You are on page 1of 54

27

Experimental Evidence on the Strength of Soil Treated with


Single and Double Fluid Jet Grouting

Authors:
Wanik, Lidia
Mascolo, Maria Cristina
Bzowka, Joanna
Modoni, Giuseppe
Shen, Shui Long
Presentation Preference: Oral
Track: Track 1/ Grouting Case Histories
Session: 1A20/ Jet Grouting: Verification and Testing
Is this a student presentation?: None Specified
Bio: None Specified
Submission type: Technical Papers (6-10 pages in length)
Date Uploaded: 1/5/2017 12:26:40 PM
____________________________________________________________

Project Manager Submission separator pages


Experimental evidence on the strength of soil treated with single and double
fluid jet grouting

Wanik L. Author M.Sc.1, Mascolo M. C. 2 Author Ph.D., Bzwka J. 1, Author Ph.D.,


Modoni G.2 Author Ph.D., Shen J.S.L. Author Ph.D. 3

1
Faculty of Civil Engineering, The Silesian University of Technology, Akademicka 5, 44-100
Gliwice, - Poland; e-mail lidia.wanik@polsl.pl and joanna.bzowka@polsl.pl
2
Department of Civil and Mechanical Engineering, University of Cassino and Southern Lazio,
via di Biasio 43, 03043 Cassino (FR) Italy; e-mail mc.mascolo@unicas.it and
modoni@unicas.it
3
Department of Civil Engineering -School of Naval Architecture, Ocean, and Civil Engineering,
Shanghai Jiao Tong University, ,800 Dong Chuan Road, Shanghai 200240, China; e:mail
slshen@sjtu.edu.cn.

ABSTRACT

The various technological solutions of jet grouting, normally grouped into single, double and
triple fluid, aim principally at increasing the efficiency of treatments obtaining larger diameter of
columns and reduce construction time and cost. Rarely attention is put on the influences of the
technology on the mechanical properties of the material, considering this as a secondary issue.
The present paper focuses on this aspect summarizing the observation of a comprehensive field
trial where sixteen columns have been created with single and double fluid jet grouting varying
the injection parameters. The large number of laboratory tests performed on samples cored from
the columns show a predominant role of the technology, being the material created with single
fluid systematically stronger than the one formed with double fluid. This aspect has been herein
explored on samples of original soil compared with materials cored from single and double fluid
jet columns: the morphology of the samples has been investigated by scanning electron
microscopy (SEM), while the mineralogical composition has been inferred with X-ray diffraction
(XRD), differential thermal (DTA) and thermo-gravimetric analysis (TGA). The analysis shows
that, if on one side the wrapping of the jet with an annulus of compressed air enhances the
cutting capacity of the double fluid jet and is this able to create larger diameters, on the other
side it leads to a lower proportion of cement in the columns, that is responsible for worse
mechanical properties of the material.

INTRODUCTION

The jet grouting technology can be nowadays found in a variety of technological solutions,
grouped for the sake of simplicity as single, double and triple fluid depending on the number of
injected fluids (Croce et al., 2014). It is recalled that much of the effort of producers of

Proceedings Paper Formatting Instructions 1 Rev. 1/2016


equipment is addressed to obtain larger columns, in order to reduce the number of boreholes and
gain efficiency in terms of time and costs. In particular, the coating of the cutting jet with air
typical of double and triple fluid systems (Modoni et al., 2006; Flora et al, 2013, Ochmanski et
al., 2015, Shen et al., 2013, Modoni et al., 2016) allows to reach diameters as large as two and
half meters that are very convenient to speed up the creation of structures like from foundation
reinforcement (Modoni & Bzowka, 2013), impervious cut offs (Croce &Modoni, 2006), bottom
plugs (Modoni et al., 2016; Eramo et al., 2012), tunnels (Ochamanski et al., 2015). However, in
most cases limited attention is paid to the mechanical properties of the jet grouted material. In
particular, jet grouting is often considered as a provisional solution not necessitating of
particularly high properties, and thus the effects of the injection technology are neglected. As a
result, apart from few examples (e.g. Tinoco et al., 2014) strength has been rarely considered as a
topic of interest from the scientific literature. On the contrary, an accurate knowledge of the
mechanical properties of the jet grouted material and of their relation with technological issues
would lead to optimize the use of jet grouting and possibly indicate that it can be used with
permanent function. One of the few examples reported in the literature (Croce et al., 2015) shows
that columns created with double fluid system possess a lower strength of the in comparison with
the single fluid system. Willing to explore this effect, the present paper summarizes the
observation of a comprehensive field trial where sixteen columns have been created with single
and double fluid jet grouting, varying the set of injection parameters and the properties of the
material investigated with laboratory tests.

THE CASE STUDY

The trial field is situated in the municipality of Bojszowy Nowy, in the Upper Silesia (Poland).
The subsoil was investigated by performing two exploratory boreholes and four Piezocone Tests
(CPTU) reaching in all cases the depth of 10 m. From top to bottom (Figure 1.a), the subsoil
presents a typical alluvial sequence with an upper 2.3 m thick layer of medium sand overlapped
to a 2 m thick layer of organic clay and to a subbase layer of coarse sand.
The field trial consists of a total number of 16 jet grouting columns, 8 created with single-fluid
and 8 with double-fluid jet grouting system (indicated as respectively 1S8S and 1D8D in
Figure 1.b). Each column is about 4 m long, ranging from 0.5 m to 4.5 m below the ground level.
After treatment the columns were discovered up to the depth of 1.5 below the ground level and
the diameter was directly measured (see Figure 2.a and Table 1).
The columns were formed varying the set of parameters from one to another in order to see the
different effects of injection. Apart from the differences among the different columns, it is soon
evident that for similar values of the parameters (namely diameter of the nozzle, grout pressure
and flow and monitor lifting speed) the use of double fluid jet grouting leads to larger diameters.
After excavation, samples of material were cored from the top part of columns falling in the
upper medium sand layer (see Figure 1.a) and subjected to uniaxial compression tests in the
laboratory.

Proceedings Paper Formatting Instructions 2 Rev. 1/2016


(b)

(a)

Figure 1. Layout of the field trial at BojszowyNowy: subsoil layering (a); columns
arrangement (b).

Column Nr and Grout Grout Lifting Average Cement Uniaxial


# diameter pressure flow rate speed diameter per unit compressive
[mm] of [bar] [l/min] [m/s] [m] volume strength
nozzles [kg/m3] [MPa]
1S 2x4.0 360 316 0.0083 1.13 474 7,8
2S 2x2.8 360 155 0.0083 0.94 337 4,55
3S 2x4.0 180 223 0.0083 0.90 528 20,8
4S 2x2.8 250 129 0.0083 0.78 407 -
5S 2x4.0 360 316 0.0083 1.13 475 11,6
6S 2x2.8 360 155 0.0083 0.98 310 11,08
7S 2x4.0 180 223 0.0083 0.90 528 11,31
8S 2x2.8 250 129 0.0083 0.86 335 11,07
1D 2x4.0 360 316 0.0083 1.75 198 -
2D 2x4.0 180 223 0.0083 1.61 165 1,6
3D 2x4.0 360 316 0.0083 1.73 203 3,4
4D 2x4.0 180 223 0.0083 1.58 171 3,1
5D 2x2.8 360 155 0.0083 1.54 125 3,3
6D 2x2.8 260 131 0.0083 - - 1,6
7D 2x2.8 360 155 0.0083 1.88 84 5,3
8D 2x2.8 250 129 0.0083 1.63 93 1,3
Table 1. Technological parameters of jet grouting columns performed in BojszowyNowy.

The laboratory tests, performed on samples of single a double fluid jet grouting of different
shape (h/d ratio of 1 and 2) are summarized reporting the statistics of the uniaxial compressive

Proceedings Paper Formatting Instructions 3 Rev. 1/2016


strength in Table 2. To perform a more detailed analysis, the mean values obtained on each
column are reported in Table 1. It is immediate to see that the samples of double fluid jet
grouting possess a systematically much lower compressive strength in comparison with the
sample of columns injected with single fluid jet grouting system, and this effect is independent
on the slenderness ratio (h/d) of the sample.

(a) (b)

Figure 2. Picture of excavated columns (a) and of the cored samples (b).

Sample shape h/d h/d=1.0 h/d= 2.0


Jet grouting system SF DF SF DF
Number of data 22 15 19 8
Mean strength [Mpa] 11.7 3.2 11.8 1.1

Standard deviation [Mpa] 1.6 0.5 1.1 0.1

Table 2. Statistics of the uniaxial compressive strength obtained from the different tests.

Considering that in most cases, jet grouting columns contribute to the stability of the reinforced
structures with their strength, this aspect deserves a deeper investigation. With this aim, the
uniaxial compressive strength obtained on each column has been related to the amount of cement
injected per unit volume of cemented material, this latter proportional to the measured diameter
of the column. Although scattered, the data in Figure 3 clearly explain the lower strength of the
material formed with double fluid as an effect of the lower proportion of cement in the columns.
In fact, it must be considered that, despite similar values of the flow rates and lifting speeds are
adopted for the single and double fluid columns (see Table 1), the double fluid system tends to
spread the injected cement over larger columns, with the results of producing a weaker cemented
material.

Proceedings Paper Formatting Instructions 4 Rev. 1/2016


Figure 3. Uniaxial compressive strength as function of the cement injected in the unit
volume of column.

MINERALOGICAL COMPOSITION

The mineralogical characterization of the original soil and of the material obtained with single
and double jet grouting was performed by means of X-ray diffraction (XRD), differential thermal
analysis (DTA) and thermogravimetric analysis (TGA).
The XRD performed on the natural sand (Figure 4.a) reveals the presence of quartz (SiO2,
JPCDS card no. 46-1045) as a prevalent crystalline phase together with a secondary phase of
dolomite ferroan, Ca(Mg, Fe)(CO3)2, (JPCDS card no. 34-517) and traces of muscovite,
KAl3Si3O10(OH)2, (JPCDS card no. 75-948) and microcline, KAlSi3O8, (JPCDS card no. 19-
932).
The same analysis performed on soil treated with single fluid jet grouting (Figure 4.b) reveals
again the presence of quartz, SiO2, (JPCDS card no. 46-1045) with traces of muscovite
KAl3Si3O10(OH)2, (JPCDS card no. 75-948), dolomite ferroan, Ca(Mg, Fe)(CO3)2, (JPCDS card
no. 34-517) and microcline, KAlSi3O8, (JPCDS card no. 19-932), but also portlandite, Ca(OH)2,
(JPCDS card no. 44-1481), vaterite, CaCO3, (JPCDS card no. 13-192), tetracalcium
monosulfoaluminate hydrate, 3CaO.Al2O3. CaSO4 xH2O, (JPCDS card no. 18-275) and
tetracalcium monocarboaluminate hydrate, 3CaO.Al2O3. CaCO3 xH2O. It is worth noting that the
last two phases belong to AFm phases which are normally generated as hydration products of
Portland cement.

Proceedings Paper Formatting Instructions 5 Rev. 1/2016


These phases are negligible or totally missing in the XRD pattern of soil treated with double
fluid jet grouting (Figure 4.c), that shows the presence of crystalline quartz with traces of
muscovite, microclin and calcite, CaCO3 (JPCDS card no. 5-586).

(a) (b)

(c)

Figure 4. XRD powder pattern of the natural soil (a), of the materiel cored from single fluid
(b) and from double fluid (c) columns (q=quartz; D=dolomite ferroan; M=Microcline;
Mu=muscovite; C=vaterite; Mo=tetracalciummonosulfoaluminate hydrate; P=portlandite;
* = tetracalciummonocarboaluminate hydrate).

