TC On GI GS First Report PDF

Download as pdf or txt
Download as pdf or txt
You are on page 1of 446

First Report of the

Technical Committee on
Ground Improvement and Geosynthetics
Indian Geotechnical Society, New Delhi

Members of the Committee:

Prof. G. L. Sivakumar Babu (Chairman), IISc, Bangalore

Prof. G. Madhavi Latha (Convenor), IISc, Bangalore

Prof. M. R. Madhav, J.N.T.U. College of Engineering, Hyderabad


Prof. K. Rajagopal, IIT Madras
Prof. Satyendra Mittal, IIT Roorkee
Prof. D L Shah, M.S.U., Vadodara

Mr. Y. Hari Krishna, Keller Ground Engineering India Pvt. Ltd.

Ms. Minimol Korulla, Maccaferri Environmental Solutions Pvt. Ltd.

14 November 2016
This report contains 446 pages including the cover page
Contents

Technical Note on Ground Improvement using Vibro


Compaction and Stone Columns: Theory & Practice
-by Keller Ground Engineering India Pvt. Ltd.
Monograph on Granular Piles and Granular Pile Anchors
-by Madhav Madhira and Vidyaranya Bandi
Guidelines for Soil Nailing Technique in Highway Engineering
Applications
-by G L Sivakumar Babu
Technical Note on Ground Improvement using Deep Soil Mixing:
Theory & Practice
-by Keller Ground Engineering India Pvt. Ltd.
Technical Note on Ground Improvement using Grouting
Techniques: Theory & Practice
-by Keller Ground Engineering India Pvt. Ltd.

GEOSYNTHETICS – Glossary & Description


-by K. Rajagopal
Design Monograph for Design of Shallow Foundation with
Geosynthetics
-by Satyendra Mittal
Cellular Confinement Systems
-by Gali Madhavi Latha
Ground improvement Case studies
-by Maccaferri Environmental Solutions Pvt. Ltd.
Indian Geotechnical Society
TC on Ground Improvement and
Geosynthetics
Technical Note on Ground
Improvement using Vibro
Compaction and Stone Columns:
Theory & Practice
-by Keller Ground Engineering India Pvt. Ltd

Revision Details:

00 20-01-2016 For Circulation Arun Madan Hari


Init. Sign. Init. Sign. Init. Sign.
Rev. Date Document Details
Prepared Checked Approved
TABLE OF CONTENTS
1 Background .................................................................................................................................... 1
1.1 General ..................................................................................................................................... 1
1.2 Technical Committee ................................................................................................................ 1
1.3 Brainstorming Session .............................................................................................................. 1
2 Deep Vibro Techniques ................................................................................................................. 1
2.1 Ground improvement in cohesive and mixed soil..................................................................... 1
2.2 Ground Improvement in granular soil ....................................................................................... 2
3 Theory and design approach ........................................................................................................ 3
3.1 Vibro Replacement Technique ................................................................................................. 3
3.1.1 Corrections for improvement factor ................................................................................ 4
3.1.2 Correction for column compressibility ............................................................................ 5
3.1.3 Correction for Overburden.............................................................................................. 5
3.2 Vibro Compaction ..................................................................................................................... 6
4 Construction methodology ........................................................................................................... 7
4.1 Vibro Replacement ................................................................................................................... 7
4.2 Vibro Compaction ..................................................................................................................... 7
5 Applications and Limitations ........................................................................................................ 8
6 Proven performance ...................................................................................................................... 9
6.1 Case study 1: Settlement control.............................................................................................. 9
6.2 Case study 2: To improve bearing capacity ........................................................................... 10
6.3 Case study 3: Liquefaction mitigation ..................................................................................... 12
6.4 Case study 4: Enhancement of Lateral Capacity ................................................................... 13
7 Observations & suggestions on IS 15284 – Part 1 ................................................................... 14
8 Bibliography ................................................................................................................................. 14

LIST OF FIGURES
Figure 1: Basic principle of Vibro replacement technique ....................................................................... 2
Figure 2: Basic principle of Vibro compaction technique ......................................................................... 3
Figure 3: Principle of Ground Improvement ............................................................................................. 3
Figure 4: Unit cell of SC and typical arrangement of triangular and square grid ..................................... 4
Figure 5: Priebe's basic improvement factor (reproduced from Priebe, 1995) ........................................ 4
Figure 6: Consideration of column compressibility .................................................................................. 5
Figure 7: Determination of depth factor ................................................................................................... 6
Figure 8 Proposed initial trial points ......................................................................................................... 6
Figure 9 Sequence of installation of vibro stone columns (dry bottom feed method).............................. 7
Figure 10 Sequence of vibro compaction process................................................................................... 8
Figure 11 Results of Single column load test and post construction settlements ................................. 10
Figure 12 Completed view of building .................................................................................................... 10
Figure 13 Typical soil profile and Cross section of Stone column below the tank ................................. 11
Figure 14 Long term observed settlement ............................................................................................. 11
Figure 15 Completed view of plant ........................................................................................................ 12
Figure 16 Compaction in progress & post treatment subsidence .......................................................... 12
Figure 17 Pre and post treatment CPT results & plate load test results................................................ 13
Figure 18: Typical soil profile ................................................................................................................. 13
Figure 19: Vibro stone columns surrounding BCIS piles ....................................................................... 14
Figure 20: Lateral load test on BCIS piles ............................................................................................. 14

LIST OF TABLES
No table of figures entries found.
LIST OF ENCLOSURES

Annexure 1 Technical paper on “Design of Vibro Stone Columns” (1995) &


Technical paper on “Vibro replacement to prevent earthquake induced
liquefaction” by Heinz J. Priebe (1998)
Annexure 2 Technical paper on “Optimal Foundations in Soft Ground: An Innovative
Approach for Economizing Cost and Time”
Annexure 3 Technical paper on “Vibro Stone Columns to Support Large Oil Storage
Tank Farm on West Coast of India”
Annexure 4 Technical paper on “Ground Improvement Solutions to Mitigate
Liquefaction: Case Studies”
Annexure 5 Technical paper on “Ground Improvement Using Vibro Techniques in
Flyash Deposits”
Annexure 6 Observations on IS 15284 part 1 Rev 1
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice

1 Background
1.1 General
The President Indian Geotechnical Society (IGS) has constituted several Technical
Committees (TCs) in order to contribute substantial technical innovations to serve the
geotechnical community by publishing guidelines in the field of ground engineering. In this
endeavour, IGS formed various TCs to seek support in the preparation of guidelines and
publish them on behalf of IGS. In order to form the guidelines, the modus operandi
suggested by IGS was to conduct brain-storming sessions in local chapters in each of the
selected themes and topics and further to record the proceedings. Each member of the
committee shall have to make a presentation followed by a detailed discussion. The
chairman of each TC will decide the sub-topics on which the theme paper will be presented
by a particular member of the committee, followed by a thorough discussion. The individual
TC will develop guidelines with regard to various fields of Geotechnology on behalf of IGS
who will contribute in a meaningful way to better geotechnical practices in India.

1.2 Technical Committee


With the above background, IGS has identified Ground Improvement and Geosynthetics is
one of the TC and the main objective is to prepare an implementable document for practicing
engineers covering Ground Improvement technology, limitations, codal provisions, case
histories esp. in India with their performances.

1.3 Brainstorming Session


IGS Hyderabad Chapter has taken initiative to support IGS and conducted one day National
Workshop on Ground Improvement and Geosynthetics on 29th August 2015 in JNTU
premises. Minutes of meeting was prepared and circulated among the TC members. It was
agreed in the meeting that design and construction aspects of ground improvement using
deep vibro techniques shall be addressed by Keller Ground Engineering Pvt. Limited (Keller).
This document describes concept, theory (developed by Keller), design & construction,
performance of ground improvement (esp. deep vibro techniques) for variety of projects
executed in India.

2 Deep Vibro Techniques


Ground Improvement is a technique that improves the engineering properties of the weak
soil mass treated. Usually, the engineering properties that are improved due to ground
improvement are shear strength, stiffness and permeability. Ground improvement has been
developed into a sophisticated tool to support foundations for a wide variety of structures.
Properly applied, i.e. after giving due consideration to the nature of the ground being
improved and the type and sensitivity of the structures being built, ground improvement often
reduces direct costs and saves precious construction time.

2.1 Ground improvement in cohesive and mixed soil


Vibro techniques are accepted method of subsoil improvement, in which large-size columns
of coarse grained material are installed in the soil by means of high capacity depth vibrators.
Performance of this composite system consisting of stone columns as reinforcing elements
and the weak soil mass that can be established theoretically can be established by full size
field plate load tests.

1
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice
Ground improvement using vibro stone columns have been profoundly increased on local
soils (cohesive and mixed soils) which are unable to take large foundation loads. Typically
stone column consists of a vertical reinforcement introduced by constructing a column of
densely packed stones partially or fully replacing the local weak soil. The construction can
either by wet or by dry method. The inclusion of stone columns in a specific grid pattern
allows the soil mass behaves like a homogenous layer of improved density and stiffness.
This process yields enhancement of load bearing capacity and minimizes the settlements of
the treated ground compared to the untreated ground.
Stone columns acts as drainage path allowing for rapid consolidation which in turn improves
the strength and deformation characteristics of the ground at a much faster rate. Stone
columns constructed using vibro techniques allows full or partial displacement instead of
partial or full replacement of the weak soil and then leads to further improvement of displaced
weak soil by faster dissipation of construction pore water pressure. Also, the improved
drainage capabilities of the stone column treated ground provide a much better resistance to
liquefaction of the surrounding soil. The resistance to liquefaction is achieved by densification
of surrounding weak soil and also by the much increased capacity for faster dissipation of
excess pore water pressure.

Figure 1: Basic principle of Vibro replacement technique

2.2 Ground Improvement in granular soil


Vibro Compaction is a technique developed by Keller in the 1930's, designed to induce
compaction of granular materials at depth. The basic principle behind the process is that
particles of non-cohesive soils can be rearranged into a denser state by means of vibration.
A Schematic showing Vibro Compaction technique is shown in Figure 2.

2
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice

Figure 2: Basic principle of Vibro compaction technique

3 Theory and design approach


3.1 Vibro Replacement Technique
Priebe (1995) developed design of vibro stone columns considering a unit cell loaded
vertically which can be adopted to different soil conditions. It overcomes the traditional
limitation of a stone column being analysed as an isolated column, loaded only on its head.

Figure 3: Principle of Ground Improvement


Contrary to vibro compaction which densifies non-cohesive soils due to vibrations, vibro
replacement improves cohesive and non-cohesive soils by reinforcing the weak soil with load
bearing columns of well compacted, coarse grained material. When the entire weak soil is
replaced with a well compacted coarser material, there is no complexity in the understanding
of its improved load carrying capacity and corresponding deformations. But, when the weak
soil is partially replaced and displaced by the introduction of these stiffer reinforcing elements
at regular grid patterns, response of this modified ground becomes complex. There are ways
for arriving at an equivalent stiffness matrix of a system that replaces some part with a
material of larger stiffness.
Similarly, there are ways and means to establish the modified density and stiffness when the
entire soil mass is densified. When the improvement is attributed to both displacement and
replacement, the quantification of improvement is difficult to determine. Considerable efforts
like large-scale load tests can only prove the effectiveness of the installed stone columns. In
a first step, an improvement factor is established by which stone columns improve the
performance of the subsoil in comparison to the state without columns just by increasing the
overall stiffness. The grid patterns and concept of unit cell is illustrated in Figure 4. Basic
improvement factor can be arrived based on the area replacement ratio and the reinforcing
material used for stone columns.

3
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice

Figure 4: Unit cell of SC and typical arrangement of triangular and square grid
Improvement factor is presented in Figure 5. According to this improvement factor, the
deformation modulus of the composite system can be established due to which settlements
will be reduced. Priebe’s method is a unit cell approach, which takes into account oedometric
conditions. This is very important because the direct use of Priebe’s composite parameters
for slope stability results in an un-conservative safety factor.

Figure 5: Priebe's basic improvement factor (reproduced from Priebe, 1995)


The deformation modulus of the composite system is one of the basic inputs for finalizing the
design of stone columns. However, the reality is that in many practical cases the reinforcing
effect of stone columns installed by vibro replacement is superposed with the densifying
effect of vibro compaction, i.e. the installation of stone columns densifies the soil between
grids increasing its k0 and kp. In such case, the densification of the soil has to be evaluated
on the basis of original soil data and correspondingly the design of vibro replacement can be
modified to suit particular improved site condition.
The basic improvement factor (n0) shall be calculated using the formula.

3.1.1 Corrections for improvement factor


Since the column cannot fail in end bearing and any settlement of the load area results in a
bulging of the column which remains constant all over its length. The following two
corrections need to be applied for improvement factor to get appropriate value.

4
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice
• Correction for column compressibility
• Correction for overburden

3.1.2 Correction for column compressibility


In the case of soil replacement, the actual improvement factor does not achieve an infinite
value as determined theoretically for non-compressible material, but it coincides with at best
with the ratio of the constrained moduli of column material and soil. Due to compressibility of
column material the area of column may get increase and the improvement factor will be
reduced. The improvement factor after compressibility correction can be calculated using
following relation and Figure 6.

Figure 6: Consideration of column compressibility

3.1.3 Correction for Overburden


As a result of column installation, the external loads the weights of the columns W C and of
the soil W S which possibly exceed the external loads considerably, has to be added. Under
consideration of these additional loads the initial pressure difference decreases
asymptotically and the bulging is reduced correspondingly. In other words, with increasing
overburden the columns are better supported laterally and therefore, can provide more
bearing capacity.
Since the pressure difference is a linear parameter in the derivations of the improvement
factor, the ratio of the initial pressure difference and the one depending on depth expressed
as depth factor fd delivers a value by which the improvement factor n1 increases to the final
improvement factor n2 = fd×n1 on account of the overburden pressure.

݊ ൌ ݂ ή ݊

5
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice

Figure 7: Determination of depth factor


The technical paper on “Design of Vibro Replacement” by Heinz J Priebe is enclosed in
Annexure 1.

3.2 Vibro Compaction


Field trials will be carried out prior to the main works to determine the working parameters for
Vibro compaction process. A level area of 50m x 50m, to carry out a field trial and necessary
area to set-up the plant & equipment is required. The site should preferably be in the vicinity
of the main works area. The site will be levelled (by others) prior to commencement of trial
works.
The vibro compaction will be carried out at selected trial area by trail & error method with
different design parameters like spacings, vibration (amplitude) time & depth of treatment.
The typical plan view of initial trial area is shown in Figure 8.

Figure 8 Proposed initial trial points


Based on the trial results, the required operating parameters, spacings, quality control
procedures, etc. for the vibro compaction will be established and same shall be adopted for
main works.

6
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice

4 Construction methodology
4.1 Vibro Replacement
Keller has developed the system of custom-built machine called the Vibrocat for installation
of vibro stone columns without using water. The Vibrocat comprises a specially constructed
track mounted supporting unit, attached with high capacity depth vibrator, which incorporates
a stone tube with compression chamber and stone feed hopper ensures properly formed
compacted stone columns to the required diameter and depth. A special feature of the dry
method is that it does not require water jetting for penetration and hence eliminates the need
to handle the collected water.
Furthermore this method can be used most successfully where limited working space is
available, especially in developed or urban areas or where no near water source can be
found. This technique provides effective drainage paths to ensure rapid consolidation. It also
has a built-in real time computer monitoring system to provide quality control on compaction
effort throughout the construction process. Sequence of installation of vibro stone columns
using dry bottom feed method is illustrated in Figure 9.

Figure 9 Sequence of installation of vibro stone columns (dry bottom feed method)

4.2 Vibro Compaction


The essential equipment for this process is a depth vibrator - a long, heavy tube enclosing
eccentric weights, driven by an electric motor. The vibrator is connected to a source of
electric power and a high-pressure water pump. Extension tubes are added as necessary,
depending on the treatment depth, and the whole assemblage is suspended from a crane.

7
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice

Figure 10 Sequence of vibro compaction process


With the electric power and water supply switched on, the vibrator is lowered into the ground.
The combination of vibration and high-pressure water jetting causes liquefaction of the soils
surrounding the vibrator, which assists in the penetration process. When the required depth
is reached, the water pressure is reduced and the vibrator pulled up in short steps. With the
inter-particle friction temporarily reduced, the surrounding soil particles then fall back below
the vibrator and, subjected to vibratory energy, are rearranged into a denser state. This
process is repeated back up to the ground level, leaving on completion, a column of very
dense material surrounded by material of enhanced density. The degree of compaction
achieved at a particular point depends on the properties of the soil being treated, the amount
of time spent at each compaction step and the distance from the vibrator. Vibro compaction
is suitable for treating sands with a fines content of less than 10 to 15%. The spacing of
probes is designed to ensure that the zones of influence overlap sufficiently to achieve
minimum requirements throughout the treated area.
Generally, the effect of the compaction becomes visible at the ground surface in the form of a
cone-shaped depression. The depression formed around the vibrator or the extension tubes
is continually in-filled with granular materials, which is either imported or obtained from the
natural granular deposits at the site. Water required for the penetration and compaction
process is obtained either by direct pumping from nearby water source or ground water using
well points.
To check that minimum requirements are being met, the normal procedure is to carry out a
series of post-compaction deep sounding tests.

5 Applications and Limitations


The ground improvement techniques can be adopted for the following ground engineering
applications.
• Enhancing the bearing capacity of in-situ soil
• Controlling the larger total & differential settlement

8
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice
• Mitigating liquefaction potential
• Enhancing lateral confinement for deep foundations
The Ground improvement using vibro replacement / vibro compaction methods being carried
out by stone columns / sand, the following limitations are to be thought during design.
• The stone column / vibro compaction is brittle material which will not take any tensile
or lateral forces and will take care of only vertical compressive loads.
• Though stone columns /vibro compaction points are carried out at regular intervals, it
shall be assumed as the whole soil mass has been improved and same property
need to be considered for design.

6 Proven performance
6.1 Case study 1: Settlement control
M/s Urban Tree Infrastructure Private Limited (Urban Tree), Chennai, proposed to develop a
residential project in Chennai. The project comprises of 198 units of Stilt + 4 floors and the
approximate area of development is about 2.5 acres.
The sub-soil in the project site comprises desiccated clay and medium dense sand up to
about 3.50m followed by relatively weak clay and sandy clay up to 6.0m depth. This top 6.0m
soil with highly varying consistency is followed by about 8.0m with medium dense sand stiff
clay deposits after which there is a 6m thick layer of medium stiff consistency. Denser sand
layers and hard clay layers are forming the remaining sub-soil profile. Required loading
intensity of the proposed structure on the soft soil is 100kPa.
Considering the project boundary conditions, vibro replacement technique with 20% area
replacement ratio (stone columns with dry bottom feed method) up to 6m depth was adopted
as a viable method for subsoil improvement and a full raft foundation supported by the
treated ground as an alternative foundation system.
Keeping the importance of the post construction performance of the structure, plate load test
has been conducted on improved ground and also about 14 locations were identified on the
raft foundation to monitor settlements during and post construction.
The results of post construction are shown below.
• Achieved bearing capacity : > 150kPa
• Long term settlement : < 50mm
The measured settlements are substantially lower than the predicted settlement, which
proved the efficiency of the raft foundation resting on improved ground.

9
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice
100
Load vs Time Curve
Load Intensity, kN/m2 90

Super structure load (kPa)


0 20 40 60 80 100 120 140 80

0 70

60

50
4 40

30

20
Settlement, mm

8
10

0
0 0 4 8 12 16 20 24 28 32 36 40 44 48 52 56 60 64 68 72 76 80 84 88 92 96 100 104 108
12
10 Point: P1S1 Point: P1S2
Point: P3S2 Point: P4S1
20
16 Point: P4S3 Point: P6S3
30 Predicted settlement

Settlement 'mm'
40
20
50

60

24 70

80
0 4 8 12 16 20 24 28 32 36 40 44 48 52 56 60 64 68 72 76 80 84 88 92 96 100 104 108
28 Time in Weeks

Figure 11 Results of Single column load test and post construction settlements

Figure 12 Completed view of building


The technical paper explaining project case study on settlement control for residential
apartment is enclosed in Annexure 2.

6.2 Case study 2: To improve bearing capacity


M/s Hindustan Petroleum Corporation Ltd. proposed to develop a tank farm for 18 nos. of
floating roof storage tank, 4 nos. of fixed roof storage tank and 3 nos. of fire water tank. The
diameter of tank is 32m and height of tank is 15m.
The subsoil consist of top 2.5m flyash fill with the SPT N value NIL followed by soft silty clay
of SPT N value of 7 up to 9m depth and this layer is underlain by medium dense sand of SPT
15 up to 25m depth.
The bearing capacity of virgin soil at foundation level of tank is 100kPa. The required bearing
capacity of a tank foundation is 200kPa.
To improve the bearing capacity of virgin soil vibro replacement method 25% area
replacement ratio up to the depth of 11.5m (stone column using wet top feed method) was
proposed.
The typical soil profile and stone column cross section below the tank is shown in Figure 13.

10
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice

Figure 13 Typical soil profile and Cross section of Stone column below the tank
The results of post construction tests are described below.
• Improved bearing capacity : > 200kPa
• Hydro test with full load (water) was carried out and the settlement was recorded. The
observed settlement is less than the allowable settlement of tank (< 300mm)

14
 
 
12


10
 


 8





 6

4
              
                





 
 


$

$

$

$   
(' 
(' 
(' 
(' 
+ 
+
+ + + +
   ( ( * * * *
2 $ $ # # # # ' ' * *

 
 
 # # " " " " &
 &
 &
 &
 )
 )
 )
 )

 !  " " % % &
 &
 % % )
 )
  ,  ,
 %   %    ,  
0
0
1 -50


0-100

/

 -150

.
-200
-

-250

-300
2345

Figure 14 Long term observed settlement

11
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice

Figure 15 Completed view of plant


The technical paper published on “Vibro stone columns to support large oil storage tank
farms in India” is enclosed in Annexure 3.

6.3 Case study 3: Liquefaction mitigation


M/s. GVK Power Ltd. developed the 2x270MW Coal based thermal power station at
Goindwal sahib, Punjab and functioning since 2015. The power plant consist of various
structures such as power house, boilers, ESPs, switch yard, cooling tower etc. and location
comes under the seismic zone of IV with PGA of 2.4g. The subsoil consist of loose sand
(relative density < 40%) with fines content 4% - 6% which has a chance of liquefaction.
Being sandy soil and fines content < 10%, vibro compaction method to a treatment depth of
10m was proposed to mitigate mainly liquefaction potential and enhancing bearing capacity
of in-situ soil.
Post plate load test & post cone penetration test (CPT) has been carried out at main working
area to access the performance of improved ground and results are satisfied as per design
requirements.

Figure 16 Compaction in progress & post treatment subsidence


The results of improved ground are described as below.
• Relative density after improvement : > 70%
• Improved bearing capacity : > 200kPa
• Backfill consumption : 10%
• Observed subsidence : 1m

12
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice

Figure 17 Pre and post treatment CPT results & plate load test results
The technical paper describing “Ground improvement solutions to mitigate liquefaction”
adopted in India is enclosed in Annexure 4.

6.4 Case study 4: Enhancement of Lateral Capacity


2x500MW Thermal power plant was proposed at Anpara, Uttar Pradesh. The plant units
such as switchyard, crusher house, conveyor and stacker reclaimer are planned to construct
on old fly ash pond. The proposed site comprises of top 3 to 13m flyash deposit followed by
silty clay layer.

Figure 18: Typical soil profile


Because of flyash deposit the required lateral capacity of pile i.e 7MT is not achieved.
Ground improvement using Vibro stone columns are installed at specified pattern
surrounding the bored cast-in-situ piles to enhance the density of fly ash deposits
which in turn can improve the lateral load carrying capacity. It was required to achieve a
design lateral load capacity of 7T with ultimate load of 21T. After the installation of bored
cast-in-situ piles and vibro stone columns by bottom feed method, initial lateral load test are
conducted on these two grid patterns as illustrated in Figure 19.

13
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice

Figure 19: Vibro stone columns surrounding BCIS piles


The results of lateral load test on piles are presented in

Figure 20: Lateral load test on BCIS piles


The results ground improvement are summarised below.
• Achieved lateral capacity : 10 MT
• Stone Column around the pile increases the lateral capacity.
The technical paper describing “Ground Improvement Using Vibro Techniques in
Fly Ash Deposits” adopted in India is enclosed in Annexure 5.

7 Observations & suggestions on IS 15284 – Part 1


In the construction industry as ground improvement has become most common practice in
the last 5 to 6 years, this is high time to prepare / revise standard for all types of ground
improvement methods. All vibration methods like vibro compaction, vibro replacement and
vibro displacement to be added in the present standard IS 15284 (part 1) or a separate
standard on “Ground Treatment by Deep Vibrations” may be brought out (similar to BS). In
the present standard (IS 15284 – part 1), the few points as appended in Annexure 6 to be
incorporated, which will give more clarity and useful for practicing engineers.

8 Bibliography
1. Hughes JMO & Withers N J, 1975, ‘Reinforcing of Soft Cohesive Soils with Stone
Columns’, Ground Engineering, Volume 7, Issue No. 3.
2. Seed H.B., and Idriss, I.M (2001). Simplified procedure for evaluating soil
liquefaction potential, J. Geotech Engineering. Div., ASCE 97 (9), 1249-1273

14
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice
3. Seed et. al (2003). Recent advances in soil liquefaction engineering: A unified and
consistent framework.
4. Priebe, H.J (1995), The Design of Vibro Replacement, Ground Engineering,
December 1995.
5. Priebe, H.J (1998), Vibro Replacement to Prevent Earthquake Induced
Liquefaction, Ground Engineering, September 1998.
6. Raju et. al. (2010). Some Environmental Benefits of Dry Vibro Stone Columns in a
Gas Based Power Plant Project, Indian Geotechnical Conference, December
2010.
7. Article by Bedanga Bordoloi and Etali Sarmah on “Fly Ash Pond Reclaimation”
dated May 2010 for Agribusiness Forum.
8. “Study of Effectiveness of Ground Improvement Techniques and Possible
Liquefaction Potential” for Anpara-D Thermal Power Project, Department of
Earthquake Engineering, IIT Roorkee
9. Raju, V.R. & Sondermann, W., 2005. Ground Improvement using Deep Vibro
Techniques. Ground Improvement Case Histories, Indraratna, B & Chu., J. (eds.),
601-638
10. Indian Standard Code (2003). “Design and construction for ground improvement -
Guidelines, Part 1: Stone columns”, IS 15284 (Part-1): 2003.
11. British Standard (2005). “Execution of Special Geotechnical Works - Ground
Treatment By Deep Vibration”, BS EN 14731-2005.
12. Building Research Establishment (2000) “Specifying Vibro Stone Columns”, BRE -
391.

15
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice

Annexure 1 Technical paper on “Design of Vibro Stone Columns”


(1995) & Technical paper on “Vibro replacement to prevent
earthquake induced liquefaction” by Heinz J. Priebe (1998)

16
The design of vibro
replacement

Dipl.-Ing. Heinz J. Priebe


Keller Grundbau GmbH

Presented by
Keller Grundbau GmbH
Kaiserleistr. 44
D-63067 Offenbach Reprint from:
Tel. 069 / 80 51 - 0 GROUND ENGINEERING
Fax 069 / 80 51 - 244 December 1995
E-mail [email protected]
www.KellerGrundbau.com Technical paper 12-61 E
The Design of Vibro Replacement

The Design of Vibro Replacement

Heinz J. Priebe
Keller Grundbau GmbH

Vibro Replacement is an accepted method for subsoil improvement, at which large-sized co-
lumns of coarse backfill material are installed in the soil by means of special depth vibrators.
The performance of this composite system consisting of stone columns and soil, is not deter-
minable by simple investigation methods like soundings, and therefore, such methods are not
suitable for design purposes. However, theoretically, the efficiency of Vibro Replacement can
be reliably evaluated.The method elaborated on a theoretical basis and described in this contri-
bution, is easy to survey and adaptable to different conditions due to the separate consideration
of significant parameters. Practically, it comprises design criteria for all frequently occurring
applications.

1 Introduction

Vibro replacement is part of the deep vibratory compaction techniques whereby loose or soft
soil is improved for building purposes by means of special depth vibrators. These techniques as
well as the equipment required is comprehensively described elsewhere [1] [1].
Contrary to vibro compaction which densifies noncohesive soil by the aid of vibrations and improves
it thereby directly, vibro replacement improves non compactible cohesive soil by the installation
of load bearing columns of well compacted, coarse grained backfill material.
The question to what extent the density of compactible soil will be improved by vibro compaction,
depends not only on the parameters of the soil being difficult to determine, but also on the
procedure adopted and the equipment provided. However, the difficulty of a reliable prognosis is
balanced by the fact that the improvement achieved can be determined easily by soundings.
With vibro replacement the conditions are more or less revers. Considerable efforts only like
large-scale load tests can prove the benefit of stone columns. However, a reliable conclusion can
be drawn about the degree of improvement which results from the existence of the stone columns
only without any densification of the soil between. This is possible because the essential parameters
attributable to the geometry of the layout and the backfill material can be determined fairly good.
In such a prognosis the properties of the soil, the equipment and the procedure play an indirect
role only and that is mainly in the estimation of the column diameter.
Basically, the design method described herewith was developed some twenty years ago and
published already [3][3]. However, in the meantime it came to several adaptions, extensions and
supplements which justify a new and comprehensive description of the method. Nevertheless,
the derivation of the formulae is renounced with reference to literature.

–1–
Heinz J. Priebe

It may be emphasized: The design method refers to the improving effect of stone columns in a
soil which is otherwise unaltered in comparison to the initial state. In a first step a factor is
established by which stone columns improve the performance of the subsoil in comparison to
the state without columns. According to this improvement factor the deformation modulus of
the composite system is increased respectively settlements are reduced. All further design steps
refer to this basic value.
In many practical cases the reinforcing effect of stone columns installed by vibro replacement is
superposed with the densifying effect of vibro compaction, i.e. the installation of stone columns
densifies the soil between. In this cases, first of all the densification of the soil has to be evaluated
and only then - on the basis of soil data adapted correspondingly - the design of vibro replacement
follows.

Notation

A grid area p area load resp. foundation pressure


b foundation width s settlement
c cohesion W weight
d improvement depth α reduction faktor in earthquake design
dGr depth of ground failure γ unit weight
D constrained modulus η safety against ground failure
fd depth factor µ Poisson´s ratio
K coefficient of earth pressure σ0f bearing capacity
m proportional load on stone columns ϕ friction angle
n improvement factor

Used subscripts, dashes and apostrophes follow from the context. Generally, subscript C means column and S
means soil. With the exception of K0 as coefficient for earth pressure at rest (Ka for active earth pressure)
subscript 0 means a basic respectively an initial value.

2 Determination of the Basic Improvement Factor

The fairly complex system of vibro replacement allows a more or less accurate evaluation only
for the well defined case of an unlimited load area on an unlimited column grid. In this case a unit
cell with the area A is considered consisting of a single column with the cross section A C and the
attributable surrounding soil.
Furthermore the following idealized conditions are assumed:
• The column is based on a rigid layer
• The column material is uncompressible
• The bulk density of column and soil is neglected

Hence, the column can not fail in end bearing and any settlement of the load area results in a
bulging of the column which remains constant all over its length.

