Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/254549213

Rate Effects in Vane Shear Testing

Article

CITATIONS READS

13 68

2 authors, including:

Paul W. Mayne
Georgia Institute of Technology
192 PUBLICATIONS 2,471 CITATIONS

SEE PROFILE

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

Centrifuge modelling of cyclic loaded monopile foundation for offshore windfarms View project

All content following this page was uploaded by Paul W. Mayne on 09 May 2016.

The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document
and are linked to publications on ResearchGate, letting you access and read them immediately.
Proceedings of the 6th International Offshore Site Investigation and Geotechnics Conference:
Confronting New Challenges and Sharing Knowledge, 11–13 September 2007, London, UK

RATE EFFECTS IN VANE SHEAR TESTING


Joek Peuchen
Fugro Offshore Geotechnics, Leidschendam, The Netherlands
Paul Mayne
Georgia Institute of Technology, Atlanta, USA

Abstract
This paper presents recommendations for a conformity assessment for vane shear tests (VST) at rotation
rates faster than standardised by ASTM and equivalent societies/institutes. The proposed rationale is
based on compensating the standard requirements by recording downhole torque and rotation versus
time. Such supplementary measurements are commonly available when using advanced onshore and
offshore VST systems. These systems allow a good check on use of ‘time to failure’ for correcting a
faster rotation rate to the standard rate. An obvious benefit is a reduced overall field testing time.

1. Introduction or ball penetration tests. This paper presents recommen-


The vane shear test (VST), or field vane test, is one of sev- dations for a conformity assessment4 for VST at rotation
eral in situ test methods commonly used to determine und- rates faster than the standards of ASTM International and
rained shear strength and strength sensitivity in water-satu- equivalent societies/institutes5, 6, 7. The proposed rationale
rated clay and silt soils. It has a long history of use1, 2. Figure is based on compensating the standard requirements by re-
1 presents basic test principles. cording downhole torque and rotation versus time. Such
supplementary measurements are commonly available
A perceived advantage of the VST is its theoretical model for when using advanced onshore and offshore VST systems.
data interpretation, i.e. lower bound limit analysis for well-
defined failure planes. Basic output consists of undrained All current VST standards fall in the group ‘method
shear strength, suv, for undisturbed conditions. Undrained specifications’, i.e. if the method complies, then some
shear strength for remoulded (or ‘residual’), sur, conditions undefined measurement accuracy may be presumed. In
can also be obtained and, hence, the strength sensitivity is comparison, the more recent CPT standards provide for
defined by St = suv/sur. Empiricism is still required to link ‘performance specifications’, i.e. requirements for accu-
this output with the parameter values required for common racy of measurement.
geotechnical calculation models. These models generally
include simplifications so that factors such as strength ani- 2. Advanced VST Systems
sotropy and time-dependency are accounted for only in an Advanced VST systems emerged in the time of rapid
approximate manner. growth of offshore oil and gas developments8. They are now
in common use, mainly for offshore site characterisation
In geotechnical practice, the VST method competes pri- (www.fugro.com) but also for advanced onshore applica-
marily with continuous in situ penetration test methods. tions (www.envi.se; www.geotech.se). As for conventional
The cone penetration test (CPT) is the most widely used VST systems, advanced systems allow tests to be conducted
method. The use of related continuous full-flow penetra- beneath the terminus of an open borehole during rotary
tion tests such as T-bar penetration and ball penetration drilling operations, or alternatively, by using direct-push
tests is increasing3. technology where the vane is deployed at the front end of
The VST is a slow test in comparison with the CPT, T-bar pushrods with or without a housing chamber.

259
Peuchen and Mayne. Rate Effects in Vane Shear Testing

of continuous recording of torque and time measurements


for a full 360° cycle of an in-place strength sensitivity, St =
2.2. The test included the determination of peak and re-
moulded strength data, from 0° to 90° at 0.2°/s; remould-
ing over 180° at 0.6°/s; and the remoulded/residual strength
phase over a final 90° at 0.2°/s. Thus, the full VST required
450s + 300s + 450s = 1200s total (20min). Nonetheless,
this shortened procedure is quite long for an in situ test by
today’s standards.
It can be noted that variation in applied rotation rate, as
shown in Figure 2, can provide additional valuable infor-
mation on soil response to rotation of a vane blade.

