Zhang, J., and Juang, C. H., Jamiolkowski, M., Ladd, C. C., Germaine, J., and Lancellotta, R. New devel-
Comparing liquefaction evaluation methods using penetration--Vs relation- opments in field and lab testing of soils. In Proceedings of the 11th
ships. Soil Dynam. Earthquake Eng., 2004, 24, 713--721. International Conference on Soil Mechanics and Foundation
Atkinson, J. H., Foundations and Slopes: An Introduction to Applications of Engineering, 1985, Vol. 1, pp. 57--154.
Critical State Soil Mechanics, 1981 (McGraw-Hill, London). Jamiolkowski, M., LoPresti, D. C. F., and Manassero, M., Evaluation of relative
Baldi, G., Bellotti, R., Ghionna, V., Jamiolkowski, M., and Pasqualini, E., Cone density and shear strength of sands from cone penetration test and flat
resistance in Dry NC and OC sands. In Cone Penetration Testing and dilatometer test. In Soil Behavior and Soft Ground Construction (GSP
Experience, pp. 145--177, 1981 (ASCE, Reston,VA). 119), pp. 201--238, 2001 (ASCE, Reston, VA).
Been, K., Lingnau, B. E., Crooks, J. H. A. and Leach, B., Cone penetration test Janbu, N. and Senneset, K., Effective stress interpretation of in-situ static
calibration for Erksak Beaufort Sea sand. Can. Geotech. J., 1987, 24(4), penetration tests. In Proceedings of the European Symposium on
601--610. Penetration Testing, 1974, Vol. 2.2, pp. 181--193.
Bolton, M. D., The strength and dilatancy of sands. Geotechnique, 1986, 36(1), Jardine, R. J., One perspective of the prefailure deformation characteristics of
65--78. some geomaterials. In Pre-Failure Deformation of Geomaterials, Vol. 2,
Burns, S. E. and Mayne, P. W., Analytical cavity expansion: critical state model pp. 855--885, 1995 (Balkema, Rotterdam).
for piezocone dissipation in fine-grained soils. Soils Found., 2002, 42(2), Jardine, R. J., Standing, J. R. and Kovacevic, N. Lessons learned from full scale
131--137. observations and the practical application of advanced testing and model-
Campanella, R. G., Robertson, P. K., and Gillespie, D., Seismic cone penetra- ling. In Deformation Characteristics of Geomaterials, Vol. 2, pp. 201--245,
tion test. In Use of In-Situ Tests in Geotechnical Engineering (GSP 6), pp. 2005 ((Taylor & Francis, London).
116--130, 1986 (ASCE, Reston, VA). Jefferies, M. G., Norsand, a simple critical state model for sand. Geotechnique,
Carter, J. P., Booker, J. R., and Yeung, S. K., Cavity expansion in cohesive 1993, 43(1), 91--102.
frictional soils. Geotechnique, 1986, 36(3), 349--358. Jefferies, M. G. and Shuttle, D. A. Norsand, features, calibration, and use. In
Chen, J. W. and Juang, C. H., Determination of drained friction angle of sands Soil Constitutive Models, Evaluation, Selection, and Calibration (GSP
from CPT. J. Geotech. Eng., 1996, 122(5), 374--381. 128), pp. 204--235, 2005, (ASCE, Reston, VA).
Cho, G-C., Dodds, J. and Santamarina, J. C., Particle shape effects on packing Koerner, R. M. Effect of particle characteristics on soil strength. J. Soil Mech.
density, stiffness, and strength, natural and crushed sands. J. Geotech. Found. Div.--ASCE, 1970, 96(SM4), 1221--1234.
Geoenviron. Eng., 2006, 132(5), 591--602. Konrad, J.-M. and Law, K. T., Undrained shear strength from piezocone tests.
Cubrinovski, M. and Ishihara, K., Empirical correlation between SPT N--value Can. Geotech. J., 1987, 24(3), 392--405.
and relative density for sandy soils. Soils Found., 1999, 39(5), 61--72. Kulhawy, F. H. and Mayne, P. W. Manual on Estimating Soil Properties for
Cubrinovski, M. and Ishihara, K., Maximum and minimum void ratio charac- Foundation Design. Report EL-6800, Electric Power Research Institute,
teristics of sands. Soils Found., 2002, 42(6), 65--78. Palo Alto, Ca, 1990.