–2–
The Design of Vibro Replacement

The improvement of a soil achieved at these conditions by the existence of stone columns is
evaluated on the assumption that the column material shears from the beginning whilst the
surrounding soil reacts elastically. Furthermore, the soil is assumed to be displaced already during
the column installation to such an extent that its initial resistance corresponds to the liquid state,
i. e. the coefficient of earth pressure amounts to K = . The result of the evaluation is expressed
as basic improvement factor n0.

A C  1 2 + f ( µS , A C A ) 
n0 = 1+ ⋅ − 1
A  K aC ⋅ f ( µS , A C A ) 

f (µ S , A C A ) =
(1 − µ S ) ⋅ (1 − A C A )
1 − 2 µS + AC A

K aC = tan 2 ( 45°− ϕ C 2 )

A poisson’s ratio of µS =   which is adequate for the state of final settlement in most cases,
leads to a simple expression.

AC  5 − AC A 
n0 = 1+ ⋅ − 1
A  4 ⋅ K aC ⋅ (1 − A C A ) 

The relation between the improvement factor n0, the reciprocal area ratio A/AC and the friction
angle of the backfill material ϕC which enters the derivation, is illustrated in the well known
diagram of Figure 1
1.

5
ϕ
ϕ S == 45.0˚
45.0˚
Improvement Factor n

ϕϕSC == 42.5˚
42.5˚ µµS B==1/3
1/3
4
ϕS == 40.0˚
ϕ
C
40.0˚

ϕCS =
= 37.5˚
37.5˚
3
ϕSC == 35.0˚
ϕ 35.0˚

1
1 2 3 4 5 6 7 8 9 10
Area Ratio A /A C

Figure 1
1: Design chart for vibro replacement

–3–
Heinz J. Priebe

3 Consideration of the Column Compressibility

2,0

ϕϕCS ==45.0°
Addition to the Area Ratio ∆ (A /A C )

45.0°
1,6
ϕϕCS == 42.5°
42.5° µµsB== 1/3
1/3

ϕϕCS== 40.0°
40.0°
1,2
ϕϕCS == 37.5°
37.5°

ϕϕS ==35.0°
35.0°
C
0,8

0,4

0,0
1 2 3 4 6 8 10 20 30 40 60 80 100

Constrained Modulus Ratio DC /DS

Figure 2:
2 Consideration of column compressibility

The compacted backfill material of the columns is still compressible. Therefore, any load causes
settlements which are not connected with bulging of the columns. Accordingly, in the case of soil
replacement where the area ratio amounts to A/AC = 1, the actual improvement factor does not
achieve an infinite value as determined theoretically for non compressible material, but it coincides
at best with the ratio of the constrained moduli of column material and soil. In this case for
compacted backfill material as well as for soil a constrained modulus is meant as found by large
scale oedometer tests. Unfortunately, in many cases soundings are carried out within the columns
and wrong conclusions about the modulus are drawn from the results which are somtimes very
moderate only.
It is relatively easy to determine at which area ratio of column cross section and grid size (AC /A)1
the basic improvement factor n0 corresponds to the ratio of the constrained moduli of columns
and soil DC /DS. For example, at µS = 1/3 the lower positive result of the following expression
(with n0 = DC /DS ) delivers the area ratio (AC /A)1 concerned.

2
 AC  4 ⋅ K aC ⋅ ( n0 − 2) + 5 1  4 ⋅ K aC ⋅ ( n0 − 2) + 5  16 ⋅ K aC ⋅ ( n0 − 1)
  =− ± ⋅   +
 A 1 2 ⋅ (4 ⋅ K aC − 1) 2  4 ⋅ K aC − 1  4 ⋅ K aC − 1

–4–
The Design of Vibro Replacement

As an approximation, the compressibility of the column material can be considered in using a


reduced improvement factor n1 which results from the formula developed for the basic
improvement factor n0 when the given reciprocal area ratio A/AC is increased by an additional
amount of ∆(A/A C).

A C  1 2 + f ( µS , A C A )  AC 1
n1 = 1 + ⋅ − 1 =
A  K aC ⋅ f ( µS , A C A )  A A AC + ∆ (A A C )

1
∆ (A AC ) = −1
( A C A )1

In using the diagram in Figure 1 this procedure corresponds to such a shifting of the origin of the
coordinates on the abscissa which denotes the area ratio A/A C that the improvement factor n1
to be drawn from the diagram, begins with the ratio of the constrained moduli and not with just
an infinite value. The additional amount on the area ratio ∆(A /AC) depending on the ratio of the
constrained moduli DC /DS can be readily taken from the diagram in Figure 22.

4 Consideration of the Overburden

The neglect of the bulk densities of columns and soil means that the initial pressure difference
between the columns and the soil which creates bulging, depends solely on the distribution of the
foundation load p on columns and soil, and that it is constant all over the column length. As a
matter of fact, to the external loads the weights of the columns WC and of the soil WS which
possibly exceed the external loads considerably, has to be added. Under consideration of these
additional loads the initial pressure difference decreases asymptotically and the bulging is reduced
correspondingly. In other words, with increasing overburden the columns are better supported
laterally and therefore, can provide more bearing capacity.
Since the pressure difference is a linear parameter in the derivations of the improvement factor,
the ratio of the initial pressure difference and the one depending on depth - expressed as depth
factor fd - delivers a value by which the improvement factor n1 increases to the final improvement
factor n2 = fd × n1 on account of the overburden pressure. For example, at a depth where the
pressure difference amounts to 50 % only of the initial value, the depth factor comes to fd= 2.
The depth factor fd is calculated on the assumption of a linear decrease of the pressure difference
as it results from the pressure lines (pC + γC·d)·KaC and (pS + γS·d) (KS = 1). However, it has to be
considered that with decreasing lateral deformations the coefficient of earth pressure from the
columns changes from the active value KaC to the value at rest K0C. Up to the depth where the
straight line assumed for the pressure difference, meets the actual asymptotic line, the depth
factor lies on the safe side. In practical cases the treatment depth is mostly less. However, safety
considerations advise not to include the advantageous external load on the soil pS in the derivations.

–5–
Heinz J. Priebe

1 p
fd = pC =
K 0C − WS WC WC AC 1 − AC A
1+ ⋅ +
K0C pC A p C pS

p C 1 2 + f ( µS , A C A )
=
pS K aC ⋅ f ( µS , A C A )

WC = Σ( γ C ⋅ ∆d ) , WS = Σ( γ S ⋅ ∆d )

KoC = 1 − sin ϕC

The simplified diagram in Figure 3 considers the same bulk density γ for columns and soil which is
not on the safe side.Therefore for safety reasons, the lower value of the soil γ S should be considered
in this diagram always.

1
fd =
K 0 C − 1 Σ( γ S ⋅ ∆d )
1+ ⋅
K0C pC

1,3

[1 -- yy. .ΣΣ((γ
ffdt == 11 // [1 γγBS ✎
.∆
∆ d)
t) // p]
p]
1,1
ϕϕSC == 45.0°
45.0°
Influence Factor y

ϕ
ϕ CS =
= 42.5°
42.5° µµSB == 1/3
1/3
0,9
ϕϕS == 40.0°
40.0°
C

ϕϕSC ==37.5°
37.5°
0,7
ϕϕSC == 35.0°
35.0°

0,5

0,3
1 2 3 4 5 6 7 8 9 10
Area Ratio A/AC

Figure 3:
3 Determination of the depth factor

–6–
The Design of Vibro Replacement

5 Compatibility Controls

The single steps of the design procedure are not connected mathematically and they contain
simplifications and approximations.Therefore, at marginal cases compatibility controls have to be
performed which guarantee that no more load is assigned to the columns than they can bear at
all in accordance with their compressibility.
At increasing depths, the support by the soil reaches such an extent that the columns do not
bulge anymore. However, even then the depth factor will not increase to infinity as results from
the assumption of a linearly decreasing pressure difference.Therefore, the first compatibility control
limits the depth factor and thereby the load assigned to the columns so that the settlement of
the columns resulting from their inherent compressibility does not exceed the settlement of the
composite system. In the first place this control applies when the existing soil is considered pretty
dense or stiff.

D C DS
fd ≤
p C pS

0,20

0,16
Influence Factor y

0,12
ϕ S = 35.0°
ϕ C = 35.0°

0,08
ϕϕCS == 37.5°
37.5°

ϕ = 40.0°
ϕCS = 40.0°
µµsB== 1/3
1 /3

ϕ
ϕ S == 42.5°
42.5° fftd <
< yy. .DECS/ /DESB, ,
0,04 C

ϕCS = 45.0°
= 45.0° be r ff t >>1 1
abut
d

0,00
1 2 3 4 5 6 7 8 9 10

Area Ratio A /AC

Figure 4:
4 Limit value of the depth factor

The maximum value of the depth factor can be drawn also from the diagram in Figure 4 4. By the
way, a depth factor fd < 1 should not be considered, even though it may result from the calculation.
In this case the second compatibility control is imperatively required which relates to the maximum
value of the improvement factor. In a certain way this control resembles the first one. It guarantees
that the settlement of the columns resulting from their inherent compressibility does not exceed
the settlement of the surrounding soil resulting from its compressibility by the loads which are

–7–
Heinz J. Priebe

assigned to each. In the first place this second control applies when the existing soil is encountered
pretty loose or soft.

AC DC
n max = 1 + ⋅( − 1)
A DS

It has to be observed that the actual area ratio AC /A has to be appointed in the formula and not
––––
the modified value AC / A. Because of the simple equation, an independent Diagram is not required.

6 Shear Values of Improved Ground

The shear performance of ground improved by vibro replacement is outmost favourable. Whilst
under shear stress rigid elements may break successively, stone columns deform until any overload
has been transferred to neighbouring columns. For example, a landslide will not occur before the
bearing capacity of the total group of columns installed has been activated. The stone columns
receive an increased portion of the total load m thereby which depends on the area ratio AC /A
und the improvement factor n.

m = (n − 1 + A C A) n

Simplifying, the recommended design procedure does not consider the volume decrease of the
surrounding soil caused by the bulging of the columns. Therefore and particularly at a high area
ratio, the soil receive a greater portion of the total load than actually calculated. In order not to
overestimate the shear resistance of the columns when averaging on the basis of load distribution
on columns and soil, the proportional load on the columns has to be reduced. The following
approximation seems to be adequate:

m′= ( n − 1) n

The diagram in Figure 5 shows in solid lines the proportional load of the columns m´ and in
dashed lines the not reduced one m.
According to the proportional loads on columns and soil, the shear resistance from friction of
the composite system can be readily averaged.

tan ϕ = m′⋅ tan ϕ C + (1 − m′ ) ⋅ tan ϕS

Since in most practical cases possible lines of sliding cover different depths which is difficult to
survey, it is recommended to consider the depth factor in clear-cut cases only, i. e. to calculate
usually with a load portion of the stone columns m1´ related to n1 and not with m2´ related to
the increased factor n2 = fd·n1.
The cohesion of the composite system depends on the proportional area of the soil.

c = (1 − A C A ) ⋅ cS

–8–
The Design of Vibro Replacement

The installation of stone columns possibly creates damages to the soil structure which are difficult
to survey. For safety reasons, it seems to be advisable to consider the cohesion also proportional
to the loads, i. e. pretty low, although this proposal is not based on soil mechanical aspects.

c′= (1 − m′ ) ⋅ cS

1,0
Dashed Lines:
m = (n - 1 + A C /A) / n
0,8
µµ sB == 11/3
/3
Proportional load m

0,6 ϕϕCS==45.0°
45.0°

ϕϕC S= 42.5°
= 42.5°

ϕ S = 40.0°
0,4 ϕC = 40.0°
ϕ = 37.5°
ϕ C S= 37.5°
Solid Lines: ϕϕCS == 35.0°
35.0°
0,2
m = (n - 1) / n

0,0
1 2 3 4 5 6 7 8 9 10
Area Ratio A/AC

Figure 5:
5 Proportional load on stone columns

7 Settlement of Single and Strip Footings

It is not (yet) possible to determine directly the performance of single or strip footings on vibro
replacement. The design ensues from the performance of an unlimited column grid below an
unlimited load area. The total settlement s∞ which results for this case at homogeneous conditions,
is readily to determine on the basis of the foregoing description with n2 as an average value over
the depth d.

d
s∞ = p ⋅
DS ⋅ n 2

Diagrams which are given in Figure 6 and Figure 7 7, allow to conclude from this value the sett-
lements of single or strip footings on groups of columns. These diagrams - with the diameter of
the stone columns D as one parameter - are based on numerous calculations which considered
load distribution on one side and a lower bearing capacity of the outer columns of the column
group below the footing on the other side.

–9–
Heinz J. Priebe

0,8
Settlement Ratio s/s∞

1600
900

No. of Stone Columns


0,6
400
225
0,4 100
64
36
0,2 16
9
4
1
0
0 4 8 12 16 20 24 28 32
Depth/Diameter Ratio d/D

Figure 6:
6 Settlement of single Footings

0,8
Settlement Ratio s/s∞

0,6
10 No. of Stone Column Rows
8
6
0,4
4
3
2
0,2
1

0
0 4 8 12 16 20 24 28 32

Depth/Diameter Ratio d/D

Figure 7:
7 Settlement of strip Footings

The diagrams do not refer directly to footing extensions as to be expected. However, there exists
an indirect reference in that the grid area A required to determine the improvement factor n, has

– 10 –
The Design of Vibro Replacement

to be derived as quotient of the footing area and the number of columns. For example, the
settlement reduction which a larger footing experiences normally at the same load, is compensated
widely by the lower improvement factor which results from an increased area ratio as follows
from a larger footing area on the same number of stone columns. The approximation given for
the diagrams by this assumed compensation seems to be acceptable for usually considered area
ratios, i. e. up to some A/AC = 10.
Quite clear that the diagrams are valid for homogeneous conditions only and refer to the settlement
s up to a depth d which is the second parameter counting from foundation level. The settlement
∆s of any layer at any depth below the footing has to be determined as difference of the settlements
up to the depths dl and du of the lower and upper bound of the layer concerned with n2 as an
average value over its thickness ∆d.

p
∆s = ⋅ [(s s ∞ )l ⋅ d l − (s s∞ )u ⋅ d u ]
DS ⋅ n 2

Since n2 increases with depth on one side due to the depth factor, but becomes less significant
with depth on the other side due to the load distribution of a limited footing, it is required even
at homogeneous conditions to subdivide greater depths.This avoids settlements being too liberally
estimated.

8 Bearing Capacity of Single and Strip Footings

A simple method to estimate the bearing capacity of single and strip footings on vibro replace-
– – of
ment exists by determining at first a fictitious width b of the footing, using the friction angle ϕ
the improved soil below the footing and the friction angle ϕS of the untreated soil on the outside,
which would develop - calculated on the basis of the friction angle ϕS of the untreated soil only -
in case of ground failure the same line of sliding outside of the improved area as the actual footing
at actual conditions. If the border line of treatment coincide with the edge of the footing - being
usually the case but not necessarily - the following formula results:

b = b ⋅ e[ ] ⋅ sin( 45 + ϕ 2 ) ⋅ sin( 90°−ϕ S )


arc( 45°−ϕ 2 )⋅tan ϕ − arc ( 45°−ϕ S 2 )⋅tan ϕ S

sin(90 − ϕ ) sin( 45°+ ϕ S 2 )

Then, for this fictitious width the bearing capacity is determined by using the friction angle of the
untreated ground ϕS and an averaged cohesion according to the proportion of fictitious footing
width and failure width outside of the footing. In pure cohesive soil the failure width equals the
footing width, thus leading to an average cohesion of c´´ = (c´ + cS) / 2.
For foundations on layered ground the shear values change with depth also.The determination of
the bearing capacity, e. g. according to the German Standard DIN 4017, becomes rather complicated
with the fictitious width since this width changes at each layer.

– 11 –
Heinz J. Priebe

A practical approximation can be achieved as follows. At first, safeties η0 and maximum depths of
ground failure lines dGr,0 are calculated applying one after another the soil parameters of every
individual layer, e. g. according to DIN 4017.

η0 = σ0 f p σ0f = ( cS ⋅ N c ⋅ νc + q ⋅ N d ⋅ νd + γ S ⋅ b ⋅ N b ⋅ νb ) ⋅ b b

d Gr , 0 = b ⋅ sin( 45°+ ϕ S 2 ) ⋅ e[ arc ( 45°+ϕ S 2 )⋅tan ϕ S ]

In a second step, the final safety η and maximum depth dGr is averaged successively with the
values of the individual layers as long as dGr(n-1) exceeds du(n) being the upper bound of the layer
concerned (dl(n) being the lower bound).

[ ]
d o( n )
η( n ) = η0( n ) + η( n −1) − η0( n ) ⋅
d Gr ( n −1)

[ ]
d o( n )
d Gr ( n ) = d Gr , 0( n ) + d Gr ( n −1) − d Gr , 0 ( n ) ⋅
d Gr ( n −1)

n≥2 η(1) = η0(1) d Gr (1) = d Gr , 0(1) When d Gr ( n −1) > d l( n ) then d Gr ( n −1) = d l ( n )

Though little bit uncomfortable, this procedure can still be performed manually in contrast to the
iteration as outlined in DIN 4017.The results of both the procedures do not differ much.

9 Liquefaction Potential of Improved Ground

Vibro replacement is suitable particularly for ground improvement in seismic areas since stone
columns possess a certain flexibility on one side and prevent liquefaction on the other side. The
stabilizing effect results from the frictional resistance of the columns which carry a considerable
amount of the external load and of the weight of the soil, and their capability to reduce excess
porewater pressure in the soil - at least in close vicinity - almost instantly.The steep reduction of
porewater pressure towards the column is in so far important as it creates kind of a filter cake
effect which maintains the lateral support required for the bearing capacity of the columns and
which prevents a higher degree of soil infiltration into the columns although the column material
does not fulfill any established filter criteria.
The complex conditions in a seismic event are investigated frequently for more or less
homogeneous ground. Nevertheless, practical criteria to evaluate the liquefaction potential were
developed rather empirically. For vibro replacement although carried out already many times
against earthquake vibrations, even an empirical evaluation is difficult since - fortunately - no
damages have been observed so far.
Usually, safety against liquefaction is concluded from the comparison of so-called cyclic stress
ratios, namely the one which is provided by the soil on the basis of its density and the one which
probably develops in a seismic event.

– 12 –
The Design of Vibro Replacement

For a rough estimation of the efficiency of vibro replacement it is proposed to reduce the cyclic
stress ratio probably developed in a seismic event, in the same ratio as the load on the soil
between the columns is reduced by vibro replacement, i. e. to use a corresponding reduction
factor α.
α = pS p = 1 n

Such a reduction seems to be adequate with regard to the favourable performance of vibro
replacement in seismic events. However, from soil mechanical aspects this is not proved and has
to be verified ultimately by the increasing number of projects carried out world-wide.
For similar reasons as outlined at the determination of the shear values, it is recommended to
use in the formula n1 rather than n2.
A diagram for the reduction factor α is given in Figure 8
8.

0,8
ϕϕCS== 35.0°
35.0°
Stress Ratio pS /p

ϕϕCS==37.5°
37.5°
0,6
ϕϕC S==40.0°
40.0°

ϕϕCS== 42.5°
42.5°
0,4
ϕ Cϕ S= =45.0°
45.0°

µµSΒ == 1/3
1 /3
0,2

0
1 2 3 4 5 6 7 8 9 10

Area Ratio A/AC

Figure 8:
8 Residual pressure on the soil after vibro replacement

10 Case Study Worked Example

The design method has been used already frequently in determining the expected behaviour of
structures on treated ground. However, in most cases the application is based on parameters
indirectly derived from field tests or even just assumed. As long as the actual performance of
vibro replacement excels such forecasts, more accurate verifications are usually omitted.
Some full scale field experiments about vibro replacement which comprise measurements beyond
common practice are outlined in [2][2]. For example, enough details of a tank foundation at Canvey
Island are given so that the design method can be applied and the results verified.

– 13 –
Heinz J. Priebe

The diameter of the tank concerned is 36 m. It is founded on a pad of approximately 1 m thickness


above soil reinforced by 10 m long stone columns in a grid with triangular spacing of 1.52 m and
an average diameter of 0.75 m measured near surface. Including some 0.4 m of top soil the
treated strata consist up to 9 m depth of silty and clayey soil occasionally with pockets of peat
followed by medium dense silty fine sand in which the columns are embedded. Referred to depths,
the given coefficients of volume change m v and the constrained moduli DS (= 1 / mv) as used in
the design computations are as follows:

Depth [m] mv [m²/MN] DS [MN/m²] Remarks


-1.0 – 50 pad
0.0 20 top soil
0.4 0.8 - 0.5 2 soft soil
1.0 1 very soft soil
1.6 1.2 - 0.5 1 very soft soil below ground water
8.2 0.3 - 0.06 10 firm soil
9.0 20 medium dense sand

At full loading of 130 kN/m² settlements were observed in the range of some 40 cm.
A computation according to the design method (s. appendix) shows a final settlement of
approximately 38 cm. Taking into consideration the pockets of peat or a possible reduction of
column diameter with depth, the value would be higher and in really good agreement.
The improvement factors n as computed on the basis of formulae, can be taken readily also from
the diagrams as follows with reference to the first layer below the ground water table which
contributes most to the settlements:

A /AC = 4.53 → Fig. 1 → n0 ≈ 2.35


DC/DS = 100 → Fig. 2 → ∆ A/AC ≈ 0.05 → A /AC = 4.58
A /AC = 4.58 → Fig. 1 → n1 ≈ 2.30

A /AC = 4.58, Σ (γ·d) = 19 ·1.0 + 18 · 0.4 + 16 · 0.6 + 15·0.6 + 5 · 6.6/2 = 61.3 kN/m²,
p = 130 kN/m² → Fig. 3 → fd ≈ 1.38 → n2 = fd·n1 = 3.17

The discrepancy to the computed value of n2 = 2.94 is due to the difference between formulae
and diagram as outlined in paragraph 4.

– 14 –
The Design of Vibro Replacement

11 Conclusions

Out of the deep vibratory compaction techniques vibro replacement covers the widest range
with regard to the application in different soils. Whilst vibro compaction is restricted to compactible
sand and gravel, the application of vibro replacement extends principally over the total range in
grain size of loose soils. Even in most of the noncohesive natural soils suitable for vibro compaction,
backfilling with coarse grained material is recommended to increase the compaction efforts - and
this means stone column installation. Pure vibro compaction has advanced just lately at gigantic
artificial deposits in different coastal regions of the world.
Notwithstanding the importance of vibro replacement, the efficiency of stone columns in soil
improvement must not be overestimated. As long as the existing soil is suitable to be densified,
this should be the preceding aim of any deep compaction treatment including vibro replacement.
However, the achievable densification depends on too many parameters to be calculable. On the
contrary the improving effect of stone columns - possibly supplementary to an achieved densification
- can be determined pretty reliably.
The application of vibro replacement which was introduced end of the fifties, relied for a long
time upon the experience of the contractors. Not before the middle of the seventies first theoretical
approaches were submitted. In its fundamentals also the design method outlined afore originates
from this time. It has proved its reliability since then. Subsequent supplements imply refinements
or extensions of the application range but not a radical alteration on the fundamentals. In respect
of the complexity of the matter the design criteria have the advantage to be easy to use and to
cover in a closed package all cases practically occurring.

References

[1] Kirsch, K.: Die Baugrundverbesserung mit Tiefenrüttlern, 40 Jahre Spezialtiefbau: 1953-1993,
Festschrift,Werner-Verlag GmbH, Düsseldorf, 1993.
[2] Greenwood, D. A.: Load Tests on Stone Columns, ASTM Publication STP 1089, Deep
Foundation Improvements: Design, Construction, and Testing, 1991.

Publications of the author to the design method:


[3] Abschätzung des Setzungsverhaltens eines durch Stopfverdichtung verbesserten Baugrundes,
Die Bautechnik 53, H.5, 1976.
[4] Zur Abschätzung des Setzungsverhaltens eines durch Stopfverdichtung verbesserten Bau-
grundes, Die Bautechnik 65, H.1, 1988.
[5] Abschätzung des Scherwiderstandes eines durch Stopfverdichtung verbesserten Baugrundes,
Die Bautechnik 55, H.1, 1978.
[6] Vibro Replacement Design Criteria and Quality Control, ASTM Publication STP 1089, Deep
Foundation Improvements: Design, Construction, and Testing, 1991.
[7] The Prevention of Liquefaction by Vibro Replacement, Proc. of the Int. Conf. on Earthquake
Resistant Construction and Design, 1990 Balkema, Rotterdam.
[8] Die Bemessung von Rüttelstopfverdichtungen, Die Bautechnik 72, H.3, 1995

– 15 –
Heinz J. Priebe
Appendix

Keller Grundbau GmbH


Kaiserleistr. 44, 63067 Offenbach, Tel. 069/8051210, Fax. 069/8051221
Program VIBRI, Version 950904, Copyright by KELLER Grundbau GmbH

Vibro Replacement at Canvey Island, Reported 1991 by Greenwood


***********************************************************
Evaluation of the Soil Improvement by Vibro Replacement
acc. to Priebe, H.: Die Bautechnik 72, 3/1995
below an Area Load on a Regular Triangular Column Grid

Foundation Pressure 130.00 kN/m² Column Material


Column Distance 1.52 m Unit Weight 19.00 kN/m³, below 1.60 m Depth 12.00 kN/m³
Row Distance 1.32 m Constrained Modulus 100.00 MN/m²
Grid Area 2.00 m² Friction Angle 40.00 Degrees
Load Level -1.00 m Press. Coefficient .22
Column Depth 10.00 m
Considered Depth 20.00 m

Subsoil Strata
Ground Water Table 1.60 m
No. TopL. Dia. A/AC DS DC/DS gamma my phi c
[m] [m] [MN/m²] [kN/m³] [degree] [kN/m²] Top L. = Top Level of Stratum Concerned
Dia. = Column Diameter
1 -1.00 .00 **** 50.00 2.00 19.00 .33 35.00 .00
A = Grid Area Resp. Reference Area
2 .00 .75 4.53 20.00 5.00 18.00 .33 25.00 5.00
AC = Cross-sectional Area of Column
3 .40 .75 4.53 2.00 50.00 16.00 .33 .00 25.00
DC = Constrained Modulus of Backfill
4 1.00 .75 4.53 1.00 100.00 15.00 .33 .00 20.00
DS = Constrained Modulus )
5 1.60 .75 4.53 1.00 100.00 5.00 .33 .00 20.00
gamma = Unit Weight )
6 8.20 .60 7.08 10.00 10.00 7.00 .33 .00 30.00
my = Poisson’s Ratio ) of Soil
7 9.00 .60 7.08 20.00 5.00 9.00 .33 30.00 .00
phi = Friction Angle )
8 10.00 .00 **** 20.00 5.00 9.00 .33 30.00 .00
c = Cohesion )
9 20.00 .00 **** 20.00 5.00 9.00 .33 30.00 .00

Soil Improvement
No. n0 d(A/AC) n1 m1 phi1 c1 fd n2 m2 phi2 c2
[degree] [kN/m²] [degree] [kN/m²]
1 Layer without Stone Columns!
2 2.34 1.17 2.01 .50 33.16 2.49 ***** 1.88 .47 32.67 2.66
3 2.34 .09 2.31 .57 25.41 10.84 1.16 2.68 .63 27.73 9.34
4 2.34 .05 2.32 .57 25.54 8.61 1.21 2.82 .65 28.44 7.09
5 2.34 .05 2.32 .57 25.54 8.61 1.27 2.94 .66 28.98 6.80
6 1.78 .52 1.72 .42 19.35 17.45 1.24 2.13 .53 24.04 14.05
7 1.78 1.17 1.65 .40 34.25 .00 ***** 1.57 .36 33.90 .00
8 Layer without Stone Columns!

The Proportional Loads on Columns are Approximated to m = 1 - 1/n Vibro Replacement


by the Keller Group
n0 = Basic Improvement Factor
- the experienced contractors
d(A/AC) = Addition to the Area Ratio (Column Compressibility)
which invented and developed
n1 = Improvement Factor (with Column Compressibility)
the basic features of the deep
(—> Recommended for Failure Analyses if n1 < n2)
vibratory compaction methods.
fd = Depth Factor (Overburden Constraint)
(***** —> Overridden by Control Checking!)
VIBRI - The only software for the
n2 = Improvement Factor (Add. with Overburden Constraint)
design of vibro replacement
m1,2 = Proportional Load on Columns )
developed by the author (Priebe)
phi1,2 = Friction Angle of Compound ) Attributable to n1 resp. n2 of the design method.
c1,2 = Cohesion of Compound )
- user friendly
by graphically supported input
Settlement Depth Infinite w/o Over-
Load Area Impr. burden - easy to survey
by alphanum. and graphic output
[m] [cm] [cm] [kN/m²]
-1.00 .26 .26 .0

.00 .14 .26 19.0 For details please contact


.40 1.37 3.66 26.2 Keller Grundbau GmbH
1.00 2.45 6.90 35.8 Technical Department
Kaiserleistr. 44
1.60 25.81 75.93 44.8
63067 Offenbach/Main
8.20 .48 1.03 77.8
Tel: 069-8051-218
9.00 .41 .65 83.4
Fax: 069-8051-221
10.00 6.46 6.46 92.4
GERMANY
37.37 95.14

– 16 –
Vibro Replacement
to Prevent
Earthquake Induced
Liquefaction

Dipl.-Ing. Heinz J. Priebe


Keller Grundbau GmbH

Presented by
Keller Grundbau GmbH
Kaiserleistr. 44
D-63067 Offenbach Proceedings of the Geotechnique-Colloquium
Tel. 069 / 80 51 - 0 at Darmstadt, Germany, on March 19th, 1998
Fax 069 / 80 51 - 244 (also published in Ground Engineering, September 1998)
E-mail [email protected]
www.KellerGrundbau.com Technical paper 12- 57 E
Vibro Replacement
to Prevent Earthquake Induced Liquefaction

Dipl.-Ing. Heinz-Joachim Priebe


Keller Grundbau GmbH, Offenbach

1 Introduction

Fortunately, Germany has only a few, localised regions that are at risk from earthquakes, and even
here the risk is small. The situation is, however, quite different in the south and south-east of
Europe. For example, Assisi in central Italy was hit by an earthquake last year.
Elsewhere, there are regions on earth, often densely populated, where the danger of both severe
and very severe earthquakes is quite real. This is of primary concern in south-east Asia and other
areas of the Pacific Rim, where approximately 85 per cent of all the world´s seismic movements
are recorded.
When a building becomes unusable, it does not necessarily mean that the structure has suffered
severe damage or a total collapse as a result of seismic shaking. Frequently the ground under and
around the building fails and the facility is lost, even though the structure itself might have sur-
vived with minor damage. In these cases, failure results from liquefaction.