3. Vane Shear Strength


Figure 1: Procedures for field vane testing during boring 3.1 Use and limitations
advancement At first glance, the vane appears to be a simple device for
The advanced VST systems surpass the measuring require- the direct in situ measurement of undrained shear strength
ments set by ASTM International5, British Standards in clay soils. Yet, over more than a half-century of research
Institute (BSI)6 and the International Organization for and utilisation in geotechnical practice, vane testing has
Standardization (ISO)7. Particularly, torque is measured been shown to reflect several complex facets of soil behav-
downhole rather than at ground surface. In addition, iour9, 10, 11. Only the peak torque and blade geometry are
torque and time are recorded continuously, typically at a used to calculate suv, despite the known effects of other
rate of 1Hz, rather than recording only a peak torque value factors, such as disturbance during insertion1, 12; waiting
and, if specified, a remoulded torque value. time after insertion13, 14; induced pore water pressures15, 16;
vane shape17, 18; vane size19, 20; blade thickness13; progressive
Use of a downhole torque sensor has obvious advantages in failure21; coefficient of consolidation22, 23; rate of shear24, 25;
reducing uncertainties for rod-friction correction. Build-up organic content of the soil26; strength anisotropy17, 27; and
of rod friction can be significant for conventional VST sys- other influences9.
tems. It should be noted that most advanced VST systems
have a torque sensor a short distance, but not directly, above Nevertheless, despite these complications, the vane shear
the vane blade. This means that some accounting for rod test provides a convenient and historic index with which
friction may be necessary. to reference other strength values obtained during geo-
technical site characterisation. The suv profile can serve as
Most advanced VST systems derive rotation information a benchmark for comparing undrained shear strengths de-
from calibrated rotation rates and time measurement. The rived from in situ penetration tests, as well as laboratory
motor driving the rotation provides a near-constant rotation triaxial and simple shear test results.
rate, independent of actual resistance. Some systems provide
additional point checks on rotation, while others provide full In limit equilibrium analyses, the measured torque, T T, is
rotation measurement. Whatever the means of acquiring ro- used to calculate the undrained shear strength that acts
tation information, all systems are affected by small resisting along a cylindrical body of soil defined by the vane geom-
rotation of the system itself as torque is applied to the vane etry (see Figure 1). The maximum torque reading is used to
blade. Such resisting rotation is derived from a rod and/or determine the peak undrained shear strength, suv, while the
reaction blade in the soil. Accurate correction for resisting residual torque reading applies to the remoulded strength,
rotation is difficult and not normally done. The rotation sur. In some research studies, attempts have been made to
information is thus tool- and soil-dependent and should be discern strength anisotropy using parallel series of vanes of
regarded as indicative. It should not be routinely used for
interpretation of accurate soil stiffness characteristics. Figure 2: Vane torque-rotation curve

Energy for applying torque is at a premium for slimline


downhole vane systems. This is particularly true for bat-
tery-operated systems. A faster rotation rate requires a
larger, more powerful motor. The rotation rate is most
significant in determining undrained shear strength for
remoulded conditions. Here, ASTM specifies a preceding
phase consisting of rapid rotation for 5 to 10 revolutions,
but this would be time-consuming. In common practice,
remoulded undrained shear strength is instead determined
from torque versus rotation data showing constant shear
stress with increasing rotation. Figure 2 shows an example

260
Proceedings of the 6th International Offshore Site Investigation and Geotechnics Conference:
Confronting New Challenges and Sharing Knowledge, 11–13 September 2007, London, UK