The DTA performed on the original soil (Figure 5.a) shows three endothermic peaks at 571.5 ,
760.5 and 867.4 C, the first indicative of the displacive and polymorphic transformation of
quartz, the other two peaks due to the decomposition of dolomite (CaCO3MgCO3) (Mackenzie,
I970). The TGA curve reported in the same plot, shows a small (2 wt %) and continuous weight
loss from room temperature up to 700 C, while the two weight losses in correspondence of the
two endothermic effects are related to the decomposition of MgCO3 and CaCO3 of dolomite. As
expected, no change in the weight loss results for the polymorphic transformation of quartz.
The material treated with single fluid jet grouting shows a larger number of less intense
endothermic peaks at 235 , 501 , 574 , 700 and 811 C (Fig. 5.b). In particular, the effect at
501C, characterized by a meaningful weight loss, has been attributed to the thermal
decomposition of portlandite, Ca(OH)2, while that at 574 C is due to the presence of quartz and
the two peaks at 700 and 811 C outcomes from thermal decomposition of dolomite.
Significant shifts toward lower temperatures are seen for these peaks compared with those of
untreated soil (Figure 5.a). Additionally, the dolomite content is lower compared to that found on
the natural soil, being the weight loss connected to the dolomite decomposition smaller. Finally,

Proceedings Paper Formatting Instructions 6 Rev. 1/2016


the soil treated with single fluid jet grouting reveals a relatively higher weight loss (4 %) at
temperatures lower than 400 C compared to 1 % seen on the untreated soil.
The material coming from double fluid jet grouting shows in DTA two endothermic peaks at
572,5 and 787,1 C (Figure 5.c), the former related to the presence of quartz, whereas the latter
due to the decomposition of dolomite (CaCO3MgCO3). The continuous weight loss observed up
to 400 C is more similar to the natural soil than to the material treated with single fluid.
Comparing the main diffraction peak of quartz for the jet grouted materials (Figure 6), it is seen
that the soil treated with double fluid has a higher content of quartz compared to that of the
single fluid.

(a) (b)

(c)

Figure 5. Differential thermal analysis (DTA) and thermogravimetric analysis (TGA) for
natural soil (a) and for the soil treated with single (b) and double (c) jet grouting.

On the other hand, the typical crystalline phases of the hydration of cement are absent in the
double fluid case, while the presence of AFM phases, portlandite and vaterite in the single fluid
justifies the positive effect of cement. It is well known that the binding effectiveness of the
cement is mainly related to the formation of a calcium silicate hydrate, CSH the principal
binding agent in cement chemistry (Taylor, 1997). The presence of this phase, not easy to be

Proceedings Paper Formatting Instructions 7 Rev. 1/2016


b)

a)

Figure 6. Main diffraction peak of quartz for the soil treated with single (a) and double (b)
jet grouting

distinguished by XRD because of its very low crystallinity (Hewlett, 2003), is confirmed in the
single fluid jet grouted soils by the typical weight loss connected to the dehydration of CSH
phase in the typical temperature range (150-450 C) during the thermal decomposition (Figure
5.b); on the other side this weight loss in much smaller for the soil treated with double fluid
system (Figure 5.b). The tetracalciummonocarboaluminate hydrate in the sample treated with
single fluid is formed at early age of cement hydration in presence of limestone or dolomite,
while the tetracalciummonosulfoaluminate hydrate is a result of the transformation of ettringite.
The conspicuous reduction of the amounts of both quartz and dolomite in the single fluid case
compared with untreated soil confirms the efficacy of jet grouting.

a) b)

Figure 7 SEM micrographs of the samples treated with single (a) or double (b) fluid

A further confirmation of the effectiveness of jet grouting with single fluid compared with that
with double fluid has been also obtained by SEM analysis (Fig. 7). The sample treated with
single fluid (a) presents platelike aggregates of portlandite crystals (Mascolo et al, 2010) together
with presence, on the quartz grains micrometric in size, of very small particles probably due to
CSH phase. Portlandite and CSH phases represent the typical hydration phases of cement which

Proceedings Paper Formatting Instructions 8 Rev. 1/2016


are absent in the micrograph of the sample treated with double fluid (b) prevalently characterized
by micrometric quartz particles.

CONCLUSIVE REMARKS
The results of a field trial have been herein summarily presented to show the effects of jet
grouting carried out with different technologies, alternating single and double fluid with variable
set of injection parameters. As expected, the direct measurement of diameter on the excavated
columns shows that the air used in double fluid jet grouting to shroud the inner jet of grout and
protect it from the energy exchange with the surrounding fluid, has the positive role of increasing
the cutting length of the jet and produce larger columns. This effect has obvious positive
consequences on the productivity of the jet grouting system allowing to cover larger areas or,
conversely to need less columns. On the other hand, the uniaxial compressive strength of the
material shows a meaningful role of the technology, being the resistance of the material formed
with single fluid much higher than that seen on the material formed with double fluid system. A
careful investigation leads to ascribe the reasons of such difference on the different amount of
cement present in the columns. In fact, considering that similar amount of cement are normally
injected per unit length of columns, the larger extension of the double fluid columns tends to
diffuse this binder over larger volumes, with the results that the obtained material is weaker.
The same conclusion comes out also from the chemical analysis of the materials, that shows a
limited presence of the typical crystalline phases of the hydration of cement in the double fluid
columns.

REFERENCES

Croce P., Flora A., Modoni G. (2014). Jet grouting: Technology, Design and Control. Taylor &
Francis Group, ISBN 978-0-415-52640-1.
Croce P., Modoni G. (2007). Design of Jet Grouting Cut-offs. Ground Improvement Thomas
Telford, vol. 11-1, pp. 11-19.
Eramo N., Modoni G., Arroyo Alvarez de Toledo M., (2012), Design control and monitoring of a
jet grouted excavation bottom plug, Proc. of the 7th Int. Symposium on Geotechnical
Aspects of Underground Construction in Soft Ground, TC28 IS Rome, Viggiani ed.,
Taylor & Francis Group London, 16-18 May 2011, pp.611-618.
Flora A., G. Modoni, S. Lirer, Croce P. (2012). The diameter of single, double and triple fluid jet
grouting columns: prediction method and field trial results. Gotechnique, Volume 63,
Issue 11, pp.934-945
Ghosh S. N., Advances in Cement Technology: Chemistry, manufacture and Testing, 2nd Ed.
2002, Tech. Books International, New Delhi, India, pag 687-697.
Hewlett P., Lea's Chemistry of Cement and Concrete, Butterworth-Heinemann, 2003

Proceedings Paper Formatting Instructions 9 Rev. 1/2016


Kakali G., Tsivilis S., Aggeli E. and Bati M., Hydration products of C3A, C3S and portland
cement in the presence of CaCO3. Cement and Concrete Research, vol. 30, Issue 7, July
2000, pages 1073-1077.
Mackenzie R.C. (editor), Differential Thermal Analysis, Vol. I. Academic Press, New York,
I970, pag. 479).
Mascolo G. Mascolo M.C., Vitale A, Marino O, Microstructure evolution of lime putty upon
aging, Journal of Crystal Growth 312 (2010) 23632368
Modoni G., Croce P., Mongiov L. (2006). Theoretical modelling of jet grouting. Gotechnique
Thomas Telford, vol. 56-5, pp. 335-347.
Modoni, G., Bzwka, J., (2012), Analysis of Foundations Reinforced with Jet Grouting, J.
Geotech. Geoenviron. Eng., ASCE, 138(12), 14421454.
Modoni G., Wanik L., Giovinco G., Bzwka J., Leopardi A., 2016, Numerical Analysis of
submerged flows for jet grouting, Proceedings of the ICE, Ground Improvement, Volume
169 Issue 1,pp. 42-53.
Modoni G., Flora A., Lirer S., Ochmanski M., Croce P., (2016) Design of jet grouted excavation
bottom plugs, Journal of Geotechnical and Geoenvironmental Engineering (ASCE),
142(7).
Ochmaski M., Modoni G., Bzwka J. , 2015, Prediction of the diameter of Jet Grouting
columns with Artificial Neural Networks, Soils and Foundations, Vol.55, No.2, pp.425-
436.
Ochmaski M., Modoni G., Bzwka J., (2015), Numerical analysis of tunnelling with jet-grouted
canopy, Soils and Foundations, Volume 55, Issue 5, October 2015, Pp. 929942.
Shen, S., Wang, Z., Yang, J., and Ho, C., (2013). Generalized Approach for Prediction of Jet
Grout Column Diameter. Journal of Geotechnical and Geo-Environmental Engineering,
ASCE, Mar. 22, 2013).
Taylor H.F.W., (1997) Cement Chemistry, Thomas Telford.
Tinoco J., Gomes Correia A., Cortez P., 2014, Support vector machines applied to uniaxial
compressive strength prediction of jet grouting columns, Computers and Geotechnics,
Vol. 55, January 2014, Pages 132140.

Proceedings Paper Formatting Instructions 10 Rev. 1/2016


107
Assessment of Jet Grout Column Diameter during Construction
Using Electrical Resistivity Imaging

Authors:
Mooney, Michael
Bearce, Richard
Presentation Preference: Oral
Track: Track 1/ Grouting Case Histories
Session: 1A20/ Jet Grouting: Verification and Testing
Is this a student presentation?: None Specified
Bio: None Specified
Submission type: Technical Papers (6-10 pages in length)
Date Uploaded: 3/12/2017 7:41:13 PM
____________________________________________________________

Project Manager Submission separator pages


Assessment of Jet Grout Column Diameter during Construction
using Electrical Resistivity Imaging

Michael A. Mooney, Ph.D., M.ASCE1 and Richard G. Bearce, Ph.D.2


1
Center for Underground Construction & Tunneling, Colorado School of Mines, 1500 Illinois
St., Golden, CO 80401; e-mail: mooney@mines.edu
2
United States Bureau of Reclamation, Denver Federal Center, P.O. Box 25007(86-68530),
Denver, CO 80225; e-mail: rbearce@usbr.gov

ABSTRACT

This paper presents the implementation of an electrical resistivity push probe to estimate the
diameter of jet grout columns. The probe is pushed into the jet grout column immediately after
mixing and removed such that no damage is done to the column. The probe allows direct
coupling of electrodes to the soilcrete, providing a demonstrated improvement over existing
techniques. Field testing was conducted on columns with diameters ranging from 0.9 m to 2.5 m
at two sites, in Berlin Germany and Miami Florida. The results show that the technique routinely
estimates the diameter to be within 5% of the as-built diameter.

INTRODUCTION

Jet grouting is an in-situ ground improvement technique used to strengthen weak/unstable


ground into foundation systems and to create hydraulic barriers via columns of soilcrete (i.e., a
mixture of grout and in-situ soil). Successful performance of jet grout columns and column
assemblies requires constructing precise column geometries. The diameter of jet grout columns
is influenced considerably by in-situ soil properties, and therefore, the as-constructed column
geometry has a high degree of uncertainty. On-site inspection of geometry during construction is
therefore critically important. However, the ground improvement industry is lacking in cost
effective techniques to determine jet grout geometry during or immediately after construction.
This paper first presents an overview of techniques that can monitor jet grout diameter,
discussing their capabilities and limitations. The paper then describes a nondestructive probe
technique to determine jet grout column diameter immediately after jet grouting (i.e., within 10-
15 minutes). The technique developed employs electrical resistivity profiling that exploits the
electrical conductivity contrast between the fresh soilcrete and the surrounding soil. The paper
then presents results from field testing of the probe on two sites a deep soil mixing site in
Florida and a jet grout site in Germany.