2 The Phenomenon of Liquefaction

Earthquake induced soil liquefaction occurs relatively frequently in areas of more or less level
ground. The term phenomenon is used not because the events seldom occur, but because the
expression highlights the often astonishing result, such as the overturning of whole apartment
blocks or the floating of buried facilities. Figure 1a shows a capsized apartment block in Niigata,
Japan, resulting from the famous 1964 earthquake (Herzog, 1980).

Figure 1a: Capsized apartment


block in Niigata,
Japan, 1964

–1–
Heinz-Joachim Priebe

Figure 1b: Subsided mosque near Sungai Penuh, Figure 1c: Tilted tank in Kobe,
Indonesia,1994 Japan, 1995

Other spectacular and more recent examples are shown in figure1b, where a mosque in Indone-
sia subsided during an earthquake in 1994, and in figure 1c, where a tank in Kobe, Japan, tilted at
the 1995 earthqake (Wenk /Schwarz, 1995).
Liquefaction of this type only occurs in loose to medium dense, saturated soils with fairly uni-
form grain size distributions, covering the silty sandy range. The most critical soil is fine sandy
grained with some silt content. Figure 2 shows the bandwidth susceptible to liquefaction. It ex-
tends from medium to coarse sand to medium to coarse silt. In loose to medium dense condi-
tions, the dynamic forces of a seismic event lead to an adjustment of the grain structure to a
denser state. Initially, if the soil does not drain sufficiently, due to a low permeability or long

Clay Silt Sand Gravel St.


100

90
Percentage Passing [by Weight]

80

70
Vibro Replacement
60
Liquefiable
50
Soils
40
Vibro Compaction
30

20

10

0
0.002 0.006 0.02 0.06 0.2 0.6 2 6 20 60
Grain Size [mm]

Figure 2: Application ranges of the deep vibratory compaction techniques

–2–
Vibro Replacement to Prevent Earthquake Induced Liquefaction

drainage paths, effective stresses in the grain structure start to pass to the pore water. The pres-
sure in the pore water, therefore, rises accordingly, and the previously existing shear resistance of
the soil diminishes.
In a limit state, the subsoil behaves like a liquid, and loses its bearing capacity. Even if this state
lasts only for a short period, extreme deformations may occur, which have little to do with nor-
mal ground failure.

3 Evaluation of the Liquefaction Potential

In the mid-sixties, extensive efforts were beginning to be made to deal with the engineering
problems of soil liquefaction. The reason might have been the 1964 Niigata earthquake, where
the phenomenon occurred extensively. However, there were also demands to reduce the risk of
liquefaction at the growing number of industrial plants, not only nuclear power stations but also
barrages, refineries or chemical plants, where construction or foundation failure could have cata-
strophic consequences.
Options for evaluating the liquefaction potential at a given earthquake magnitude can be roughly
subdivided into three substantial groups: theoretical approaches, laboratory tests and statistical
analyses.
With theoretical approaches, the problem that arises is how to consider the prevailing ground
conditions properly. In the second group, laboratory tests can at least deal with the material
concerned, but problems lie in the difficulty of realistically modelling natural conditions. Although
the soil density is the main parameter in evaluating the liquefaction potential, the influence of
ageing and naturally occurring anomalies must not be underestimated. This is equally applicable
when transmitting model behaviour to actual conditions.

10
Earthquake Intensity

X
Richter-Scale
MM-Scale (Modified Mercalli)
IX 9
MSK-Scale (Medwedew, Sponheuer, Karnik)
JMA-Scale (Japan Meteorological Agency)
VIII Earthquake Zones acc. to DIN 4149 8
Earthquake Zones acc. to UBC (Uniform Building Code)
Magnitude M (Richter-Scale)

VII 7

VI 6

V 5

4 IV 4
Earthquake Zones

3 III 3

2 II 2

1 I 1

0
0
0,01 0,02 0,03 0,04 0,05 0,1 0,2 0,3 0,4 0,5 1
Earthquake Acceleration in g

Figure 3: Correlation between common earthquake scales and earthquake acceleration

–3–
Heinz-Joachim Priebe

So far, the simplest and probably most reliable method seems to be the evaluation of the soil
liquefaction potential on the basis of statistical analyses. For this, the forces expected during a
seismic event are compared with the forces that the subsoil under consideration can actually
resist. Generally, the maximum surface acceleration on level ground is used as the characteristic
value for the forces developed by an earthquake. In figure 3, common earthquake scales, based on
subjective perceptions and damage observations, are correlated to each other and to the more
objective Richter scale. Figure 3 also gives an idea of the maximum surface acceleration which is
attributable to the various scales and, as such, it can be used as support for design purposes. The
diagram shows values at the conservative side (Klein, 1990), at least for higher intensities. This is
intentional for safety reasons, as the statistical analyses do not contain any further safety factors.
First, in order to evaluate the liquefaction potential, the Seismic Stress Ratio SSR = τ h /σ´v0 devel-
oped in the field at level ground conditions is determined. This is, depending on depth, the ratio
between the shear force created by the assumed earthquake magnitude and the effective over-
burden pressure (Seed/Idriss/Arango, 1983).

τh a σ
= 0.65 ⋅ max ⋅ v 0 ⋅ rd
σv′ 0 g σv′ 0
where
τh = seismic shear force [kN/m²]

σv0 = total overburden pressure [kN/m²]

σ´v0 = effective overburden pressure [kN/m²]

a max = maximum earthquake acceleration [m/sec2]

0.65 = average value of accelaration in comparison to amax

g = acceleration due to gravity [m/sec2]

rd = reduction value depending on depth [≈ 1 – 0.012 · z]

z = depth [m]

The stress ratio that the soil can resist, is found by statistical analyses of a multitude of soundings
at earthquake affected subsoils. The first of these analyses, related to a magnitude of M = 7.5, was
originally based on Standard Penetration Tests (Seed/Idriss/Arango, 1983). It was then continued,
attuned to the more reliable Static Cone Penetration Tests and refined with regard to the silt
content of soils, by other investigators (Robertson/Campanella, 1985; Stark /Olson, 1995).
Figure 4 shows the relationship between seismic stress ratio and modified cone resistance qc1 at
the boundary state.

–4–
Vibro Replacement to Prevent Earthquake Induced Liquefaction

0,5

Sandy Silt Silty Sand


0,4 D50 [mm] < 0.10 0.10 < D 50 [mm] < 0.25
F. C. [%] > 35 5 < F. C. [%] < 35

0,3 Magnitude M = 7.5

0,2
Clean Sand
0.25 < D 50 [mm] < 2.0
F. C. [%] < 5

0,1

qc1 = 180 qc / (80 + σ´v0)

0
0 2 4 6 8 10 12 14 16
Corrected CPT Tip Resistance qc1 [MPa] Stark and Olson, 1995

Figure 4: Relationship between seismic stress ratio and CPT tip resistance for sandy soils

1. 8
q c1 = Cq ⋅ q c Cq ≈
0. 8 + σ ′v 0 100
where qc = cone resistance [MPa]

The evaluation proposed in the German manual Grundbau -Taschenbuch coincides principally with
the procedure outlined above (Klein, 1990).
The method was further modified with the introduction of the friction ratio between sleeve
friction and cone resistance. In this case, the determination of the grain size distribution can be
omitted (Suzuki/Koyamada/Tokimatsu, 1997). In the publication concerned a slightly modified for-
mula, as given below, is used to determine the relevant seimic stress ratio (abbreviations as be-
fore).
τh a max σv 0
= 01
. ⋅ ( M − 1) ⋅ ⋅ ⋅ (1 − 0.015 ⋅ z )
σ ′v 0 g σ′v 0
where M = earthquake magnitude
In this formula, the average value of the acceleration is related to the magnitude M, and the re-
duction value rd is slightly modified. The diagram shown in figure 5 should be used with modified
parameters given below, where sleeve friction and cone resistance is considered via a soil behavior
type index Ic .

–5–
Heinz-Joachim Priebe

0,5
4
Q = (qc – σv0) / σ´v0
3 F = 100 fs / (qc – σv0) [%]
0,4
Ic = [(3.47 – log Q) + (log F + 1.22)2 ]0.5
2
F(Ic)

2
Shear Stress Ratio τ d / σ´v

0,3 1

0
1 1,2 1,4 1,6 1,8 2 2,2 2,4 2,6
0,2 Soil Behaviour Type Index, Ic

0,1

0
0 2 4 6 8 10 12 14 16 18 20
Adjusted Tip Resistance, qt1 x F(Ic) [MPa] Suzuki et al., 1997

Figure 5: Relationship between seismic stress ratio and CPT tip resistance

Q = ( q c − σ v 0 ) σ′v0

F = 100 ⋅ fs (q c − σv 0 ) [%] fs = sleeve friction

[
Ic = (3.47 − log Q ) + ( log F + 1.22 )
2
]
2 0. 5

In conclusion, it should be noted that the evaluation of the liquefaction potential by comparing
the expected forces of an earthquake with those that the soil can actually resist, is not a safety
analysis in the customary way, as the magnitude of an earthquake is a random variable.

4 Performance of Vibro Replacement

Vibro Replacement is more suitable than most other foundation techniques in preventing lique-
faction during a seismic event. Generally, the problem can be solved by three alternatives, for
which vibro replacement offers the best conditions: soil compaction, drainage and increase in
shear resistance.
Which of these prevails with regard to vibro replacement, depends on the grain size distribution
of the soil concerned. Figure 2 shows a dashed transition zone relating to deep vibratory
compaction techniques, which is located more or less in the middle of the area susceptible to
liquefaction, and plays a dominant role.

–6–
Vibro Replacement to Prevent Earthquake Induced Liquefaction

It is assumed here that the two main compaction techniques, vibro compaction and vibro re-
placement, are already well known. Therefore, only the main features are described here, in order
to understand their performance in soils susceptible to liquefaction. Vibro compaction
increases soil density, while vibro replacement, in its pure application, reinforces cohesive, non-
compactible soils, using load carrying columns of imported coarse aggregate such as gravel or
crushed material. Often, vibro replacement has the combined effect of considerable densification
of the surrounding soil by vibrational effects during the installation of stone columns. As in static
cases, it is also useful in dynamic events to consider both components of improvement sepa-
rately, namely the soil densification and the improvement by the stone columns.
Soils with a grain size distribution curve entirely outside the transition zone at the sandy, gravelly
side are suitable for treatment solely by vibro compaction. If the distribution curve is totally within
the dashed zone, or partly within it with the remainder on the coarse side, it is advisable to add
backfill material, that is, install a stone column. The reason for this is to improve transmission of
vibrations via the imported material, as the main target is still the densification of the surround-
ing soil. In all other cases, where the grain size distribution curve is partly or entirely to the left
of the transition zone, that is, in the silty clayey area, substantial improvement in the densification
due to vibration can not be expected, and stone columns have to be installed which provide
considerable improvement due to their stiffness and shear resistance.
Therefore, as shown in figure 2, the transition zone between vibro compaction and vibro replace-
ment subdivides the range of soils susceptible to liquefaction into three parts. Generally, as de-
scribed above, for soils that are within the sandy range, vibro compaction without imported backfill
material is suitable for preventing earthquake induced liquefaction. This can be substantiated by
cone penetration tests as outlined in section 3 above. In addition, in both the other cases, where
vibro replacement is either recommended or definitely required, the improvement by densification
can also be substantiated by the same method. However, if the favourable effects of the stone
columns with respect to drainage capability and increase in shear resistance are ignored, the re-
sult is a very conservative treatment design.
The type of drainage which contributes considerably to the reduction of pore water pressure
during the usually very short period of an earthquake, and which thereby reduces the risk of
liquefaction, can only be expected in densifiable soils where vibro compaction alone is recom-
mended. In finer grained soils with relatively low permeability, substantiating a reduction in
porewater pressure, and hence proving that the risk of liquefaction has been reduced, becomes
difficult. However, even if a considerable reduction in the pore water pressure cannot be proved,
even at short distances from the column, the high permeability of the columns themselves is
decisive for their bearing performance. A high gradient at the periphery of the columns provides
the required lateral support, and hence the bearing capacity, even at moments where the soil
between the columns could tend to liquefy.
Columns which are laterally supported in the limiting state by the difference in hydraulic pres-
sure provide considerable shear resistance, especially due to load concentration. However, com-
pared to piles, they have the advantage of high flexibility which can absorb the amplitudes occur-
ring during an earthquake, without losing their bearing capacity. This implies that seismic shocks
are not reduced significantly by vibro replacement. Therefore, the treatment does not necessarily
provide protection for the buildings, but is primarily concerned with the prevention of liquefac-
tion. Any conclusion should not discount that in some cases, buildings which might have other-
wise been destroyed by the shaking of non-liquefied soil have remained relatively undamaged
solely because of liquefaction.

–7–
Heinz-Joachim Priebe

Since it is difficult to quantify the favourable effects of stone columns in seismic events, in some
cases they were not considered in the design at all. In some other cases, where the densification
of the soil just missed the specifications, they were given a qualitative consideration only, in the
sense that they would compensate for the deficit in soil densification. Such an approach is not
satisfying and even rough approximations on a static basis are preferable.

5 Evaluation of Liquefaction Potential with Vibro Replacement

It is not possible to estimate, by statistical analyses, the extent to which the risk of liquefaction is
reduced by vibro replacement. Fortunately, no failure has occurred so far, even though the appli-
cation of this method of improvement has been steadily growing. Nevertheless, such analyses for
vibro replacement would be difficult to be implemented due to additional variable parameters
attributable to the arrangement of the columns.
As the existing analyses collected so far have already proved reliable, a different approach can be
made in the design in order to include them. The question does not then become, to what ex-
tent have the shear resistance and load bearing capacity been changed by the stone columns, but
rather, which part of the forces exerted by an earthquake are borne by the columns without any
damages? Such a step, namely reducing the acting forces instead of increasing the resisting ones,
is permissible as long as the evaluation, as previously mentioned, is not a customary determina-
tion of safety.
It is difficult to determine what amount of the acting forces in a seismic event is taken by the
stone columns. Evaluations on the basis of computer simulations or theoretical approaches are
more suitable than laboratory tests, in which possibilities are limited. A relatively simple proce-

5
ϕ C = 45.0˚

ϕ C = 42.5˚ µS = 0,5
Improvement Factor n

4
ϕ C = 40.0˚

ϕ C = 37.5˚
3
ϕ C = 35.0˚

1
1 2 3 4 5 6 7 8 9 10
Area Ratio A / AC Priebe 1998

Figure 6: Design chart for vibro replacement

–8–
Vibro Replacement to Prevent Earthquake Induced Liquefaction

dure for the design of vibro replacement in static cases was outlined comprehensively in Ground
Engineering (Priebe, 1995). This method was introduced a while ago in Germany and has since
proved reliable. Static cases mostly comprise long-term processes, and therefore, as a general
approximation, a Poisson´s ratio of µ S = 1/3 was proposed.
On the other hand, for short-term seismic events, it seems more realistic to consider deforma-
tions of the soil with the volume remaining constant, that is, to calculate with µS = 0.5 which also
simplifies the formulae. In the above mentioned procedure the improvement factor n0 , which is
the basic value of improvement by vibro replacement, is determined initially using some
simplifications and approximations. It is shown in figure 6.

K aC = tan 2 ( 45o − ϕ C 2)
AC 1
n0 = 1+ ⋅ −1
A K aC ⋅ ( 1 − A C A )

where A = attributable area within the compaction grid


AC = cross section of stone columns
ϕC = friction angle of column material

The reciprocal value of this improvement factor is merely the ratio between the remaining stress
on the soil between the columns pS, and the total overburden pressure p taken as being uni-
formly distributed without soil improvement and, as such, can be used as a reduction factor. It is
simple to calculate this value or it can be taken from figure 7.

0,8
ϕ C = 35.0˚

ϕ C = 37.5˚
Stress Ratio pS / p

0,6
ϕ C = 40.0˚

ϕ C = 42.5˚
0,4
ϕ C = 45.0˚

µS = 0,5
0,2

0
1 2 3 4 5 6 7 8 9 10
Area Ratio A / AC Priebe 1998

Figure 7: Residual stress on soil between stone columns

–9–
Heinz-Joachim Priebe

On the understanding that the loads taken by the columns from both the structure and the soil
do not contribute to liquefaction, it is proposed to use this factor α to reduce the Seismic Stress
Ratio created by an earthquake, and hence evaluate the remaining liquefaction potential as out-
lined in section 3.
This procedure represents an approximation which, although not being completely satisfactory
from a geotechnical point of view, nevertheless does realistically consider the stabilising effect of
vibro compaction.

6 Case Histories

The first project were vibro replacement was used to reduce the risk of liquefaction in a seismic
event was at the waste water treatment plant of Santa Barbara, in California, USA. This project is
relatively well documented and especially interesting, as it has already withstood an earthquake
where the acceleration corresponded to the one considered in the design (Mitchell, 1986)
More recently, this improvement technique has been increasingly applied against the risk of earth-
quakes in a wide range of applications, from bridges and barrages to industrial and marine facili-
ties (Dobson, 1987).
To date, only a few facilities constructed on ground improved by vibro replacement have actually
been subjected to earthquakes. However, in such cases, no serious results have been observed,
and therefore, at least qualitatively, vibro replacement has proved to be a success (Mitchell / Wentz,
1991).
On a more recent project for oil storage tanks on the Black Sea coast in Georgia, vibro replace-
ment has been carried out to reduce both the settlements and the risk of liquefaction. A soft
silty clayey top layer, with thickness ranging from 1m to 4 m, is underlain by mostly sandy strata
with varying silt content up to a depth of approximately 22 m which, in turn, is underlain by
calcareous hard clay. The groundwater table lies at around 2 m below ground level.

The tanks, already under construction, have a diameter of 66 m and, in service condition, exert a
foundation pressure of up to 150 kN/m². With regard to earthquakes, the design was based on a
maximum acceleration of a max = 0.25 g which corresponds to a magnitude of M ≈ 6.5 according
to figure 3.
For the tank used in the design example below, the soft top layer was completely removed by
soil replacement up to a depth of 2.7 m. Vibro replacement was then carried out down to a
maximum depth of 11.5 m, with up to 1 m diameter columns placed on a square grid at 3 m
spacing.
Calculations for the design, utilising the relevant formulae and diagrams, were performed using a
computer program, and the results are shown in figure 8 below. The upper level of the layers
concerned is of main interest with regard to the required cone resistance. For the silty sands
beneath the soil replacement, fines contents of 15 per cent were assumed, and the bulk densities

– 10 –
Vibro Replacement to Prevent Earthquake Induced Liquefaction

estimated. The measured cone resistance decreased considerably at 8 m depth, most probably
due to a silt content higher than the 15 per cent considered, which is also demonstrated by
densification of the soil not being achieved by the treatment below this depth. Therefore, in this
instance, it is acceptable that the required cone resistance slightly exceeds the measured one.
The same also applies at the full depth of treatment. For the rest, the measured values are well
above those required and, therefore, there is no concern at all with regard to liquefaction of the
upper strata which are generally more crucial.

SUPSA Terminal, Tank D

Maximum Earthquake Acceleration amax = 0.25 g Magnitude M = 6,5


Parameters of Vibro Replacement Grid Size A 9.00 m²
Column Diameter D 1.00 m
Area Ratio AC /A 0.09
Backfill Material ϕ 42.50˚
Coeff. of Act. Earth Pressure KaS 0.19
Reduction Factor α 1) 0.71
1) Priebe, Kolloquium Darmstadt, 1998
General Parameters Ground Water Table 2.00 m
2) Stark and Olson, Journal of
Surface Load 0.00 kN/m² Geotechnical Engineering, 1995

2)
Calculation
F. C. Depth γ γ´ σv0 σ´v0 rd SSR α * SSR qc1 Cq qc, requ. qc, meas.
[%] [m] [kN/m³] [kN/m²] [MPa] [MPa]

5 0.00 19.0 11.0 0.0 0.0 1.00 0.000 0.000 0.0 2.250 0.0 0.0
5 2.00 19.0 11.0 38.0 38.0 0.98 0.159 0.113 6.8 1.525 4.5 10.0
15 2.70 18.0 9.0 52.7 45.7 0.97 0.181 0.129 4.7 1.432 3.3 10.0
15 4.00 18.0 9.0 77.4 57.4 0.95 0.209 0.148 5.5 1.310 4.2 10.0
15 8.00 18.0 9.0 153.4 93.4 0.90 0.241 0.172 6.4 1.038 6.2 6.0
15 11.50 18.0 9.0 219.9 124.9 0.86 0.247 0.175 6.6 0.878 7.5 7.0

Figure 8: Evaluation of the liquefaction potential at SUPSA Terminal, Tank D

– 11 –
Heinz-Joachim Priebe

References

Department of the Navy, Naval Facilities Engineering Command, Design Manual 7.3 (1983)
Soil Dynamics, Deep Stabilization, and Special Geotechnical Construction

Dobson, T. (1987)
Case Histories of the Vibro Systems to Minimize the Risk of Liquefaction.
Soil Improvement - A Ten Year Update, Geotechnical Special Publication No. 12

Engelhardt, K., Golding, H. C. (1975)


Field testing to evaluate stone column performance in a seismic area. Geotechnique 25

Herzog, M. (1980)
Zwei Versagensarten des Baugrundes unter Erdbebenwirkung. Bauingenieur 55

Klein, G. (1990)
Grundbau-Taschenbuch, 4. Auflage, Teil 1

Mitchell, J. K. (1986)
Material Behaviour and Ground Improvement.
International Conference on Deep Foundations, Beijing, Vol. 2

Mitchell, J. K., Wentz, F. J. 1991


Performance of Improved Ground during the Loma Prieta Earthquake.
Report No. UCB/EERC-91/12

Peck, R. B. (1979)
Liquefaction Potential: Science Versus Practice.
ASCE, Journal of the Geotechnical Engineering Division, Vol. 105

Prater, E. G. (1977)
Verflüssigung von Bodenschichten infolge Erdbeben.
Mitteilungen der Schweizerischen Gesellschaft für Boden- und Feldmechanik, Nr. 97

Priebe, H. J. (1989)
The prevention of liquefaction by vibro replacement.
Conference on Earthquake Resistant Construction and Design, Berlin

Priebe, H. J. (1995)
Die design of vibro replacement. Ground Engineering, Vol. 28, No. 10

Robertson, P. K., Campanella, R. G. (1985)


Liquefaction Potential of Sands Using the CPT.
ASCE, Journal of Geotechnical Engineering, Vol. 111

Seed, H. B., Booker, J. R. (1977)


Stabilization of Potentially Liquefiable Sand Deposits Using Gravel Drains.
ASCE, Journal of the Geotechnical Engineering Division, Vol. 103

Seed, H. B., Idriss, I. M., Arango, I. (1983)


Evaluation of Liquefaction Potential Using Field Performance Data.
ASCE, Journal of Geotechnical Engineering, Vol. 109

– 12 –
Vibro Replacement to Prevent Earthquake Induced Liquefaction

Singh, S. (1996)
Liquefaction characteristics of silts. Geotechnical and Geological Engineering

Stark, T. D., Olson, S. M. (1995)


Liquefaction Resistance Using CPT and Field Case Histories.
ASCE, Journal of Geotechnical Engineering, Vol. 121

Suzuki, Y., Koyamada, K., Tokimatsu, K. (1997)


Prediction of liquefaction resistance based on CPT tip resistance and sleeve friction.
International Conference on Soil Mechanics and Foundation Engineering, Hamburg

Wenk, T., Schwarz, J. (1995)


Das Große Hanshin-Erdbeben vom 17. Januar 1995. Bautechnik 72

– 13 –
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice

Annexure 2 Technical paper on “Optimal Foundations in Soft Ground:


An Innovative Approach for Economizing Cost and Time”

17
November 19, 2015 20:10 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 164

OPTIMAL FOUNDATIONS IN SOFT GROUND:


AN INNOVATIVE APPROACH FOR ECONOMIZING
COST AND TIME

I.V. ANIRUDHAN1 , A. MADAN KUMAR2 and Y. HARI KRISHNA2


1 Consultant, Geotechnical Solutions, Chennai, India
2 Keller Ground Engineering India Pvt Ltd, Chennai, India

Improvements of productivity and effective project management procedures have


become extremely important for speedy completion of projects with positive results.
This paper discusses a milestone project comprising 198 residential units of stilt plus
four levels that highlighted the benefits of these key aspects. Success of ground
modification procedure benefitting the entire cycle involving End Users, Suppliers,
Bankers and Developer is discussed.
Soil profile of the present study comprises weak layers of silty clay and sandy
clay with varying consistency for a considerable depth. Deep foundations have
been the automatic choice of foundation for major buildings constructed in similar
environment. After serious geotechnical appraisal, reconsideration of the foundation
system with an alternative solution of Ground Improvement was found to be optimal
in terms of Cost and Time. Full raft foundation over the ground improved by
installation of Vibro Stone Columns using dry bottom feed technique was found to
be the most suitable alternative foundation system.
The ground improvement works were completed within six weeks as against six
months to that of pile foundations. This was made possible through effective project
management and the raft foundation was cast simultaneously. Full size plate load
tests were conducted to ascertain effectiveness of the ground improvement works.
Success of the foundation system was proved by full scale monitoring of foundation
settlement during and after completion of the project over a span of 2 years. This
paper describes design considerations, quality control in the construction of Vibro
Stone Columns using dry bottom feed method and performance monitoring of the
project thereafter.

1. Introduction
Deep foundations like bored cast-in-situ piles and driven piles have historically been
the foundation choice for major buildings and other structures constructed in the
weak soil deposits. Construction of pile foundations is becoming a challenge due
to their high cost, large construction time and also due to severe environmental
issues (noise pollution, ground vibrations and carbon footprint). However, given
their acceptance in the construction community, driven pile foundations were initially
selected as the foundations for the proposed residential project. Since the proposed
building complex is located within the close proximity of a well-developed residential
locality, severe resistance by the neighborhood to the pile driving activities called for
a rethinking on the foundation system.
According to the soil investigations conducted at the proposed project site, the
top 6m to 8m of soil profile comprises of silty clay / sandy clay having highly
varying consistency. Ground water table was encountered at about 3.0m below the
existing ground level. Under the present scenario, the reconsideration of the type of

Advances in Soft Ground Engineering.


Edited by C. F. Leung, T. Ku and S. C. Chian
Copyright c 2015 by ICSGE15 Organizers. Published by Research Publishing
ISBN: 978-981-09-7520-3 :: doi: 10.3850/978-981-09-7520-3 164 459
November 19, 2015 20:10 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 164

460 I.V. Anirudhan et al.

foundation system with respect to the geotechnical analysis gave a thought to assess
feasibility of alternative foundation system using appropriate ground improvement
system.
Ground Improvement refers to a technique that improves the engineering proper-
ties of the weak soil mass treated. Usually, the properties that are modified are shear
strength, stiffness and permeability. Ground improvement has been developed into
a sophisticated tool to support foundations for a wide variety of structures. Properly
applied, i.e. after giving due consideration to the nature of the ground being improved
and the type and sensitivity of the structures being built, ground improvement often
reduces direct costs and saves precious construction time

2. Acceptability of Ground Improvement Procedures


For the last two decades, ground improvement using stone columns have been
profoundly increased on local soils which are unable to take high foundation loads.
Typically stone column consists of a vertical reinforcement introduced by constructing
a column of densely packed stones partially or fully replacing the local weak soil.
The construction can either by wet or by dry method. The inclusion of stone
columns in a specific grid pattern allows the soil mass behaves like a homogenous
layer of improved density and stiffness. This process yields enhancement of load
bearing capacity and minimizes the settlements of the treated ground compared to
the untreated ground.
Stone columns acts as drainage path allowing for rapid consolidation which in
turn improves the strength and deformation characteristics of the ground at a much
faster rate. Stone columns constructed using vibro techniques allows full or partial
displacement instead of partial or full replacement of the weak soil and then leads
to further improvement of displaced weak soil by faster dissipation of construction
pore water pressure. Also, the improved drainage capabilities of the stone column
treated ground provide a much better resistance to liquefaction of the surrounding
soil. The resistance to liquefaction is achieved by densification of surrounding weak
soil and also by the much increased capacity for faster dissipation of excess pore water
pressure.

3. Ground Improvement by Vibro Techniques


Vibro Replacement is an accepted method for subsoil improvement, in which large-
size columns of coarse grained material are installed in the soil by means of high
capacity depth vibrators. Performance of this composite system consisting of stone
columns as reinforcing elements and the weak soil mass that can be established
theoretically can also be established by full size field plate load tests. Priebe (1995)
developed design of vibro stone columns on theoretical basis which can be adopted
to different soil conditions.
Contrary to vibro compaction which densifies non-cohesive soils due to vibrations,
vibro replacement improves cohesive and non-cohesive soils by reinforcing the weak
soil with load bearing columns of well compacted, coarse grained material. When
the entire weak soil is replaced with a well compacted coarser material, there is no
complexity in the understanding of its improved load carrying capacity and corre-
sponding deformations. But, when the weak soil is partially replaced and displaced
November 19, 2015 20:10 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 164

Optimal Foundations in Soft Ground: An Innovative Approach 461

Fig. 1. Unit cell of stone columns and typical arrangement of triangular grid and square grid.

by the introduction of these stiffer reinforcing elements at regular grid patterns,


response of this modified ground becomes complex. There are ways for arriving at
an equivalent stiffness matrix of a system that replaces some part with a material
of larger stiffness. Similarly, there are ways and means to establish the modified
density and stiffness when the entire soil mass is densified. When the improvement is
attributed to both displacement and replacement, the quantification of improvement
is difficult to determine. Considerable efforts like large-scale load tests can only prove
the effectiveness of the installed stone columns. In a first step, an improvement factor
is established by which stone columns improve the performance of the subsoil in
comparison to the state without columns just by increasing the overall stiffness. The
grid patterns and concept of unit cell is illustrated in Figure 1. Basic improvement
factor can be arrived based on the area replacement ratio and the reinforcing material
used for stone columns.

Fig. 2. Priebe’s basic improvement factor (reproduced from Priebe, 1995).


November 19, 2015 20:10 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 164

462 I.V. Anirudhan et al.

Improvement factor (a factor is established by which stone columns improve the


performance of the subsoil in comparison to the state without columns) is presented
in Figure 2. According to this improvement factor, the deformation modulus of the
composite system can be established due to which settlements will be reduced.
The deformation modulus of the composite system is one of the basic inputs
for finalizing the design of stone columns. However, the reality is that in many
practical cases the reinforcing effect of stone columns installed by vibro replacement
is superposed with the densifying effect of vibro compaction, i.e. the installation of
stone columns densifies the soil between grids increasing its k0 (coefficient of earth
pressure at rest) and k p (coefficient of passive earth pressure). In such case, the
densification of the soil has to be evaluated on the basis of original soil data and
correspondingly the design of vibro replacement can be modified to suit particular
improved site condition.