Figure 5: Vane shear stress–rotation curves at different loading


Figure 3: Vane correction factor for stability calculations9 rates for Backebol clay14

differing geometries to determine strengths along different 4. Strain Rate Effects


directions17, 18, 27. It is widely accepted that the undrained strength of clay
3.2 Correction factors increases with rate of loading. To illustrate this point, a se-
ries of isotropically consolidated undrained compression
The history of vane use in stability evaluations of embank-
(CIUC) triaxial data conducted at different rates of strain
ments, footings and excavations has indicated the need for
for Haney clay40 are shown in Figure 4. Interestingly, the
(a) correction factor(s)21, 27, 28, 29, 30, 31, 32, for clays located
peak axial strain value, εf, is rather constant and appears to
primarily in Scandinavia and Canada. The correction takes
be independent of strain rate.
the form
(1) Vane tests are conducted at rates of strain much greater
than standard laboratory reference tests. A conventional
where τmax is the mobilised shear strength for the stability isotropically consolidated triaxial compression test with
calculations and the correction factor, µv, has been empiri- pore water pressure measurements is run at 1%/hr. If the
cally related to either plasticity index21, liquid limit31 or the axial strain at failure is 2%, the total time to reach peak
normalised strength28. Figure 3 depicts a relationship de- shear strength is thus on the order of 2hr (120min). In con-
tailed by Chandler9. trast, a field vane test conducted at a rate of 0.1°/s (6°/min)
A correction of the vane strength according to Bjerrum21 will reach peak strength in approximately 1 to 2min, or at
was justified on considerations of three factors: (i) strain rate least 1 order of magnitude faster than a companion triaxial
effects; (ii) strength anisotropy; and (iii) progressive failure. shear test.
Offshore, a vane correction factor is not normally adopted
An illustrative example of vane strain rate effects is shown
for shear strength evaluations3, 33. In Japan, the conven-
in Figure 5 for Backebol clay14. As the definition of strain is
tional correction of vane strengths based on plasticity index
ambiguous in VST, the rate of the testing is reported as time
is apparently too conservative, as discussed by Hanzawa34,
to failure, tf. Over the tested range of tf from 10 000min
Ohta35 and Tanaka36. Some uncertainty also appears to exist
to 1s, the vane shear strength can be seen to increase from
in the use of correction factors in some Canadian clays32.
13 to 28kPa, or over a factor of 2. As with the aforemen-
The discrepancies noted above in the use of measured and/ tioned triaxial results, the peak value appears at the same
or corrected vane strengths have led some to re-analyse the equivalent ‘failure strain’, here the angle of rotation, θ. For
stress states around the vane and formulate new methods Backebol clay, the rotation angle at failure averages about θf
for its interpretation35, 37, 38, 39. = 4°. An equivalent rotation rate, dθd /dt
/dt, for these tests can
be defined as
Figure 4: Undrained stress-strain response for Haney clay at
different strain rates40 (2)

Similarly, the time to failure for the triaxial tests can be de-
termined by dividing the strain at failure by the applied rate
of strain, so that
(3)

One convenience in the use of time to failure, tf, to describe


rate effects is that the concept can be applied to a variety
of laboratory and field tests, as well as full-scale loading of
foundations and earthworks. For example, laboratory triaxial
strength data for three clays, where loading rates were varied,
were reviewed by Lacasse41 and are shown in terms of time

261
Peuchen and Mayne. Rate Effects in Vane Shear Testing

Figure 6: Rate effects in triaxial tests on Figure 7: Rate effects for in situ cone and plate load tests in
three clays in terms of time to failure41 terms of time to failure at Cowden clay site, UK42