Duncan, J. M., The role of advanced constitutive relations in practical applica- Kulhawy, F. H. and Mayne, P. W., Relative density, SPT, and CPT inter-
tions. In Proceedings of the 13th International Conference on Soil relationships. In Calibration Chamber Testing, pp. 197--211, 1991
Mechanics and Foundation Engineering, 1994, pp. 31--48 (Balkema, (Elsevier, New York).
Rotterdam). Lade, P. V., Overview of constitutive models for soils. In Calibration of
Durgunoglu, H. T. and Mitchell, J. K., Static penetration resistance of soils. I: Constitutive Models (GSP 139), pp. 1--34, 2005 (ASCE, Reston, VA).
Analysis. In In-Situ Measurement of Soil Properties, Vol. I, pp. 151--171, Lade, P. V. and Bopp, P.A. Relative density effects on sand behavior at high
1975a (ASCE, Reston, VA). pressures. Soils Found., 2005, 45(1), 1--26.
Durgunoglu, H. T. and Mitchell, J. K. (1975b). Static penetration resistance of Lee, J., Salgado, R., and Carraro, J. A. H., Stiffness degradation and shear
soils. II: Evaluation of theory and implications for practice. In In-Situ strength of silty sands. Can. Geotech. J., 2004, 41(5), 831--843.
The Second James K. Mitchell Lecture 257

Lee, K. M., Shen, C. K., Leung, D. H. K. and Mitchell, J. K., Effects of Pestana, J. and Whittle, A. J., Formulation of a unified constitutive model for
placement method on geotechnical behavior of hydraulic fill sands. J. clays and sands. Int. J. Numer. Anal. Methods Geomech., 1999, 23(12),
Geotech. Geoenviron. Eng, 1999, 125(10), 832--846. 1215--1243.
Lehane, B. and Cosgrove, E., Applying triaxial compression stiffness data to Porcino, D. and Ghionna, V. N., Comparison between in-situ and lab tests on
settlement prediction model for foundations on sand. Geotech. Eng., 2000, undisturbed frozen samples for a natural coarse sand. In Geotechnical and
142, 191--200. Geophysical Site Characterization, Vol. 2, pp. 1843--1850, 2004
Lehane, B. and Fahey, M., A simplified nonlinear settlement prediction model (Millpress, Rotterdam).
for foundations on sand. Can. Geotech. J., 2002, 39(2), 293--303. Robertson, P. K., Soil classification using the cone penetration test. Can.
Locat, J., Tanaka, H., Tan, T. S., Dasari, G. R., and Lee, H., Natural soils, Geotech. J., 1990, 27(1), 151--158.
geotechnical behavior and geological knowledge. In Characterisation and Robertson, P. K. and Campanella, R. G., Interpretation of cone penetration tests.
Engineering Properties of Natural Soils, Vol. 1, pp. 3--28, 2003 (Swets & Part I: Sand. Can. Geotech. J., 1983, 20(4), 734--745.
Zeitlinger, Lisse). Robertson, P. K. and Wride, C. E., Evaluating cyclic liquefaction potential
Lunne, T., Eidsmoen, T., Gillespie, D., and Howland, J. D., Laboratory and field using the cone penetration test. Can. Geotech. J., 1998, 35(3), 442--459.
evaluation of cone penetrometers. In Use of In-Situ Tests in Geotechnical Robertson, P. K., Wride, C. E., et al., The Canadian liquefaction experiment, an
Engineering (GSP 6), pp. 714--729, 1986 (ASCE, Reston, VA). overview. Can. Geotech. J., 2000a, 37(3), 499--504.
Lunne, T., Lacasse, S., and Rad, N. S., General report, SPT, CPT, PMT, and Robertson, P. K., Wride, C. E., et al., The CANLEX project, summary and
recent developments in in-situ testing. Proceedings of the International conclusions. Can. Geotech. J., 2000b, 37(3), 563--591.