4. Vibro Stone Columns (Dry Bottom Feed Method)


Keller has developed the system of custom-built machine called the Vibrocat for
installation of vibro stone columns without using water. The Vibrocat comprises a
specially constructed track mounted supporting unit, attached with high capacity
depth vibrator, which incorporates a stone tube with compression chamber and
stone feed hopper ensures properly formed compacted stone columns to the required
diameter and depth. A special feature of the dry method is that it does not require
water jetting for penetration and hence eliminates the need to handle the collected
water. Furthermore this method can be used most successfully where limited working

Fig. 3. Sequence of installation of vibro stone columns (dry bottom feed method).
November 19, 2015 20:10 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 164

Optimal Foundations in Soft Ground: An Innovative Approach 463

space is available, especially in developed or urban areas or where no near water


source can be found. This technique provides effective drainage paths to ensure rapid
consolidation. It also has a built-in real time computer monitoring system to provide
quality control on compaction effort throughout the construction process. Sequence
of installation of vibro stone columns using dry bottom feed method is illustrated in
Figure 3.

5. About the Project and Subsoil Conditions


Urban Tree Infrastructure Private Limited (Urban Tree), Chennai, proposed to develop
a residential project in Chennai. The project comprises of 198 units of Stilt + 4 floors
and the approximate area of development is about 2.5 acres. Typical project layout is
shown in Figure 4.
The sub-soil in the project site comprises desiccated clay and medium dense sand
up to about 3.50m followed by relatively weak clay and sandy clay up to 6.0m depth.
This top 6.0m soil with highly varying consistency is followed by about 8.0m with
medium dense sand stiff clay deposits after which there is a 6m thick layer of medium
stiff consistency. Denser sand layers and hard clay layers are forming the remaining
sub-soil profile.

Fig. 4. Overall layout of the proposed project site.


November 19, 2015 20:10 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 164

464 I.V. Anirudhan et al.

6. Design of Foundation System


The performance of individual footings (shallow foundations) would not be effective
due to weak to very weak soil layers till 6m depth below the existing ground level.
Larger settlements were expected in case of raft foundation due to high structural
loads. The range of settlement estimated in the investigation report was about
220mm out of which the top 6m to 8m soil attributing about 115mm. Based on the
subsoil conditions, driven cast-in-situ piles resting in hard clay layers below 25m
were adopted and, as described earlier, the construction of piles were stopped due
to environmental issues.
In this context a possible alternative solution of suitable ground improvement
technique in place of already chosen driven cast-in-situ pile foundations resting in
hard/competent strata available at about 25m as foundation system is very much ad-
vantageous. Under these circumstances, the developer has contacted M/s Keller India
to undertake design and execution of the ground improvement works. Considering
the project boundary conditions, vibro replacement (stone columns with dry bottom
feed method) was selected as a viable method for subsoil improvement and a full raft
foundation supported by the treated ground as an alternative foundation system. The
selected method of ground improvement satisfied in addressing environmental issues
raised at project site. In this method, the stone columns are installed by displacement
technique (without removing any soil). Hence, the site environment would be
comparatively clean and tidy. In addition, benefits with regard to economizing the
foundation cost and optimizing construction time was proven invaluable for the
project.

Fig. 5. Footprint of the raft foundation (divided into pours for ease of construction).
November 19, 2015 20:10 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 164

Optimal Foundations in Soft Ground: An Innovative Approach 465

As described above, the construction will have stilt plus four floors with average
load intensity at the foundation level which is be approximately 75 to 85 kPa. A
total footprint area of about 5500m2 under the raft foundation is to be treated. The
foundation raft was divided into six pours for ease in construction. Layout of the
proposed building is shown in Figure 5. Individual column loads from the super
structure vary from 25 T to 185 T. Though the raft foundation transmits uniform
pressure to the bearing soil, denser grid was adopted for pour having large column
loads.
Critical review in terms of strength and deformation characteristics for the pro-
posed loading conditions were made using Priebe (1995) design methodology and
appropriate geometry (stone column diameter, spacing, pattern and depth) has been
chosen. The final scheme was reviewed and vetted by M/s Geotechnical Solutions,
Chennai, a third party specialist Geotechnical Consultancy firm for Geotechnical
compliance. Typical scheme and cross section of ground improvement using vibro
stone columns using dry bottom feed method adopted for the present project is
illustrated in Figure 6.

6.1. Quality Control and Monitoring


In order to measure and assure the quality of stone columns being constructed, it is
necessary to adopt stringent quality control and quality assurance procedures to meet
the specifications and to satisfy the client’s requirement at various stages of execution
of the project.

6.2. During Execution of Ground Improvement


The installation of each stone column was recorded by the use of an automated com-
puterized recording device fitted to the Vibrocat. This instrument yields a computer
record (M4 Graph) of the installation process in a continuous graphical mode, plotting
depth versus time and power consumption (compaction effort) versus time. The
information provided includes:
– Stone column reference number
– Date of installation
– Start and finish times of installation
– Period required for installation
– Maximum depth
– Compaction effort during penetration and compaction process
The above parameters allow to monitor the quality of the stone columns being
installed. Further, diameter of the stone column and consumption of backfill are con-
tinuously monitored by the site personnel to estimate the in-situ achieved diameter.

6.3. Post Construction


Full size field plate load test is one of the accepted ways to assess the performance
of the improved soil treated with stone columns. The size of the test pad and the
magnitude of the test load can vary according to the stone column layout, treatment
depth, load and type of structure. Routine Stone Column Load tests were performed
November 19, 2015 20:10 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 164

466 I.V. Anirudhan et al.

Fig. 6. Typical soil profile showing ground improvement arrangement.

Load Intensity, kN/m2


0 20 40 60 80 100 120 140
0

4
Settlement, mm

12

16

20

Fig. 7. Results of single column routine plate load test.

to ascertain the effectiveness of design and performance of the ground improvement


works. The observed settlements are within the acceptable limits of 75 to 100 mm
for raft foundations according to the stipulations specified in Indian Standard Code
November 19, 2015 20:10 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 164

Optimal Foundations in Soft Ground: An Innovative Approach 467

of Practice (IS 1904-1986) for the applied design load intensity of 100 kPa. Load test
results are presented in Figure 7.

7. Real Time Settlement Monitoring


Success of the foundation system needs to be proved by full scale monitoring of
foundation settlement during and post completion of the project. Post construction
real time monitoring offers confidence on the engineering judgment taken at various
stages of the project completion. In this section the predicted design settlements
which are calculated using conventional methods are being compared with the actual
settlement occurred at the site by adopting proper monitoring systems.
Keeping the importance of the post construction performance of the structure,
about 14 locations were identified on the raft foundation to monitor settlements
during and post construction. After completion of the installation of the ground
improvement works, raft foundation is laid on the treated ground. The entire building
foundation area was divided into 6 zones which are delineated based on concrete
pour-1 to pour-6. Typical arrangement of the selected locations is shown in Figure 8.
Selection of 14 numbers is based on the number of concrete pours in overall raft
foundation i.e., 2 locations per each concrete pour and reduced levels were recorded
in regular intervals. Summary of the predicted and observed settlements are shown
in Figure 9.
The measured settlements are substantially lower than the predicted settlement,
which proved the efficiency of the raft foundation resting on improved ground. It can

Fig. 8. Settlement monitoring points.


November 19, 2015 20:10 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 164

468 I.V. Anirudhan et al.

Settlement Designed Observed Settlement


Points Settlement Settlement recently
(mm) (mm) observed on
P1S1 64 50
P1S2 64 45
P2S1 64 13
P2S2 64 23
P3S1 64 47
P3S2 64 51
P4S1 64 51
29th May 2015
P4S2 64 47
P4S3 64 47
P5S1 64 24
P5S2 64 28
P6S1 64 39
P6S2 64 40
P6S3 64 52

Fig. 9. Summary of predicted and observed settlements.

Settlement Results - All Pours

100
Load vs Time Curve

90
Super structure load (kPa)

80

70

60

50

40

30

20

10

0
0 0 4 8 12 16 20 24 28 32 36 40 44 48 52 56 60 64 68 72 76 80 84 88 92 96 100 104 108

10 Point: P1S1 Point: P1S2


Point: P3S2 Point: P4S1
20
Point: P4S3 Point: P6S3
30 Predicted settlement
Settlement 'mm'

40

50

60

70

80
0 4 8 12 16 20 24 28 32 36 40 44 48 52 56 60 64 68 72 76 80 84 88 92 96 100 104 108
Time in Weeks

Settlement vs Time Curve

Fig. 10. Results of observed settlements.


November 19, 2015 20:10 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 164

Optimal Foundations in Soft Ground: An Innovative Approach 469

Fig. 11. Completed structure (foundation resting on vibro stone columns, dry bottom feed).

be seen from Figure 10 that the load in the super structure increases with increase in
number of floors and the corresponding settlements are increased.
The superstructure load is increased from 0 to 80 kPa in 20 weeks and corre-
spondingly predicted settlements (analytical method) increased from the 0 to 64
mm. However, the observed settlements are considerably less than the predicted
settlements as well as the allowable settlements of 75 to 100 mm for raft foundations
resting in clayey soils. Pattern of the observed settlements at one of the pour is
presented in Figure 10. The total settlement observed at the start of maximum loading
(after 18 weeks) was about 30mm that was gradually increased to about 50 mm during
the next 17 weeks and remained more or less uniform thereafter. It is suggesting that
the long term settlements will be of much smaller range than that was expected.

8. Conclusions
Application of vibro replacement proved to be an effective ground improvement
solution in varying soil conditions. It is also proven from the results of extensive
monitoring results that the required performance was achieved. Vibro stone columns
made it possible to support residential buildings on weak deposits. In addition
to improving shear strength and compressibility parameters, offered acceleration in
the overall construction schedule and enabled the project to be completed within
stipulated duration.
The ground improvement works were completed within 6 weeks (as against 6
months to that of pile foundations) that was made possible through effective project
management. The project is getting delivered to the end users ahead of time as a result
of construction speed of alternative foundation solution (i.e. 6 months vs. 6 weeks)
marking a milestone in ground modification. The savings in time is key to success
of ground improvement benefitting the entire cycle involving End Users, Suppliers,
Bankers and Developer.
November 19, 2015 20:10 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 164

470 I.V. Anirudhan et al.

Acknowledgements
A project of this nature is possible only with the cooperation and combined efforts
of many persons. The authors wish to thank the developer of the project for their
kind permission to publish the information contained in this paper. In particular, the
authors wish to acknowledge Keller management for their continuous support and
encouragement in finalization of theme of this paper including valuable comments
and suggestions that helped the authors to refine and improve the quality of the paper.
Thanks are also due to Keller India site team for providing photographs and other
execution data.

References
1. Hughes JMO & Withers N J, 1975, ‘Reinforcing of Soft Cohesive Soils with Stone Columns’,
Ground Engineering, Volume 7, Issue No. 3.
2. Heinz J. Priebe, 1995, The Design of vibro replacement, Ground Engineering, GT 037-13 E.
3. IS: 1904 – 1986 (2006), Code of Practice for Design and Construction of foundations in Soils: General
Requirements.
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice

Annexure 3 Technical paper on “Vibro Stone Columns to Support Large


Oil Storage Tank Farm on West Coast of India”

18
November 19, 2015 20:2 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 094

VIBRO STONE COLUMNS TO SUPPORT LARGE OIL STORAGE


TANK FARM ON WEST COAST OF INDIA

SRIDHAR VALLURI1 , DEEPAK RAJ2 , and VISHAL SHUKLA3


1 Geotechnical Manager, Keller Ground Engineering (I) Pvt. Ltd, Mumbai, India.
E-mail: [email protected]
2 Director, Keller Ground Engineering (I) Pvt. Ltd, Mumbai, India.

E-mail: [email protected]
3 Civil Manager, Vijay Tanks and Vessels Pvt. Ltd, Vadodara, India.

E-mail: [email protected]

A large oil storage terminal is being constructed on the west coast of India at Pipavav
port, in the state of Gujarat along the Arabian Sea. Proposed terminal has a storage
capacity of 250,000 kilolitres to handle classified petroleum products A, B, C as well as
non-classified products. The terminal comprises of 48 number of steel storage tanks
with diameter varying from 12m to 25m and heights varying from 18 m to 20 m with
slenderness ratio varying from 0.8 to 1.5. The storage tanks comprises of floating and
fixed roof tanks with ring beam foundation to support the tank shell.
Extensive soil investigation has been carried out by exploring boreholes, conduct-
ing standard penetration tests and cone penetration tests. The subsoil consists of
six to ten meters thick soft to firm silty clay layer. Ground improvement using
Vibro Stone Columns is done to reduce the total settlements, control the differential
settlements and increase the bearing capacity of the subsoil. Area replacement ratios
varying from 16 to 23% were used with column diameters in the range of 1 m and
treatment depths up to 10 m to support the tank foundations. More than twenty
field load tests are conducted across the site in the foot print of the tank as one
of the quality control measure. Settlement monitoring was carried out during the
hydrostatic tests for all the 48 tanks.
This paper summarizes the details of proposed tanks, subsoil conditions, design
scheme of vibro stone columns, quality control measures taken during the construc-
tion, hydrostatic test results and their analysis.

Keywords: Storage tanks, Slenderness ratio, Soft clay, Vibro stone columns, Quality
control, Load tests, Hydrostatic tests.

1. Introduction and Project Background


Demand for petroleum and non-petroleum products is on rise due to rapid indus-
trialization and growth in Indian subcontinent. A number of refineries and storage
terminals are being constructed across the country, in order to cater to India’s growing
demand.
Gulf Petrochem (I) Pvt. Ltd. is developing an ’Oil Storage Terminal’ at Pipavav
Port. The terminal is located in Saurashtra in the state of Gujarat, 152 nautical miles
northwest of Mumbai on the west coast of India and 140 km southwest of Bhavnagar.
The proposed oil storage terminal will handle Class A, B and C as well as non-
classified products with an annual storage capacity of 250,000kL. The terminal on
commissioning will have complete flexibility for storing any type of oil product. The
proposed terminal site is spread over eight (8) hectares area. The subsoil at the site

Advances in Soft Ground Engineering.


Edited by C. F. Leung, T. Ku and S. C. Chian
Copyright c 2015 by ICSGE15 Organizers. Published by Research Publishing
ISBN: 978-981-09-7520-3 :: doi: 10.3850/978-981-09-7520-3 094 365
November 19, 2015 20:2 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 094

366 Sridhar Valluri et al.

Pipavav

Fig. 1. Project Location - Oil Storage Terminal at Pipavav, Gujarat.

consists soft to firm clayey silt followed by stiff soils. The subsoil soil is expected to
settle excessively under the imposed foundation loads. Hence, ground improvement
using vibro stone columns is done to support the tank and the associated utility
building foundations to increase the shear resistance of the soil, to control post
construction long term and differential settlements.

2. Details of the Structures at Proposed Tank Farm


The proposed Oil Storage Terminal consists of construction of 48 no’s of steel storage
tanks and associated utility structures like Office buildings, truck loading facilities
and weigh bridges. Based on the storage type of the liquid, the tanks are arranged
under six (6) enclosures as shown in Figure 2. Enclosure-1 comprises of floating roof
tanks while the rest of the enclosures are having fixed roof tanks including two (2)
no’s of fire water tanks. The diameter (D) and height (H) of the proposed tanks are
varying from 12 m to 26 m and 18 m to 20 m, respectively. Due to higher slenderness
ratio (H/D) and the uplift forces arising from tank design, ring beam foundation with
anchor bolts were proposed as foundation interface for all the tanks. Details of tanks
and utility structures proposed on ground improvement at the facility are provided
in Table 1.
November 19, 2015 20:2 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 094

Vibro Stone Columns to Support Large Oil Storage Tank Farm on West Coast of India 367

Fig. 2. Layout of Oil Storage Terminal at Pipavav, Gujarat.

Table 1. Details of the Tank Farm and Structures

Enclosure No D / L (m) H / B (m) Type of Storage Storage


Roof Product Capacity (KL)

Enclosure-1 8 18.5 20 Floating Class A 56,648


7 12 18
1 13 18
Enclosure-2 6 26 20 Fixed Class C 60,000
Enclosure-3 6 18 20 Fixed Non - 30,000
Classified
Enclosure-4 6 18 20 Fixed 30,000
Enclosure-5 6 25 20 Fixed Class B 55,950
Enclosure-6 6 14 20 Fixed Class C 17,538
Fire Water Tanks 2 20 20 Fixed Water 12,560
Engineering office 1 20 10 RCC (G+3) — —
Weigh Bridge 2 12 3 — — —
November 19, 2015 20:2 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 094

368 Sridhar Valluri et al.

Fig. 3. Typical sub soil profile.

3. Subsoil Profile
The existing ground level (EGL) at the proposed site is varying from RL +4.6 m to
RL +3.8 m. Extensive soil investigation was carried out at the proposed oil storage
terminal site by exploring 15 number of boreholes and 46 number of electric Cone
Penetration Tests (CPT) up to the refusal levels. This included an extensive field and
laboratory tests on the disturbed and undisturbed soil samples obtained during the
investigation. The subsoil in general consists of 1 m to 1.5 m thick fill from EGL,
followed by 6.5 m to 9 m thick soft to firm silty clay/clayey silt layer with SPT, N
values ranging from 4 to 10. This layer is underlain by 1.5 m to 3 m stiff to very stiff
clayey silt/silty clay with SPT - N values ranging from 20 to 50, followed by the 1.5 m
to 3 m thick clayey sand layer and below this rock is encountered. Typical subsoil
stratification is shown in Figure 3.

4. Design Criteria
The design loading intensity of the proposed steel storage tanks is varying from 20
to 22 T/m2. Allowable long term total settlement at edge is 300 mm and 150 mm
for fixed roof tanks and floating roof tanks, respectively. The allowable maximum
differential settlement along the periphery between any two points is 1 in 300 and 1 in
500 for fixed roof tanks and floating roof tanks, respectively. For the utility buildings
and associated structures the required design loading intensity was 15 T/m2 with an
allowable post construction long term settlements of 40 mm.
November 19, 2015 20:2 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 094

Vibro Stone Columns to Support Large Oil Storage Tank Farm on West Coast of India 369

5. Proposed Ground Improvement Solution


The silty clay/ clayey silt in the top layers were expected to settlement excessively
under the imposed foundation loads and also the soil capacity was not adequate to
support the tank foundations. Hence, ground improvement using Vibro stone column
is proposed to support the foundations of the storage tanks and associated terminal
structures.
Using Vibro stone column technique, the following geotechnical improvements are
achievable:
– Improvement in the stiffness of the subsoil to decrease settlements
– Improvement in the shear strength of the subsoil to increase bearing capacity
– Rapid consolidation of the subsoil
– To mitigate liquefaction potential

5.1. Concept of Vibro Stone Columns


Vibro Stone Column (Vibro replacement) technique introduces a coarse grained ma-
terial as load bearing elements consisting of gravel or stone aggregate as a backfill
medium. The stone column and the in situ soil form an integrated system having low
compressibility and high shear strength. The excess pore water pressure can dissipate
through the stone column, which also acts as a vertical drain. The settlement expected
for the treated soil is reduced while the rate of settlement is increased when compared
with the untreated soils.

5.2. Design of Vibro Stone Column Works


The design analysis of Vibro stone columns is carried out according to Priebe’s (1995)
design methodology to meet the technical performance criteria as summarized in

Fig. 4. Typical cross section of tank with vibro stone columns.


November 19, 2015 20:2 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 094

370 Sridhar Valluri et al.

section 4. Vibro stone columns of 1 m diameter with an area replacement ratio varying
from 16% to 23% were used across the site to meet the design requirements of floating
as well as fixed roof tanks. Columns were designed to be terminated in the stiff layers.
Depth of the columns was varying from 9.5 m to 10.5 m below EGL across the site.
Treatment of stone columns was extended beyond the foot print area of the ring beam
to provide confinement and to take care of the edge stability. Figure 4 shows typical
cross section of tank with vibro stone columns foundation.

6. Construction Methodology
Top feed wet method was used for the installation of the vibro stone columns. In this
method, the depth vibrator and extension tubes are suspended from a crawler crane.
The vibrator penetrates the ground with the help of water jets at the side of vibrator,

Fig. 5. Step wise illustration of vibro stone column installation by top feed wet method.

ll i f ib l b f d h d
Fig. 6. Installation of Vibro stone columns by top feed wet method.
November 19, 2015 20:2 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 094

Vibro Stone Columns to Support Large Oil Storage Tank Farm on West Coast of India 371

its self-weight and horizontal vibrations. An annular space is created between the
vibrator and borehole walls through which stone is fed from the top, to the tip of the
vibrator. The up-down motion of the vibrator compacts the stone laterally into the
surrounding soil. This results in a well compacted stone column that has a diameter
larger than original hole. Wheel loaders were used to continuously supply the stone
from the stockpiles at site. Figure 5 shows step wise installation process of vibro stone
columns by top feed wet method.
Two (2) vibro stone column rigs were used to treat the areas proposed on ground
improvement. Works were carried out during the period from November, 2012 to July,
2013 to treat 16,600 m2. A typical picture taken during the installation of vibro stone
columns is presented in Figure 6.

7. Quality Control and Quality Assurance


In the execution of vibro stone column works quality is assured by implementing
various control measures at different stages as per the comprehensive field quality
control procedures outlined and submitted during the design stage.
The same has been summarized below:
– Pre-Construction: Soil investigation, stone aggregate source approval and testing.
– During Construction : Monitoring of construction parameters
– After Construction : Testing by means of load tests (single and group)
Prior to commencement of works, an extensive soil investigation program was
carried out as detailed in section 3 of the paper. The design is carried out considering
the subsoil data in the foot print of the tank to arrive at the optimum design of vibro
stone columns. The material used for the stone columns was checked from the source
prior to start of works at approved and accredited national laboratories. Tests include
included crushing, abrasion, sulphate resistance, water absorption and grain size
analysis as per the specifications outlined in BS EN 14731:2005 and BRE-391. Material
testing was carried at periodic intervals for the samples collected from site and source
to see that sound stone aggregate material is used for the construction of vibro stone
columns. Typical stone aggregate specifications are outlined in Table 2.
During construction the vibro stone column, installation process is monitored using
real time computerized monitoring system. The vertical position and the current
drawn by the depth vibrator is continuously measured (in real time) and displayed
to the operator. This data is printed in the form of graph and which was reviewed

Table 2. Stone aggregate specification for vibro stone columns.

S.No Tests Criteria

1 Specific Gravity > 2.5


2 Aggregate Crushing Value < 30%
3 Los Angeles Abrasion Value < 30%
4 Water Absorption < 2%
5 Soundness < 12%
6 Aggregate Size 75 mm to 12 mm
November 19, 2015 20:2 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 094

372 Sridhar Valluri et al.

Fig. 7. Typical quality control graph output of vibro stone column.

by the engineer daily. The quality control graph provides the information of stone
column number, date of installation, start and finish date of installation, period taken
to install the column, maximum depth and compaction effort during penetration
and compaction process. A typical quality control graph of vibro stone column for
enclosure 1 is presented in Figure 7.
After installation of vibro stone columns load tests are conducted to check the load
carrying capacity of the improved ground. Total fifteen (15) routine single column
load tests and five (5) routine three column group load tests were done on the installed
vibro stone columns, in order to check the design capacity of the stone columns in turn
to check the safe bearing capacity of soil after ground improvement. The tests were
conducted as per the guidelines specified in the Indian standard code, IS 15284 (Part-
1): 2003.

Load in Tons
0 50 100 150 200 250 300
0

4
Settlement in mm

12

16

20

Fig. 8. Load vs settlement plot for routine three column group load test.
November 19, 2015 20:2 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 094

Vibro Stone Columns to Support Large Oil Storage Tank Farm on West Coast of India 373

A typical result showing the load vs settlement plot for a routine group column
load test is presented in Figure 8. The result of the load tests were well within the
permissible limit indicating the columns were built with good quality and work man
ship.

8. Performance of Tanks - Hydrotest Results


Hydrotest was carried out for all the 48 no’s of tanks and settlement monitoring
was carried out as per the procedure outlined in the design report. Slow stage
hydrotest was carried out for the tanks founded on vibro stone columns to control

Fig. 9. Details of water filling rate for slow stage hydro test and location of settlement markers
along the tank periphery.
p y

Fig. 10. Average Peripheral Settlement plot during Hydrotest for Enclosure 1 (Floating Roof).
November 19, 2015 20:2 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 094

374 Sridhar Valluri et al.

the differential settlements and control the long term post construction settlements.
Settlement of tanks was measured using a series of survey points established on the
tank shell prior start of hydrotest. Details of the filling rate adopted for the hydrotest
and the settlement observation points for the tanks are presented in the Figure 9.
The final load (full water height) was maintained till the rate of settlements was
stabilized. A clear trend of stabilization of settlement was seen in the entire tank farm
within seven (7) days of reaching full stage water load. Typically, the duration for

Fig. 11. Average Peripheral Settlement plot during Hydrotest for Enclosure 3 (Fixed Roof).

Fig. 12. Completed Oil Storage Terminal.


November 19, 2015 20:2 RPS/Trim Size: 24cm x 17cm for Proceedings/Edited Book 094

Vibro Stone Columns to Support Large Oil Storage Tank Farm on West Coast of India 375

the hydrotest for each tank was in the range of 35 to 50 days which was sufficient to
reach the estimated degree of consolidation and to control the long term settlements
post hydrotest. The average settlement of the tanks recorded during hydrotest for
enclosure E-1 and E-5 are presented in Figure 10 and Figure 11, respectively.
Measured peripheral settlements of tanks ranged from 50 to 75 mm in enclosure-1
(floating roof tanks) while the same was 125 to 175 mm for the tanks in enclosure-
3 (fixed roof). Settlements were also monitored for the Engineering office building
(G+3 storied) building. Measured settlements were in the range of 18 to 23 mm for the
building.
The measured settlements of the tanks and the buildings were well within the
allowable limits in terms total and differential settlements indicating very good
performance of vibro stone columns.

9. Conclusions
An Oil Storage Terminal is being developed in Pipavav in Gujarat. The terminal
development consists of construction of 48 No’s of storage tanks and associated utility
facilities. Ground improvement using vibro stone columns is proposed to support
the foundations at the terminal. Vibro stone columns are proposed to enhance the
shear strength and compressibility parameters of the subsoil and also accelerate the
consolidation of the soft soils.
Success for a project of this kind is only possible with quality control at each stage
of construction which is implemented for this project. Settlement monitoring results
have shown the effectiveness of vibro stone columns to support the tank foundations
and ancillary buildings. The settlements were uniform and well within the tolerable
limits. Ground improvement with Vibro stone columns has resulted in homogenizing
the subsoil which helped in controlling the differential settlements.
The Oil storage terminal is commissioned successfully in the month of February,
2015.

Acknowledgements
The authors wish to thank the developer M/s Gulf Petrochem and the EPC contractor
M/s Vijay Tanks & Vessels for their kind permission to publish the information
contained in this paper. The authors also wish to acknowledge Keller management
for their continuous support and encouragement in finalization of theme of this paper
including valuable comments and suggestions that helped the authors to refine and
improve the quality of the paper. Thanks are also due to design and execution team
of M/s Keller and M/s Vijay Tank & Vessels for providing hydrotest monitoring data
at regular intervals.

References
1. Building Research Establishment (2000). “Specifying Vibro Stone Columns”, BRE - 391.
2. British Standard (2005). “Execution of Special Geotechnical Works - Ground Treatment By
Deep Vibration”, BS EN 14731-2005.
3. Indian Standard Code (2003). “Design and construction for ground improvement - Guide-
lines, Part 1: Stone columns”, IS 15284 (Part-1): 2003.
4. Priebe, H.J. (1996). “The Design of Vibro Replacement”, Ground Engineering, December
1995.
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice

Annexure 4 Technical paper on “Ground Improvement Solutions to


Mitigate Liquefaction: Case Studies”

19
Proceedings of Indian Geotechnical Conference
December 13-15,2012, Delhi (KN 6)
GROUND IMPROVEMENT SOLUTIONS TO MITIGATE LIQUEFACTION: CASE STUDIES
Madan Kumar Annam, Technical Manager, Keller India, [email protected]
V. R. Raju, Managing Director, Keller Asia, [email protected]

ABSTRACT: There have been major advances occurred in the past in understanding as well as practicing of engineering
treatment of seismic soil liquefaction and assessment of seismic site response. While research on liquefaction continues, the
geotechnical engineering practice has developed various techniques for site improvement that can mitigate the potential effects
of liquefaction. The first part of this paper address soil liquefaction and second part concentrates on the case histories where
ground improvement methods using vibro techniques were implemented to mitigate liquefaction-induced damages in major
infrastructure projects.

INTRODUCTION consideration can actually resist. Generally, the maximum


Liquefaction is defined as the transformation of a granular surface acceleration on level ground is used as the
material from a solid to a liquefied state as a consequence of characteristic value for the forces developed by an
increased pore-water pressure and reduced effective stress. earthquake.
Liquefaction is one of the critical problems in geotechnical
engineering. High ground water levels and alluvial soils have Estimation of two variables for evaluation of liquefaction
a high potential risk for damage due to liquefaction, resistance of soils is expressed in terms of Cyclic Stress Ratio
especially in seismically active regions. The most critical soil (CSR), the seismic demand on a soil layer and Cyclic
is fine sand with some silt content. Resistance Ratio (CRR), the capacity of the soil to resist
liquefaction. One of the most widely accepted and used SPT
Evaluation of the Liquefaction Potential based correlations is the “deterministic” relationship
A large part of India lies in potentially hazardous earthquake proposed by Seed, et al (1984, 1985) represented in Fig. 2.
prone zones. A large portion of eastern, western and Seed and Idriss (1971) formulated the following equation for
northeastern part of the country comes under Zone V and calculation of the Cyclic Stress Ratio:
Zone IV (Refer Fig. 1).
CSR = (τav/σvo’) = 0.65(amax/g)(σvo/σvo’)rd

Fig. 1 Seismic Zone Map of India


Fig. 2 SPT Clean-Sand Base Curve for Magnitude 7.5
The simplest and probably most reliable method to evaluate Earthquakes
the soil liquefaction potential is statistical analyses from the Where, amax = peak horizontal acceleration at the ground
past history. For this, the forces expected during a seismic surface generated by the earthquake (discussed later); g =
event are compared with the forces that the subsoil under acceleration of gravity; σvo and σvo’ are total and effective
Madan Kumar Annam; V. R. Raju
vertical over burden stresses, respectively; and rd = stress Vibration is achieved by means of powerful vibrator at
reduction coefficient which accounts for flexibility of the soil deeper depths. The vibrator is connected to a source of
profile. Seed and Idriss (2001) used the approximated electric power and a high-pressure water pump. Extension
equation for CRR for clean-sands that used a base curve and tubes are added as necessary, depending on the treatment
fitting the following equation: depth, and the whole assemblage is suspended from a crane.
A Schematic showing Vibro Compaction technique is
CRR7.5 = [1/ {34-(N1)60}] + [(N1)60/135] + [50/ (10(N1)60+45] presented in Fig. 4.

Where (N1)60 is SPT N value projected for clean sand Vibro Replacement
obtained after corrections on measured field SPT N value. The stabilization of weak deposits by displacing the soil
The above equation is valid for (N1)60< 30. For (N1)60>30, radially with the help of a depth vibrator, refilling the
clean granular soils are too dense to liquefy and are classed as resulting space with granular material and compacting the
non-liquefiable (Refer Seed and Idriss 2001). same with the vibrator is referred to as Vibro Replacement.