to failure in Figure 6. The relative undrained shear strengths ticity index (see Figure 10).
were normalised to a common reference su value correspond- Contrary to the arguments made by Perlow and Richards20
ing to a tf = 140min. For each log cycle of time, there was a for use of peripheral velocity, Randolph3 evaluated the thin
relative change in shear strength on the order of 10 to 15%. shear zone at the blade edges and concluded that ‘it is the ro-
Time to failure also allows a cross-comparison of undrained tation rate of the vane, rather than the peripheral velocity, that
shear strengths obtained from different test types, as illustrat- controls the rate dependency of the torque’. Furthermore,
ed by Figure 7, where su values from a variety of in situ tests there is an apparent conflict in Perlow and Richards’ logic
are shown together, including cone penetrometer, screwplate in that they recommended that the testing rate be scaled in
and bearing load tests at the Cowden site in the UK42. proportion to vane diameter, dθ d /dt ∝ dd, but the soil affect-
Despite the simplicity of time to failure in representing ed by consideration of consolidation is proportional to the
rate effects among different test types, the value of su is also square of the vane diameter, or dθ d /dt ∝ d2 (see Chandler9
dependent upon the direction of loading, boundary con- and Mahmoud . The available results in Figure 10 show
11)

ditions and method of theoretical interpretation (i.e. limit the marked data points from Perlow and Richards20 for San
equilibrium, cavity expansion theory, plasticity theorems). Diego silt (α = 36; β = 0.13) and Gulf of Maine clay (α = 60;
Thus, a more fundamental examination is to look at one β = 0.20), both of which have the highest rate parameters of
test type at variable strain rates. any of the considered soils. Are the soils indeed unusual, or
was the Perlow and Richards’ approach flawed?
4.1 Vane shear database
To assist in the evaluation of strain rate effects on the meas- Two approaches for representing viscous rate effects on the
ured vane strengths, a database of field and laboratory undrained vane strength of fine-grained materials will be
vane data has been compiled to review the rotation rates, adopted here (semi-logarithmic and log-log), as given in
and times to failure, tf. Alternatively, the vane Table 1 for rotation rate and time to failure. The expressions
data can be processed in terms of a peripheral velocity, vp, for the latter are approximate as assumed by a constant level
as suggested by Perlow and Richards20 of strain at failure, or corresponding rotation angle at fail-
ure (as discussed previously with equations 2 and 3).
(4)
The database considers a total of 27 field vane series and
where d is the blade diameter. The reasoning here is that in seven laboratory miniature vane studies44. In most in-
order to compare miniature laboratory and field vane data, stances, complete specifications of the vane equipment (e.g.
faster rotation rates are required for the smaller blades to
achieve the edge velocities of the larger field vane. Biscontin
Table 1:
and Pestana24, 43 evaluated the effect of vp on the vane
Formats used to represent viscous rate effects in vane shear data
strength of a bentonitic kaolin mix under laboratory con-
ditions, which showed general agreement with field vane Rate Measurement Semi-Logarithmic Power Law
test data (Figure 8). Both the log strain and power function Form Format
formats were used to fit the data (Table 1). For the latter, Rotation Rate:
a range of rate exponent values 0.05 ≤ β ≤ 0.10 generally
bounded the data considered. The log parameter, α, and
power function exponent, β, can be interrelated, as seen in Time to Failure: tf
Figure 9. There is, however, no discernable trend between
either of these rate parameters and the corresponding plas-

262
Proceedings of the 6th International Offshore Site Investigation and Geotechnics Conference:
Confronting New Challenges and Sharing Knowledge, 11–13 September 2007, London, UK