Conference on Soil Mechanics and Foundation Engineering, 1994, Vol. 4, Rowe, P. W., The relation between the shear strength of sands in triaxial
pp. 2339--2403. compression, plane strain, and direct shear. Geotechnique, 1969, 19(1),
Lunne, T., Robertson, P. K., and Powell, J. J. M., Cone Penetration Testing in 75--86.
Geotechnical Practice, 1997 (Chapman & Hall, London). Salgado, R., Mitchell, J. K., and Jamiolkowski, M., Calibration chamber size
Lunne, T., Long, M. and Forsberg, C. F., Characterization and engineering effects on penetration resistance in sand. J. Geotech. Geoenviron. Eng.,
properties of Holmen, Drammen sand. In Characterisation and 1998, 124(9), 878--888.
Engineering Properties of Natural Soils, Vol. 2, pp. 1121--1148, 2003 Salgado, R., Bandini, P., and Karim, A., Shear strength and stiffness of silty
(Swets & Zeitlinger, Lisse). sand. J. Geotech. Geoenviron. Eng., 2000, 126(5), 451--462.
Downloaded by [65.209.12.68] at 12:29 22 March 2016

Matsuo, O. and Tsutsumi, T., Evaluation of cone penetration testing as in in-situ Sandven, R., Senneset, K., and Janbu, N., Interpretation of piezocone tests in
liquefaction resistance measurement. In Geotechnical Site cohesive soils. In Penetration Testing 1988, Vol. 2, pp. 939--953, 1988
Characterization, Vol. 2, pp. 1309--1315, 1998 (Balkema, Rotterdam). (Balkema, Rotterdam).
Mayne, P. W., Determination of OCR in clays by piezocone tests using Santamarina, J. C. and Cho, G. C., Soil behavior, the role of particle shape. In
cavity expansion and critical state concepts. Soils Found., 1991, Advances in Geotechnical Engineering: The Skempton Conference, Vol. 1,
31(1), 65--76. pp. 604--617, 2004 (Thomas Telford, London).
Mayne, P. W., Stress--strain--strength--flow parameters from enhanced in-situ Schmertmann, J. H., Measurement of in-situ shear strength. In In-Situ
tests. In Proceedings of the International Conference on In-Situ Measurement of Soil Properties, Vol. 2, pp. 57--138, 1975, (ASCE,
Measurement of Soil Properties and Case Histories, 2001, pp. 27--48. Reston, VA).
Mayne, P. W., Integrated ground behavior, in-situ and lab tests. In Deformation Schmertmann, J. H., Guidelines for Cone Penetration Test, Performance and
Characteristics of Geomaterials, Vol. 2, pp. 155--177, 2005 (Taylor & Design. Report FHWA-TS-78-209, Federal Highway Administration,
Francis, London). Washington, DC, 1978.
Mayne, P. W. and Campanella, R. G., Versatile site characterization by seismic Senneset, K., Sandven, R., Lunne, T., By, T., and Amundsen, T., Piezocone tests
piezocone. In Proceedings of the 16th International Conference on Soil in silty soils. In Penetration Testing 1988, Vol. 2, pp. 955--974, 1988
Mechanics and Geotechnical Engineering, Vol. 2, pp. 721--724, 2005 (Balkema, Rotterdam).
(Millpress, Rotterdam). Senneset, K., Sandven, R., and Janbu, N., Evaluation of soil parameters from
Mayne, P. W. and Kulhawy, F. H., K0--OCR relationships in soil. J. Geotech. piezocone tests. In Transportation Research Record 1235 (In-Situ Testing
Eng., 1982, 108(GT6), 851--872. of Soil Properties for Transportation), pp. 24--37, 1989 (National
Mayne, P. W., Mitchell, J. K., Auxt, J. A., and Yilmaz, R. US National Report Academy Press, Washington, DC).
on CPT. In Proceedings of the International Symposium on Cone Skempton, A. W., Standard penetration test procedures and the effects in sands
Penetration Testing, Vol. 1, Swedish Geotechnical Society Report 3,95, of overburden stress, relative density, particle size, ageing, and overcon-
1995, pp. 263--276, (Swedish Geotechnical Society, Linköping). solidation. Geotechnique, 1986, 36(3), 425--447.