GROUND IMPROVEMENT TECHNIQUES The resulting matrix of compacted soil and stone columns has
The liquefaction potential of weak deposits can be mitigated improved load bearing and settlement characteristics. A
with ground improvement techniques such as vibro schematic showing the basic principle of the vibro
replacement (vibro stone columns) and vibro compaction. replacement technique, explained in Fig. 5. Keeping the site
These techniques use vibratory energy to densify loose soils conditions in view vibro stone columns can be installed either
at depth by backfilling. Principles of vibro replacement and wet method (top feed) or dry method (bottom feed).
vibro compaction techniques are discussed in this paper. Technically and functionally, vibro stone columns installed in
both methods serve similar.
Deep Vibro Techniques
Vibro technique offers the weak deposits to get compaction,
drainage and increase in shear resistance. Fig. 3 shows
transition zone of soils tends to liquefiable and possible
techniques of ground improvement with deep vibro
compaction or replacement.

Fig. 5: Schematic showing vibro replacement technique

The above ground improvement techniques were adopted in


various infrastructure projects to mitigate liquefaction
potential across India.

Few case studies of the executed projects falls under seismic


Fig. 3 Application ranges of the deep vibro techniques Zone IV and V as per IS 1893 Part 1 (2002) are discussed in
the following sections.
Vibro Compaction
The basic principle behind the vibro compaction process is CASE STUDIES
that particles of non-cohesive soils can be rearranged into a Power Plant at Goindwal Saheb, Punjab
denser state by means of vibration. A power plant of capacity 2 x 270 MW coal based Thermal
Power Plant was built at Goindwal Sahib, near Amritsar,
Punjab. Power plant structures such as Boiler, Electro Static
Precipitator (ESP), Switch Yard, Power House Building, etc.
were planned as part of development of power plant.

Soil at this project site is primarily sandy silt / silty sand to


about 1.5m to 2m depth, followed by fine sands with fines
content of about 6% to the final explored depth of about 30m.
The average SPT N value is 10 up to a depth 4m to 6m from
the existing ground level and SPT N value ranges from 15 to
Fig. 4 Schematic showing vibro compaction technique 25 to a depth of about 15m. Medium dense to dense sand
Ground improvement solutions to mitigate liquefaction and its application
layers were encountered beyond 15m depth with SPT N the above, 2nos plate load tests (20 t/m2 & 40 t/m2) were
values are generally > 25. conducted at cooling tower I & II area to assess the load vs
settlement behavior of the improved ground for required safe
The existing natural soils (fine sands) at the proposed site bearing capacity of 10 t/m2. The plate load test results
being loose were susceptible to liquefaction in an event of an indicate that the settlements are less than 5mm in both cases.
earthquake. Hence, Ground Improvement by Vibro
Compaction Technique was proposed to mitigate liquefaction
and to enhance the bearing capacity.

Vibro Compaction for main works has been carried out to a


depth of 8m.

Fig. 6 Twin Vibrators in action. Fig. 8 Pre and Post CPT results at ESP area.

A School Building, Noida


One of the reputed educational societies in Noida, Uttar
Pradesh has proposed to build a school building. The soil
investigation at the proposed site revealed that loose to
medium dense fine sand exists to a depth of 9m which is
susceptible to liquefaction. Since the project location falls
under Seismic Zone IV, the required field SPT (Standard
Subsidence (1.0m) Penetration Test) shall be more than 20 (performance
criteria requirement by the designer), to mitigate liquefaction
and to achieve bearing capacity.

Fig. 7 The difference in levels achieved (about 1m) as a Ground improvement technique using vibro compaction was
result of vibro compaction proposed as treatment to mitigate liquefaction and to enhance
the safe bearing capacity of the loose sand deposit till 9 m
Post Cone Penetration Tests (CPT) were conducted after depth below the existing ground level.
completion of the vibro compaction for various structures, as
part of QA/QC procedures. Based on the analysis of the post About three boreholes prior to and two boreholes after the
CPT results, a Relative Density of more than 70% was commencement of vibro compaction works were carried out
achieved. Pre and post cone resistance (Qc) values are shown in the project area to assess the effect of vibro compaction. It
in Fig. 8 in which the target relative density is also presented. can be seen from Fig. 9 that the post treatment SPT N values
The targe relative density is evaluated based on correlations (shown in discontinuous lines) are larger than the
proposed by Schmertmann with respect to Qc. In addition to
Madan Kumar Annam; V. R. Raju
performance line confirming the effect of the ground to medium dense sand up to a depth of 10m below the
improvement. existing ground level with SPT N ranging from 11 to 18. A
SPT N [ ] clayey silt / silty clay layer was encountered up to 20 m depth
0 10 20 30 40 50 60 with SPT N more than 20. A load intensity of 10 t/m2 to 15
0
t/m2 was anticipated due to various structures. The top soil
Pre SI BH1
1
layers up to 10m depth were susceptible to liquefaction.
Pre SI BH2
Pre SI BH3
2
Perfrm. Line
Post SI BH4
3
Post SI BH5

4
Depth [m]

10
Fig. 9 Graph showing performance of vibro compaction
Load Intensity, t/m2
0 5 10 15 20 25 30 35 40
0

2
Settlement, mm

Fig. 11 Installation of vibro stone columns at STP, Noida


4
Vibro Stone Columns were installed in triangular grids of
different spacing under strip and raft foundation to a depth of
6
10m from existing ground level to mitigate the liquefaction
and to enhance the bearing capacity.
8
Load Intensity [t/m2 ]
0 5 10 15 20
10 0
Fig. 10 Graph showing Load Intensity vs Settlement
2
Further, a field plate load test was also performed at site to
assess bearing capacity of the treated ground. The observed
settlements were within the acceptable limits for the applied
Settlement [mm]

4
load intensity as illustrated in Fig. 10.
6
Sewage Treatment Plant, Noida
Greater Noida Development Authority was constructing a
137 MLD Sewage Treatment Plant in Greater Noida, Uttar 8
Pradesh. Various structures such as Chlorination tanks, SBR
basins, Air blower, Grit chambers etc. were proposed. The
project location falls under seismic Zone IV, however as per 10
the requirements of the Client the ground improvement
techniques were adopted to satisfy liquefaction effects of
12
Zone V conditions.
Fig. 12 Load Intensity vs. Settlement curve at STP, Noida
The soil profile comprise of silty sand layer of 3m to 4m
thick with SPT N ranging from 6 to 8 followed by fine, loose
Ground improvement solutions to mitigate liquefaction and its application
A field plate load test was performed at site to assess bearing Pre and Post CPT at site on trial stone columns were carried
capacity of the treated ground. The observed settlements were out. Post treatment results showed a two-fold increase in the
found within the acceptable limits for the applied load CPT values of the sand zones as shown in Fig. 13. Hydro
intensity as shown in Fig. 12. tests were performed on the installed tanks and the expected
settlements were in the range of 120 mm, well within the
LNG Terminal, Hazira limits at the centre of tank under full tank load of 23.0 t/m2.
Two liquefied natural gas storage tanks were constructed in
Hazira LNG Terminal Project of each 84 m in diameter and
with a filling level of approximately 35 m. The site is located
at an estuary on the coast of the Khambhat Gulf in India.

The subsoil profile at the project site consists of loose to


medium dense silty sand up to a depth of 16 m below the
existing ground level. The upper 4 to 5 m was recently
reclaimed material. The fines content of the sand was in the
range of 15% on average, sometimes slightly higher. Very
dense sand with SPT > 50 was encountered below 16m from
Fig. 14 Tanks commissioned after successful Hydro Test and
existing ground level.
in operation for last 8 years
A peak ground acceleration (PGA) of a = 0.24g confirming to
Power Plant in North Delhi
seismic Zone IV conditions were assumed in the design of
A gas based combined cycle power generating capacity of
ground improvement system. Analysis and design was
108 MW was constructed at North Delhi. Main plant
carried out using the method stipulated by Priebe (1998) for
structures were proposed to be built on deep foundations
the initial in-situ density conditions. Ground improvement
whereas lightly loaded ancillary structures such as
using vibro replacement technique was adopted to mitigate
clariflocculator, storage tanks, switchyard etc. (loading
liquefaction.
intensity of about 10 t/m2) was proposed to be placed on
shallow foundations.
Vibro stone columns of 16 m long were installed in a square
grid pattern. Additional strips of stone columns were installed
The soil profile at the project site in general consists of loose
around the periphery of the tank to provide additional
to medium dense sandy soils with N values ranging between
stability to the treatment area in case of a seismic event.
5 and 10 to a depth of about 10 to 12m followed by dense
silty sands / sandy silt (N > 15 to 30) to about 30m. This
project site falls under Zone IV with peak ground
acceleration of 0.24g. The native loose sandy soil deposits
were susceptible to liquefaction to a depth of about 10 to 12m
below existing ground level.

Vibro replacement technique using dry vibro stone columns


(bottom-feed displacement method) was adopted to ensure
required bearing capacity to eliminate pile foundations. In
addition, the proposed ground improvement technique was
designed to mitigate the liquefaction potential in the event of
earthquake.

Fig. 13 Pre and Post CPT results


Fig. 15 Load Intensity vs. Settlement Curve
Madan Kumar Annam; V. R. Raju
The proposed technique allowed fast construction in a
congested site avoiding usage of water and subsequent muck
removal. Single column plate load test was conducted on the
installed dry vibro stone columns to a maximum load
intensity 50 t/m2. Settlement under the ultimate test load was
observed to be less than 16mm (Refer Fig. 15).

In addition to the technical performance and commercial


benefits, an embodied CO2 calculation showed an
environmental benefit (lower greenhouse gas emissions) of
the Vibro stone column solution as shown in Table 1.

Table 1 Embodied Carbon Comparison


Embodied CO2 [kg]
Sl.
Item Stone
No. BCIS Piles
Columns
1 Concrete
667,516 -
(incl. wastage)
2 Stone aggregates
- 42,412
(incl. wastage)
3 Reinforcement
382,373 -
Steel
4 Fuel consumption
for installation 199,354 118,350

Total 1,249 T 161 T

Product Packaging Unit at Babrala, Uttar Pradesh


Expansion of the existing Product Packaging Unit was under Fig. 16 Typical borehole log at Babrala
development at Babrala, Uttar Pradesh. As part of expansion,
various structures such as Conveyor Belts, MCC cum Control Plate load tests were performed over the treated area and
room and the Wagon Loading Platform were to be results confirmed that the settlements at design loads were
constructed. within the allowable limits.

The soils at the project site consisted of loose to medium


dense sand with fines less than 10% to about 12 m depth
from EGL with top 3.0 to 4.0m of clayey silt. The ground
water table was encountered at a depth of 3.0 m from EGL.
The site was prone to liquefaction during an event of
earthquake as it falls under seismic Zone IV.

Ground improvement using vibro compaction/vibro


replacement technique was proposed to mitigate the
liquefaction potential of the soil to enhance safe bearing
capacity of soil and reduce the estimated total and differential
settlement of the soil to a depth of 12 m below EGL. Fig. 17 Sketch showing ground improvement with
combination of vibro techniques at MCC Room of the plant
Vibro stone columns were installed for conveyor belt
foundations and the transfer towers. A combination vibro Post treatment CPTs’ were also executed which proved an
stone columns and vibro compaction has been used to support average improvement of two folds in the Qc values and of
the foundations of MCC room which is the most interesting 70% relative density was achieved. Also, post treatment SPTs
aspect of the project. showed a considerable improvement in SPT N values to
about 2 to 3 times in the treated area as shown below in Fig.
The vibro stone columns were constructed in the top 4m 18.
followed by vibro compaction up to 12m. An illustrative
sketch is shown in Fig. 17.
Ground improvement solutions to mitigate liquefaction and its application

Fig. 18 Pre and Post SPT comparison

CONCLUSIONS
The presence of liquefaction soil does not mean that one has
to abandon the site or to install deep foundations. In seismic
zones with liquefiable soils, ground improvement technique
provides technically sound and cost effective solutions.

Efficient and economic solutions to problems caused by soil


conditions require a thorough evaluation of project
conditions, project needs, method capabilities and a field test
program.
Six projects have been described in which vibro methods
have been successfully used to accomplish the required
ground improvement to mitigate liquefaction and to enchance
bearing capacity requirements. The success is measured in
terms of either load tests or by the comparison of results of
pre and post soil investigation done at particular sites.

REFERENCES

1. Seed H.B., and Idriss, I.M (2001). Simplified procedure


for evaluating soil liquefaction potential, J. Geotech
Engineering. Div., ASCE 97 (9), 1249-1273
2. Seed et. al (2003). Recent advances in soil liquefaction
engineering: A unified and consistent framework.
3. Priebe, H.J (1995), The Design of Vibro Replacement,
Ground Engineering, December 1995
4. Priebe, H.J (1998), Vibro Replacement to Prevent
Earthquake Induced Liquefaction, Ground Engineering,
September 1998
5. Raju et. al. (2010). Some Environmental Benefits of Dry
Vibro Stone Columns in a Gas Based Power Plant
Project, Indian Geotechnical Conference, December
2010
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice

Annexure 5 Technical paper on “Ground Improvement Using Vibro


Techniques in Flyash Deposits”

20
National Conference on Recent Advances in Ground Improvement Techniques
February 24-25, 2011, CBRI Roorkee, India

Ground Improvement Using Vibro Techniques in


FlyAsh Deposits
Dr V R Raju
Keller Ground Engineering (I) Pvt Ltd, India

ABSTRACT
Ash ponds are extensively located in India. Unavailability of suitable construction site for extension of existing
power plants or to build a new plant makes it worthwhile to consider the ash pond as one of the options. Ash
ponds in general are not consistent with the depth and density characteristics of the hydraulically deposited fly
ash across the site. The traditional methods of foundation design in such situations may result in commercially
unviable solution. Ground improvement in such case provides a techno-commercially feasible solution. Anpara
Thermal Power Plant by Uttar Pradesh Rajya Vidyut Utpadan Nigam Ltd (UPRVUNL) is one such classic
example. The site allocated for the proposed development of Unit D of the power plant is an abandoned ash
pond. An extensive research has been done and was established that ground improvement using stone columns
(dry bottom feed method) shall be adopted to not only mitigate the liquefaction potential but also to enhance the
bearing capacity of the hydraulically deposited fly ash deposits. The stone columns are also installed to enhance
the lateral capacity of bored cast-in-situ piles. This paper illustrates the soil conditions, proposed ground
improvement technique to address the geotechnical applications of bearing capacity, liquefaction mitigation and
enhancing the lateral capacity of piles and discuss the pre and post treatment testing.

1. INTRODUCTION
Ash ponds are extensively located in India. The area occupied by the ash ponds is more than
250sq.km and it is likely to cross 1,000sq.km by 2012 (Bedanga Bordoloi and Etali Sarmah,
2010). Unavailability of suitable construction site for extension of existing power plants or to
build a new plant makes it worthwhile to consider the ash pond as one of the options. Ash
ponds in general are not consistent with the depth and density characteristics of the
hydraulically deposited fly ash across the site. This results in inadequate bearing capacity and
lateral capacities of deep pile foundations. The traditional methods of foundation design in
such situations may result in commercially unviable solution. Ground improvement in such
case provides a techno-commercially feasible solution.

Anpara Thermal Power Plant in Uttar Pradesh is a classic example for such case. Uttar
Pradesh Rajya Vidyut Utpadan Nigam Ltd (UPRVUNL) is expanding the existing power
plant by setting up Unit-D of 2 x 500 MW capacity at Anpara, near Sonebhadra. The site
allocated for the proposed development is an abandoned ash pond of area approximately
5,400 acres. The depth of ash varies across the site and ranges between 3m and 13m and is
loose to medium dense in condition. It was found out during the initial soil investigation that
the existing bearing capacity of the fly ash deposits is the less than the required i.e., 10T/m2

1
for open foundations of structures like pump house, cable gallery etc at coal handling plant.
Also, site falls under Zone –III according to the IS 1893 (Part 1):1982, making it susceptible
to liquefaction in an event of an earthquake. An extensive research has been done (study of
effectiveness of ground improvement techniques and possible liquefaction potential for
Anpara D Thermal Power Project, IIT Roorkee) and was established that ground
improvement using stone columns (dry bottom feed method) shall be adopted to not only
enhance the bearing capacity but also to mitigate the liquefaction potential of the fly ash
deposits. Further, the stone columns are also installed surrounding the bored cast-in-situ piles
to enhance the lateral capacity for structures like stacker reclaimer, crusher house etc of coal
handling plant, which otherwise was giving low lateral capacity.

2. SOIL CONDITIONS
The project site is an old ash pond. Depth and density characteristics of fly ash vary across
the site. The depth generally ranges between 3m and 13m underlain by clayey silt / silty clay
to about 23m depth. Below this dense sandy silt or hard clayey silt was found and
occasionally weathered rock (granitic gneiss) is encountered.

Density characteristics vary considerably within the depth of fly ash. The SPT N values
recorded are as low as 2 to as high as 30, but generally vary from 3 to 8. This is followed by
stiff to hard clayey silt with SPT N values ranging between 9 and 30. The following figure-1
illustrates the cross sectional profile of the site indicating the variation in depth and density
characteristics of fly ash.

Fig. 1 Typical sectional profile illustrating the soil conditions at Coal Handling Plant location

2
3. GROUND IMPROVEMENT TECHNIQUE
Vibro Stone columns using bottom feed method is adopted as a ground improvement
technique. This method does not require water for penetration, which avoids the need to
handle and dispose large quantities of muck and also makes it environment friendly. It is also
well suited for a congested site, with many simultaneous activities. For this method of
installation, a rig called Vibrocat is used. It consists of a bottom-feed depth vibrator mounted
on a crawler-rig. An operational advantage of the Vibrocat is that it is able to exert a pull-
down force, improving penetration speed and hence productivity. A typical Vibrocat unit,
used on site, is shown in Fig. 2.

Fig. 2 A typical Vibrocat unit


The Vibrocat feeds the coarse granular material to the tip of the vibrator with the aid of
pressurized air. The installation method consists of alternative steps of penetration and
retraction. During the retraction, gravel runs from the vibrator tip into the annular space
created and are then compacted using vibrator thrusts and compressed air. Fig. 3 illustrates
the schematic of this process.

Fig. 3 Schematic of stone column installation (Dry bottom feed method)

3
4. GEOTECHNICAL APPLICATION
Ground improvement using vibro stone columns by bottom feed has been adopted to achieve
the following objectives:

4.1 Improve Bearing Capacity of Open Foundations

The density characteristics of fly ash vary across the site as a result the net safe bearing
capacity for open foundations is less than the desired value of 10T/m2. It is proposed to install
stone columns to at least 0.5m into the underlying stiff clayey silt / silty clay layer to achieve
the desired bearing capacity for open foundations.

4.2 Enhance Lateral Capacity of Piles

The existence of loose fly ash deposits resulted in less than the desired lateral capacity of
bored cast-in-situ piles. Stone columns were installed around the bored cast-in-situ piles to
enhance the density characteristics of the fly ash surrounding the piles there by improving the
lateral capacity to 7T (working load).

4.3 Mitigate Liquefaction Potential

According to IS 1893 (Part 1):1982, the site falls under Zone –III making it susceptible to
liquefaction in an event of an earthquake under the possible excitation or peak ground
acceleration of 0.16g. According to Table 1, Note 4 of IS 1893 (Part 1):1982, soils with SPT
N values less than 20 for Zone III are liable to liquefy.

The SPT N values obtained range between 3 and 8 within the fly ash depth indicating the
possibility of liquefaction in an event of an earthquake. The proposed stone columns
increases the density characteristics of the fly ash, there by not only enhance the bearing
capacity (section 3.1) but also mitigate the liquefaction potential.

It was proposed to adopt dry bottom feed method for installation of stone columns to achieve
above applications. Extensive initial field trials were carried out before carrying out the main
works to assess the suitability of the technique as well as to assess the required grid pattern to
achieve post performance criteria.

4
5. INITIAL FIELD TRIALS
Initial field trials were carried out to assess the bearing capacity and also the lateral capacity
of bored cast-in-situ pile foundations as a result of stone column installation. The following
sections illustrate the field trials carried out elaborately to address the above listed
geotechnical applications.

5.1 Bearing Capacity

Vibro stone columns of 0.9m diameter at 2m centre to centre spacing in a triangular grid
pattern, terminating at least 0.5m into the underlying stiff silty clay or clayey silt is proposed
as treatment scheme to achieve the target bearing capacity of 10T/m2. According to the guide
lines stipulated in IS 15284 (Part 1): 2003 – “Design and Construction for Ground
Improvement – Guide Lines”, single and group column initial load tests are performed at site
to assess the increase in bearing capacity as well as the settlements characteristics of stone
columns. The following figure 4 illustrate the results of plate load tests conducted on single
and group of 3 columns at coal handling plant location.

(a) (b)
Fig. 4 Load Vs Settlement plot of (a) Load Test on Single Column (b) Load Test on Group of 3 columns

5.2 Lateral Capacity of Piles

Vibro stone columns are installed at specified pattern (as illustrated in figure 5) surrounding
the bored cast-in-situ piles to enhance the density of fly ash deposits which in turn can
improve the lateral load carrying capacity. It was required to achieve a design lateral load
capacity of 7T with ultimate load of 21T. After the installation of bored cast-in-situ piles and
vibro stone columns by bottom feed method, initial lateral load test are conducted on these
two grid patterns.

5
Fig. 5 Initial field trials layout for lateral load (a) Stone Column of 500mm dia. (b) Stone Columns of 750mm
diameter surrounding the piles with 0.5m dia stone column at the centre

The results indicated that the deformations are within the allowable limits of 5mm at the
design load of 7T (according to IS 2911, Part 4, Cl. 7.4) even for 0.5m grid pattern shown in
fig 5(b). The following figure 6 illustrates the observations made during the initial lateral load
tests (load vs deflection plot).

Load Vs Deflection

Load [T]
0 10 20 30
0.0

5.0
Settlement [mm]

10.0
0.5m dia. Stone Column
grid1
15.0

0.75m dia. Stone Column


20.0 Grid

25.0

Fig. 6 Load Vs deflection plot of initial lateral load test on bored cast-in-situ piles

6. EXECUTION OF MAIN WORKS


Upon successful execution of initial field trials to assess the improvement in density
characteristics of fly ash deposits after ground improvement, main works have been carried
out. Ground improvement using vibro stone columns (dry bottom feed method) is carried out
for open foundations of the structures like pump house, cable gallery, drive house etc at coal
handling plant. About 34,000lin.m of vibro stone columns 0.9m dia are installed for open
foundations of various structures of coal handling plant.

6
The following figure 7 illustrates the typical drawing of stone columns installed for open
foundations of pump house structure of coal handling plant.

Fig. 7 Details of Stone Columns installed for open foundations of pump house Structure

Similarly, to enhance the lateral load carrying capacity of bored cast-in-situ piles of structures
like conveyor, crusher house etc of coal handling plant, 45,000 lin.m of 0.5m diameter vibro
stone columns are installed. The following figure 8 illustrates the schematic of stacker
reclaimer at coal handling plant, where stone columns are used to enhance the lateral capacity
of piles.

Fig. 8 Typical details of Stone columns installed surrounding the piles to enhance the lateral capacity

7
The following pictures (figure 9) illustrate the works in progress at coal handling plant
locations.

(a) (b)
Fig. 9 (a) Installation of Stone Columns and Bored Cast-in-situ Piles using Hydraulic rigs in progress
(b) Conveyor 9A Structure under construction – here piles in combination with 500mm dia piles are used as
foundation

6. QUALITY CONTROL AND QUALITY ASSURANCE


Quality control procedures are important firstly to assure the client that the product he
receives is of a high standard, secondly to prevent costly re-work for the contractor and most
importantly to ensure public safety. Generally, quality control is applied pre-construction,
during construction and post-construction. Various standards can be used to aid in the
formulation of good contract specifications and quality control procedures.

Fig. 10 Typical quality control record print out generated simultaneously during installation of stone
columns

8
For Vibro Stone Columns, it is essential to ensure that columns are built to the right depth, to
the right diameter and are properly compacted. Computerized monitoring (as shown in figure
10) of the penetration depth of the vibrator easily ensures that the design depth is reached.
Sensors within the depth vibrator can readily measure the amperage drawn by the motor,
giving an indication of the compaction effort of the depth vibrator. IS 15284 (Part 1): 2003
gives guidelines on the estimation of the column diameter based on fill consumption. In the
case of dry bottom-feed stone columns (See Raju & Sondermann, 2005), even the location of
each charge of stone along the depth of the column may be determined from the record of
depth vs. amperage. Post-construction, load tests are routinely performed as a quality control
measure. Another useful general standard for stone column construction and testing is EN
14731:2005.

7. CONCLUSIONS
In view of the unavailability or scarcity of the suitable construction site ash ponds form one
of the options to consider for the proposed development. Detailed study is required to carry
out to check the suitability of the ash pond for the proposed development. In general it has
been noticed that the geotechnical characteristics i.e., density of hydraulically deposited fly
ash is not consistent with depth which may pose challenges with regards to bearing capacity
and pile lateral load carrying capacity. In such a situation, ground improvement using dry
vibro stone columns provides a techno-commercially feasible solution. It is evident from the
experience in Coal Handling Plant structures at Anpara D Thermal Power Plant in Uttar
Pradesh that the ground improvement using stone columns (dry bottom feed method) can be
successfully adopted to enhance the bearing capacity of the fly ash deposits. Stone columns
also ensured mitigation of liquefaction potential of the site under an event of earthquake as
the site falls under zone III. Further, the stone columns also helped in enhancing the lateral
capacity of deep pile foundations. With this experience, similar application is adopted at
other structures of the power plant such as Switchyard and for the Water Treatment Plant
structures, which is currently under construction.

References
1. Article by Bedanga Bordoloi and Etali Sarmah on “Fly Ash Pond Reclaimation” dated May 2010 for
Agribusiness Forum.
2. IS 1893 (Part 1):1982 “Criteria for Earthquake Resistant Design of Structures - Part 1 : General Provisions
and Buildings”

9
3. “Study of Effectiveness of Ground Improvement Techniques and Possible Liquefaction Potential” for
Anpara-D Thermal Power Project, Department of Earthquake Engineering, IIT Roorkee

4. IS 15284 (Part 1): 2003 – “Design and Construction for Ground Improvement – Guide Lines”
5. IS 2911 (Part 1) Section 2- “Code of practice for design and construction of pile foundations: Part 1
Concrete piles, Section 2 Bored cast-in-situ piles”
6. IS 2911, Part 4, “Code of practice for design and construction of pile foundations: Part 4 Load test on
piles”
7. Raju, V.R. & Sondermann, W., 2005. Ground Improvement using Deep Vibro Techniques. Ground
Improvement Case Histories, Indraratna, B & Chu., J. (eds.), 601-638
8. EN 14731:2005, “Execution of special geotechnical works. Group treatment by deep vibration”

10
Technical Note on Ground Improvement using Vibro Compaction and Stone Columns: Theory & Practice

Annexure 6 Observations on IS 15284 part 1 Rev 0

21
  
          
        

 
                !" #   $

•  %&' ()*%+,(%-) -.,*%+/ 0* 1+),. -23+)4'2'% &0* 5'()2' 2)*% ()22) 3+0(%-(' - %&'

60*% 7 %) 8 /'0+*9 %&-* -* &-1& %-2' %) 3+'30+' :+'4-*' *%0.0+. ;)+ 066 %/3'* ); 1+),.

-23+)4'2'% 2'%&).*<

• =66 4-5+0%-) 2'%&).* 6->' 4-5+) ()230(%-)9 4-5+) +'360('2'% 0. 4-5+) .-*360('2'% %) 5'

0..'. - %&' 3+'*'% *%0.0+. ? @7ABC D30+% @E )+ 0 *'30+0%' *%0.0+. ) FG+),. H+'0%2'%

5/ I''3 J-5+0%-)*K 20/ 5' 5+),1&% ),% D*-2-60+ %) L?E<

•  %&' 3+'*'% *%0.0+. D ? @7ABC M 30+% @E N' ;''6 %&0% %&' ;)66)N-1 3)-%* %) 5' -()+3)+0%'.9

N&-(& N-66 1-4' 2)+' (60+-%/ 0. ,*';,6 ;)+ 3+0(%-(-1 '1-''+*<

\WTUT[Z YW]VZXVT
OP QRP STUVWXYZXR[ iTVRjjT[kTk
^]U YTW _O `abcd ^Y]WZ `ef bgghe

@ I-02'%'+ ); ?%)'  %&' 3+'*'% *%0.0+. Dl60,*' % -* +'()22'.'. %&0% %&'

()6,2 m<@E %&'+' -* ) *3'(-;-( .-02'%'+ ); 4-5+0%'. *%)'

+'()22'.0%-) ;)+ %&' .-02'%'+ ()6,2 *&066 5' - %&' +01' );

); *%)' ()6,2* n<B2 %) @<A2

A =+'0 ); %+'0%2'% o+'*'% *%0.0+. Dl60,*' 8<CE  3+0(%-(' 4-5+) *%)' ()6,2*

*3'(-;-'* %&0% %&' 0336-(0%-) ); 0+' 5'-1 ,*'. 'p%'*-4'6/ ;)+

*%)' ()6,2* -* 6-2-%'. %) 5,-6.-1*9 -.,*%+-06 *%+,(%,+'*9

'250>2'%*9 %0>* '%(< -;+0*%+,(%,+' 3+)q'(%* 6->'

+0-6N0/*9 0-+3)+%*9 3)+% ;0(-6-%-'*9

3)N'+ 360% *%+,(%,+'* '%(< %)

-23+)4' 5'0+-1 (030(-%/9 %)

+'.,(' *'%%6'2'%* 0. %)

2-%-10%' 6-r,';0(%-) 3)%'%-06

s H+'0%2'% .'3%& o+'*'% *%0.0+. *0/* %&0% %&'  3+0(%-(' N-%& %&' 3+'*'%

04'+01' .'3%& ); %+'0%2'% 20/ 'r,-32'% 040-605-6-%/9 %&'

5' 0+),. @72 .'3%& ); *%)' ()6,2 *&066 5'

0* 1-4' 5'6)N

@< H&' .'3%& ); *%)'

()6,2 ,*-1 N'% %)3

;''. 2'%&). -* A72 %)

sn2

A< t*-1 .+/ 5)%%)2u;''.

2'%&). (0 ()*%+,(% 0

()6,2 ); @72 %) An2

.'3%&

C ?30(-1  l60,*' m<s ); 3+'*'% *%0.0+.9 % -* ,*,06 3+0(%-(' %&0% *30(-1

-% -* 2'%-)'. %&0% %&' ()6,2 ); *%)' ()6,2 5+)0.6/ +01'*

*30(-1 5+)0.6/ +01'* ;+)2 A %) ;+)2 @<7 %) A<7 %-2'* ); *%)'

s< % -* )% (6'0+ ()6,2 .-02'%'+ 50*'. )

.'*-1 =+'0 v'360('2'% v0%-)


\WTUT[Z YW]VZXVT
OP QRP STUVWXYZXR[ iTVRjjT[kTk
^]U YTW _O `abcd ^Y]WZ `ef bgghe

7 =+'0 v'360('2'%  %&' 3+'*'% ().' %&'+' -* )  3+0(%-(' %&' =+'0

v0%-) *3'(-;-( +01'* ); =+'0 v'360('2'% v0%-) ); @nw %)

v'360('2'% v0%-) snw -* 5'-1 ,*'. N-.'6/ 50*'.