Figure 11: Normalised vane strengths versus time to failure


Figure 8: Normalised vane shear strengths as a function of pe- and fitted semi-log
ripheral velocity24, 43
and other important characteristics were missing. Therefore,
only a general trend of the effects of rate on the measured
vane strengths was viable. Further correlation of the rate
parameters with clay indices could not reliably be done.
The normalised vane strengths versus time to failure, tf, are
presented in Figure 11 for 23 different clays. The full range
of times spans over 7 orders of magnitude, from 0.0005
to 10 000min. Of interest is the slope of the relationship
between normalised vane shear strength and time to failure
for assessing the relative rate effect on measured suv. At a
consistent reference for time to failure of tf0 = 1min, most
data appear represented by a log rate parameter between
0.10 ≤ α ≤ 0.14 (or 10 to 14%). However, at very high rates
of testing where tf < 0.1min, a marked increase in strength
Figure 9: Interrelationship between log rate parameter, αR, increase is evident with each log cycle of time. At very long
and power function exponent rate parameter, β, from vane times to failure, (e.g. tf > 100min), it is unclear whether
data (from Biscontin and Pestana43in Table 1) full or partial drainage effects have occurred and how these
influence the conclusions derived from this simplistic look
at vane rate effects.
rectangular versus tapered); manufacturer type (e.g. Geonor
versus Nilcon versus Wykeham Farrance); vane size, d, d H/d;
H/d Figure 12 presents data on vane shear strength in terms of
blade thickness, e, rod friction; and other information was rate of rotation from laboratory and field data on ten clays.
not provided by the data sources. Also, procedural differences Here, over 9 orders of magnitude are subtended in the full
such as insertion waiting period, rate of rotation, revolutions rotation rates applied. At low rotation rates, <
before remoulded readings, borehole size and field/labora- 1°/min, it is unclear whether these data are affected by par-
tory conditions were not reported. Likewise, full details on tial consolidation effects and/or waiting times applied after
the clay index properties including water content, plasticity, vane insertion. Nevertheless, using a power law format, it is
over-consolidation ratio, mineralogy, degree of cementation, shown that the variation in normalised strengths is reason-
clay sensitivity from either fall cone or vane measurements able with a rate exponent 0.05 ≤ β ≤ 0.10 for most clays.

Figure 10:
Vane vp rate
parameters versus
clay plasticity
index (data from
Biscontin and
Pestana24 in
Table 1)

263
Peuchen and Mayne. Rate Effects in Vane Shear Testing

velop. The tendency for partial (or full) drainage depends


upon the coefficient of consolidation, cvh, of the material
(5)
where k = hydraulic conductivity (coefficient of permeabil-
ity), D’ = constrained modulus and γw = unit weight of wa-
ter. In laboratory tests such as the triaxial apparatus, plane
strain device, hollow cylinder and torsional shear, the drain-
age paths can be physically closed to ensure that undrained
conditions prevail. In field tests, however, a truly undrained
condition is not only uncontrollable but also difficult to
Figure 12: Normalised vane strengths versus rotation rate assess. By testing at fast rates, the likelihood of undrained
conditions prevailing in the field is increased, with added
rheological effect due to rate-dependency effects.
Selected results from field and laboratory vane tests where
partial drainage effects occur at slow loading rates are pre-
sented in Figure 1345. The reference rate for field vane tests is
0.1°/s (6°/min), whereas for laboratory miniature vanes the
reference rate is 1°/s (60°/min). At very low rates of shearing
(e.g. dθ
d /dt < 0.1°/s), the normalised vane strength, sv/sv0, in-
creases above the observed minimum as the effects of partial
drainage approach the drained strength of the soil. A nor-
malised time factor, T* T*, was proposed by Blight22 to guide
the calculation of the minimum rate of shearing necessary
to ensure undrained conditions. This approach was re-exam-
Figure 13: Effects of partial drainage on measured vane
ined by Morris and Williams23 to improve the accuracy of the
strengths for selected clays (field data from Roy and Leblanc13
evaluations. Randolph3 devised a unified formula that spans
and laboratory data from Davies et al.45
drained to partially drained to undrained response of in situ
tests (cone, T-bar, vane) over different rates of loading.
4.2 Partial drainage considerations
For field loading conducted at very ‘fast’ shearing rates, 5. Conclusions and Recommendations
constant volume is maintained, ∆V/V V/V0 = 0, such that the
Advanced VST systems surpass the measuring requirements
failure state corresponds to the mobilisation of the und-
set by ASTM and others because
rained shear strength of the soil. However, for field tests at
‘medium’ to ‘slow’ rates of loading, it must be considered • Torque is measured downhole rather than at ground
that some amount of drainage could occur with a corre- surface
sponding dissipation of excess pore water pressures. In fact, • Torque and time are recorded continuously, typically at
if the loading is applied slowly enough, a drained strength a rate of 1Hz, rather than recording only a peak torque
can be measured17 since no excess pore water pressures de- value and, if specified, a remoulded torque value.