Proceedings Paper Formatting Instructions 1


STATE OF THE PRACTICE IN JET GROUT MONITORING

A number of techniques have been used for quality control/quality assurance QC/QA assessment
of jet grouted columns. The density and geometry of a column can be estimated from design
parameters and the actual volume of grout injected for each borehole with some expectation of
grout/soil ratio (Ho 2011). During injection, properties such as grout pressure and drill string
rotation rate are controlled via the machine operator. This type of QC/QA is commonly used in
practice, and is the current definition of "monitoring" (Larsson 2005). However, Stark et al.
(2009) concludes that the grout:soil ratios for field mixed columns is often highly variable and
not consistent with mix design parameters.
Other common approaches used for jet grout geometry assessment in industry include
radial coring/probing or column excavation (e.g., Duzceer and Gokalp 2004, Olgun and Martin
2008, Rollins et al. 2010, Yoshida 2010, Burke 2012, Bruce 2012, Wang et al. 2012, etc.), but
these approaches require 2+ days for adequate curing and are difficult/unfeasible to perform
below the water table. Furthermore, these tests are destructive and can only be performed on test
columns (and not production columns).
A number of nondestructive approaches have been proposed to measure jet grout column
geometry, including mechanical downhole devices (Passlick and Doerendahl 2006). These
devices, however, are cumbersome and potentially not recoverable from the column.
Temperature monitoring has also been proposed (Meinhard 2002, Mullins 2010, Sellountou and
Rausche 2013), but the inherent heterogeneity of field soils and soilcrete can cause considerable
error in the ability of these techniques to estimate diameter. In a related field, thermal imaging
has been used with some success to assess diaphragm walls and diaphragm wall joints
(Doornenbal et al. 2011, Spruit et al. 2011).
Several wave propagation approaches have been applied to estimate geometry of jet grout
columns and similar cement grout applications. The earliest attempt to characterize grout
injections in the field using seismic methods was conducted by E. L. Majer (1989). Several
downhole seismic arrays were placed around the injection area, which was composed of
fractured crystalline rock. The goal of Majers study was to measure acoustic emission events
caused by the pressurized injection of grout to track the flow of the grout in the subsurface. All
active sensors detected acoustic emissions during the grout injection, and incoming events had
frequency content ranging from 2-10 kHz. However, the largest acoustic emission events
occurred after the grout injection pressure had been released. Majer concluded these significant
acoustic emissions are the result of the fractured rock and grout mixture setting to a final
configuration in the absence of grout pressure.
The most common electrical method used to evaluate chemically stabilized soil is DC
electrical resistivity. This technique is based on Ohms' Law and Earth material resistance to
current flow. There is limited literature applying DC resistivity to jet grout columns, so
additional DC resistivity studies related to similar cement grout applications are also discussed.
The electric cylinder method (ECM) is a commercially-available DC resistivity technique used to
estimate the geometry of a jet grouted column (Frappin and Morey 2001, Frappin 2011). The
ECM employs a central borehole with a slotted casing in the center of the column (either pushed
into the fresh column or drilled in after 1-2 days of curing). After casing placement, a chain of
electrodes is lowered into the water-filled casing to allow electrical coupling between the jet
grout and the electrodes (i.e., the electrodes are coupled to the water, which is coupled to the jet
grout through the slots in the casing). This approach uses a type of pole-pole electrode array

Proceedings Paper Formatting Instructions 2


configuration that requires reference electrodes on the ground surface. Frappin and Morey (2001)
conclude that the ECM can estimate to within 10% of the column diameter. However, in regions
where geometry changes are the result of changing soil conditions, there is an additional 0.5m
error. This can result in considerable uncertainty.
A laboratory-based study by Bearce et al. (2015) showed that electrodes directly coupled
to the soilcrete combined with a Wenner- electrode configuration can provide compelling
advantages over indirect electrode coupling through a fluid-filled, slotted casing (such as that
used in the ECM). Namely, the use of electrodes directly coupled to the soilcrete (as opposed to
indirect coupling via a fluid-filled slotted casing) provides an in-situ estimate of soilcrete
resistivity that can vary with depth. Accurate prediction of column diameter requires accurately
known soil and soilcrete resistivities. To this end, obtaining an in-situ estimate of soilcrete
resistivity increases the accuracy of diameter predictions. Furthermore, the borehole based
Wenner- approach does not require reference electrodes on the surface (an approach that is
sometimes unfeasible in real construction environments) and constrains the current injection to
the vicinity of the jet grout column.

ELECTRICAL RESISTIVITY PUSH PROBE

Probe. The probe development builds upon the findings of Bearce et al. (2015) that
demonstrated the benefit of direct electrode coupling. The probe, shown in Figures 1 and 2, is
composed of 1.5m long sections constructed from PVC pipe with ring electrodes spaced every
0.3m (Figure 1a). Any number of sections may be used. In this study, four sections with 20
electrodes were used. Each section is comprised of a full-length 1.5m long inner PVC pipe and
sleeved with 10cm outer diameter short pieces that act as spacers between the ring electrodes.
AWJ drill rod is used to provide structural integrity to the sections (Figures 1a, b). The aluminum
ring electrodes are 10 cm in diameter and 0.8 cm in height. They are spaced at 30 cm, providing
5 per 1.5m section. (Figure 1a, c). The space between the drill rod and the inner PVC pipe is
hollow and contains instrumentation wiring (Figure 1b). This air void also serves to electrically
isolate the outer PVC and ring electrodes from the electrically conductive support drill rod.
Complete details of the probe design and assembly are provided in Bearce (2015). The sections
are assembled on-site to form the full 6m probe shown in Figure 2b. The probe was designed for
a Wenner- electrode protocol with equally spaced electrodes throughout (Figure 2a); however,
other protocols, e.g., Wenner-beta/gamma/Schlumberger, can be used. With surface reference
electrodes, pole-pole configurations could also be employed.
The assembled probe was fastened to a boom mast (Figure 2b) and lowered into the
freshly mixed jet grout or DSM columns (Figures 2c, d). Because the soilcrete was still in slurry
form, the probe was easily inserted into the column and held in place with gravity. The probe
was centered by on-site personnel hand-guiding the probe into the column as it was lowered
from the boom mast (Figure 2c). The probe is removed within 30 minutes, while the soilcrete is
still in slurry form, and therefore does not adversely impact the column.

Proceedings Paper Formatting Instructions 3


Figure 1: Illustration of the electrical push probe geometry and components.

Figure 2: (a) Illustration of the electrical push probe and measurement points obtained
from the Wenner- protocol for the electrode spacings used in this study; (b) (d) the
electrical push probe attached to a placement rig, submerged into a fresh jet grout column.

Proceedings Paper Formatting Instructions 4


Method of Analysis. A detailed presentation of the underlying physics and the analysis approach
is beyond the length capacity of this paper. This can be found in Bearce (2015) and Bearce et al.
(2016). The electrical resistivity technique is governed by Ohms law. Current ( ) is injected
across a pair of electrodes (A and B) to create an electric field in the subsurface that is sampled
by measuring the potential difference ( ) across two measurement electrodes (M and N).
Each measurement yields a value of resistance R () that is converted to an apparent resistivity
(m) using a geometric correction factor k (m),

= = (1)

where is the potential difference (V) measured across electrodes M and N, iAB is the current
(A) injected across electrodes A and B, and k is the geometric correction factor (m). Each
measurement corresponds to at a depth zMN, i.e., the midpoint of electrodes M and N.
Geometric factors (k) are estimated with FE modeling using COMSOL Multiphysics. The k
factors for the probe depend on probe depth and are also affected by the finite volume ring
electrodes used to inject and measure current (as opposed to the point electrodes conventionally
used in electrical resistivity testing). Resistance values obtained from the probe are converted to
apparent resistivity using FE k factors.
To estimate column diameter from probe measurements, FE modeling is performed with
multiple column diameters. The soil resistivity profile is known from the vertical electrical
survey and the soilcrete resistivity is estimated from the a = 0.3 m measurements as validated in
Bearce (2015). Using the resistivity values for all materials in the system, a manual inversion
technique is used to estimate column diameter. The push probe electrical resistivity measurement
protocol is simulated in the FE model to obtain an FE-estimated termed . A linear
correlation is formed between and the diameter of the FE-modeled column. values from
the push probe tests are plotted on the correlation line to estimate the diameter of the field-
constructed column. (and ) depend on probe depth, electrode configuration, and
soil/soilcrete resistivities, and thus an individual correlation is required for each measurement
point. This analysis method is described in detail in Bearce (2015), and Bearce et al. (2015).

RESULTS

Jet Grout Column. The jet grout columns evaluated in this study were constructed at a depth (z)
of 3-10m at a test site south of Berlin, Germany operated by the BAM Federal Institute for
Materials Research and Testing (see Bearce et al. 2016a for a detailed presentation). The site
contains post glacial sediments consisting of sandy layers of varying grain size with some silts
and organic materials. At the time of data acquisition, the groundwater table was measured at
3m. To obtain a resistivity background profile for the in-situ soil (prior to jet grouting), crosshole
dipole-dipole measurements were conducted using permanently embedded ring electrode casings
with a 14m length, 3m horizontal separation, and 0.5m minimum electrode spacing. While the
depth interval evaluated is composed predominantly of sand, crosshole DC resistivity tests
indicate that the sand has stratified resistivity with depth (Figure 3). A profile of the constructed
columns within the background soil profile is illustrated in Figure 3. One column was grouted at
a constant pressure of 40MPa (Figure 3a) with an estimated diameter to be between 1.2 and 1.3m
(Galindo-Guerreros et al. 2017). A second column was grouted at a pressure of 30MPa from z =

Proceedings Paper Formatting Instructions 5


3-6.5m and 40MPa from z = 6.5-10m (Figure 3b) with an estimated diameter of 1.25-1.3m from
z = 6.5-7m and a reduced column diameter of 0.9-1.1m from z = 3-6.5m.

Figure 3: Berlin, Germany jet grout project soil profile, grouting pressure, probe positions,
and measured region for (a) column 1, and (b) column 2.
Figure 4 presents the resulting estimated diameter with depth for both columns. The
estimated diameter are averaged from apparent resistivities measured at a = 0.6, 0.9, and 1.2m
(Figure 2a). Measurements of apparent resistivity at a = 0.3 m are only sensitive to the soilcrete.
The estimated diameter for column 1 was around 1.25 m and in agreement with the contractor-
estimated diameter based on grout volume used as well as in comparison with crosshole seismic
testing performed days later (Galindo-Guerreros et al. 2017). The estimated diameter of column
2 ranged between 0.8 1.0 m until a depth of 6.5 m wherein it increased to 1.2 m. These
estimates as well match contractor and independent testing results.

Figure 4: Berlin, Germany jet grout project Estimated diameter with depth for (a)
column 1 and (b) column 2.

Proceedings Paper Formatting Instructions 6


Deep Soil Mix Column. Testing was also performed on two 2.4 m deep soil mix (DSM)
columns constructed at a construction site in Miami, Florida. The use of a DSM column was
purposeful so that the ground truth diameter would be known. The site geology is shown in
Figure 5a and consisted of fully saturated poorly graded sand. The water table was at the surface
and was consistent with saltwater. The probe was placed to a depth of 5.5 m for testing. The
resulting estimated diameter profile with depth is shown in Figure 5b. As shown in Figure 5b, the
estimated diameter is well within 5% of the actual diameter.

Figure 5: Miami, Florida deep soil mix column: (a) stratigraphy and probe depth, (b)
estimated diameter with depth.

Proceedings Paper Formatting Instructions 7


CONCLUSION

Direct couple electrical resistivity imaging is ideally suited for imaging jet grout column
diameter immediately after construction. The technique relies upon the electrical conductivity
contrast between soilcrete and the surrounding soil (and groundwater), and this contrast is the
greatest in freshly mixed soilcrete. Field implementation of the probe for 0.9 1.25m diameter
jet grout column monitoring in Germany and 2.5 m diameter deep soil mix column monitoring in
Florida revealed that the technique to be accurate within 5% of the actual diameter.

ACKNOWLEDGEMENTS

Funding for this study was provided by the National Science Foundation under the Partnership
for International Research and Education (PIRE) Program (OISE-1243539). The authors also
wish to thank Hayward Baker for providing field test sites and on-site support during Florida
field testing, and the BAM Federal Institute for Materials Research for providing the Berlin
Germany jet grout field site tested in this research. The authors would particularly like to thank
BAM collaborators Ernst Niederleithinger and Julio Galindo-Guerreros for their assistance in
probe implementation and data acquisition during the Germany field test phase.
REFERENCES

Abu-Zeid, N., Balducci, M., Bartocci, F., Regni, R., and Santarato, G. (2009). Indirect estimation
of injected mortar volume in historical walls using the electrical resistivity tomography.
Journal of Cultural Heritage 11, pp 220-227.
Abu-Zeid, N., Botteon, D., Cocco, G., and Santarato, G. (2006). Non-invasive characterization of
ancient foundations in Venice using the electrical resistivity imaging technique. NDT&E
International 39, pp 67-75.
Araji, A. H., Revil, A., Jardani, A., Minsley, B. J., and Karaoulis, M. (2012). Imaging with cross-
hole seismoelectric tomography. Geophys. J. Int. Vol 188 (3), pp. 1285-1302
Backe, K., Lile, O., and Lyomov, S. (2001). Characterizing Curing Cement Slurries by Electrical
Conductivity. SPE Drilling & Completion, Dec. 2001. pp. 201-207
Bearce, R.G., Mooney, M.A., and Kessouri, P. (2016). Direct Couple Electrical Resistivity
Imaging of Freshly Constructed Deep Soil Mix Column Diameter. J. Geotech.
Geoenviron. Eng., in review.
Bearce, R.G., Mooney, M.A., and Kessouri, P. (2015). Electrical Resistivity Imaging of
Laboratory Soilcrete Column Geometry. J. Geotech. Geoenviron. Eng., 142(3),
10.1061/(ASCE)GT.1943-5606.0001404, 04015088.
Bearce, R. G. (2015). Geometry Assessment and Strength/Stiffness Monitoring of Lime and
Cement Modified Soils via Characterization of Curing-Induced Property Changes
Estimated from Seismic Wave Propagation Techniques and Electrical Resistivity.
Doctoral Disseration. Colorado School of Mines.
Bruce, D. A., (2012). Specialty Construction Techniques for Dam and Levee Remediation. CRC
Press. pp 81-91.
Burke, G. K. (2012). The State of Practice in Jet Grouting. Proceedings of the Fourth
International Conference on Grouting and Deep Mixing, New Orleans, LA.
COMSOL Multiphysics (2014). COMSOL Multiphysics Users Guide.