) %&' *)-6 ().-%-)9 6)0.-1

().-%-) 0. .'*-1

3'+;)+20(' (+-%'+-0

8 *%0660%-) %'(&-r,'*  ='p,+' lu@<A ); 3+'*'% H&' *3'(-;-'. 2'%&). -*

*%0.0+. 'p360-'. 05),% %&' 3+0(%-(066/ )% ;'0*-56' *) -% *&066

-*%0660%-) ); *%)' ()6,2 ,*-1 5' +'2)4'.

FSXWTVZ jxk VXWVxy]ZXR[ jTZzRkK

- N&-(& 5'%)-%' *6,++/ -* ,*'.

;)+ UZ]{XyX|X[} ZzT {RWTzRyT 0. -%

*&066 5' YxjYTk RxZP

m *%0660%-) %'(&-r,'* ='p,+' l ); 3+'*'% *%0.0+. % *&066 5' 2).-;-'. 0* 5'6)N

1-4'* %&' ;)66)N-1 2'%&).* ;)+

%&' -*%0660%-) ); *%)' ()6,2 lu@ ~)u .-*360('2'% 2'%&).

luA I-*360('2'% 2'%&).

lu@ ~)u .-*360('2'% 2'%&). lus J-5+) 2'%&).

luA I-*360('2'% 2'%&). s<@< I-*360('2'% 2'%&).

lus J-5+) +'360('2'% 2'%&). s<A< J-5+) +'360('2'%

0E '% 3+)('** 2'%&).

5E I+/ 3+)('** 0E '% 3+)('**

Dv'360('2'%€ %)3 ;''.

2'%&).E

5E I+/ 3+)('**

DI-*360('2'%€ 5)%%)2

;''. 2'%&).E

B J-5+) +'360('2'% ='p,+' lus D5E ); 3+'*'% I+/ 3+)('** D5)%%)2 ;''.

2'%&). *%0.0+. 'p360-* %&0% I+/ 3+)('** 2'%&).E -* .-*360('2'%

-* *,-%056' ;)+ *)-6* +'60%-4'6/ &-1& 2'%&). 0. %&' *%)' *&066 5'

--%-06 *%+'1%& N-%& +'60%-4'6/ 6)N .'6-4'+'. 0% %&' %-3 ); 4-5+0%)+

N0%'+ %056' N&'+' %&' &)6' (0 5' %&+),1& 5)%%)2 ;''.

*%0. ); -%* )N ,3) 'p%+0(%-) 2'(&0-*2 0. *&066 5' ,*'. ;)+

); %&' 3+)5'9 *,(& 0* ,*0%,+0%'. 066 %/3' ); *)-6* 0. -++'*3'(%-4'

;-66* ); N0%'+ %056' 0* %&' 4-5+0%)+

*%+-1 N-66 1) -%) %&' +'r,-+'.

.'3%& ); %+'0%2'% 0. +'20-*

- %&' &)6' .,+-1 ()6,2

()*%+,(%-)

 ‚'0*,+'2'% ); *%)' =;%'+ -*%0660%-) ); *%)' ()6,2 % -* ()22) 3+0(%-(' %&0% %&'

()6,2 .-02'%'+ %&'+' -* ) *3'(-;-( *%)' 4)6,2' ()*,2'. ;)+

2'%&).:3+0(%-(' -* 2'%-)'. %) 20>-1 0 *-16' ()6,2 -* ,*'.

2'0*,+' %&' .-02'%'+ ); ()6,2 ;)+ ();-+2 %&' .-02'%'+ 0* %&'

.'3%& ); ()6,2 -* >)N


\WTUT[Z YW]VZXVT
OP QRP STUVWXYZXR[ iTVRjjT[kTk
^]U YTW _O `abcd ^Y]WZ `ef bgghe

@n ƒ*%-20%-) ); 6)0.  %&' 3+'*'% *%0.0+. Dl60,*' =6%'+0%-4' 2'%&).* 6->' „'-… †<

(030(-%/ <s<AE -% -* 2'%-)'. %&0% 0/ o+-'5' 2'%&).* D%&' .'*-1 );

)%&'+ 06%'+0%-4' 2'%&).* (0 5' 4-5+) +'360('2'%E (0 5' ,*'.

,*'. ;)+ %&' '*%-20%-) ); 6)0.

(030(-%/ N&-(& -* 0(('3%'.

-%'+0%-)066/

@@ *%0660%-) %'(&-r,'* ‡)+ N'% 0. .+/ 2'%&).9 %&' H&' -*%0660%-) 3+)('** 0.

-*%0660%-) 3+)('** -* )% 2'%&). 'p360-'. - L+-%-*&

'p360-'. - %&' *%0.0+. *%0.0+. D5'6)N 3-(%,+'E *&066 5'

-(6,.'. ;)+ 2)+' (60+-%/

ˆ‰Š ‹‰ŠŒŽ ‘ ’‰‰Ž“”

•–— ‹‰ŠŒŽ ˜ŠŠ‹ ’‰‰Ž“

@A G+0.-1 ); *%)'* H&' 3+'*'% *%0.0+. %&'+' -* ) H&' ()22) 3+0(%-(' 0. 0* 3'+

*3'(-;-( 1+0.-1 ;)+ .-;;'+'% L? %&' ;)66)N-1 1+0.-1 *&066 5'

2'%&).* ); *%)' ()6,2 0.)3%'.

@< ‡)+ N'% 3+)('** %&'

1+0.' ); *%)' -* @A %)

m722

A< ‡)+ .+/ 3+)('** D5)%%)2

;''.E %&' 1+0.' ); *%)'

-* @A %) Cn22
iT™TWT[VTUf

@< L? ƒ~ @Cms@€ Ann79 Fƒp'(,%-) ); *3'(-06 1')%'(&-(06 N)+>* MG+),. %+'0%2'% 5/ .''3

4-5+0%-)*K<

A< Lvƒ G0+*%)9 0%;)+.9 IA m †v9 F?3'(-;/-1 4-5+) *%)' ()6,2*K<
MONOGRAPH ON

GRANULAR PILES AND GRANULAR PILE ANCHORS

1
Madhav Madhiraand 2Vidyaranya Bandi
1 Prof. Emeritus, JNT University and Visiting Professor, IITH, Hyderabad, 500085, India
2 Engineering Manager, L&T TI-IC, Mumbai 400097, India.

1. INTRODUCTION

Granular piles are often constructed through soft soils fully penetrating to an end

bearing stratum or as floating piles in deep deposits, the tips restingat depths where the

strength of the soil is adequate.

Failure mechanisms (Fig. 1) for a single granular pile are bulging, general shear and pile

failure though probable failure is by bulging or pile failure. Methods to estimate the

ultimate capacity of granular piles corresponding to general shear, bulging and pile failures

are presented in Table 1. In pile failure mode, the total load applied on the granular pile is

resisted by shaft resistance generated along the shaft length and the bearing resistance at

the base of the GP while the resistance generated by lateral confinement of the granular fill

material near the top in GP limits its bulging capacity.

The functional utility of the granular pile to carry the compressive load is extended to

resist the uplift or pullout forces generated in foundations by a simple modification of

connecting the base of the foundation to a plate, pedestal or geogrid at the tip of the

granular pile by a cable or rod to transfer the pullout load (Fig. 2).

1
Fig. 1 (a) Bulging, (b) General Shear and (c) Pile Failure Mechanisms for Single
Granular Pile

Uplift Force, Po
G.L

Cable
Dense
Granular Fill
L d
In Situ Soil

Plate/Pedestal

Fig. 2 Granular Pile Anchor.

2
Table 1 Estimation of Ultimate Load (Aboshi and Suematsu, 1985)
Mode of
Derived Formula References
Failure
sin 

q ult   c zk pc  2c o k pc 11  sin 
s
,
s
Greenwood
where kpc is the passive earth pressure coefficient of column (1970)
and s is the frictional resistance of soil


qult  Fc1C o  Fq1 Qo 11  sin 
sin 
s
, Vesic (1972),
s
Datye &
where Fc1 and Fq1 are the cavity expansion factors and Qo Nagaraju
is the surcharge stress (1975)
Bulging
1  sin  s Hughes and
q ult   ro  4C o  ,
1  sin  s Withers
(1974)
1  sin  s
qult 
1  sin  s
 2 2
4C o   ro  K o q s  W B  1  W B q s
 
 
Madhav et al.
where W and B are diameters of stone column and footing (1979)
respectively.

1 
q ult  C o N c    c BN     c D f N q ,
2  Madhav and
where Nc, Nq and Nγ are the dimensionless parameters that Vitkar (1978)
General depend on the trench and soil parameters.
Shear 1 
qult    c B tan 3    2Co tan 2   2(1  a s )C o tan
2  Barksdale and
tan  s a s tan  s 
1
Bachus (1983)
  45 0 
2

 
q ult  1  a s C o   s z   s z a s tan  s cos 2 
n ,
Sliding s  Aboshi et al.
1  n  1a s
Surface (1979)
where ar is area replacement ratio and s & γs are column
parameters

3
2. ULTIMATE CAPACITY OF GRANULAR PILE (GP) IN HOMOGENOUS
GROUND

Ultimate capacity of granular pile in compression is estimated for homogenous ground,

i.e., the undrained strength, cu, of the in situ soil is constant with depth.

2.1 Single Granular Pile in Compression

A granular pile of diameter, d, and length L, is considered (Fig.3). The saturated unit

weight, the undrained strength and the shear modulus of the in situ soil assumed constant

with depth are s, cu and G respectively while  gp and gp are respectively the angle of

shearing resistance and unit weight of the granular pile material.

The ultimate pile capacity, Pcomp, is limited by the interface shear stresses, τ, acting on

the cylindrical boundary and the ultimate bearing stress, q b, at the base of the GP (Fig. 4a).

The ultimate shear stresses,  equals the undrained shearstrength, cu, while the limiting

bearing stress, q b, equals Nc.cu (Fig. 4b).The ultimate capacity, P ult, of GP in compression

 .d 2
by pile capacity after normalization with .cu reduces to
4

* L
Ppf  4  Nc (1)
d
where Ppf * = Pult, pf/{d2/4}cu and Nc – bearing capacity factor that varies from 6.2 to 9

for L/d increasing from 0 to 5 or more.

4
Compressive Load, P

G.L

gp, gp

L d
G, s, cu

Fig. 3 Granular Pile under Compression.

Compressive Load, P

Base Resistance, qb

(a) (b)
Fig. 4(a) Pile & (b) Bulging Failures for GP.

5
For bulging failure, following Gibson and Anderson (1961), Hughes and Withers (1974)

and Hughes et al. (1975), for expansion of a cavity near the top (at a depth of d/2 from the

top) of the GP, Pult, bf is

.d 2
Pult, bf  N cu .Nc*   ho  (2)
4

where the lateral confining pressure  h o , is the horizontal total stress at depth equal to half

(1  sin  gp )
the diameter (d/2) of GP, N c*  1  ln  G  and N   . Normalizing Pult, bf
c  (1  sin  gp )
 u 

 .d 2
with .cu Eq. (2) reduces to
4

4 Pcomp
P* 
d 2 cu

 N  N *c    (3)

  w .d  K o . sub 
where      1
 cu   w 

The critical length, (L/d)cr defined as is the length at which the ultimate capacities by

pile and bulging failures equal. The ultimate capacity is governed by pile failure for L/d

smaller than the critical length and by bulging falure for L/d greater than the critical length.

2.2 Ultimate Pullout Capacity of Granular Pile Anchor (GPA) – Homogenous Ground

The applied pullout load is transferred to the base through the cable or steel rod attached

to the base plate, pad or sheet placed prior to the installation of the granular pile material

(Fig. 5).

The ultimate pullout capacity of the GPA is the lesser of the loads at which it is either

pulled out by pile (Fig. 6a) or by bulging (Fig. 6b) failure. The normalized ultimate

capacity, P* of GPA by pile capacity is


6
Pullout Load, Pult

G.L

gp, gp

d
L
G,s, cu

Fig. 5 GPA under Pullout

τ Pult
τ
τ
W ‘ = gp.V

Pult ‘ Pullout Load
L-d/2



τ
τ σh
(a) (b)

Fig. 6 Pullout (a) and Bulging (b) Failures of GPA.

4 Pult L
P*  2
 (4   ) (4)
d cu d

7
 gp.d
where 
cu
Bulging is considered likely to occur at a distance of half-diameter of the GPA from the

tip instead of from the top as was considered for bulging capacity of granular piles in

compression. The bulging capacity of the GPA is

 .d 2
Pult  .N cu .Nc*   ho  (5)
4

The total horizontal stress, h0, is considered at depth z  L  d  2


 assuming groundwater
level to be at ground level. The normalized ultimate pullout load by bulging, P* is

4 Pult   L 1 
P*  2
 N   N c*   .    (6)
d cu   d 2 

where    w .d  K . sub  1 - a lateral confining stress parameter that depends particularly
cu   w 

on lateral earth pressure coefficient, K, of the in situ soil. The critical length, (L/d)cr

defined as is the length at which the ultimate capacities by pile and bulging failures equal.

The ultimate capacity is governed by pile failure for L/d smaller than the critical length and

by bulging falure for L/d greater than the critical length.

2.3 Results

The ultimate capacity of GP in compression and the ultimate pullout capacityof GPA

are estimated for both the pile and bulging failure mechanisms using Eqs. 1 & 3 for GP and

4 & 6 for GPA for the following ranges of the parameters: s: 14 to 16 kN/m3; gp: 18 to 21

kN/m3; cu: 10 to 60 kPa; L/d: 1 to 25; gp: 300 to 450; G/cu: 50 to 500; sd/cu: 0.1-2;

(=gpd/cu ): 0.1-2.5; γsubd/cu: 0.03 to 0.7; γwd/cu: 0.08 to 1.2, β = 0.1 - 1.6 and K0=0.5-1.0.

8
The ultimate capacity of GP is presented in Fig. 7 as a function of L/d for φgp in the

range of 300 to 450, for G/cu = 100 & 200, and β = 1.0.

45

450

400
30

350
P*

gp=300

15
G/cu=100
200

0
0 10 L/d 20 30

Fig. 7 Ultimate capacity, P* of GP - Effect of φgp for G/cu = 100 and 200 & β = 1.0.

9
30

2.6

1.9
P*

1.45
25 1
0.55
 = 0.1

20
0 15 30
L/d

Fig. 8 Ultimate capacity, P*of GP - Effect of for gp = 350 & G/cu=200.

Fig. 8 depicts the effect of the lateral stress parameter, , on the ultimate capacity of the
GP.

10
450

400
(L/d)cr

5 350

gp =300

0
0 300 600
G/cu

Fig.9 Critical length, (L/d)cr for GP in compression – Effect of gp for β = 1.0.
10
The effects of G/cu and gp on the critical length, (L/d)cr of GP are shown in Fig. 9.

The variation of ultimate pullout capacity, P * with L/d showing the effect of for = 1,

G/cu=200 & φgp=350 is depicted in Fig. 11.

The ultimate pullout capacity of GPA is presented in Fig. 10 as a function of L/d and

includes the effects of G/cu &  gp for 1.3 & =1.0. It may be noted that the ultimate

pullout capacity increases with L/d even for bulging failure mode since bulging is expected

to occur near the tip of GPA.

200

450

400
P*

350
100
gp=300

G/cu =50
200
500
0
0 15 30
L/d

Fig. 10 Ultimate pullout capacity, P* vs L/d for GPA – Effect of G/cu & gp for 1.3 &
 = 1.0.

11
120

=0.1

Bulging Failure
P*

60 0.7
1.3
1.9
2.5

0
0 15 30
L/d

Fig. 11 Ultimate pullout capacity, P* vs L/d for GPA – Effect of for = 1, G/cu=200 &
φgp=35 0.

120
1
p*

0.5
60

=0.1

0
0 15 30
L/d

Fig. 12 Ultimate pullout capacity, P* vs L/d for GPA – Effect of  for = 1.3, G/cu=200 &
φgp=35 0.
12
The effect of the lateal stress parameter, , on ultimate pullout capacity of GPA is

depicted in Fig. 12.

20

350
(L/d)cr

10
gp=300

0
0 300 600
G/cu

Fig. 13 Critical length, (L/d)cr vs G/cu for GPA -Effect of gp for  1.0 in
GPA.

13
30

=0.7
(L/d)cr

15 1.3

1.9
2.5

0
0 300 600
G/cu

Fig. 14 Critical length, (L/d)cr vs G/cu for GPA - Effect of for &gp

16
1
(L/d)cr

8 0.55

=0.1

0
0 300 600
G/cu

Fig. 15 Critical length, (L/d)cr vs G/cu for GPA - Effect of for gp 

14
The variation of critical length, (L/d)cr of GPA with G/cu for gp varying from 300 to 450 is

shown in Fig. 13 while the variations with  and  in Figs 14 and 15 respectively.

3. NON-HOMOGENOUS GROUND

The undrained shear strength of in situ soil is considered (Fig.16) to increase linearly

with depth (non-homogenous ground), and the ultimate capacities of the GP and GPA

estimated. The variation of undrained shear strength of normally consolidated soil with

depth normalized with length of the granular pile, is expressed as

 z
cu ( z )  cuo 1   c  (7)
 L

where c, non-homogeneity strength parameter expresses the rate of increase of undrained

shear strength with depth.

cu

Non-homogenous
Ground
 z
cu  cuo 1   c 
 L

z
L

Fig. 16 Profile of undrained shear strength of the soil with normalized depth

15
3.1 Ultimate Capacity of Granular Pile (GP)

The ultimate compressive capacity, Pult of GP by pile capacity normalized with

 .d 2
.cuo is
4

 L   
P *   4.  N c .1  c  (8)
 d  2 

where the normalized compressive capacity, P*,of GP, P*  Pult is


.d 2 
 cuo
 4 
The normalized ultimate capacity, P*, of GP for bulging failure in non-homogeneous

ground is

4 Pcomp  d  * 
P*  2
 N  1   c  N c  0.5.  (9)
d cu 0  2.L  

3.2 Ultimate Pullout Capacity of Granular Pile Anchor (GPA)

The normalized of ultimate pullout capacity, P*of GPA for pile failure is

L
P*  {4(1  0.5. c )  } (10)
d

 gp.d
where   - function of the density of the granular fill material.
cuo

The undrained strength of the soil at distance d/2 from the tip of GPA where bulging is

expected to occur, is

 ( L  d / 2)   c  L 1 
cu  cuo .1   c   cuo .1  .    (11)
 L   (L / d )  d 2  

The normalized ultimate pullout load, P*, of GPA for bulging failure, is

16
4 Pult  c  L 1  *  L 1 
P*   N   1  .    N c   .    (12)
 d 2 c uo  (L / d )  d 2  d 2 

where    w .d   sub . K o  1 - lateral confining pressure parameter


c uo  w 

3.3. Results

The ultimate compressive and pullout resistances of GP & GPA in non-homogenous

ground are estimated for both the pile and bulging failure mechanisms using Eqs. 8 &10,

and 12 & 14 respectively for the following ranges of the parameters: s: 14 to 16 kN/m3;

gp: 18 to 21 kN/m3; cuo: 10 to 60 kPa; L/d: 1 to 25; gp: 300 to 450; G/cuo: 50 to 500;

sd/cuo: 0.1-2; (=gpd/cuo): 0.1-2.5; γsubd/cuo: 0.03 to 0.7; γwd/cuo: 0.08 to 1.2, αc =0.5 –

1.0, β = 0.1 - 1.6 and K0=0.5-1.0.

500

26 200

100
P*

G/cuo=50

10
0 15 30
L/d

Fig. 17 Ultimate compressive capacity, P* for GP vs. L/d - Effect of G/cu for φgp = 350, β
= 1.0 & c=0.5 in non-homogenous ground.

17
The variations of ultimate capacity of GP in non-homogeneous ground (undrained strength

increasing linearly with depth) with L/d for different G/cu and non-homogeneity parameter,

c are given in Figs. 17 and 18.

30

1
P*

0.7
27
0.5

c=0

24
0 12.5 25
L/d

Fig. 18 Ultimate compressive capacity, P* for GP vs L/d– Effect of c for G/cuo = 200, φgp
= 350 & β = 1.0 in non-homogenous ground.

18
5.5

c=0
(L/d)cr

0.5
0.7
4
1

2.5
0 300 600
G/cuo

Fig. 19 Critical length, (L/d)cr vs. G/cuo for GP – Effect of cfor φgp = 350 & β = 1.0 in
non-homogenous ground.

8
c=0

0.5
0.7
(L/d)cr

2
28 37 46
Angle of Shearing Resistance,  gp

Fig. 20 Critical length, (L/d)cr vs. gpfor GP– Effect of c for G/cuo=200 & β = 1.0 in
non-homogenous ground.

19
The effects of non-homogeneity parameter, c, as effecting the variations of critical length,

(L/d)cr of GP with G/cuo and φgp are given in Figs. 19 and 20.

4.8

c=0

4.2 0.5
0.6
0.7
(L/d)cr

3.6

3
0 1.5 3

Fig. 21 (L/d)cr vs.  for GP–Effect of cfor G/cuo=200 &φgp=35 0 in non-homogenous


ground.

Fig. 21 depicts the effect of non-homogeneity parameter, con the variation of (L/d)crfor

GP with  for G/cuo=200 & φgp=35 0.

20
150

500
200
100
P* G/cuo=50

75

0
0 15 30
L/d

Fig. 22 Ultimate pullout capacity, P* for GPA vs. L/d -Effect of G/cu for φgp=350, =1.3, β
= 1.0 & c=0.5 in non-homogenous ground.

Figs. 22 and 23 present the variations of ultimate pullout load of GPA with L/d for

different G/cu and c.

21
140
1
0.8
0.5
c=0
P*

70

0
0 15 30
L/d

Fig. 23 Ultimate pullout capacity, P* for GPA vs. L/d– Effect of c for G/cuo=200, φgp=350,
λ=1.3 & β = 1.0 in non-homogenous ground.

Figs. 24 and 25 show variations of (L/d)cr with (G/cu0) and φgp and show the effect of non-

homogeneity parameter, c.

22
16
c=0

0.5
0.7
(L/d)cr 1

12.5

9
0 300 600
G/cuo

Fig. 24 Critical length, (L/d)cr vs. G/cuo for GPA – Effect of cfor φgp = 350, λ=1.3 & β =
1.0 in non-homogenous ground.

27
c=0
0.5
0.7
(L/d)cr

17

7
28 35 42
Angle of Shearing Resistance,  gp

Fig. 25 Critical length, (L/d)cr vs. gp for GPA–Effect of c for G/cuo=200, λ=1.3 & β = 1.0
in non-homogenous ground.

23
27
c=0

0.5
(L/d)cr
0.7

1
17

7
0 1.25 2.5

Fig. 26 Critical length, (L/d)cr vs.  for GPA –Effect of cfor G/cuo =200, φgp =350 &
β=1.0 in non-homogenous ground.

The variations of the critical length, (L/d)cr of GPA with the parameters and β for

different non-homogeneity parameter, c, can be seen in Figs. 26 and 27 respectively.

24
50

c=0

0.5
(L/d)cr

25 0.7
1

0
0 0.8 1.6

Fig. 27 Critical length, (L/d)cr vs. for GPA –Effect of c for G/cuo =200, φgp =35 0 &
=1.3 in non-homogenous ground.

4. LOAD - DISPLACEMENT RESPONSE OF GP ANDGPA

Settlement of a granular pile under working loads is similar to that ofincompressible

floating pile in a half space, with correction for the effect of pile compressibility and is

given (Poulos and Davis 1980) as

P .I
  (13)
K s .d

where I  I o . R k . R h . R  , ρ - settlement at top of GP, P - applied axial load,

Io- settlement influence factor for incompressible pile in semi-infinite mass, for νs=0.5, RK

- correction factor for pile compressibility, Rh - correction factor for finite depth of layer on

25
a rigid base, and Rν - correction factor for soil Poisson’s ratio, νs. Plot of Io is given in Fig.

28 while those for RK, Rh, and Rv in Figs. 29, 30 & 31.

1.00

db/d=1
2
Io

0.10
3

0.01
0 10 20 L/d 30 40 50

Fig. 28 Settlement Factor Io for L/d=100 and νs=0.5 for Incompressible Pile (Poulos and
Davis, 1980)

26
3

Values of

100
50
RK 2
25

10

2
1
1
10 100 1000 10000
K

Fig. 29 Correction factor for compressibility, RK for νs=0.5 (Poulos and Davis, 1980)

1
Values of

0.8 50

25
0.6 10
Rh 5
0.4

2
0.2
1

0
1 2 0
0.5
Fig. 30 Correction factor for Finite Layer, Rh (Poulos & Davis, 1980).

27
1

0.95

0.9
K=100
R
500
0.85
2000
1000
0.8

0.75
0.0 0.1 0.2 0.3 0.4 0.5
s

Fig. 31 Correction factor for Poisson’s ratio, R (Poulos & Davis, 1980).

Displacements, u, of GPA under working loads are estimated in a similar manner as

that for GP, i.e., treating it as a compressible pile subjected to pullout.

P .I
 u  (14)
E s .d

where P is the pullout load and I the influence coefficient for upward displacement. The

top ρu0 and tip displacements, ρuL are represented respectively by the displacement

influence coefficients IUO and IUL.

28
2

50

25

1.5

10
IUO

L/d=5
1

0.5
10 100 1000
K
Fig. 32 Normalized tip displacement, IUO vs. K, for νs=0.5 – Effect of K.

13
IUL

6.5
50

25

10

L/d=5
0
10 100 1000
K
Fig. 33 Normalized top displacement, IUL, vs. K, for νs=0.5 – Effect of L/d.

29
The variations of displacement influence coefficients of GPA at the tip , IUO and at the top,

IUL, with relative pile stiffness, K for different L/d are given in Figs. 32 and 33.

CONCLUSIONS

Solutions and results for the ultimate capacities of GP in compression and GPA in

pullout are presented for homogenous (undrained shear strength constant with depth) and

non-homogenous (undrained shear strength increasing linearly with depth) ground

conditions. The ultimate capacities are reported as the lesser of the pile and bulging

capacities. The ultimate capacities of GP and GPA are functions of granular pile and in-

situ ground properties, viz., the unit weight, gp and angle of shearing resistance, gp of

granular pile material; and unit weight, s, undrained strength, cu, the rigidity modulus, G

and the non-homogeneity strength parameter, c of the soil.

Variations of ultimate capacities in GP and GPA with L/d are presented as functions of

G/cu, gp, unit weight parameter, and lateral confinement pressure parameter, for

homogenous and non-homogeneous ground conditions. The transition from pile to bulging

capacity with the L/d is termed as the critical length, (L/d)cr. Variations of (L/d)cr as

functions of relevant parameters including the non-homogeneity strength parameter, c are

presented.

Displacements of GP and GPA are presented considering them as compressible pile, in

situ soil to behave linearly and the in situ ground to be homogeneous. The elastic

continuum approach of Poulos & Davis (1980) is extended to predict displacement

responses of GP and GPA. The variations of normalized displacement influence

coefficients with L/d and K are presented.

30
REFERENCES

Aboshi H., Ichimoto E., Enoki M. and Harda K. (1979).The Composer-A Method to
Improve Characteristics of Soft Clays by Inclusion of Large Diameter Sand Columns.
Proceedings of International Conference on Soil Reinforcement: Reinforced Earth and
Other Techniques, Paris, Vol.1: pp. 211-216.

Aboshi, H. and Suemastu N. (1985). The Sand Compaction Pile Method: State-of-The-Art-
Paper. Proceedings of 3 rd International Geotechnical Seminar on Soil Improvement
Methods, Nanyang Technological Institution, Singapore.

Barksdale R. D. and Bachus R. C. (1983).Design and Construction of Stone Columns.


Report No. FHWA/RD-83/026, U. S. Department of Transportation, Federal Highway
Administration, Washington, D. C., pp. 194.

Datye K.R. and Nagaraju S.S. (1975).Installation and Testing of Rammed Stone Columns.
Proceedings of Indian Geotechnical Society Specialty Sessions, 5th ARC on SMFE,
Bangalore, pp. 101-104.

Gibson R. E. and Anderson W. F. (1961).In Situ Measurement of Soil Properties with the
Pressuremeter. Civil Engineering Publication Works Revision 56, pp. 615-618.

Greenwood D.A. (1970).Mechanical Improvement of Soft Soils below Ground Surface.


Proceedings of the Ground Engineering Conference, Institute of Civil Engineers,
London, June 11-12, pp. 9-20.

Hughes J.M.O. and Withers N.J. (1974). Reinforcing of Soft Cohesive Soils with Stone
Columns. Ground Engineering Journal, Vol. 7, No. 3, pp. 42-49.

Hughes J.M.O., Withers N.J. and Greenwood D.A. (1975). A Field Trial of the Reinforced
Effect of Stone Column in Soil. Geotechnique, Vol. 25, No.1, pp. 31-44.

Madhav, M.R. and Vitkar, P.P. (1978). Strip Footing on Weak Clay Stabilized with
Granular Trench or Piles. Canadian Geotechnical Journal, Vol.15, No.2, pp. 605-609.

Madhav M. R., Iyengar N. G. R., Vitkar P. P. and Nandi S. A. (1979). Increased Bearing
Capacity and Reduced Settlements due to Inclusions in Soil. Symposium on
Reinforcements of Soils, Paris, Vol. 2: pp. 329-333.

Vesic A.S. (1975). Bearing Capacity of Shallow Foundations. Foundation Engineering


Handbook, Van Nostrand Reinhold, pp. 121-144.

31
Guidelines for Soil Nailing Technique in Highway
Engineering Applications

Submitted to

Technical Committee
on Ground Improvement and geosynthetics
Indian Geotechnical Society, New Delhi

Dr. GL Sivakumar Babu


Professor

Department of Civil Engineering


Indian Institute of Science, Bangalore – 560 012
Karnataka, India

Date: November , 2015


PREFACE

This document provides interim guidelines for the analysis, design and construction of
soil nail walls in highway engineering applications. Recommended guidelines are the
outcome of the research carried out at Civil Engineering Department, Indian Institute of
Science, Bangalore towards research project “Guidelines for Soil Nailing Technique in
Highway Engineering, Research Scheme (R-86)” financially supported by Ministry of
Shipping, Road Transport and Highways, Government of India New Delhi. Document is
divided into 6 main sections addressing different issues related to soil nailing technique in
the following manner:

Section 1: Provides a customary introduction to the soil nailing technique.

Section 2: Provides information and guidelines about methods of construction and


selection of materials for soil nail walls.

Section 3: Discusses allowable stress design method based analysis of failure modes of
soil nail walls and provide recommendations and methodology to evaluate
factors of safety against various failure modes.

Section 4: Provides general steps and other considerations for the design soil nail walls.