Table 2: Recommendations for advanced VST systems

No Recommendations
1 Quick test ( < 2min) for peak undrained shear strength: (1) Push vane to required depth; (2) Wait for 1min; (3) Rotate at maxi-
a b c
mum speed (<2o/s) to just beyond peak shear stress – if offline then to about 25° rotation ; (4) End of test; (5) Apply fixed (or
stratum specific) correction for rotation rate.
2 Rapid test (< 5min) for peak undrained shear strength: (1) Push vane to required depth; (2) Wait for 1min (3) Rotate at maxi-
mum speed (< 2°/s) to well beyond peak shear stress (e.g. 45°); (4) Reduce rotation rate to 0.1 or 0.2°/s for about 1min (5)
Increase rotation rate to < 2°/s for about 1min (6) End of test; (7) Apply test-specific correction for rotation rate.
3 Rapid test (< 10min) for peak and remoulded undrained shear strength: (1) Push vane to required depth; (2) Wait for 1min; (3)
Rotate at maximum speed (< 2°/s) to well beyond peak shear stress (e.g. 45°); (4) Reduce rotation rate to 0.1 or 0.2°/s for about
1min; (5) Increase rotation rate to < 2°/s until no further drop in shear stress is observed, with a maximum of 360°; (6) Reduce
rotation rate to 0.1 or 0.2o/s for about 1min; (7) End of test; (8) Apply test-specific correction(s) for rotation rate and estimate (if
necessary) large-strain shear strength.
a The limit of < 2°/s restricts the departure from the standard rate and hence the magnitude of a required correction factor for strain rate
b The term ‘offline’ refers to VST systems where data read-out and checking takes place after recovery of the tool from the soil
c A value of 25° rotation should adequately capture peak undrained shear strength
d The variation in strain rate within a test allows a test-specific check on the correction factor for strain rate<query, is d placement right?

not placed in table in paper>

264
Proceedings of the 6th International Offshore Site Investigation and Geotechnics Conference:
Confronting New Challenges and Sharing Knowledge, 11–13 September 2007, London, UK