Proceedings Paper Formatting Instructions 8


Chung, D. L. L. (2004). Electrically conductive cement-based materials. Advances in Cement
Research 16, No. 4, pp. 167-176.
Daily, W. and Ramirez, A.L.: 2000, Electrical imaging of engineered hydraulic barriers,
Geophysics 65, pp 8394.
Doornenbal, P., Hopman, V., and Spruit, R. (2011). High resolution monitoring of temperature in
diaphragm wall concrete
Duzceer, R., and Gokalp, A. (2004). Construction and Quality Control of Jet Grouting
Applications in Turkey. Third International Conference on Grouting and Ground
Treatment, New Orleans, LA. Feb. 10-12, 2003
Essler R., and Yoshida, H. (2004). Jet Grouting. Ground Improvement, 2nd edition. CRC Press.
Chapter 5, pp. 160-195.
Frappin, P. (2011) CYLJET An Innovative Method for Jet Grouting Column Diameter
Measurement. First International Conference of Engineering Geophysics. Al Ain, United
Arab Emirates.
Frappin, P. and Morey, J. (2001) Jet Grouted Column Diameter Measurement using the Electric
Cylinder Method. Soletanche Bachy. Internal Publication.
Galindo -Guerreros, J., Niederleithinger, E., Mackens, S., and Fechner, T. (2017). A new quality
assurance tool for jet grout columns. Near Surface Geophyics, in press.
Guo, K., Milkreit, B., and Qian, W. (2014). Geometry factor for near surface borehole resistivity
surveys: a key to accurate imaging and monitoring. GeoConvention 2014: FOCUS. May
2014, Calgary, Canada.
Huang, Q. and Lin Y. (2010). Selectivity of seismic electric signal (SES) of the 2000 Izu
earthquake swarm: a 3D FEM numerical simulation model. Proc. Jpn. Acad., Ser. B 86
Kim, J. H., Yi, M. J., Park, S. G., and Kim, J. G. (2009) 4-D inversion of DC resistivity
monitoring data acquired over a dynamically changing earth model, Journal of Applied
Geophysics, 68, pp. 522-535.
Kumar, V. S., Kelekanjeri, G., and Gerhardt, R. (2008). A closed-form solution for the
computation of geometric correction factors for four-point resistivity measurements on
cylindrical specimens. Meas. Sci. Technol. 19.
Meinhard, K. (2002). Sizing for Strength. Euporean Foundations. Summer 2002, pp. 18.
Mullins, G. (2010). Thermal Integrity Profiling of Drilled Shafts. Journal of the Deep
Foundations Institute. Vol 4, pp 54-64.
Niederleithinger, E., Hbner, M., and Amir, J.M. (2010). Crosshole sonic logging of secant pile
walls - a feasibility study. In Proceedings of the Symposium on the Application of
Geophysics to Environmental and Engineering Problems (SAGEEP), Keystone, Colo.
Vol. 2, pp. 685693.
Passlick and Doerendahl (2006). Quality assurance in Jet Grouting for a deep seated slab in
Amsterdam. Piling and Deep Foundations Conference, Amsterdam. 2006.
Rajabipour, F., Sant, G., and Weiss, J. (2007). Development of Electrical Conducvitity-Based
Sensors for Health Monitoring of Concrete Materials. Transportation Research Board
2007 Annual Meeting CD-ROM. 16 pages.
Revil, A., Karoulis, M., Johnson, T., and Kemna, A. (2012). Review: Some low-frequency
electrical methods for subsurface characterization and monitoring in hydrogeology.
Hydrogeological Journal (10 Feb 2012) pp. 1-42.

Proceedings Paper Formatting Instructions 9


Rcker, C., Gnther, T., and Spitzer, K. (2006). Three-dimensional modelling and inversion of
DC resistivity data incorporating topography I. Modelling. Geophys. J. Int. 166, pp.
495-505.
Santarato, G., Ranieri, G., Occhi, M., Morelli, G., Fischanger, F., Gualerzi, D. (2011). Three
dimensional electrical resistivity tomography to control the injection of expanding resins
for the treatment and stabilization of foundation soils. Engineering Geology 119, pp. 18-
30.
Sellountou, E. A. and Rausche, F. (2013). Quality management by means of load testing and
integrity testing of deep foundations. Pile Dynamics Inc., Cleveland, OH. Internal
publication.
Spruit, R., Hopman, V., van Tol, F., and Broere, W. (2011). Detecting defects in diaphragm
walls prior to excavation. Proceedings of the 8th International Symposium on Field
Measurements in GeoMechanics.
Spruit, R., van Tol, F., Broere, W., Slob, E., and Niederleithinger, N. (2014). Detection of
anomalies in diaphragm walls with crosshole sonic logging. Canadian Geotechnical
Journal, Vol. 51(4), pp. 369-380.
T&A Survey (2013). 3D Borehole Radar Determination of Jet Grout Column Diameter. T&A
Survey internal publication, Amsterdam, The Netherlands.
Wang, S., Mu, M., Chen, D., and Ren, G. (2012) Field design of Jet Grouting Parameters on
Soilcrete Columns. Applied Mechanics and Materials, Vol 170-173, pp 3068-3071.
Wang, X., Yue, H., Liu, G., Zhao, Z. (2011). The Application of COMSOL Multiphysics in
Direct Current Method Forward Modeling. Xian International Conference on Fine
Geological Exploration and Groundwater & Gas Hazards Control in Coal Mines.
Yi, M.J., Kim, J.H. and Son, J.S., (2009). Borehole deviation effect in electrical resistivity
tomography. Geosciences Journal, vol. 13, No. 1, pp. 87-102.
Yoshida, H. (2010). The progress in jet grouting in the last 10 years in the Japanese market.
Proceedings of the 35th Annual Conference on Deep Foundations, Hollywood, CA.

Proceedings Paper Formatting Instructions 10


127
An Unconventional Application of Jet Grouting to Install
4900 kN Ground Anchors in Loose Alluvial Soil

Authors: Manassero, Vittorio


Presentation Preference: Oral
Track: Track 1/ Grouting Case Histories
Session: 1A20/ Jet Grouting: Verification and Testing
Is this a student presentation?: None Specified
Bio: None Specified
Submission type: Technical Papers (6-10 pages in length)
Date Uploaded: 1/10/2017 5:46:43 AM
____________________________________________________________

Project Manager Submission separator pages


An Unconventional Application of Jet Grouting to
Install 4900 kN Ground Anchors in Loose Alluvial Soil

Vittorio Manassero1

1
Underground Consulting s.a.s, via Certosa 1/i; I-27010 San Genesio ed Uniti (Pavia), Italy;
e-mail: v.manassero@undergroundconsulting.it

ABSTRACT

This paper addresses an unconventional application of jet grouting, where improved soil columns
acted as the bonding section of ground anchors installed in loose alluvial soil with an unusually
high capacity. A retaining wall, 25m high, which had been constructed to support refuse and
waste material, began to suffer displacements that became so large as to reach very critical
conditions in terms of stability safety. To improve its stability 120 ground anchors were installed,
each capable of withstanding a 4900 kN working load. The bonding section was achieved by a
1.5m diameter jet-grouting column. Suitability tests were performed to confirm the adequacy of
the design and acceptance tests were carried out on all the ground anchors, both at a loading
force of 5900 kN. All the anchors performed very well, showing good elastic behavior at each
loading level and no symptoms of incipient failure.

INTRODUCTION

A hollow gravity retaining wall 25m high and 300m long was constructed to support refuse and
waste material within the Cerro Maggiore municipal waste landfill, located in the north of Italy.
The resistant reinforced concrete section was made of two parallel longitudinal walls and
transverse buttresses to form a series of 60 adjacent cells with a 5 x 6m foot-print (Figures 1 and 2).
After its construction, as the waste backfill thickness increased, the wall began to suffer
displacements and rotations that became so large as to reach very critical conditions in terms of
stability safety. Furthermore, these displacements caused the failure of the HDPE geomembrane
installed against the rear face of the wall, with subsequent leakage of leachate.
It was then necessary to design remedial works to increase the safety factor, preventing
the wall from sliding and falling over, and to repair the leachate leakages.
At first, the design of the remedial works was made by the client and foresaw the increase
of the wall weight and the application of external stabilizing forces by standard ground anchors.
In a second stage, the specialized contractor in charge proposed an improvement and
optimization of the design, introducing a soil improvement below the foundation and ground
anchors of unusually high capacity to apply the required stabilizing forces to the wall.

1
Figure 1. General view of the hollow gravity wall.

THE DESIGN OF REMEDIAL WORKS

In the design phase of the remedial works the following techniques were envisioned (Manassero
2011):
- improvement of the foundation soil;
- increase of the wall weight;
- application of stabilizing forces through a series of permanent ground anchors.
The foundation soil was composed of alluvial medium-dense sand and gravel while the retained
material applying the earth pressure to the wall was refuse and waste backfill.
The soil improvement was planned below the toe of the foundation at the front face of the
wall (Figure 2) and was carried out by permeation grouting through tubes manchettes.
The increase of the weight was obtained by partial backfilling with ballast concrete as
high as 10m, out of 25m of total height of the wall (Figure 2). This backfilling was achieved in
two phases: the first one, 5m high, was made before the installation of the ground anchors and
the second one, again 5m high, after the installation.
The design of the ground anchors faced two main problems: the anchor could not go
through the walls rear vertical face because of the presence of the retaining geomembrane but
had to penetrate through the bottom horizontal surface of the foundation. Secondly, the statics
and boundary conditions determined the installation of all the anchors at the same level, high
enough to guarantee an adequate lever arm for the horizontal component of the stabilizing forces.
At the first stage of the design the application of the stabilizing forces was anticipated by means
of 10 ground anchors per cell, each one with a capacity of 980 kN, with a total of 600 anchors.
But it was complex to house 10 anchors per cell because of:
- difficulties in the surface set-up of 10 anchors at the same level inside a cell only 5m wide;
- a high risk of interference between the boreholes while drilling due to the undesirable and
unavoidable deviations from the theoretical axis;

2
- interference between the bonded sections at depth, thus reducing the group anchor
capacity.
For these reasons the application of a limited number of very high capacity ground anchors was
preferred to a much larger number of anchors of standard capacity. Per each cell, two high
capacity anchors were foreseen. They were all fitted with 29 steel strands and capable of
withstanding a 4900 kN working load.
But a standard anchor, even if grouted by numerous rounds of post grouting, could never
have borne such a load if installed in a loose alluvial soil. Therefore, the decision was to
preliminarily improve the soil at each anchor axis with a 1.5m diameter jet-grouted column. This
would be used later on as the bonding section of the ground anchor to be installed.
The anchor head position was designed at 5.0m above ground level and the anchor axis
inclination alternated between 50 and 59 below the horizontal. The total length was 33m, with
17m of free length and 16m of bonding length.

Figure 2. Cross section of the hollow gravity wall.