Section 5: Illustrates step-by-step the design methodology for soil nail walls (discussed in
sections 3 and 4) with the help of a design example.

Section 6: Provides procedures and requirements for conducting field pullout testing of
soil nails.

Additionally, frequently required information regarding design of soil nail walls is


presented in tabular form in Appendix A. A format for preparing database of typical soil
nail wall project is also presented in Appendix B.

Recommendations of the Federal Highway Administration (FHWA 2003) report


“Geotechnical Engineering Circular No. 7: Soil Nail Walls”, being an exhaustive and
most prevalently used manual for soil nail walls in practice; has been appraised
analytically and using finite element based rigorous computational tool and are suitably
considered as the key reference in the preparation of these interim guidelines. It is
believed that the document fulfills the requirements in professional works.

GL Sivakumar Babu

ii
CONTENTS

Page No.

Acknowledgments ii
Preface iii
Contents iv
List of Tables vii
List of Figures viii
List of Symbols ix

SECTION 1: INTRODUCTION

1.1 DEFINITION 01
1.2 ORIGIN 01
1.3 DESIGN METHODS 02
1.4 SOIL NAILING APPLICATIONS IN HIGHWAY ENGINEERING 02
1.5 ADVANTAGES AND LIMITATIONS 04
1.5.1 Advantages 04
1.5.2 Limitations 04

SECTION 2: CONSTRUCTION MATERIALS AND METHODS

2.1 INTRODUCTION 05
2.2 SOIL NAIL INSTALLATION METHODS 05
2.3 COMPONENTS OF A TYPICAL SOIL NAIL WALL 05
2.4 CONSTRUCTION SEQUENCE 07
2.5 GEOTECHNICAL ASPECTS 09
2.5.1 Soil investigation 09
2.5.2 Bond strength 10
2.5.3 Suitable in-situ ground conditions 11

SECTION 3: ANALYSIS OF FAILURE MODES

3.1 INTRODUCTION 12
3.2 FAILURE MODES OF SOIL NAIL WALLS 12
3.3 EXTERNAL FAILURE MODES 13
3.3.1 Global stability 13
3.3.2 Sliding Stability 16

iii
3.3.3 Bearing capacity (or base heave) failure 18
3.4 INTERNAL FAILURE MODES 19
3.4.1 Soil nail pullout failure mode 20
3.4.2 Nail tensile strength failure mode 21
3.5 FACING DESIGN AND FAILURE MODES 21
3.5.1 Facings design procedure 22

SECTION 4: GENERAL DESIGN CONSIDERATIONS

4.1 INTRODUCTION 29
4.2 GENERAL DESIGN STEPS 29
4.2.1 Initial soil nail wall considerations 29
4.2.2 Preliminary design 31
4.2.3 Final design 32
4.3 OTHER DESIGN CONSIDERATIONS 33
4.3.1 Loads and load combinations 33
4.3.2 Permissible soil nail wall deformations 33
4.3.3 Drainage measures 33
4.3.4 Corrosion protection 34

SECTION 5: DESIGN EXAMPLE

5.1 INTRODUCTION 35
5.2 INITIAL DESIGN CONSIDERATIONS 35
5.3 PRELIMINARY DESIGN 36
5.4 FINAL DESIGN 38
5.4.1 EXTERNAL FAILURE MODES 38
5.4.1.1 Global stability 38
5.4.1.2 Sliding stability 40
5.4.1.3 Bearing capacity (or basal heave) failure 42
5.4.2 INTERNAL FAILURE MODES 42
5.4.2.1 Soil nail pullout failure 42
5.4.2.2 Soil nail tensile strength failure 42
5.4.3 FACING DESIGN AND CHECKS 44
5.4.4 GENERAL CONSIDERATIONS AND DESIGN SUMMARY 49
5.4.4.1 General considerations 49
5.4.4.2 Design summary 49

iv
SECTION 6: FIELD PULLOUT TESTING OF SOIL NAILS

6.1 INTRODUCTION 52
6.2 FIELD PULLOUT TEST APPARATUS 52
6.3 TYPES OF FIELD PULLOUT TESTS 53
6.3.1 Verification test 54
6.3.2 Proof test 54
6.3.3 Creep test 54
6.4 TEST PROCEDURE 55
6.4.1 Procedure for verification test 55
6.4.2 Procedure for proof test 56
6.5 TEST ACCEPTANCE CRITERIA 57
6.5.1 Acceptance criteria for verification test 57
6.5.2 Acceptance criteria for proof test 58
6.5.3 Acceptance criteria for creep test 58
6.6 TYPICAL TEST DATA SHEET 58

APPENDIX A: GENERAL DESIGN TABLES 60


APPENDIX B: TYPICAL SOIL NAIL WALL PROJECT SUMMARY 65

v
LIST OF TABLES

Table caption Pg. No.

Table 5.1 Desired minimum recommended factors of safety for permanent


soil nail walls. 36
Table 5.2 Maximum axial tensile force developed in each nail from top. 37
Table 5.3 Allowable axial force carrying capacity of nails at different levels. 39
Table 5.4a Factor of safety against soil nail pullout failure (static case). 43
Table 5.4b Factor of safety against soil nail pullout failure (seismic case). 43
Table 5.5 Factor of safety against soil nail tensile strength failure. 44
Table 5.6 Summary of important design parameters. 50
Table 5.7 Summary of factors of safety attained for various failure modes. 50
Table 5.8 Summary of facing design (temporary and permanent). 50
Table 6.1 Loading schedule for verification test. 56
Table 6.2 Loading schedule for proof test. 57

vi
LIST OF FIGURES

Figure caption Pg. No.

Fig. 1.1 Example applications of soil nail walls in highway engineering. 03


Fig. 2.1 Components of a typical soil nail wall (FHWA 2003). 06
Fig. 2.2 Typical construction sequence of soil nail walls (FHWA 2003). 08
Fig. 2.3 Soil investigation boring layout for soil nail walls (FHWA 2003). 10
Fig. 3.1 Principal failure modes of soil nail walls. 12
Fig. 3.2 Global stability of soil nail walls: (a) simplified single wedge failure
13
mechanism; (b) forces acting on the failure plane.
Fig. 3.3 Failure wedge triangle. 15
Fig. 3.4 Sliding stability of soil nail walls. 17
Fig. 3.5 Excavation base heave failure for soil nail walls (FHWA 2003). 19
Fig. 3.6 Pullout failure of soil nails. 20
Fig. 3.7 Punching shear failure mode: Bearing plate connection used in
26
temporary facing (FHWA 2003).
Fig. 3.8 Punching shear failure mode: Headed stud connection used in
26
permanent facing (FHWA 2003).
Fig. 5.1 General outline of designed soil nail wall. 38
Fig. 5.2 Details of facing reinforcement. 51
Fig. 6.1 Field pullout test set up (FHWA 2003). 53
Fig. 6.2 Typical format of soil nail field pullout test data sheet. 59

vii
LIST OF SYMBOLS

∆L min Minimum acceptable movement


(LP)z Pullout (effective bond) length of nail at depth z
(T)z Axial tensile force in nail at depth z
AH Cross-sectional area of the stud head
ASH Cross-sectional area of the headed-stud shaft
Ast Cross-sectional area of tensile reinforcement (i.e. nail)
At Nail bar cross-sectional area
Reinforcement cross sectional area per unit width in the vertical and
avm, ahm
horizontal directions at the at mid-span
Reinforcement cross sectional areas per unit width in the vertical and
avn, ahn
horizontal directions at the nail head
Total cross sectional area of waler bars in the vertical and horizontal
Avw, Ahw
directions
B’ Width of influence in an excavation
Be Width of excavation
BL Base width of sliding failure surface
c In-situ soil cohesion
CF Factor that considers the non-uniform soil pressures behind the facing
D Maximum of DN and DDH
Effective diameter of assumed conical failure surface at the center of the
D’c
section
DB Depth of strong deposit from bottom of excavation
DDH Grout (drill) hole diameter
DH Diameter of the stud head
DS Diameter of the headed-stud shaft
DTL Design test load
E Young’s modulus of steel
fck Characteristic yield strength of concrete
Fh Horizontal seismic inertia force
FSFF Factor of safety against facing flexure failure
FSFP Factor of safety against facing punching failure
FSFT Factor of safety against facing tensile failure
FSG Factor of safety for global stability
FSH Factor of safety against excavation base heave (bearing capacity) failure
FSHT Factor of safety against headed-stud tensile failure
FSP Factor of safety against nail pullout failure
FSSL Factor of safety for sliding stability
FST Factor of safety against nail tensile strength failure
Fv Vertical seismic inertia force
fy Characteristic yield strength of steel

viii
h Temporary/permanent facing thickness
H Vertical wall height
H1 Effective height of soil mass that considers sloping ground
hc Effective depth of conical surface
Heq Equivalent wall height
i Nail inclination with respect to horizontal
K Active earth pressure coefficient
KA Static active earth pressure coefficient
KAE Seismic active earth pressure coefficient
kh Horizontal seismic coefficient
kv Vertical seismic coefficient
L Length of soil nail
L1 Length of top nail required to cross failure plane
L2 Length of top nail required to satisfy pullout criterion
LB Soil nail bonded length (for load testing)
LBP Length of bearing plate
Le Length of excavation
LF Length of inclined failure surface
LP Pullout (effective bond) length
LS Headed-stud length
Nc Bearing capacity factor
NF Normal force acting on the failure surface
NH Number of headed-studs in the connection
P Maximum applied test load
P Active lateral earth force
PA Active lateral earth force for static case
PAE Active lateral earth force for seismic case
Qall Allowable pullout resistance per unit length
qs Uniformly distributed surface loading
QT Total surcharge load
qu Ultimate bond strength
Qu Ultimate pullout capacity per unit length
RFP Facing punching shear capacity
RHT Tensile failure of the headed-studs
RP Nail pullout resistance (capacity)
RT Nail bar tensile capacity
SF Shear force on failure surface
Sh Horizontal nail spacing
SHS Headed-stud spacing
Su Undrained shear strength of fine grained soils
Sv Vertical nail spacing

ix
Sv1 Vertical spacing of first nail (topmost nail)
T1 Axial tensile force in first nail (topmost nail)
Tall Allowable axial tensile nail force
Teq Equivalent axial tensile nail force
tH Headed-stud head thickness
Tmax Maximum axial tensile nail force
To Axial tensile nail force at nail head
tp Thickness of bearing plate
UL unbonded length of nail (test)
W Weight of rigid failure block ABC
z Arbitrary depth measured from ground surface
α Wall face batter angle (from vertical)
β Backslope angle
βeq Equivalent backslope angle
γ In-situ soil unit weight
ΔH Slope rise up to bench (if present)
ΔH Equivalent overburden
ρ Provided reinforcement ratio
ρmax Maximum limit of reinforcement ratio
ρmin Minimum limit of reinforcement ratio
ΣD Sum of driving forces along failure plane
ΣR Sum of resisting forces along failure plane
φ In-situ soil angle of internal friction
ψ Inclination of failure plane
ω An angle relating the horizontal and vertical seismic coefficients

x
SECTION 1: INTRODUCTION

1.1 DEFINITION

Soil nailing is an innovative earth retaining technique. It is innovative in the sense that,

unlike other earth retaining techniques, such as conventional retaining walls and

geosynthetic reinforced slopes, its construction proceeds from the top to bottom with

reinforced soil being in-situ. In principle, soil nailing consists of the passive

reinforcement of existing ground by installing closely spaced steel bars (i.e. nails), which

may be subsequently encased in grout. A shotcrete facing is then applied at the wall face

to provide continuity. This ground support technique relies on the mobilization of the

tensile strength of the steel reinforcement at relatively small deformations in the

surrounding ground. In a soil nail wall, the properties and material behaviour of three

components—the native soil, the reinforcement (nails) and the facing element—and their

mutual interactions significantly affect the performance of the structure. Additionally,

various other factors such as the construction sequence, the installation method of nails,

the connection between the nails and the facing are also likely to influence the behaviour

of soil nail walls.

1.2 ORIGIN

The origin of soil nailing can be traced since 1960s when it was used as a support system

for underground excavations in rock referred to as the New Austrian Tunneling Method.

Since then, soil nailing technique has been widely utilised for various slope stability

applications in developed countries like USA, France, UK, etc. In India, over the last

decade, soil nailing technique has been considerably used for the applications such as,

stabilization of road/rail side slopes, basement excavations, additional support for bridge

abutments and side walls of the approaches for subways.

1
1.3 DESIGN METHODS

Since its inception, various analysis approaches such as: limit equilibrium analysis, multi-

criteria analysis, kinematical limit analysis, strain compatibility analysis, discrete element

analysis, nonlinear programming and finite element analysis have been developed to

study behaviour of soil nail walls. However, for practical purposes, the available analysis

and design methods for soil nail walls can be broadly classified into two main categories:

(a) Limit equilibrium design methods, which are used to evaluate the global safety factor

of the nailed structures with respect to a rotational or translational failure along

potential sliding surfaces, taking into account the shearing, tension, or pull-out

resistance of the inclusions crossing the potential failure surface.

(b) Working stress design methods which are used to estimate the tension and shear forces

generated in the nails during construction under the design loading conditions and

evaluate the local stability at each level of nails.

Limit equilibrium based methods have attracted the attention of the researchers

because of their simplicity, reasonable accuracy and popularity among the practicing

engineers. The main shortcoming of limit equilibrium based methods is that they fail to

address deformation behaviour of soil nail walls adequately. Deformation estimates are

generally obtained empirical relations based on the past experiences. Rigorous

computational tools based on numerical techniques such as: finite element and finite

difference methods are often employed to study overall stability and deformation

estimates of soil nail walls. In this document, allowable stress design method (ASD) for

the analysis and design of soil nail structures is used and recommended.

1.4 SOIL NAILING APPLICATIONS IN HIGHWAY ENGINEERING

Soil nail walls are particularly well suited to excavation applications for ground

conditions that require vertical or near-vertical cuts. Soil nail walls have been used

2
successfully in highway cuts; end slope removal under existing bridge abutments during

underpass widening; for the repair, stabilization, and reconstruction of existing retaining

structures; and tunnel portals. Some of the example applications of soil nail walls in

highway engineering are shown in Fig. 1.1.

(a) Roadside slope (b) Approaches to subway

(c) Bridge abutment (d) Highway widening

(e) Widening under existing bridge (f) Temporary shoring

Fig. 1.1 Example applications of soil nail walls in highway engineering.

3
1.5 ADVANTAGES AND LIMITATIONS

1.5.1 Advantages

Soil nail walls exhibit numerous advantages when compared to ground anchors and

alternative top down construction techniques. Soil nailing is less disruptive to traffic and

causes less environmental impact, require smaller right-of-way, relatively flexible and can

accommodate relatively large total and differential settlements. Soil nail walls have

performed well during seismic events owing to overall system flexibility; and are more

economical than conventional concrete gravity walls when conventional soil nailing

construction procedures are used.

1.5.2 Limitations

Soil nail walls may not be appropriate for applications where very strict deformation

control is required for structures and utilities located behind the proposed wall, as the

system requires some soil deformation to mobilize resistance. Also, occurrence of utilities

may place restrictions on the location, inclination, and length of soil nails in the upper

rows. Soil nail walls are not well-suited where large amounts of groundwater seep into

the excavation because of the requirement to maintain a temporary unsupported

excavation face. Moreover, construction of soil nail walls requires specialized and

experienced contractors.

4
SECTION 2: CONSTRUCTION MATERIALS AND METHODS

2.1 INTRODUCTION

Various aspects such as: basic elements of the typical soil nail wall; construction

methodology; construction materials; extent of geotechnical investigation and favorable

conditions have been presented in the subsequent sections.

2.2 SOIL NAIL INSTALLATION METHODS

Two most prevalently used soil nail installation techniques in practice are: (i) drilled and

grouted soil nails and (ii) driven soil nails.

Drilled and grouted soil nails, also known as grouted nails, are approximately 100

mm to 200 mm diameter nail holes drilled in the soil mass to be retained, typically, at a

spacing of about 1.5 m apart. Steel bars are then placed and the holes are grouted.

Grouted nails can be used for both temporary and permanent applications. On the other

hand, driven soil nails are relatively small in diameter (20 mm to 25 mm) and are

mechanically driven into the ground. They are usually spaced approximately 0.5 m to 1.0

m apart. The use of driven soil nails allows for a faster installation (as compared to drilled

and grouted soil nails); however, this method of installation cannot provide good

corrosion protection other than by sacrificial bar thickness.

Grouted nails are recommended for all types of soil nail walls applications; and in

particular, for walls with vertical height more than 5 m. Driven nails shall only be used

for temporary excavation support purposes, and when wall heights are smaller (less than

or upto 5.0 m).

2.3 COMPONENTS OF A TYPICAL SOIL NAIL WALL

Various components of a typical soil nail wall are shown in Fig. 2.1.

5
(a) Nail bars: Threaded solid or hollow steel reinforcing bars are commonly used for soil

nails. Bars generally have a nominal tensile strength of 415 MPa or higher. Minimum

recommended diameter of reinforcement bar (tendon) is 20 mm.

Fig. 2.1 Components of a typical soil nail wall (FHWA 2003).

(b) Nail head: The nail head comprises of following main components: the bearing-plate,

hex nut, and washers; and the headed-stud (see Fig. 2.1). The bearing plate is made of

Grade 250 MPa steel and is typically square, 200 to 250 mm side dimension and 19

mm thick. The purpose of the bearing plate is to distribute the force at the nail end to

the temporary shotcrete facing and the ground behind the facing. The bearing plate

has a central hole, which is inserted over the nail bar. Beveled washers are then placed

and the nail bar is secured with a hex nut or with a spherical seat nut. Washers and

nuts are steel with a grade consistent with that of the nail bar commonly of 415 MPa

or higher. Nuts are tightened with a hand-wrench. The headed-stud connection may

6
consist of four headed studs that are welded near the four corners of the bearing plate

to provide anchorage of the nail head into the permanent facing. For temporary walls,

the bearing plate is on the outside face of the shotcrete facing.

(c) Grout: Grout for soil nails is required to fill the annular space between the nail bar

and the surrounding ground, and for shotcreting of the temporary facing. Grout for

soil nail walls is commonly a neat cement grout with the water/cement ratio typically

ranging from 0.4 to 0.5. Grout mix shall be prepared in accordance with IS: 9012 –

1978 “Recommended practice for shotcreting”. Grout shall have a minimum 28 days

characteristic strength of 20 MPa. For filling up nail holes, grout shall be pumped

shortly after the nail bar is placed in the drillhole to reduce the potential for hole

squeezing or caving. In solid nail bar applications, the grout may be injected by tremie

methods through a grout pipe, which is previously inserted to the bottom of the

drillhole, until the grout completely fills the drillhole.

(d) Centralizers: Centralizers are devices made of polyvinyl chloride (PVC) or other

synthetic materials that are installed at various locations along the length of each nail

bar to ensure that a minimum thickness of grout completely covers the nail bar. They

are installed at regular intervals, typically not exceeding 2.5 m, along the length of the

nail and at a distance of about 0.5 m from each end of the nail.

(d) Other components: Other main component of soil nail wall includes corrosion

protection elements; wall facings and drainage system (see Sections 3 and 4 for

details).

2.4 CONSTRUCTION SEQUENCE

Typical sequence of construction of a soil nail wall include following steps as shown in

Fig. 2.2:

Step 1: Excavation of initial cut;

7
Step 2: Drilling hole for nail;

Step 3: Installation of nails followed by grouting and placing of drainage strip;

Step 4: Placing of construction facing and installation of bearing plates;

Step 5: Repetition of process till final level is reached; and

Step 6: Placing of final facing.

Fig. 2.2 Typical construction sequence of soil nail walls (FHWA 2003).

8
2.5 GEOTECHNICAL ASPECTS

2.5.1 Soil investigation

Exploratory borings shall be performed for the design of soil nail walls to obtain: (1) SPT

N-values to classify soil and delineate the stratigraphy, (2) both disturbed and undisturbed

soil samples for laboratory soil testing, and (3) observations of groundwater. Laboratory

testing of soil samples collected during borings enable judicious selection of soil

parameters for the design of soil nail walls. Fig. 2.3 can be used as a preliminary guide to

help designers plan the number, location, and frequency of borings for soil nail wall

applications.

For soil nail walls more than 30 m long, borings should be spaced between 30 to

60 m along the proposed centerline of the wall. For walls less than 30 m long, at least one

boring is necessary along the proposed centerline of the wall. Borings are also necessary

in front and behind the proposed wall. Borings behind the wall should be located within a

distance up to 1 to 1.5 times the height of the wall behind the wall and should be spaced

up to 45 m along the wall alignment. If the ground behind the proposed wall is sloping,

the potentially sliding mass behind the wall is expected to be larger than for horizontal

ground. Therefore, borings behind the proposed wall should be located farther behind the

wall, up to approximately 1.5 to 2 times the wall height. Borings in front of the wall

should be located within a distance up to 0.75 times the wall height in front of the wall

and should be spaced up to 60 m along the wall alignment.

The depth of borings should extend at least one full wall height below the bottom

of the excavation (see Fig. 2.3). Borings should be deeper when highly compressible soils

(i.e., soft to very soft fine-grained soils, organic silt, and peat) occur at the site behind or

under the proposed soil nail wall. The required boring depths for soil nail wall projects

may be greater if deep loose, saturated, cohesionless soils occur behind and under the

9
proposed soil nail wall and the seismic risk at the site require that the liquefaction

potential be evaluated. The subsurface investigation depths may need to be deep at

proposed sites of soil nail walls where seismic amplification is of concern, particularly in

deep, soft soils. If rock is encountered within the selected depth, a core at least 3 m long

retrieved in two 1.5 m long runs should be obtained to inspect the nature of the rock and

its discontinuities.

(a) Typical plan (distances shown are recommended maximum)

(b) Section A-A

Fig. 2.3 Soil investigation boring layout for soil nail walls (FHWA 2003).

2.5.2 Bond strength

Bond strength of the nail-soil interface is an extremely important parameter required for

the design, analysis and performance assessment of soil nail walls. The pullout capacity

of a soil nail installed in a grouted nail hole is affected by the size of the nail (i.e.,

perimeter and length) and the ultimate bond strength qu. The bond strength is the

10
mobilized shear resistance along the soil-grout interface. For drilled and grouted nails, the

bond strength is a function of ground conditions around the nail (i.e. soil type and

conditions); soil nail installation type including drilling method, grouting procedure, grout

nature, grout injection (e.g. gravity or under pressure); and the size of the grouted zone.

The bond strength adopted for the design of soil nails is commonly based on

conservative estimates obtained from field correlation studies and local experience in

similar conditions. Consequently, some percentage of the soil nails shall be load tested

(pullout tests) in the field to verify bond strength design. Procedure for conducting field

pullout tests on soil nails is described in Section 6. As a preliminary estimate bond

strength for various soil conditions and construction methods based on the literature may

be adopted from Table A.2 of Appendix A.

2.5.3 Suitable in-situ ground conditions

Following are the in-situ conditions considered favorable for the prospective use of soil

nailing technique.

(a) Soil shall be able to stand unsupported to a depth of about 1 m – 2 m high vertical or

nearly vertical cut for 24-48 hours.

(b) Groundwater table shall be sufficiently below level of the lowermost soil nail at all

cross-sections.

(c) Favorable soils: Stiff to hard fine –grained soils, dense to very dense granular soils

with some apparent cohesion, weathered rock with no weakness planes and glacial

soils.

11
SECTION 3: ANALYSIS OF FAILURE MODES

3.1 INTRODUCTION

Allowable stress design (ASD) methodology is recommended for the analysis of potential

failure modes of soil nail walls. Governing equations for the analysis and design of soil

nail walls in consideration of each potential failure mode have been described for the

most general case, and hence, designer may suitably simplify the governing equations to

suit project specific conditions. Design example presented in Section 5 may be referred

for better understanding.

3.2 FAILURE MODES OF SOIL NAIL WALLS

Failure modes of soil nail walls can be broadly classified into three distinct groups as

external, internal and facing failure modes (see Fig. 3.1). Recommended minimum

factors of safety for the design of soil nail walls using ASD method are presented in

Table A.1.

Fig. 3.1 Principal failure modes of soil nail walls.

12
3.3 EXTERNAL FAILURE MODES

3.3.1 Global stability

Global stability of the soil nail walls refers to the overall stability of the reinforced soil

nail wall mass. In this failure mode, along the slip surface (i.e. potential failure plane), the

driving force due the self weight and external loading on the retained mass exceeds the

resisting force provided by the in-situ soil and the nails.

Simplified global stability analysis

Global stability analysis for soil nail walls can be carried out using a simple, single wedge

failure mechanism (see Fig. 3.2) with failure plane assumed to be inclined at an angle

ψ= 45 + ( φ / 2 ) (in degrees) with respect to the horizontal. The factor of safety for global

stability FSG is expressed as the ratio of the resisting forces ΣR and driving forces ΣD,

which acts tangentially to the potential failure plane:

FSG =
∑R (3.1)
∑D

Fig. 3.2 Global stability of soil nail walls: (a) simplified single wedge failure mechanism;
(b) forces acting on the failure plane.

13
Considering the equilibrium of forces (see Fig. 3.2b) acting along the failure plane and

rearranging the terms, factor of safety for global stability FSG can be determined as

cL F + Teq cos ( ψ − i ) + ( W + QT − Fv ) cos ψ + Teq sin ( ψ − i ) − Fh sin ψ  tan φ


FSG = (3.2)
( W + QT − Fv ) sin ψ + Fh cos ψ
where: c [kPa] = in-situ soil cohesion;

φ [degrees] = in-situ soil angle of internal friction;

i [degrees] = nail declination with respect to horizontal;

α [degrees] = wall face batter with respect to vertical;

β [degrees] = backslope angle with respect to horizontal;

H [m] = vertical height of the soil nail wall;

H cos(α + β)
LF [ m] = = length of the potential failure plane;
cos α sin(ψ − β)

q s H cos ( ψ + α )
QT [ kN / m ] = = total surcharge load; and
cos α sin ( ψ − β )

qs [kPa] = distributed surcharge loading.

Determination of weight of failure wedge, W

Using sine rule for ΔABC (see Fig. 3.2a and Fig. 3.3), we get

a b c
= = (3.3)
cos(ψ + α) cos(α + β) sin(ψ − β)

Using Hero’s formula

Area of ∆ABC = s(s − a)(s − b)(s − c) where s = (a + b + c) / 2 (3.4)

Therefore, weight of failure wedge, W [kN/ m] = γ x ΔABC (3.5)

where: γ [kN/m3] = in-situ soil unit weight.

14
Fig. 3.3 Failure wedge triangle.

Determination of equivalent nail force Teq

Allowable axial force carrying capacity Tall of any particular soil nail embedded at depth

z below the ground surface can be determined as the minimum of its pullout capacity RP

and tensile capacity RT given by Eqs. (3.6) and (3.7) respectively.

( R P )z [ kN ] = ( πD LP q u ) /1000 (3.6)

( R T )z [ =
kN ] ( 0.25πd f ) /1000
2
y (3.7)

where: D [mm] = effective diameter of nail (equal to d [mm] i.e. diameter of

reinforcement bar (nail) for driven nails and equal to DDH [mm] i.e. drillhole diameter in

case of grouted nails); qu [kPa] = ultimate bond strength (see Table A.2); fy [MPa] =

characteristic yield strength of reinforcement bar; LP [m] = effective bond length (i.e. nail

length in the passive zone), given by Eq. (3.8).

 (H − z) cos(ψ + α) 
(L P ) z [ m ]= L −   (3.8)
 cos α sin(ψ + i) 

where: L [m] = length of the soil nail.

If, Sv [m] is the vertical spacing of soil nails, total number of nails n provided in a given

cross-section of the soil nail wall can be determined. Knowing, Sh [m] i.e. horizontal (out-

of-plane) spacing of soil nails, the equivalent nail force Teq per meter length of the soil

nail can be obtained using Eq. (3.9).

15
n
1
Teq [kN/ m] =
Sh
∑ (T )
j=1
all j (3.9)

Determination of seismic inertia forces Fh and Fv

Pseudo-static approach can be adopted for seismic stability analysis of soil nail walls. In

this method, the earthquake-induced, time-varying forces of inertia acting within a

potentially sliding rigid block involving the soil nail wall system are replaced by

equivalent pseudo-static force acting at the center of gravity of the analyzed block. The

horizontal and vertical components of this equivalent pseudo-static force are expressed

as:

Horizontal seismic inertia force, Fh [kN/=


m] k h (W + QT ) (3.10a)

Vertical seismic inertia force, Fv [kN/=


m] k v (W + QT ) (3.10b)

It is to be noted that the horizontal inertia force Fh should be directed away from the slope

and the vertical inertia force Fv should be directed upwards (see Fig. 3.2). The appropriate

values of seismic coefficients kh and kv can be evaluated using IS: 1893 Part I (2002)

“Code of practice for earthquake resistant design of structures”. For practical applications

the vertical inertia force is generally disregarded, however, for particular cases, in

accordance with IS: 1893 Part I (2002), vertical seismic coefficient kv can be taken equal

to two-thirds of the horizontal seismic coefficient kh for the evaluation of the vertical

inertia force Fv.

3.3.2 Sliding stability

Sliding stability considers the ability of the soil nail wall to resist sliding along the base of

the retained system in response to lateral earth pressures behind the soil nails. Sliding

failure may occur when additional lateral earth pressures, mobilized by the excavation,

exceed the sliding resistance along the base. To evaluate the factor of safety for sliding

stability, soil nail wall is considered as a rigid block of width BL against which lateral

16
earth force from retained soil are applied (see Fig. 3.4). It is assumed that the

displacements of the soil block along its base are large enough to mobilize the active

pressure acting on it.

Fig. 3.4 Sliding stability of soil nail walls.

The factor of safety against sliding FSSL is calculated as the ratio of horizontal resisting

forces ΣR to the applied driving horizontal forces ΣD, i.e.

∑ R c b BL + ( W + QT − Fv + P sin βeq ) tan φb


=
FS = (3.11)
∑D Fh + P cos βeq
SL

where:

c b and φb =soil strength parameters along the base of rigid sliding block (AD);

βeq = equivalent backslope angle [for broken slopes βeq = tan-1(ΔH/2H1), for infinite

slopes βeq = β];

BL [ m ] =
L + H tan α = base width of the rigid sliding block (AD);

H1 [ m ]= H + ∆H= H + L tan β = effective height over which earth pressure acts (CD);

QT [kN/ m] = q s L = total surcharge load;

17
1 2 
2
L L
W[kN/ m] = WABF + WBCE + WBEDF= γH  tan α + 2   +   tan β  = total weight
2  H H 

of the rigid sliding block (ABCD);

Fh [kN/=
m] k h (W + QT ) = horizontal seismic inertia force; and

Fv [kN/=
m] k V (W + QT ) = vertical seismic inertia force.

Determination of total active thrust P

The total lateral active thrust P acting behind the wall-nailed soil block is expressed as:

γH12  2q  cos α  
P= K (1 − k v ) 1 + S   (3.12)
2  γH1  cos ( β − α )  

where: K = coefficient of lateral active earth pressure which can be determined using Eq.