For conformity assessment, the additional features can 15. Kimura T and Saitoh K. (1983). Effect of disturbance due to in-
shorten testing times while maintaining accuracy, as in- sertion on vane shear strength of normally consolidated cohesive
soils. Soils and Foundations 23(2), 113–124.
ferred from ASTM or equivalent standards. In particular,
viscous rate effects in clays may be adequately described by 16. Matsui T and Abe N. (1981). Shear mechanisms of vane test in
soft clays. Soils and Foundations 21(4), 69–80.
either a semi-log function using a rate parameter of αR =
0.13 ± 0.03, or alternatively by using a power law format 17. Aas G. (1965). A study of the effect of vane shape and rate of
strain on the measured values ofofin situ shear strength of clays.
with a rate exponent parameter β = 0.075 ± 0.025. This ap- Proc. 6th Int. Conf. on Soil Mech. and Foundation Engng., Vol. 1,
plies for test rates close to standard rates of testing of dθ
d /dt Montreal, 141–145.
Montreal
= 6°/min, or tf = 1min. 18. Donald IB, Jordan DO, Parker RJ and Toh CT. (1977). The vane
Table 2 presents recommendations for shortening testing for test: a critical appraisal. Proc. 9th Int. Conf. on Soil Mech. and
Foundation Engng, Vol. 1, Tokyo, 81–88.
advanced VST systems to 2 to 10min. The recommenda-
tions include options for site-specific checks on rate effects. 19. Åhnberg H, Larsson R and Berglund C. (2004). Influence of vane
size and equipment on the results of field vane tests. In Richards
AF (ed.). Geotechnical and Geophysical Site Characterization, Vol.
References 1. Rotterdam: Millpress 271–277.
1. Cadling L and Odenstad S. (1950). The vane borer: an appa-
ratus for determining the shear strength of clay soils directly in 20. Perlow M and Richards AF. (1977). Influence of shear velocity
the ground. Royal Swedish Geotechnical Institute Proceedings No. 2, on vane shear strength. J. of the Geotech. Engng Div. 103(GT1),
Stockholm, 88p. 19–32.
2. Skempton AW. (1948). Vane tests in the alluvial plain of the River 21. Bjerrum L. (1973). Problems of soil mechanics and construction
Forth near Grangemouth. Géotechnique 1(2), 111–124. on soft clays. Proc. 8th Int. Conf. on Soil Mech. and Foundation
Engng, Vol. 3, Moscow, 111–159.
3. Randolph MF. (2004). Characterisation of soft sediments
for offshore applications. In da Fonseca AV and Mayne, PW 22. Blight GE. (1968). A note on field vane testing of silty soils.
(eds.). Geotechnical and Geophysical Site Characterization, Vol. 1. Canadian Geotech. J. 5(3), 142–149.
Rotterdam: Millpress, 209–232. 23. Morris, P.H. and Williams, D.J. (2000). A revision of Blight’s mod-
4. International Organization for Standardization (ISO)/ el of field vane testing. Canadian Geotech. J. 37(5), 1089–1098.
International Electrotechnical Commission (IEC). (1996). 24. Biscontin G and Pestana JM. (1999). Influence of peripheral
Conformity assessment, ISO/IEC 17000. Guide 2 1996 general velocity on undrained shear strength and deformability char-
terminology. acteristics of a bentonite-kaolinite mixture. Research Report
5. ASTM. (2001). Standard Test Method for Field Vane Shear Test UCB/GT/99-19, Dep. of Civil and Envir. Engng., University of
in Cohesive Soil, ASTM Standard D 2573, 2001. In Annual California, Berkeley.
Book of ASTM Standards. West Conshohocken, PA: ASTM 25. Wiesel CE. (1973). Some factors influencing in situ vane test re-
International. sults. Proc. 8th Int. Conf. on Soil Mech. and Foundation Engng,
6. British Standards Institute (BSI). (1990). British Standard Vol. 1.2, Moscow, 475–479.
Methods of Test for Soils for Civil Engineering Purposes, Part 26. Slunga E and Helander R. (1985). On the influence of organic
9: In situ Tests, Clause 4.4: Determination of in-situ vane shear content on the undrained shear strength of cohesive soils. Proc.
strength of weak intact cohesive soils. in BS 1377. London: BSI. 11th Int. Conf. on Soil Mech. and Foundation Engng, Vol. 4, San
7. International Organization for Standardization (ISO). (2003). Francisco, 2349–2352.
ISO/WD 22476-9, 2003: Geotechnical investigation and testing 27. Silvestri V, Aubertin M and Chapuis RP. (1993). A study of un-
– Field testing – Part 9: Field Vane Test. Geneva: ISO. drained shear strength using various vanes. Geotech. Testing J.
8. Geise JM, Hoope J and May RE. (1988). Design and offshore 16(2), 228–237.
experience with an in situ vane. In Richards AF (ed.). Vane Shear 28. Aas G, Lacasse S, Lunne T and Høeg K. (1986). Use of in situ
Strength Testing in Soils: Field and Laboratory Studies, ASTM tests for foundation design on clay. In Clemence, SP (ed.). Use
STP 1014. West Conshohocken/PA: ASTM, 318–336. of In Situ Tests in Geotechnical Engineering
Engineering, GSP 6. Reston, VA:
9. Chandler RJ. (1988). The in situ measurement of the undrained ASCE, 1–30.
shear strength of clays using the field vane. In Richards AF (ed.). 29. Azzouz AS, Baligh MM and Ladd CC. (1983). Corrected field
Vane Shear Strength Testing in Soils: Field and Laboratory Studies, vane strength for embankment design. J. of Geotech. Engng.
ASTM STP 1014. West Conshohocken, PA: ASTM, 13–44. 109(5), 730–734.
10. Flaate K. (1966). Factors influencing the results of vane tests. 30. Helenelund KV. (1977). Methods for reducing undrained shear
Canadian Geotech. J. 3(1), 18–31. strength of soft clay. Swedish Geotechnical Institute Report No. 3,
11. Mahmoud M. (1988). Vane testing in soft clays. Ground Linköping, 59p.
Linköping
Engineering 21(7), 36–40. 31. Larsson R. (1980). Undrained shear strength in stability calcula-
12. Cerato AB and Lutenegger AJ. (2004). Disturbance effects of tion of embankments and foundations on soft clays. Canadian
field vane tests in a varved clay. In da Fonseca AV and Mayne, Geotech. J. 17(4), 591–602.
PW (eds.). Geotechnical and Geophysical Site Characterization, Vol. 32. Lefebvre G, Ladd CC and Paré JJ. (1988). Comparison of field
1. Rotterdam: Millpress, 861–867. vane and laboratory undrained shear strength in soft sensitive
13. Roy M and Leblanc A. (1988). Factors affecting the measurements clays. In Richards AF (ed.). Vane Shear Strength Testing in Soils:
and interpretation of the vane strength in soft sensitive clays. In Field and Laboratory Studies, STP 1014. West Conshohocken,
Richards AF (ed.). Vane Shear Strength Testing in Soils: Field and PA: ASTM, 233–246.
Laboratory Studies, ASTM STP 1014. West Conshohocken, PA: 33. Kolk, H.J., ten Hoope, J. and Ims, B.W. (1988). Evaluation of
ASTM, 117–128. offshore in situ vane test results. In Richards AF (ed.). Vane Shear
14. Torstensson BA. (1977). Time-dependent effects in the field vane Strength Testing in Soils: Field and Laboratory Studies, STP 1014.
test. Int. Symp. on Soft Clay, Bangkok, 387–397. West Conshohocken, PA: ASTM, 339–353.