Mayne, P. W., Schneider, J. A. and Martin, G. K., Small- and large-strain soil Stroud, M. A., The standard penetration test, its application and interpretation. In
properties from seismic flat dilatometer tests. In Pre-Failure Deformation Penetration Testing in the UK, pp. 29--49, 1988 ((Thomas Telford, London).
Characteristics of Geomaterials, Vol. 1, pp. 419--426, 1999 (Balkema, Terzaghi, K., Theoretical Soil Mechanics, 1943 (Chapman & Hall, London;
Rotterdam). John Wiley, New York).
Meyerhof, G. G., The ultimate bearing capacity of foundations. Geotechnique, Vaid, Y. P. and Sivathayalan, S., Fundamental factors affecting liquefaction
1951, 2(1), 301--332. susceptibility of sands. Can. Geotech. J., 2000, 37(3), 592--606.
Meyerhof, G. G., Bearing capacity and settlement of pile foundations. Vesić, A. S., Expansion of cavities in infinite soil mass. J. Soil Mech. Found.
J. Geotech. Eng. Div.--ASCE, 1976, 102(GT3), 197--228. Div.--ASCE, 1972, 98(SM3), 265--290.
Mimura, M., Characteristics of some Japanese natural sands: data from undis- Vesić, A. S., Design of Pile Foundations. Synthesis of Highway Practice,
turbed frozen samples. In Characterisation and Engineering Properties of NCHRP 42, Transportation Research Board, Washington, DC, 1977.
Natural Soils, Vol. 2, pp. 1149--1168, 2003 (Swets & Zeitlinger, Lisse). Wride, C. E. and Robertson, P. K., CANLEX: The Canadian Liquefaction
Mitchell, J. K., New developments in penetration tests and equipment. Experiment. Data Review Report, 1999 (BiTech, Richmond, BC) (five
Penetration Testing 1988, Vol. 1 pp. 245--261, 1988 (Balkema, Rotterdam). volumes).
Mitchell, J. K. and Gardner, W.S., In-situ measurement of volume change Wride, C. E., Robertson, P. K., Biggar, K. W., et al., Interpretation of in-
characteristics. In In-Situ Measurement of Soil Properties, Vol. 2, situ test results from the CANLEX sites. Can. Geotech. J., 2000,
pp. 279--345, 1975 (ASCE, Reston, VA). 37(3), 505--529.
Mitchell, J. K. and Keaveny, J. M., Determining sand strength by cone penet- Wroth, C. P., Penetration testing, a more rigorous approach to interpretation. In
rometer. In Use of In-Situ Tests in Geotechnical Engineering (GSP 6), Penetration Testing 1988, Vol. 1, pp. 303--311, 1988 (Balkema,
pp. 823--839, 1986 (ASCE, Reston, VA). Rotterdam).
Mitchell, J. K. and Lunne, T., Cone resistance as a measure of sand strength. Yamashita, S., Hori, T., and Suzuki, T., Effects of fabric anisotropy and stress
J. Geotech. Eng., 1978, 104(GT7), 995--1012. condition on small-strain stiffness of sands. In Deformation
Mulilis, J. P., Seed, H. B., Chan, C. K., Mitchell, J. K., and Arulanandan, K. Characteristics of Geomaterials, Vol. 1, pp. 187--194, 2003 (Swets &
Effects of sample preparation on sand liquefaction. J. Geotech. Eng., 1977, Zeitlinger, Lisse).
103(GT2), 91--108. Yu, H.-S., James K. Mitchell Lecture. In-situ soil testing, from mechanics to
Olsen, R. S. and Mitchell, J. K., CPT stress normalization and prediction of soil interpretation. In Geotechnical and Geophysical Site Characterization,
classification. In Proceedings of the International Symposium on Cone Vol. 1, pp. 3--38, 2004 (Millpress, Rotterdam).
Penetration Testing, Vol. 2, Swedish Geotechnical Society Report 3,95, Yu, H.-S. and Mitchell, J. K., Analysis of cone resistance, review of methods.
1995, pp. 257--262 (Swedish Geotechnical Society, Linköping). J. Geotech. Geoenviron. Eng., 1998, 124(2), 140--149.

Anda mungkin juga menyukai