3
SOIL IMPROVEMENT BY PERMEATION GROUTING

The tubes manchettes for the soil improvement were arranged on 4 alignments (Figure 3) and
spaced 1.6m along the walls alignment. They were 5.5 to 7.5m long and equipped with sleeves
that acted as one-way valves every 0.5m. Once the borehole drilling had been completed, the
TAM was installed and the sealing carried out by pumping a sleeve cement-bentonite grout into
the annular space between pipe and soil, completely replacing the drilling fluid. Permeation
grouting was then carried out through the TAMs using a selective and repeated procedure by
injecting a cement-based grout.

Figure 3. Soil improvement arrangement.

The grout was composed of cement, bentonite and water and had the following characteristic
ratios (by weight):
- W/C = 1.33 (C/W = 0.75)
- B/W = 0.03
The cement was Type III-B (high slag blast-furnace cement) class 42.5 and the bentonite was of
low yield with limit liquid within the range of 300-400.
The main physical characteristics of the cement-based grout were the following:
= 1.43 g/cm3; Marsh viscosity 40 s; bleed 0.5%.
The grouting strategy adopted was the following:
- grouting was performed first on row No. 1, then No. 3, No. 4 and finally No. 2 (Figure 3);
- within each row the TAMs were differentiated into primary and secondary tubes;
- for each row, the secondary TAMs were drilled and grouted only upon completion of the
primary ones;
- the grouting procedure was selective and repeated;

4
- the grouting procedure was based on refusal pressure and maximum design volume;
- the grouting flow rate was always kept within the range of 8-12 l/min
- the refusal pressure was fixed at 3 MPa;
- the maximum design volume was fixed at 20% Vt (i.e. 20% of the theoretical volume of the
soil to be grouted relevant to each valve) for the first stage and 10% Vt for the second stage;
- for each valve, injection was stopped when either the design volume or the refusal
pressure was reached, whichever came first;
- if the stop was due to refusal pressure, grouting of the valve was considered completed;
- if the stop was due to design volume, grouting was resumed later on, at a second stage,
using the same criteria and refusal pressure and a design volume reduced to 10% Vt.;
- no further stages were required after the second one.
These grouting pressures are understood as net pressures, i.e. after deducting the losses along the
grouting line, packer and sleeved valve.

HIGH CAPACITY GROUND ANCHORS

To improve the stability of the hollow gravity wall, a total of 120 permanent ground anchors
capable of withstanding a 4900 kN working load were installed. They were 33m long, with a
16m bonding length, and they were fitted with 29 steel strands (steel section 150 mm2).
A diagram of the ground anchor is shown in Figure 4. The characteristics are as follows:
- a borehole of 229 mm dia.;
- uncoated strands both at the bonding length and free length;
- a corrugated HDPE pipe 175mm O.D. for the anchor protection over the whole length;
- all the requisites to assure the permanent function of the anchor;
- grout and vent pipes;
- external and internal spacers;
- a jet grouting column at the bonding section (installed prior to the anchors installation);
- cement grout in all the annular spaces between the corrugated HDPE pipe and soil or jet
grouting, and inside the HDPE pipe;
- a permanent bag packer at the interface between free and bonding sections;
- steel anchor head embedded within the reinforced concrete wailing beam.
At each anchor axis, a pre-placed 1.5m diameter jet grout column was installed to improve the
soil along the bonding section of the anchor which would be installed later on. As the starting
point of the drilling was located 5m above ground level, it was not possible to operate using a
conventional drilling rig. The choice was taken to use a special crawler-mounted lifting platform,
with an elevated rig mast to drill all the boreholes necessary for the installation of the scheduled
jet grouting columns.
The jet grouting columns were carried out by the double-fluid procedure. A cement grout
with W/C = 0.83 was used and the cement was a type III-B (high slag blast-furnace cement),
class 42.5.

5
Figure 4. Diagram of the 4900 kN capacity ground anchor

The operative parameters adopted were as follows:


- grout pressure 50 MPa
- grout delivery 265 l/min
- compressed air pressure 1 MPa
- compressed air delivery 3600 l/min
- withdrawal speed 19.5 cm/min
- rotation speed 12 rpm
- specific energy 67.5 MJ/m (38 MJ/m3)
Once the jet grouting material had achieved suitable strength, drilling through the column could
proceed and the anchor be installed. The same special crawler-mounted lifting platform bearing

6
the drilling rig mast utilized for jet grouting, was used to drill all the ground anchor boreholes.
It is important to note that a consideration for choosing the jet grout column diameter, in
addition to meeting the performance requirements, was to allow drilling to remain inside the
column even in the presence of deviations up to the operative tolerance limit allowed for the
inclined boreholes in project (1.5% of the length).
Drilling through the soil and through the jet grouting column was carried out by down-
the-hole hammer with a nominal diameter of 229mm.
The anchors were assembled on site in a dedicated storage area located nearby, down-hill
of the wall. They were wheeled from the storage area to the borehole to be equipped by a 33m-
long trolley designed and constructed for the purpose. Due to the significant weight of the anchor
(more than 1.5 t), the installation was enabled by a mobile crane running along the road located
at the top of the wall (Figure 5)
To lift the anchor, clamping at its upper edge was possible thanks to an anchor plate fitted
with wedges; each single strand was locked to the plate by its own wedge. Lowering of the
anchor into the borehole was carried out carefully and slowly in order to safeguard the integrity of
the corrugated HDPE pipe and of the grouting pipes installed on the external surface of that pipe.
To bond the anchor to the jet grouting column the following phases were foreseen:
- filling of the bonding length both inside and outside the HDPE;
- expansion of the bag packer at the interface between bonding length and free length;
- filling of the free length, limited to the ring-shaped section outside the corrugated pipe;
- pressure grouting of the bonding length.
The grout mixture was a cement suspension in water, with W/C ratio equal to 0.43 and
Rheobuild 1000 as plasticizer admixture (Rheo/C = 1.8%). The cement was a type I (Portland),
class 52.5. The main physical characteristics of the grout mixture were:
= 1.89 g/cm3; Marsh viscosity 90 s; bleeding 0.5%.

To expand the permanent bag packer an unstable cement suspension was used with W/C = 1, in
order to allow a quick pressure filtration of the free water through the fabric of the bag.
Filling of the bonding length was carried out simultaneously inside and outside the
corrugated HDPE pipe by pumping the grout mixture through the dedicated grouting pipes. The
permanent bag packer was expanded upon reaching the theoretical volume of the bonding length.
After expansion of the bag, grouting into the bonding length continued until a grout with
the same density of that introduced flowed out from the vent pipe.
Then filling of the external portion of the free length followed and continued until a grout
with the same density of that introduced flowed out from the top of the hole.
Pressure grouting of the bonding section followed, injecting the same grout mixture
through a grouting pipe fixed to the external surface of the corrugated pipe. Bearing in mind that
the anchor was to be bonded to a jet grouting column and that the bonding section was already
filled with grout, the following grouting strategy was devised:

7
Figure 6. Installation of a ground anchor by Figure 5. Multi-strand jack used for
a mobile crane located on top of the wall suitability and acceptance tests.

- the grouting procedure was a global and single stage injection;


- the grouting procedure was based on refusal pressure, maximum design volume and
minimum threshold pressure;
- the grouting flow rate was always kept below 3 l/min;
- the refusal pressure was fixed at 1.5 MPa;
- the maximum design volume was fixed at 100 l;
- the minimum threshold pressure was fixed at 1 MPa;
- injection was stopped when either the design volume or the refusal pressure were
reached, whichever came first;
- if the stop was due to refusal pressure, grouting was considered completed;
- if the stop was due to design volume, grouting was considered completed only if a
minimum threshold pressure of 1 MPa had also been reached. Otherwise injection was
resumed and continued up to the achievement of the minimum threshold pressure.
The grouting pressures were taken as net pressures, i.e. after deducting the losses along the
grouting line.
The filling of the internal portion of the free length was performed after the acceptance
stress test and lock-off at the working load.

8
SUITABILITY AND ACCEPTANCE TESTS

All the anchors were tested for acceptance at a loading force of 5900 kN; the standard procedure
adopted was that given by the Italian recommendations AICAP (1992). Two anchors were
chosen for testing to confirm the suitability of the anchor design to the in situ ground conditions,
adopting a special procedure in accordance with the Swiss code SIA V 191 (1995).
For all the tests the load was applied using a multi-strand hydraulic jack (Figure 6) with
the following characteristics: section of plunger 1079.5 cm2; stroke 130 mm; maximum number
of strands 31; maximum load 6670 kN.
The suitability tests given by the Swiss code required 7 loading/unloading cycles. A
typical result of the suitability test is shown in Figure 7.

Figure 7. Typical result of suitability test in accordance with SIA V 191 (1995).

For all the loading-unloading cycles we can observe behaviour very close to theoretical elastic
deformation. The analysis of the characteristic parameters confirmed such good behaviour:
- maximum displacement 78.9 mm
- creep displacement 0.2, 0.1, 0.4, 0.7, 1.0, 1.5 mm at the different steps of the test
(max 1.9% of the maximum displacement);
- tg2/tg1 0.995, 0.990, 0.992, 0.988, 0.987, 0.982 at the different steps;
- residual displacement 13.7 mm at the last unloading (17% of the max displacement);
tg2/tg1 being the ratio between the slope of the loading curve and the slope of the immediately
preceding unloading curve (according to SIA V 191, the anchor behaviour is acceptable if
tg2/tg1 > 0.90).
The second suitability test gave a very similar result to the first one. The results achieved
in the suitability tests were solidly confirmed by the acceptance tests performed on all the other
anchors following the AICAP recommendations. A typical result of an acceptance test is shown
in Figure 8.

9
Figure 8. Typical result of acceptance test in accordance with AICAP (1993).

The characteristic parameters gathered from the test shown in Figure 8 are:
- maximum displacement 77.0 mm
- creep displacement 1.0 mm at both steps of the test (1.3% of the maximum
displacement);
- tg2/tg1 0.960;
- residual displacement 8.0 mm at the end unloading (10% of the max. displacement).
The results of this test are well representative of all the acceptance tests performed on the 120
ground anchors installed within the site. All the ground anchors performed very well, showing
good elastic behaviour at each loading level, without showing any symptoms of incipient failure.

CONCLUSIONS

When the geotechnical design requires the installation of very high capacity ground anchors and
the in-situ soil is not suitable for bearing the anticipated design loads, a possible way to
overcome the problem is to preliminarily improve the soil along the anchors bonding section.
The challenging case study described in this paper shows that pre-placed, 1.5m diameter jet grout
columns at each anchor axis allowed a 4900 kN working load to be borne by ground anchors
installed in a loose alluvial soil. One hundred and twenty ground anchors were installed and
tested for suitability or acceptance at a loading force of 5900 kN, all performing very well.

REFERENCES

A.I.C.A.P. (1993). Ancoraggi nei terreni e nelle rocce Raccomandazioni.


Manassero, V. (2011). Nuove tecnologie per sistemi di drenaggio e ancoraggio. XXIII Cycle
of Conferences on Geotechnics in Turin, 23-24 November.
SIA V 191 (1995). Tirants dancrage precontraints. Societ Suisse Ingenieurs et Architectes.

10
210
Uplift Control of Jet Grouted Bottom Plug by Friction
Mobilized along Embedded Anchoring Elements

Authors: Topolnicki, Michal


Presentation Preference: Oral
Track: Track 1/ Grouting Case Histories
Session: 1A20/ Jet Grouting: Verification and Testing
Is this a student presentation?: None Specified
Bio: None Specified
Submission type: Technical Papers (6-10 pages in length)
Date Uploaded: 1/7/2017 3:04:14 PM
____________________________________________________________

Project Manager Submission separator pages


Uplift control of jet grouted bottom plug by friction mobilized along
embedded anchoring elements
Micha Topolnicki, Prof. Ph.D.