265
Peuchen and Mayne. Rate Effects in Vane Shear Testing

34. Hanzawa H. (1991). A new approach to determine the shear of undisturbed clay. J. of the Geotech. Engng Div. 103(GT7),
strength of soft clay. Proc. GeoCoast ’91, Vol. 1. Yokohama: Port 693–709.
and Harbour Research Institute, 23–28. 41. Lacasse S. (1979). Safety level of gravity platforms: effect of load
35. Ohta H. (1991). Role of constitutive model in estimating design duration on undrained behaviour of clay and sand. Internal
strength. Proc., GeoCoast ’91, Vol. 2. Yokohama: Ports and Harbor Report No. 40007-1, Oslo: Norwegian Geotechnical Institute
Research Institute, 1001–1002. (NGI), 121p.
36. Tanaka H. (1995). Applicability of Bjerrum’s correction factor 42. Powell JJM and Quarterman RST. (1988). The interpretation of
to Japanese clays. Proc. 11th European Conf. on Soil Mech. and cone penetration tests in clays, with particular reference to rate effects.
Foundation Engng, Vol. 1. Copenhagen: Danish Geotechnical Penetration Testing 1988, Vol. 2. Rotterdam: Balkema, 903–909.
Society dgf-Bulletin 11, 285–290. 43. Biscontin G and Pestana JM. (2001). Influence of peripheral ve-
37. Morris PH and Williams DJ. (1993). A new model of vane shear locity on vane shear strength of an artificial clay. Geotech. Testing
strength testing in soils. Géotechnique 43(3), 489–500. J. 24(4), 423–429.
38. Morris PH and Williams DJ. (1994). Effective stress vane shear 44. Mayne PW. (2004). Report on rate effects in vane shear testing.
strength correction factor correlations. Canadian Geotech. J. Report DTF No. 04-0170; Project R-352, Fugro Engineers B.V.,
31(3), 335–342. Leidschendam, The Netherlands, 56p.
39. Wroth CP. (1984). The interpretation of in situ soil tests. 45. Davies MCR, Almeida MSS and Parry RHG. (1989). Studies
Géotechnique 34(4), 449–489. with centrifuge vane and penetrometer apparatus in a normal
40. Vaid YP and Campanella RG. (1977). Time-dependent behavior gravity field. Geotech. Testing J. 12(3), 195–203.

266
View publication stats

You might also like