Keller Holding GmbH, c/o Keller Polska, Rdestowa 51A, PL81-577 Gdynia, Poland
e-mail: mtopolnicki@keller.com.pl

ABSTRACT

Jet grouted horizontal plug was used to seal the bottom of 20.5m deep TBM launching shaft
constructed for a new underwater road tunnel in Gdask. The plug was positioned right below
the excavation level, and anchored with hollow bar steel micropiles 76 mm in diameter and 15m
long, installed centrally in 1m diameter and 10m long jet grout (soilcrete) columns. The columns
were arranged in a triangular grid with side length of 2.1m, and constructed as part of a single
production cycle together with the plug columns that were 3.1m in diameter and 3.5m in length.
The pull-out capacity of micropiles was determined by the frictional resistance mobilized at the
tendon/soilcrete interface, and the measured average bond to soilcrete strength ratio was 0.25. A
novel aspect was a combined use of the soilcrete plug as structural and watertight element. The
adopted design solution contributed to reduction of bending moments and horizontal
deformations of diaphragm walls, to more safer and controllable execution of works, and to cost
and construction time savings.

INTRODUCTION

Unwanted inflow of ground water through the bottom of deep excavation pits may be restrained
by horizontal plugs, often executed with jet grouting (Soilcrete) technology. In typical design
situations the soilcrete plug is positioned below excavation bottom, at a depth ensuring that the
total stress of reduced overburden outbalances hydrostatic uplift pressure with a sufficient
margin of safety. The resulting depth is therefore usually significant, and often larger than the
minimum embedment needed to satisfy stability of structural walls, protecting the excavation.
Consequently, the excavation walls must be elongated to close the sealing trough. This triggers
additional costs and may lead to uneconomical design. Beyond controlling seepage, the soilcrete
plug may be also utilized as an efficient structural strut below the excavation bottom, reducing
wall bending and deflection. For this purpose the plug should be positioned as high as possible,
preferably right below the excavation level. In this position, however, the plug is usually
unstable and requires anchoring.
A combined function of an anchored soilcrete plug, acting as structural and watertight
element, was successfully utilized during construction of two deep chambers built in difficult
ground conditions prevailing at the Vistula River estuary in Gdask, and used for the assembly
of the TBM and shield drilling of a new road tunnel, comprising twin bores about 12.5m in
diameter and 1072.5m in length (Fig. 1).

(a) (b)
Figure 1. (a) Construction of a shield tunnel under the Dead Vistula river in Gdask,
Poland; (b) TBM launching shaft.

THE SOLUTION ADOPTED FOR THE TBM LAUCHING SHAFT

The design solution selected for the protection and bottom sealing of a 20.5m deep excavation
pit for the TBM launching shaft was the result of geotechnical contractors proposal to move
from working underwater to a dry operation to cut construction time and reduce costs (Fig. 2).
The design changes also helped make the scheme much safer and more controllable during the
execution of works. The ground conditions and the adopted protection system of the launching
shaft are shown in Figure 3.

Figure 2. Comparison of design solutions considered for the launching shaft.

2
Va

Vb

VIa
Vb

Vc
Cross-section - (cf. Figure 1b)

Figure 3. Ground conditions and the adopted protection system of the launching shaft.

The ground comprised alternating marine and alluvial deposits of medium-dense to dense sands
and soft organic silts with a low strength to a depth of about 23m below GL, and underlying silty
clay and sand layers. The low-permeable silty clay layer (VIa) could not be used as a natural cut-
off barrier for the launching and receiving shafts because it would be unstable against uplift
pressure developed during ground excavation works. The groundwater in the area of the tunnel
is mostly confined, and its level is stabilized at a depth of about 0.5m above sea level (ASL),
which was less than 1m below the working platform. This caused additional problems and
required elevating the guide walls during the construction of the diaphragm walls. The tunnel
structure is protected against maximum water level of +2.5m ASL during normal exploitation,
representing extreme flood level (EFL). During construction stages two design water levels of
+0.5m and +1.5m ASL have been also used in the analyses.
The adopted protection system of the launching shaft during temporary construction
stages included 1.2m thick slurry diaphragm walls along the chamber boundaries and embedded
to 31m below GL, concrete and steel strutting, and the soilcrete plug constructed at a depth of 24
to 20.5m below GL (Fig. 3). In order to balance the buoyancy force, the plug was anchored with
hollow bar steel micropiles, 76mm in diameter and 15m long, installed centrally in 1m diameter
and 10m long jet grout columns, constructed as part of a single production cycle together with
the plug columns that were 3.1m in diameter and 3.5m in length. Along the plug section the
tendon diameter was increased to 160mm with factory grouted corrugated sheathing. The
anchoring elements were positioned in a grid of equilateral triangles with side length of 2.1m,
adapted to the basic arrangement of the columns constituting the horizontal plug. Due to
simultaneous execution of the anchoring elements and plug columns the monolithic connections

3
between the two elements was improved and thus the required rigidness and water tightness of
the whole system protecting the excavation bottom was ensured.
The soilcrete plug was installed utilising the Super-Jet monitors and air-shrouded jetting
technology. The jet grouting parameters were determined after in-situ tests. Because of
significant drilling depths it was also necessary to check the vertical deviation of each drilling
hole with the use of a special inclinometer. The verticality test allowed detection of displacement
of columns and a shorter response time to fill in possible leaks in the plug. The average bore
deviation was only 0.4% throughout construction in the launching shaft, revealing a high level of
precision.
Special attention was paid to the impact of the phasing of the work with different design
water levels at each stage of the excavation. The following phases were implemented to begin
construction of the launching shaft and protect the excavation pit:
Phase I Execution from GL of the middle section of the anchored jet grouted plug, leaving a
free stripe along the planned slurry diaphragm walls. The anchoring elements and jet grouted
plug were constructed in a single production cycle reaching depths up to 34m below GL.
After grouting, the proper reinforcement was installed in fresh soilcrete columns and
disconnected about 19m below GL (Fig. 4a),
Phase II Construction of 1.2m thick slurry diaphragm walls along the chamber boundaries
and embedded to 31m below GL (Fig. 4b),
Phase III Construction of a jet grouted plug along diaphragm walls to provide proper water
tightness of the system (Fig. 4c),
Phase IV Construction of a jet grouted block and slab in front of the launching shaft,
stabilising TBM at the beginning of shield drilling operation (Fig. 4d),
Phase V Primary excavation to the depth of 4.5m below GL and construction of a
reinforced concrete top slab with wide openings, enabling lowering and installation of the
TBM at the chamber bottom,

I II III

(a) (b) (c)

V
IV VI VII+VIII

(d) (e) (f)

Figure 4. Main construction phases adopted for the TBM launching shaft.

4
Phase VI Continuation of excavation to the depth of 12m below GL and installation of a
temporary steel strutting structure (Fig. 4e),
Phase VII Excavation to the depth of 20.5m below GL and construction of the bottom slab,
anchored to steel micropiles (Fig. 4f),
Phase VIII Dismantling of the steel strutting structure installed at the depth of 11.5m below
GL and construction of the eyepiece at the headwall; the launching shaft is ready for TBM
assembly.

In order to monitor the diaphragm walls, the soilcrete plug with anchoring, installation of the
strutting and groundworks in the excavation carefully, as well as to analyse large amounts of
production data and control measurements, a virtual model of all structural elements of the
launching chamber was prepared, corresponding to the design solution (Fig. 5a). On particular
stages of groundworks it was possible to check, among others, displacement of the diaphragm
walls or stresses in the steel struts, and to compare the measurements results with permissible
values. Example records of wall displacements, illustrating the beneficial impact of the soilcrete
plug acting as a structural support below the excavation level, are shown in Figure 5b.

IN-18

IN-20

(a) (b)
Figure 5. (a) Virtual model of the launching chamber; (b) Displacements measured
perpendicular to the wall surface for two excavation depths (inclinometers 18 and 20).

FIELD VERIFICATION OF THE BOND RESISTANCE OF ANCHORING ELEMENTS

In the adopted design solution the bond strength of the micropile to soilcrete interface plays a
key role, and is especially important in temporary construction stages when the hydrostatic uplift
pressure acting on the plug needs to be counterbalanced by a sufficient anchoring capacity of the
micropiles embedded in soilcrete columns. To verify the bond resistance in the field, a series of
four pull-out test was conducted on the construction site using the arrangement shown in Figure
6. The tests A1 and A2 comprised tendons with factory grouted corrugated sheathing diameter
160mm and active bond lengths of 1 and 2m, respectively. In tests B1 and B2 a row hollow stem
bar type DSI T76-1600 was used, and the active bond sections were 3 and 4m long. All tendons

5
3

4 A2

4 A1

(a) (b) (c)


Figure 6. (a) Arrangement of four pull-out tests: 1 - soilcrete column with a diameter of
about 1m, 2 - corrugated sheathing, 3 p.v.c. sheath, 4 - hollow stem bar DSI T76-1600;
(b) Factory grouted corrugated sheathing diameter 160mm; (c) Tendons A1 and A2.

were installed in fresh soilcrete columns that were about 1m in diameter and 8.5m in length. A
standard testing set-up with a hydraulic jack and a reaction beam was used (Figs. 7a,b). After
completion of test A2 the tendon was extracted from the column for a visual inspection,
confirming that the slip occurred at the corrugated sheathing/soilcrete interface (Fig. 7c).

(a) (b)

(c)
Figure 7. (a) Installation of test tendon; (b) Testing set-up; (c) Corrugated tendon,
retrieved from the soilcrete column after completion of test A2.

6
The recorded relationships between the applied pull-out force and the net tendon displacement
(after subtracting elastic bar elongation along the free lengths) are shown in Figure 8a, noting
that the maximum applied load had to be terminated at about 1600 kN, corresponding to the
ultimate bearing capacity of the DSI T76-1600 bar (Technical Approval, 2011). Assuming a
uniform shear strain distribution along all active bond lengths the corresponding bond strength
vs. average shear strain relationships were determined, as presented in Figure 8b. The plotted
data indicate that the ultimate bond strengths can be related to about 0.6 to 0.8% shear strain
mobilisation. In case of test A1, with the shortest bond length of only 1m, it is likely that
multiple slips occurred with a progressing displacement, and that a larger force is needed to
effectively fully withdraw the tendon from the column as compared to pull-out force compatible
with 0.8% shear strain.

(a) (b)
Figure 8. (a) Pull-out force vs. tendon net displacement (after subtracting elastic bar elon-
gation along the free length); (b) Bond strength vs. average shear strain along bond length.

When analysing the ultimate bond strength of the tendon to soilcrete interface one should also
take into account the soilcrete strength, shown in Table 1. The data represent soilcrete UCSs at
28 days of curing, measured on cube wet grab samples with a side length of 15 cm. The ratio of
the ultimate bond stress to the mean UCS of soilcrete was found to be in the range of 0.2 to 0.33,
with a mean value of 0.25. No distinct difference between the bonding efficiency of the
corrugated sheathing and the row deformed bar embedded in soilcrete column was noticed.

Table 1. Summary of pull-out tests and evaluation of the ultimate bond strength.

Test Soilcrete UCS Active Ultimate Ratio of the Remarks


bonding bond ultimate bond
Cubes (15 cm), Mean surface strength strength to
at 28 days [MPa] [MPa] [m2] [MPa] mean UCS, [-]
A1 7.2, 8.2, 9.8, 10.2 8.85 0.503 2.0 (1) 0.23 at 0.8% strain
(1)

extrapolated
(2)

A2 5.4, 5.8, 6.0, 6.2 5.85 1.005 1.5 (1) 0.26 at 0.8% strain
B1 5.2, 6.4, 7.4, 7.6 6.65 0.716 2.2 (2) 0.33 Pull-out tests
conducted after
B2 8.2, 8.4, 9.2, 10.2 9.00 0.955 1.8 (2) 0.20 35 to 37 days

7
SELECTED DESIGN ASPECTS

The complete design was checked using several numerical analyses and the displacements
anticipated were in line with the predictions. Special attention was paid to the impact and
influence of the phasing of the work with different water levels at each stage of the excavation.
In relation to this study, however, the following design aspects require further discussion.

Structural (internal) pull-out capacity of the micropile. The design pull-out capacity of the
hollow bar type T76-1600 had to evaluated for temporary (up to 2 years) and permanent
situations, allowing for design service life of the tunnel structure of 100 years. The assumed
double protection against corrosion consisted of a thick soilcrete cover, and a sacrificial
thickness of steel (acc. to EN 1993-5, Tab. 4.1). The quoted value of loss of steel thickness was
3.25mm for aggressive natural soils and 100 years service life. Taking into account nominal and
reduced cross-sections of the hollow bar due to a long-term corrosion, reading Anom=25.48cm2
and Acor=17.57cm2, as well as a partial safety factor for the resistance of cross-section in tension
M2 =1.25 (cf. EN 1993-1, 6.1) the design pull-out resistance of the adopted micropile is:
fyd = fyk / M2 = 490 / 1.25 = 392 MPa = 39.2 kN/cm2,
Rd,temp = fyd Anom = 999 kN for all temporary stages, and
Rd,perm = fyd Acor = 689 kN for the permanent situation.

Tendon to soilcrete bond capacity. The characteristic compressive strength of soilcrete for the
tunnel project was set to fck,cube(28-days)= 5 MPa. Based on the test data presented above, the
ultimate bond strength can be evaluated as fbu = 0.25 fck,cube. With a partial safety factor M = 2,
the design bond strength of the tendon to soilcrete interface equals fbd = fbu /2 = 0.625 MPa.
Consequently, the design pull-out resistance of the adopted micropile equals:
- for the row bar embedded in 10m long soilcrete column: Rbd = 0.07610625 = 1,491 kN,
and
- for the corrugated part embedded in 3.5m thick soilcrete plug: Rbd = 0.163.5625 = 1,099
kN.

External pull-out capacity of soilcrete column. The ultimate baring capacity of the soilcrete
column pulled out from the ground reads:
= /
where: qci static cone tip resistance in i-th soil layer [MPa], ksi coefficient depending on soil
type, qc value and piling technology [-], ti thickness of i-th soil layer [m], D column diameter
[m]. Taking into account a design soil profile, showing average cone tip resistance of about 2.5
MPa in clay and about 20 MPa in sand, and a soilcrete column of one meter diameter the
ultimate pull-out capacity is:
Rsk = (2500/40 4.0 + 20000/150 6.0) 1.0 = 3,297 kN.
The design pull-out resistance for a 10m long soilcrete column pulled out from the ground reads:
Rsd = Rsk / M = 3297 / 2 = 1,648 kN.

Comparing the above design situations it can be concluded that the minimum pull-out resistance
of the considered anchoring system is governed by the design tensile capacity of the adopted
micropile type DSI T76-1600.

8
Global stability against uplift. Global stability of the launching shaft subjected to uplift forces
caused by hydrostatic pressures was verified for temporary and permanent construction stages
with different ground water levels. Temporary stages 1a and 1b, corresponding to buoyancy
forces acting on the undersides of jet grouted plug and of the foundation slab, respectively, were
calculated for ground water levels at +0.5m (Fig. 9a) and +1.5m ASL (Fig. 9b). In addition to
the dead weight of the structure, enclosed soil and soilcrete plug also the effective weight of the
ground adjacent to the soilcrete columns was taken into account. Frictional resistance that may
develop between the ground and the sidewalls of the shaft structure was conservatively
neglected (Ts=0). The permanent stage 2, with buoyancy force acting on the bottom surface of
foundation slab, was analysed for EFL at +2.5m ASL taking into account additional stabilising
component resulting from the dead weight of structures to be constructed inside the chamber
(e.g. technological building).

(a) (b)
Figure 9. Global stability scheme for the launching shaft; (a) Temporary stage 1a with
uplift force acting on the underside of the soilcrete plug and with GWL at +0.5m ASL; (b)
Temporary stage 1b with uplift force acting on the underside of the foundation slab and
with GWL at +1.5m ASL.

The buoyancy safety in the uplift limit state (UPL) was checked using the following general
formula:
+ (1)
where: Vk - characteristic value (index k) of uplift force at the respective level, representing
destabilising action, - characteristic value of effective dead weight component (lower
estimate), Tk - characteristic value of effective friction component (lower estimate), dst/stb -
partial safety factors for destabilising and stabilising permanent actions in the UPL, adopted
from EC7, table A15, Appendix A (dst = 1.0, stb = 0.9).

Particular form of Eq.1 depends on the analysed construction stage. In case of stage 1a (cf. Fig.
9a) it reads:
+2 + +2 (2)
while for stage 1b (cf. Fig. 10) it changes to:

9
+ + +2 (3)
where: Ve - uplift force at the underside of the jet grouting plug (index e), - dead weight of
soilcrete plug ( - under water), - friction force between soilcrete plug and tunnel wall, -
dead weight of soilcrete columns and adjacent ground under water, - friction force between
tunnel wall and the ground, Vp - uplift force at the underside of concrete slab (index p), - dead
weight of concrete slab, - dead weight of tunnel walls under water.

It was found that the global stability condition was decisive for the determination of the number
and length of anchoring elements preventing uplift of the excavation bottom, with temporary
stages being more important than the permanent situation.

CONCLUSIONS

The project involved construction of large excavations for tunnel ramps and two deep TBM
shafts as well as working with high groundwater levels and some challenging ground conditions.
The adopted design was the result of geotechnical contractors proposal to move from working
underwater to a dry operation to cut construction time and reduce costs. A novel aspect was a
combined use of anchored soilcrete plugs in the TBM shafts acting as structural and watertight
elements. In case of the launching chamber the location of soilcrete plug just below the
excavation bottom contributed to significant reduction of bending moments and horizontal
deformations of diaphragm walls.
Quality control was also a prime focus, and most of techniques were trialled before work
on the main construction started. Special attention was paid to the adopted anchoring system,
comprising hollow stem steel micropiles installed centrally in soilcrete columns forming
anchoring piles and the sealing plug in a single production cycle. Dedicated pull-out tests
revealed that the ultimate bond strength fbu of the tendon to soilcrete interface can be evaluated
in relation to the characteristic unconfined compressive strength of soilcrete fck, taking fbu =
0.25fck,cube. This finding is in a good agreement with the recommended value of fbu = fck /3,
proposed by Burke and Yoshida (2013), noting the usually adopted relation between the
compressive strength measured on cylindrical (h/d=2:1) and cube samples, i.e. fck = 0.8 fck,cube,
leading to 0.25/0.8 = 0.31.
When analysing buoyancy safety for temporary and permanent construction stages, the
limit cases investigated included the uplift capacity of the individual anchoring elements, and the
global stability of the shaft structure assuming that the anchoring elements form a uniform soil
monolith together with the surrounding ground within their zone of influence. Special attention
was paid to the impact and influence of the phasing of the work with different water levels at
each stage of the construction.

REFERENCES

Burke, G. and Yoshida, H. (2013). Jet Grouting. Ground Improvement, Chapter 6, Ed. K. Kirsch
and A. Bell, CPC Press.
Technical Approval No AT/2011-02-2691for DYWI-Drill hollow bars (2011) - in Polish.

10
214
Quantitative Interpretation of Surface Wave Testing for
Assessment of Ground Improvement by Jet Grouting

Authors:
Lin, Chih-Ping
Lin, Chun-Hung
Chang, Yu-Cheng
Chien, Chih-Jung
Presentation Preference: Oral
Track: Track 1/ Grouting Case Histories
Session: 1A20/ Jet Grouting: Verification and Testing
Is this a student presentation?: None Specified
Bio: None Specified
Submission type: Technical Papers (6-10 pages in length)
Date Uploaded: 1/10/2017 2:12:41 AM
____________________________________________________________

Project Manager Submission separator pages


Quantitative Interpretation of Surface Wave Testing for Assessment
of Ground Improvement by Jet Grouting

Chih-Ping Lin, Ph.D.,1 Chun-Hung Lin, Ph.D.,2 Yu-Cheng Chang3, Chih-Jung Chien4

1
Department of Civil Engineering, National Chiao University, Taiwan; e-mail:
cplin@mail.nctu.edu.tw
2
Disaster Prevention & Water Environment Research Center, National Chiao Tung University,
Taiwan; e-mail: chlin.ce@gmail.com
3
Shih-Ho Engineering Co., Ltd, Hsinchu, Taiwan
4
Li-Jia Engineering Co., Ltd, Taipei, Taiwan.

ABSTRACT

Jet grouting technique is special in that the ground becomes highly heterogeneous after
installing the improvement columns. The surface wave method was recently shown to hold good
promise to assess the average property of the improved ground and evaluate the area replacement
ratio. Even though the improved ground was highly heterogeneous along the seismic survey line,
the surface wave method worked well and yielded a homogenized shear wave velocity profile.
The increase of shear wave velocity after ground improvement was found to be correlated with
the area replacement ratio. This study further examined the quantitative relation between the
percentage increase of shear wave velocity and the area replacement ratio. The effect of offline
sampling (perpendicular to survey direction) was also investigated. 3D spectral element method
was utilized for the numerical investigation. Numerical results show that the offline sampling
range is about 75%-85% of the wavelength. Hence, it should be noted when planning survey
lines. Furthermore, a positive correlation between replacement ratio and velocity increase exist
and the difference caused by the velocity contrast between soils and improved columns could be
neglected as the velocity contrast becomes greater than about 5. An empirical predictive model
between the velocity increase percentage and area replacement ratio was established based on the
numerical simulations.

INTRODUCTION

Ground improvement is the primary application of many geotechnical construction


techniques, permitting construction on incompetent soils by changing their characteristics. It is
often carried out to modify the ground to increase shear strength and reduce compressibility of
soft soils. Ground improvement by stiffening columns, such as jet grouting, stone columns, and
deep soil mixing, are frequently used. These techniques are special in that the ground becomes

Proceedings Paper Formatting Instructions 1 Rev. 10/2015


highly heterogeneous after installing improvement columns. Quality assurance is important to
confirm that design objectives are achieved. The conventional methods (e.g., standard
penetration test, dynamic probing, field vane shear tests, plate loading tests, and laboratory
testing) used to assess the ground improvement ratio are time consuming and cost ineffective. Of
greater concern is that they can be performed only on erratic spots and the results are not
representative of the overall behavior. Seismic wave velocity can be used to quantify soil
improvement by measuring wave velocities before and after ground improvement. Because of
the capability of sampling a representative volume of the ground, the use of seismic surface wave
methods is a promising technique for assessment of such ground improvement.
Seismic surface wave methods are based on the dispersive feature of Rayleigh (or Love)
waves in vertically heterogeneous formation. The surface wave method involves generating an
impact source on the ground and recording the seismograms at several locations, from which the
dispersion curve (i.e., phase velocity as a function of frequency) of surface wave is extracted.
The shear wave velocity (Vs) profile can be estimated from inversion of dispersion curve based
on 1D layered model. There have been some case studies on employing surface wave methods
for assessing ground improvement, such as in dynamic compaction (Kim and Park, 1999), vibro-
stone columns (Moxhay et al., 2001), and jet grouting (Lin et al., 2012). These studies support
the prospective idea of using the surface wave methods in the quality assurance of ground
improvement. However, quantitative interpretation of surface wave testing results in terms of
engineering parameters and in the context of ground improvement is still challenging, especially
in the case of ground modification with improvement columns. Surface wave methods assume
that the propagation medium is horizontally layered. What surface wave methods measure in
highly heterogeneous improved ground remains to be investigated.
Even though the ground becomes highly heterogeneous after installing the improvement
columns, Lin et al. (2012) showed that the overall shear wave velocity of the improved ground is
effectively measured by the surface wave method. Their field case showed that the percentage
increase of shear wave velocity was very close to the design replacement ratio of ground
improvement. In order to come up with a quantitative interpretation, this study investigated the
homogenization of shear wave velocity measured by the surface wave method in the
heterogeneous ground with improved columns by 3-D numerical simulations. The extent of off-
line lateral sampling (perpendicular to survey direction) by the surface wave testing was also
examined. Based on the results, the potential of quantifying the replacement ratio from
percentage increase of shear wave velocity measured by surface wave testing is presented.

METHOD

The software package SPECFEM3D Cartesian was used to compute seismograms for
earth models. It simulates wave propagation bases upon the spectral-element method (SEM) and
in 3-D models. The SEM is a continuous Galerkin technique (Tromp et al., 2008), which can be
seen as a particular case of the discontinuous Galerkin technique with optimized efficiency

Proceedings Paper Formatting Instructions 2 Rev. 10/2015


owing to its tensorized basis functions, to simulate forward and adjoint coupled acoustic-
(an)elastic seismic wave propagation on arbitrary unstructured hexahedral meshes. Six boundary
conditions were specified: one free surface condition and five radiation conditions for simulating
the infinite earth. For radiation conditions at four sides and bottom, convolutional perfectly
matched absorbing layers (Komatitsch and Martin, 2007) was applied to reduce the reflections
from the boundaries.
Typical multi-channel configuration with 24 receivers and 2 m receiver spacing was used to
simulate the seismograms recorded by surface wave testing. The Gaussian function was taken as
the normal force point source applied to the free surface. Time step and grid size were properly
selected to meet the stability condition. Uniform half space model was used to calibrate these
numerical parameters. The parameters used for calculations are listed in Table 1.
The multi-station seismograms represent a dataset in the time-space domain which are
transformed into the frequency-velocity domain for dispersion analysis. Various algorithms of
wavefield transformation exist, but they are essentially equivalent. We followed the algorithm
proposed by Lin and Chang (2004) to process the synthetic seismograms and the amplitude
spectrum of the 2-D wavefield transform is presented in the frequencyvelocity domain.

Table 1 Parameters used for the 3-D spectrum element modeling


Parameters Value
Grid size(m) 0.25
Time steps (s) 0.0002
2

Source Wavelet exp[ t ]

100000
(s) 0.01

RESULTS AND DISSCUSSIONS

Sampling range perpendicular to survey line


Field data showed that there is no significant difference between the line through grout
columns and that through the un-grouted area in a field case where grouting spacing was 2.8 m
(Lin et al., 2012). This implies that surface waves are not simply confined to the vertical plane
passing through the geophones. Offline conditions on both sides affect propagation
characteristics of surface waves as well. In order to understand the offline sampling range and
take that into consideration while planning survey lines, vertical sandwich earth models (Fig. 1)
were constructed to investigate the offline sampling range of surface waves. A vertical soil layer
of width W was sandwiched by grout layers. A parametric study was conducted by varying the
width W, shear wave velocity contrast (VC), and Vs of soil layer as listed in Table 2. The W in
each model varied from 2 to 10 m. The survey line was located at the center of the middle soil
layer. To avoid further complication and uncertainty involved during shear wave velocity

Proceedings Paper Formatting Instructions 3 Rev. 10/2015


inversion, the offline sampling range was investigated on the dispersion characteristics of surface
waves.

Table 2 Parameters of sandwich earth models


EM VC Vps, m/s Vss, m/s Vpg, m/s Vsg, m/s
L1 4.7 330 170 1920 800
L2 2 330 170 800 340
L3 4.7 200 100 1130 470

Figure 1 Sandwich earth model and surveying configuration (Plan view)

As an example, Fig. 2a shows the synthetic seismogram for EM-L1 of Table 2 with W =
8 m. Refractions from the interface of soil and grout can be observed. The corresponding
dispersion image from the wavefield transformation is shown in Fig. 2b. The dispersive
characteristic and higher modes appear although the material under survey line is homogeneous.
Apparently the dispersive features were attributed to the heterogeneity in the direction
perpendicular to the survey line. Here we focus on the fundamental mode. The fundamental
dispersion curves of EM-L1 cases of different width W are depicted in Fig. 2c. It can be seen that
as the width of sandwiched soil layer increases, the phase velocity becomes less dispersive. Five
percent deviation from the homogenous non-dispersive phase velocity was chosen as the
threshold to determine the wavelength that is significantly affected by the side layers. The
relation between the affected wavelength and distance from the survey line to the side interface
for all cases are shown in Fig. 3. It can be seen that velocity contrast and the shear wave velocity
of soil layer would affect the relation, especially the shear wave velocity of soil layer. The offline
sampling range of surface wave is velocity dependent. However, a basic understanding could be
gained. The offline sampling range is about 75% to 85% wavelength. With this understanding,
what the surface wave testing is sensing along the survey line in the context of field condition
can be expected.

Proceedings Paper Formatting Instructions 4 Rev. 10/2015


(a)
0

0.2

Time (s)
0.4

10 20 30 40 50 60
Offset (m) W=10m
(b) 20 (c) W=8m
20

30 W=6m

Theoretical phase
30

velocity of soils
40 40 W=4m

Frequency, Hz
Frequency, Hz

50 50
5% error line
60 60
W=2m
70 70

80 80

100 150 200 140 160 180 200


Phase Velocity, m/s Phase velocity, m/s

Figure 2 (a) Synthetic seismogram of surface wave testing in the sandwich earth model (for
Model L1 with W = 8m); (b) dispersion image of (a); (c) comparison of dispersion curves
retrieved from EM-L1 cases of different width W.

7 1.35
1.30
6 1.19 1:1

5
Wavelength, m

2
VC = 2, Vs = 170m/s
1 VC = 4.7, Vs = 170m/s
VC = 4.7, Vs = 100m/s
0
0 1 2 3 4 5 6 7
Distance to the grouted soils, m
Figure 3 Relation between the affected wavelength and distance from the survey line to the
side interface

Homogenization of surface wave testing


The general earth model of the ground improvement for numerical simulation is
illustrated in Fig. 4. 20-m-long improved columns were installed 5 m below the surface. The
cross section of the improved columns was set to be 1m x 1m. The spacing between two adjacent
columns (S, center to center), Vs of soil, and Vs of grout were varied to study the effect of area
replacement ratio (Ra) and velocity contrast on velocity improvement. The survey lines were
located both right above the installed columns and between the installed columns for each case.

Proceedings Paper Formatting Instructions 5 Rev. 10/2015


Figure 4 Column-installed earth models and surveying configuration (a) plan view (b) side
view

Parameters of the column-installed earth model were listed in Table 3. EM, Vps, Vss,
Vpg, Vsg, S and Ra in Table 3 represent earth model, compression wave velocity of soil, shear
wave velocity of soil, compression wave velocity of grout, shear wave velocity of grout, column
spacing, and replacement ratio, respectively. In Group A, the velocity contrast between grout and
soil was kept constant while varying the column spacing to achieve different replacement ratio
(Ra). A few cases of replacement ratio were repeated under different velocity contrast in Group B.

Table 3 Parameters of column-installed earth models


EM VC Vps, Vss, Vpg, Vsg, S, Ra,
m/s m/s m/s m/s m %
A-1 4.7 330 170 1920 800 4 6.3
A-2 4.7 330 170 1920 800 3.5 8.2
A-3 4.7 330 170 1920 800 3 11.1
A-4 4.7 330 170 1920 800 2.75 13.3
A-5 4.7 330 170 1920 800 2.5 16
A-6 4.7 330 170 1920 800 2.25 20
A-7 4.7 330 170 1920 800 2 25
A-8 4.7 330 170 1920 800 1.75 33.3
A-9 4.7 330 170 1920 800 1.5 44.4
B-1 10 330 170 4080 1700 4 6.3
B-2 10 330 170 4080 1700 2.5 16
B-3 10 330 170 4080 1700 2 25
B-4 10 330 170 4080 1700 1.5 44.4

As an example, synthetic seismogram and dispersion image for EM-A-4 were shown in Fig.
5a and b. Similar to field experience, the shot record seems normal even though the earth
model was highly heterogeneous with grouted columns. The shear wave velocity profiles for
each shot gather were analyzed by SurfSeis 3.0 software. In order to minimize the inversion
uncertainty and better quantify the velocity change due to ground improvement, the layer
thickness was fixed by setting up a two-layer model with 5 m top unimproved layer and a half

Proceedings Paper Formatting Instructions 6 Rev. 10/2015


space representing the improved ground. The results of inversion for Group A are shown in Fig.
5c. The inverted Vs profiles were almost identical for survey lines passing through columns and
that in the middle between columns. This phenomenon agrees with the field observation in Lin et
al. (2012) and can be explained by the offline sampling range of surface wave discussed above.

Figure 5 (a) Synthetic seismogram of surface wave testing in column-installed soils (EM-A-
4 in Table 3); (b) dispersion image of (a); (c) inverted 2-layer Vs profiles for Group A in
Table 3

The percentage increase of Vs (abbreviated as VPI) due to ground improvement was further
analyzed for each case and plotted against the area replacement ratio as shown in Fig. 6. As
expected, VPI increases with area replacement ratio. Group A and B represent two conditions of
velocity contrast. Their velocity increase trends with area replacement ratio are similar. The VPI
for Group B (VC=10) is only slightly greater than Group A (VC=4.7). VPI increases with
increasing VC, but it will reach an asymptotic limit. For VC greater than about 5, VPI is
governed by area replacement ratio and not sensitive to VC. For jet grouting, this is usually the
case. Based on the results of 3-D numerical simulations, an empirical relation between VPI and
Ra can be obtained:

VPI = -2.82 Ra3 + 2.93 Ra 2 + 1.11 Ra (1)

Using Eq. (1), the overall area replacement ratio of jet grouting can be estimated by measuring
VPI using surface wave testing before and after jet grouting.

The homogenization behavior of Rayleigh wave is quite complex since it involves


elliptical particle motion including both longitudinal and transverse movements. An empirical
relation between VPI and Ra is presented here based on 3-D numerical simulations. It is believed

Proceedings Paper Formatting Instructions 7 Rev. 10/2015


that the regularly arranged columns can be treated as an equivalent transversely isotropic
medium based on long-wave assumption. The apparent shear modulus exhibited in the results of
numerical simulations should be some sort of combination of the equivalent shear modulus in the
propagating direction and depth direction. A theoretical relation between VPI and Ra may be
further discussed and derived. Before that, we believe that the relation derived from the
numerical simulations can be practically applied to assess the replacement ratio of ground
improvement by jet grouting.

Figure 6 The relation between velocity percentage increase and area replacement ratio

CONCLUSION

Compared to traditional quality assessment techniques in ground improvement, the


surface wave method was recently shown to hold good promise to assess the average property of
the improved ground and evaluate the area replacement ratio in a non-destructive, efficient, and
cost effective fashion. Even though the improved ground was highly heterogeneous along the
seismic survey line, the surface wave method worked well and yielded a homogenized shear
wave velocity profile. In order to quantify the relationship between the velocity increase due to
jet grouting and the area replacement ratio, 3-D numerical simulations of surface wave testing on
various condition of jet grouting were conducted. Numerical results show that the offline
sampling range is about 75%-85% of the wavelength, depending on the shear wave velocity
under the survey line. This should be noted when planning survey lines and interpreting the
resulting Vs profile. In all simulated conditions where replacement ratio is greater than 6%, Vs

Proceedings Paper Formatting Instructions 8 Rev. 10/2015


profiles measured by surface wave testing were almost identical for survey lines passing through
columns and that in the middle between columns. Furthermore, a positive correlation between
area replacement ratio and velocity increase exists and the difference caused by the velocity
contrast between soils and improved columns could be neglected as the velocity contrast
becomes greater than about 5. An empirical predictive model between the velocity increase
percentage and area replacement ratio was established based on the numerical simulations. A
theoretical relation between velocity increase and replacement ratio may be further discussed and
derived. Before that, it is believed that the relation derived from the numerical simulations can be
practically applied to assess the replacement ratio of ground improvement by jet grouting. More
field case studies will be conducted to further validate our recommendation.

REFERENCES

Kim, D. and Park, H. (1999) Evaluation of ground densification using spectral analysis of
surface waves (SASW) and resonant column (RC) tests. Canadian Geotechnical
Journal, 36, 291-299.
Komatitsch, D. and Martin, R..(2007) An unsplit convolutional Perfectly Matched Layer
improved at grazing incidence for the seismic wave equation. Geophysics,
72(5):SM155SM167.
Lin C-P. and Chang. T-S. (2004) Multi-station analysis of surface wave dispersion. Soil
dynamics and earthquake engineering, 24, 877-886.
Lin, C.-P., Lin, C.-H., Dai, Y.-Z, and Chien, C.-J. (2012) Assessment of Ground Improvement
with Improved Columns by Surface Wave Testing. 4th International Conference on
Grouting and Deep Mixing, New Orleans, USA.
Moxhay, A., Tinsley, R., and Sutton J. (2001) Monitoring of soil stiffness during ground
improvement using seismic surface waves. Ground Engineering Magazine, January,
3437.
Tromp J., Komatitsch, D. and Liu. Q.,(2008) Spectral-element and adjoint methods in
seismology. Communications in Computational Physics, 3(1):132.

Proceedings Paper Formatting Instructions 9 Rev. 10/2015

You might also like