(3.13).

cos 2 ( φ − α − ω)
K= 2
(3.13)
 sin(φ + β) sin(φ − β − ω) 
cos ω cos 2 α cos ( α + β + ω) 1 + 
 cos ( α + β + ω) cos(β − α) 

where: ω [degrees] = an angle relating the horizontal and vertical seismic coefficients

such that φ − β ≥ ω , given by Eq. (3.14).

 k 
ω = tan −1  h  (3.14)
 1− kv 

3.3.3 Bearing capacity (or base heave) failure

Bearing capacity failure, though not a prominent failure mode, may be of concern when a

soil nail wall is excavated in fine-grained or soft soils. Because the wall facing do not

extend below the bottom of the excavation the unbalanced load caused by the excavation

may cause the bottom the excavation to heave and cause a bearing capacity failure of the

foundation (see Fig. 3.5).

18
Fig. 3.5 Excavation base heave failure for soil nail walls (FHWA 2003).

The factor of safety against heave FSH is given by:

Su N c
FSH = (3.15)
 S 
H eq  γ − u 
 B' 

 B 
where: Su [=
kPa ] cN c 1 + e  = undrained shear strength of the soil
 Le 

  φ 
N C = cot φ e π tan φ tan 2  45 +  − 1 = bearing capacity factor
  2 

Heq [m] = equivalent wall height = H+ΔH, with ΔH [m] as an equivalent overburden;

B’ [m] = width of influence, B’ [m] = Be/ 2, where: Be [m] = width of excavation.

When a strong deposit underlying the soft layer and occurring at a depth DB < 0.71Be

below the excavation bottom is encountered (see Fig. 3.5b), B’ in Eq. (3.15) must be

replaced by DB [m].

3.4 INTERNAL FAILURE MODES

Internal stability of soil nails shall be checked at each nail level. Pullout failure and nail

tensile strength failure are the two important internal failure modes of soil nail walls.

19
Following procedure shall be followed to determine the factors of safety of soil nails

against internal failure modes.

3.4.1 Soil nail pullout failure mode

Nail pullout failure is a failure along the soil-grout or soil-nail (in case of driven nails)

interface due to insufficient intrinsic bond strength and / or insufficient nail length. It is

the primary internal failure mode in a soil nail wall. For any particular nail embedded at

depth z (see Fig. 3.6), factor of safety against pullout failure FSP can be determined as the

ratio of its pullout capacity Rp to the maximum axial force T developed in the nail.

R 
(FSP ) z =  P  (3.16)
 T z

where: (R P ) z is as determined by Eq. (3.6); and knowing the value of coefficient of lateral

earth pressure K from Eq. (3.13), maximum axial force T at depth z can be obtained as:

(T) z [ kN
= ] K(qs + γz)ShSv (3.17)

Fig. 3.6 Pullout failure of soil nails.

20
3.4.2 Nail tensile strength failure mode

Tensile failure of a soil nail takes place when the axial force along the soil nail T is

greater than the nail tensile capacity which may lead to the break-up of the tensile

member. Factor of safety against nail tensile strength failure FST for any nail embedded at

depth z can be calculated as the ratio of maximum axial tensile load capacity RT of nail to

the maximum axial force T developed in the nail at depth z.

( FST )z = 
RT 
 (3.18)
 T z

where: (R P ) z and (T) z are as determined by Eq. (3.7) and Eq. (3.17) respectively.

3.5 FACING DESIGN AND FAILURE MODES

Soil nail walls are generally provided with two types of facings: (a) temporary facing and

(b) permanent facing. Temporary facing is usually constructed by providing

reinforcement in the form of welded wire mesh throughout the wall face, and by

additional bearing plates and waler bars at the nail heads; which is, subsequently

shotcreted. On the other hand, permanent facing is usually constructed as cast-in-place

reinforced cement concrete. However, reinforcement in the permanent facing may be

adopted in the form of welded wire mesh or reinforcement bars in either direction. Most

of the times temporary facing resists major portion of the loads transferred from soil nails

at nail head at the wall face, while permanent facing serves the purpose of improving

aesthetic of the wall face. Connection between temporary facing and permanent facing is

usually provided by means of headed-studs (usually four numbers per plate) welded on

the bearing plates. Following section presents the step-by-step procedure for designing

conventional facing system for soil nail walls along with recommended checks for

various facing failure modes.

21
3.5.1 Facings design procedure

Step 1: Determine design nail head tensile force at the wall face To

To [ kN ] = Tmax [ 0.6 + 0.2(Smax − 1) ] (3.19)

where: Tmax [ kN ] = maximum axial force developed in the soil nails; and

Smax [ m ] = maximum of the Sv and Sh

Step 2: Select facing thicknesses

Temporary facing thickness h: [e.g., 100, 150, 200 mm].

Permanent facing thickness h: [e.g., 200 mm].

Step 3: Select appropriate facing materials

(a) Adopt steel reinforcement grade Fe 415 (or Fe 500) with characteristic strength

fy = 415 MPa (or fy = 500 MPa).

(b) Adopt suitable welded wire mesh (WWM) (see Table A.3) and reinforcement

bar (see Table A.4).

(c) Adopt suitable concrete/shotcrete grade between M20 to M30 with

characteristic compressive strength fck between 20-30 MPa.

(d) Adopt suitable headed-stud characteristics (see Table A.5).

(e) Adopt bearing plate geometry: minimum size 200 × 200 mm x19 mm.

Step 4: Check for minimum and maximum reinforcement requirements

(a) Calculate the minimum and the maximum reinforcement ratios as:

f ck [ MPa ]
ρmin [ % ] =
20 (3.20)
f y [ MPa ]

f [ MPa ]  600 
ρmax [ % ] =
50 ck   (3.21)
f y [ MPa ]  600 + f y [ MPa ] 

22
Therefore, the placed reinforcement shall be ρmin ≤ ρ ≤ ρmax . In addition the ratio

of the reinforcement in the nail head and mid-span zones should be less than 2.5 to

ensure comparable ratio of flexural capacities in theses areas.

(b) Select reinforcement area per unit length of WWM for temporary/permanent

facing (see Table A.3) at the nail head an and at mid-span am in both the vertical

and horizontal directions. Usually, the amount of reinforcement at the nail head is

adopted same as the amount of reinforcement at the mid-span (i.e., an = am) in

both vertical and horizontal directions.

However, for temporary facing, if waler bars are used at the nail head in addition

to the WWM, recalculate the total area of reinforcement at the nail head in the

vertical direction and horizontal direction using Eqs. (3.22) and (3.23)

respectively.

A vw
a=
vn a vm + (3.22)
Sh
A
a=
hn a hm + hw (3.23)
Sv

where: avn and ahn are the reinforcement cross sectional areas per unit width in the

vertical and horizontal directions at the nail head respectively; avm and ahm are the

reinforcement cross sectional area per unit width in the vertical and horizontal

directions at the at mid-span respectively; and Avw and Ahw are the total cross

sectional area of waler bars in the vertical and horizontal directions respectively.

Note: If rebar is used in permanent facings instead of WWM, the total area of

reinforcement must be suitably converted to a per unit length basis.

(c) Calculate the reinforcement ratio at the nail head and the mid-span as:

an
ρn [ % ] = 100 (3.24)
0.5h

23
am
ρm [ % ] = 100 (3.25)
0.5h

(d) Verify that the reinforcement ratio of the temporary and permanent facing at

the mid-span and the nail head are greater than the minimum reinforcement ratio

(i.e., ρmin ≤ ρ ), otherwise increase the amount of reinforcement (an and/or am) to

satisfy this criterion.

(e) Verify that the reinforcement ratio of the temporary and permanent facing at

the mid-span and the nail head are smaller than the maximum reinforcement ratio

(i.e., ρ ≤ ρmax ), otherwise reduce the amount of reinforcement (an and/or am) to

satisfy this criterion.

Step 5: Verify facing flexural resistance RFF for temporary and permanent facings

(a) Calculate facing flexural resistance RFF for the temporary and permanent

facing as the minimum of:

CF S 
R FF [ kN ] = × ( a vn + a vm )  mm 2 / m  ×  h h [ m ]  × f y [ MPa ] (3.26)
265  Sv 

CF S 
R FF [ kN ] = × ( a hn + a hm )  mm 2 / m  ×  v h [ m ]  × f y [ MPa ] (3.27)
265  Sh 

where:

CF = factor that considers the non-uniform soil pressures behind the facing. For

permanent facing CF is adopted taken equal to 1, whereas, for temporary facings

with thickness: 100 mm, 150 mm and 200m, CF shall be adopted as 2.0, 1.5 and

1.0 respectively.

(h) Determine the safety factor against facing flexural failure using Eq. (3.28) and

if minimum recommended factor of safety against facing flexural failure is not

achieved, redesign the facing with increased thickness of facing, steel

24
reinforcement strength, concrete strength, and/or amount of steel and repeat the

facing flexural resistance calculations.

R FF
FSFF = (3.28)
To

Step 6: Verify facing punching shear resistance RFP

Punching shear failure of the facing can occur around the nail head and must be

evaluated at: (1) bearing-plate connection (used in temporary facings), and (2)

headed-stud connection (commonly used in permanent facings).

As the nail head tensile force increases to a critical value, fractures can form a

local failure mechanism around the nail head resulting in a conical failure surface.

This failure surface extends behind the bearing plate or headed studs and punches

through the facing at an inclination of about 45 degrees (see Figs. 3.7 and 3.8).

The size of the cone depends on the facing thickness and the type of the nail-

facing connection (i.e., bearing-plate or headed-studs). For practical purposes,

punching shear capacity RFP is assessed similar to the concrete structural slabs

subjected to concentrated loading and is evaluated as:

R FP [ kN ] 330 f ck [ MPa ] πDc' [ m ] h c [ m ]


= (3.29)

where:

D'c = effective diameter of conical failure surface at the center of section (i.e.,

considering an average cylindrical failure surface)

hc = effective depth of conical surface

For temporary facing (see Fig. 3.7)

D'c = LBP + h and hc = h; where: LBP = length of bearing plate and h = thickness of

temporary facing.

25
Fig. 3.7 Punching shear failure mode: Bearing plate connection used in temporary
facing (FHWA 2003).

Fig. 3.8 Punching shear failure mode: Headed stud connection used in permanent
facing (FHWA 2003).

For permanent facing (see Fig. 3.8)

D'c = minimum of (SHS + hc and 2hc) and hc = LS – tH + tP

where: SHS = headed-stud spacing, LS = headed-stud length, tH = headed-stud head

thickness, and tP = bearing plate thickness (see Table A.5).

26
Knowing the punching shear capacity RFP of the facing and the axial force at nail

head To, factor of safety against facing punching shear failure modes FSFP can

determined as:

R FP
FSFP = (3.30)
To

If capacity for the temporary/permanent facing is not adequate, then implement

larger elements or higher material strengths and repeat the punching shear

resistance calculations.

Step 7: Verify headed-stud resistance RHT (Permanent facing)

(a) The tensile capacity of the headed-studs connectors providing anchorage of the

nail into the permanent facing (also providing connection between both temporary

and permanent facings) must be verified. The nail head capacity against tensile

failure of the headed-studs RHT can be computed as:

R HT = N H A SH f y (3.31)

where:

NH = number of headed-studs in the connection (usually 4); ASH = cross-sectional

area of the headed-stud shaft; and fy = tensile yield strength of the headed-stud.

Knowing the nail head capacity against tensile failure of the headed-studs RHT and

the axial force at nail head To, factor of safety against the tensile failure of the

headed-studs FSHT can determined as:

R HT
FSHT = (3.32)
To

(b) Also verify that compression on the concrete behind headed-stud is within

tolerable limits by assuring that:

A H ≥ 2.5A SH and t H ≥ 0.5 ( D H − DS ) (3.33)

27
where:

AH = cross-sectional area of the stud head; tH = head thickness; DH = diameter of

the stud head; and DS = diameter of the headed-stud shaft.

Note: Provide sufficient anchorage to headed-stud connectors and extended them

at least to the middle of the facing section and preferably behind the mesh

reinforcement in permanent facing. Also, provide a minimum 50 mm of cover

over headed-studs.

Step 8: Other facing design considerations

To minimise the likelihood of a failure at the nail head connection: (1) bearing

plates should be mild steel with a minimum yield stress fy equal to 250 MPa, (2)

nuts should be the heavy-duty, hexagonal type, with corrosion protection, and (3)

beveled washers (if used) should be steel or galvanized steel.

28
SECTION 4: GENERAL DESIGN CONSIDERATIONS

4.1 INTRODUCTION

This section summarises general steps for the design of soil nail walls for practical

applications. Salient design considerations have been highlighted to facilitate designers in

arriving at the desired soil nail wall configuration ensuring stability and performance

requirements.

4.2 GENERAL DESIGN STEPS

Design example presented in Section 5 shall be considered helpful in understanding the

step-by-step design procedure for the most common applications of soil nail walls in

highway engineering (such as: stabilization of vertical cuts for approaches to subways). In

general, the design procedure for soil nail walls includes the following steps:

4.2.1 Initial soil nail wall considerations

(a) Wall layout: Establish the layout of the soil nail wall, including: (1) wall height; (2)

length of the wall; (3) backslope; and (4) wall face batter. Wall face batter typically

ranges from 0º to 10º. The evaluation of the wall layout also includes developing

longitudinal profile of the wall, locating wall appurtenances (e.g., traffic barriers,

utilities, and drainage systems), and establishing ROW limitations.

(b) Soil nail vertical and horizontal spacing: Typically, same nail spacing can be

adopted in both horizontal Sh and vertical Sv directions. Nail spacing ranges from 1.25

to 2 m (commonly 1.5 m) for conventional drilled and grouted soil nails, and as low

as 0.5 m for driven nails. As a general rule, soil nail spacing in horizontal and vertical

direction must be such that each nail has an influence area Sh × Sv ≤ 4 m 2 . Though,

soil nail spacing may get affected by the presence of existing underground structures,

29
the design engineer should specify a minimum horizontal soil nail spacing of about

1.0 m.

(c) Soil nail pattern on wall face: The soil nail pattern on wall face may be adopted as

one of the following: (1) square (or rectangular); (2) staggered in a triangular pattern;

and (3) irregular (at limited locations) depending upon the ease of construction and

site-specific constraints.

(d) Soil nail inclination: Soil nails are typically installed at an inclination ranging from

10 to 20 degrees from horizontal with a typical inclination of 15 degrees. Inclination

of soil nails in this range helps in proper flow of grout as well as better provides

reinforcing effect.

(e) Soil nail length and distribution: The distribution of soil nail lengths in a soil nail

wall can be selected as either uniform (i.e., only one nail length is used for the entire

wall), or variable, where different nail lengths may be used for individual soil nail

levels within a wall cross section. Uniform nail pattern is recommended for most

applications; however, provision of longer nail lengths in upper two-thirds to three-

quarters of the wall height significantly reduces wall deformations. Also, variable nail

lengths may become inevitable in presence of underground utilities which imposes

restriction on designed nail length.

(e) Soil nail materials and soil properties: Shall be adopted as specified in Section 2.

(f) Other initial considerations: Evaluate corrosion potential; explore drilling methods

likely to be adopted by the prospective contractors (this will provide necessary help in

the appropriate selection of the design ultimate bond strength of soil nails); estimate

drillhole diameter based on previous experience in similar ground conditions; select

desired factors of safety for the different failure modes; and define loadings required

to be considered in the design.

30
4.2.2 Preliminary design

(a) For the specific project application, obtain general parameters as mentioned in Section

4.2.1. For known drilling method and soil conditions, from Table A.2 adopted

suitable value of ultimate bond strength qu, which may otherwise, be obtained by

conducting field pullout tests on soil nails as described in Section 6.

(b) Determine tensile force T developed in nail at depth z using Eq. (4.1)

1 − sin φ
T [ kN
= ] K(q s + γz)ShSV ; where: =
K K= (4.1)
1 + sin φ
a

and note the maximum value of axial tensile force Tmax.

(c) For a minimum factor of safety of against nail tensile failure FST = 1.80, determine

required cross-sectional area At of the nail bar can be determined as:

Tmax FST
A t  mm 2  = (4.2)
fy

choose closest commercially available bar size that has a cross-sectional area at least

that evaluated using Eq. (4.2).

(d) Determine minimum nail length L as the maximum of L1 and L2

=L1
( H − Sv1 ) cos ψ + 2T1 and L = 0.6H (4.3)
sin ( ψ + i )
2
πD DH q u

where: T1 and Sv1 are the axial force and vertical spacing of top nail respectively; ψ =

inclination of failure plane; i = nail inclination; L2 = minimum recommended length

and L1 is the length obtained to satisfy following two criteria:

(i) Length of the top nail (i.e. first nail) should be such that it crosses the failure

surface.

(ii) Length of the top nail (i.e. first nail) gives a factor of safety against nail pullout

failure is more than 2.0.

31
Note: Minimum nail length L shall be evaluated for the static case considering

vertical height H of the soil nail wall (i.e. assuming face batter α = 0 degrees)

irrespective of the actual wall geometry.

4.2.3 Final design

Final design of soil nail wall requires following steps:

(a) Check for external failure modes: (i) Global stability, (ii) Sliding stability and (iii)

Basal heave or bearing capacity. Global stability and sliding stability for static and

seismic (if required) conditions of the soil nail walls may be evaluated using the

procedure stated in Sections 3.3.1 and 3.3.2 respectively. Basal heave or bearing

capacity failure mode shall be considered prominent only in cases where soil nail

walls are founded on soft soils. Basal heave or bearing capacity failure mode may be

evaluated with reference to the Section 3.3.3.

Note: For the evaluation of the global stability, use of rigorous computational tools in

addition to the proposed conventional methodology is strongly recommended.

(b) Check for internal failure modes: (i) Soil-nail pullout failure and (ii) Nail tensile

failure. Both of these internal failure modes shall be evaluated at each nail level under

static and seismic (if required) conditions with reference to the Section 3.4.

(c) Facing design and checks: For both temporary and permanent soil nail wall facings,

design and recommended checks shall be conducted under static and seismic (if

required) conditions with reference to the Section 3.5.

(d) Other design considerations such as: permissible wall deformations (may be verified

using suitable computational tools), internal and surface drainage, corrosion

protection and any site-specific issues shall be adequately addressed in accordance

with relevant standards and bye-laws.

32
4.3 OTHER DESIGN CONSIDERATIONS

4.3.1 Loads and load combinations

Soil nail walls used on typical highway projects are typically subjected to the following

different loads during their service life: (i) Dead loads DL (e.g., weight of the soil nail

wall system, lateral earth pressure, weight of a nearby above-ground structure); (ii) Live

loads LL (e.g., traffic loads); (iii) impact loads IL (e.g., vehicle collision on barriers above

soil nail wall); and (iv) earthquake loads EQ. Following load combinations are

recommended to assess the most critical loading condition:

(a) DL + LL

(b) DL + LL + IL

(c) DL + EQ

For earthquake loads, allowable stresses shall be increased by 133 percent from the values

obtained with factors of safety for static loads.

4.3.2 Permissible soil nail wall deformations

The maximum permissible lateral deformation at the top of the soil nail walls constructed

in weathered rock and stiff soils is 0.1%H; sandy soils is 0.2%H and for fine-grained

soils is 0.3%H. Under no circumstances maximum permissible lateral deformation shall

exceed 0.3% H, where: H is the vertical height of the soil nail wall. Permissible vertical

deformation (i.e., settlement) shall be considered to be same as the permissible horizontal

deformation.

4.3.3 Drainage measures

(a) Short term drainage measures: Surface water and groundwater must be controlled

both during and after construction of the soil nail wall. A surface water interceptor

ditch, excavated along the crest of the excavation and lined with concrete, is a

33
recommended element for controlling surface water flows. Additionally, if ground

water impacts are temporary or localized, suitable dewatering measures may be taken

for lowering the groundwater table

(b) Long term drainage measures: By means of the drainage system comprising of: (i)

vertical geo-composite drain strips placed suitably along the face of wall; (ii) weep

holes in the form of PVC pipes of typical diameter as 300-400 mm placed through the

face at the location of expected localised seepage; (iii) provision of horizontal or

slightly inclined drain pipes of typical diameter 50 mm installed at the locations

where it is necessary to control the groundwater pressures imposed on the retained

soil mass; (iv) installation of permanent interception ditch behind the wall at its crest

to prevent surface water runoff from infiltrating behind the wall or flowing over the

wall edge; and (v) provision of a vegetative protective cap to reduce or retard water

infiltration into the soil.

4.3.4 Corrosion protection

Corrosion potential of the soil must be evaluated for all permanent soil nail walls and, in

some cases, for temporary walls. Corrosion potential of soil can be evaluated based on

tests results of the following properties: pH (potential of hydrogen); electrical resistivity;

chloride content; sulfate content; and presence of stray currents.

Corrosion protection measures: (a) Specify a minimum grout cover of 25 mm between

the reinforcement nail bar and the soil; (b) recommend epoxy coating of minimum

thickness 0.4 mm on the nail bars shall be applied by the manufacturer prior to shipment

of nails to the construction site, which is, subsequently to be encased in grout cover; and

(c) adopt other site-specific suitable corrosion protection measures.

34
SECTION 5: DESIGN EXAMPLE

5.1 INTRODUCTION

This section illustrates the step by step design methodology of soil nail walls with the

help of an example. It is intended to provide field practitioners an exposure to the design

methodology of soil nail structures with reference to the various failure modes and design

considerations described in the Sections 3 and 4.

Problem statement: A permanent soil nail is designed to support a vertical cut for the side

wall of the approach road for a subway. The maximum design vertical height of the side

wall to be supported is 8 m. Groundwater table is considered to be significantly below the

zone of influence. Backslope of the wall is assumed to be flat. Site is considered to be at

an urban location with the possibility of commercial establishments in the vicinity of the

soil nail wall. Peak horizontal ground acceleration is assumed to be 0.15g for seismic

considerations. It is also assumed that the project do not have specific construction

restrictions and particular requirements such as aesthetics, deformation control etc.

5.2 INITIAL DESIGN CONSIDERATIONS

(a) Vertical height of wall: H = 8 m

(b) Face batter: α = 0.0 degrees; Backslope angle: β = 0.0 degrees

(c) Soil nail spacing: Sh = Sv = 1.5 m (Note: vertical spacing of first nail Sv1 = 0.75 m)

(d) Soil nail spacing pattern at wall face: Square

(e) Soil nail inclination: i = 15 degrees

(f) Soil nail length distribution: Uniform

(g) Soil nail material: Grade Fe 415; fy = 415 MPa

(h) Representative soil properties from soil investigation report:

35
Soil type: dense to very dense silty sands; Cohesion: c = 5 kPa; Friction angle:

φ =350 ; Unit weight: γ =18.9 kN / m3 .

Adopt ultimate bond strength: qu = 100 kPa (see Table A.2)

(i) Loads considered:

Self weight of the structure and surcharge load: qs = 20 kN/m2

Seismic loading: kh = 0.15 and kv = 0.0

(j) Drilling method: rotary drilling; Drillhole diameter: DDH = 130 mm

(k) Desired minimum factors of safety for various failure modes under static and seismic

conditions are shown in Table 5.1.

Table 5.1 Desired minimum recommended factors of safety for permanent soil nail walls.
Factor of safety
Failure mode Resisting component Symbol
Seismic Static

External stability Global FSG 1.10 1.5


Sliding FSSL 1.10 1.5
Pull-out resistance FSP 1.50 2.0
Internal stability
Nail bar tensile strength FST 1.35 1.8
Facing flexure FSFF 1.10 1.5
Facing failure Facing punching failure FSFP 1.10 1.5
Headed stud tensile FSHT 1.50 2.0

5.3 PRELIMINARY DESIGN

(a) Determine maximum axial force Tmax

Tensile force T developed in nail at depth z is given by

1 − sin φ 1 − sin 35
T [ kN
= ] K(qs + γz)ShSV ; for static case: =
K K= = = 0.27
1 + sin φ 1 + sin 35
a

On substituting values of various parameters, T [ kN


= ] 12.2 + 11.53z (see Table 5.2)

36
Table 5.2 Maximum axial tensile force developed in each nail from top.
Nail No. j Depth of nail z [m] Axial force T [kN]
01 0.75 (z = Sv1) 20.85
02 2.25 38.14
03 3.75 55.54
04 5.25 72.73
05 6.75 90.00

Therefore, maximum axial force: Tmax = 90.0 kN

(b) Determine minimum nail length L and nail diameter d

Minimum length of soil nail L is adopted as the maximum of L1 and L2:

=L1
( H − Sv1 ) cos ψ + 2T1 and L = 0.6H
sin ( ψ + i )
2
πD DH q u

Here: ψ= 45 + (35 / 2)= 62.50 ; i = 150 ; qu = 100 kPa; DDH = 130 mm; T1 = 20.85 kN

Sv1 = 0.75 and H = 8 m. On substituting these values in above equation

=
L1
( 8 − 0.75) cos62.5 + 2 × 20.85= 4.44 m and L2 = 0.6 x 8 = 4.80 m.
sin ( 62.5 + 15 ) π × 0.13 ×100

Hence, adopt nail length: L = 4.80 m.

For a minimum factor of safety of against nail tensile failure FST = 1.80, the required

cross-sectional area At of the nail bar can be determined as:

Tmax FST 90 ×1000 ×1.80


 mm 2 
At = = = 390
fy 415

Select reinforcement bar of diameter d = 25 mm providing cross sectional area At = 490

mm2 (> 390 mm2). This bar can be installed with no difficulty in the drillhole. Available

grout cover is at least (100 - 25) / 2 = 37.5 mm > minimum cover = 25 mm.

General layout of the soil nail wall is shown in Fig. 5.1 and is used as the reference in the

following calculations.

37
5.4 FINAL DESIGN

Final design consists of (i) evaluation of external and internal failure modes under static

and seismic conditions (if considered), and (ii) facing design and checks against facing

failure modes.

Fig. 5.1 General outline of designed soil nail wall.

5.4.1 EXTERNAL FAILURE MODES

5.4.1.1 Global stability

(a) Static global stability

Determination of length of failure plane LF

H 8
[m]
L F= = = 9.02
sin ψ sin 62.5

Determination of equivalent nail force Teq

R P [ kN ] = πD DH L P q u = π × 0.13 × L P ×100 = 40.84L P

38
 ( H − z ) cos ψ   ( 8 − z ) cos62.5 
where: L P [ m ] =
L− =4.80 −  =1.02 + 0.47z
 sin ( ψ + i )   sin ( 62.5 + 15 ) 

πd 2f y π × 252 × 415
=
R T [ kN ] = = 203.71
4 ×1000 4 ×1000

Allowable axial force carrying capacity Tall [kN] of nail embedded at depth z is the

minimum of RP and RT.

For Sh = 1.5 m, equivalent nail force Teq can be determined as:

1 n 1 5
Teq [ kN / m ] =∑ ( Tall ) j = × 568.08 =
378.72 ; Here: n = 5 and ∑(T ) all j
is obtained
Sh j=1 1.5 j=1

from Table 5.3.

Table 5.3 Allowable axial force carrying capacity of nails at different levels.
Effective Allowable axial
Nail pullout Nail tensile
Nail No. j Depth of pullout force carrying
capacity RP capacity RT
(from top) nail z [m] length Lp capacity of nail
[kN] [kN]
[m] Tall [kN]
01 0.75 1.37 55.95 203.71 55.95
02 2.25 2.08 84.95 203.71 84.95
03 3.75 2.78 113.53 203.71 113.53
04 5.25 3.49 142.53 203.71 142.53
05 6.75 4.19 171.12 203.71 171.12
5

∑(T )
j=1
all j
= 568.08

Determination of weight of failure wedge W

With reference to the Fig. 5.1, weight of failure wedge ABC can be determined as:

W [ kN / m ] = γ × ( Area of ∆ABC ) = 0.5γH 2 cot ψ

Therefore, W [ kN / m ] = 0.5 ×18.9 × 82 × cot 62.5 = 314.84

Determination of total surcharge load QT

QT [ kN / m=
] q s H cot ψ= 20 × 8 × cot 62.5= 83.29
Therefore, term ( W + QT =
) 314.84 + 83.29= 398.13kN / m

39
Global stability safety factor FSG under static conditions is given by:

cL F + Teq cos ( ψ − i ) + ( W + QT ) cos ψ + Teq sin ( ψ − i )  tan φ


FSG =
( W + QT ) sin ψ
Substituting the values of various parameters FSG under static condition is obtained as:

( 5 × 9.02 ) + 378.72cos ( 62.5 − 15) + ( 398.13) cos62.5 + 378.72sin ( 62.5 − 15)  tan 35
FSG =
( 398.13) sin 62.5
FSG = 1.77 (> 1.50 safe for static case)

(b) Seismic global stability

For kh = 0.15 and kv = 0.0

Horizontal seismic inertia force, Fh [ kN / m ] = k h (W + QT ) = 0.15 × 398.13 = 59.72

Vertical seismic inertia force, Fv [ kN / m ]= k v (W + QT )= 0.0

Global stability safety factor FSG under seismic conditions is given by:

cL F + Teq cos ( ψ − i ) + ( W + QT − Fv ) cos ψ + Teq sin ( ψ − i ) − Fh sin ψ  tan φ


FSG =
( W + QT − Fv ) sin ψ + Fh cos ψ
On substitution of values of various parameters FSG under seismic condition is obtained

as:

588.10
=
FSG = 1.54 (> 1.10 safe for seismic case)
380.71

5.4.1.2 Sliding stability

(a) Static sliding stability

Factor of safety for sliding stability of soil nail wall FSSL in static condition is given by:

c b BL + ( W + QT + Psin βeq ) tan φb


FSSL =
Pcos βeq

Here: c b = c = 5 kPa; φb = φ = 350 ; βeq = β = 0.0; BL = L = 4.80 m

For static case total active lateral earth pressure P = PA can be determined as:

40
γH 2  2q S  18.9 × 82  2 × 20 
PA [ kN / m=
] K 1 + = 0.27 1 + = 206.57
2  γH  2  18.9 × 8 

where: coefficient of lateral active earth pressure: K = Ka = 0.27

W [kN/m] = Unit weight x Area of sliding wedge ABDE =18.9 x (8 x 4.8) = 725.76

QT [kN/m] = Surcharge load x Length AD = qs x BL = 20 x 4.8 = 96

Therefore, term (W + QT) = 725.76 + 96 = 821.76 kN/m

Substituting of values of various parameters, FSSL under static condition is obtained as:

FSSL
(=
5 × 4.8 ) + ( 821.76 + 206.57sin 0 ) tan 35
2.90 (> 1.50 safe for static case)
206.57cos0

(b) Seismic sliding stability

Factor of safety for sliding stability of soil nail wall FSSL in seismic condition is given by:

c b BL + ( W + QT − Fv + Psin βeq ) tan φb


FSSL =
Fh + Pcos βeq

For kh = 0.15 and assumed kv = 0.0

For seismic case total active lateral earth pressure P = PAE can be determined as:

γH 2  2q 
PAE [ kN / m=
] K (1 − k v ) 1 + S 
2  γH 

where: coefficient of lateral active earth pressure: K = Kae, which can be evaluated as:

cos 2 ( φ − ω) cos2 ( 35 − 8.53)


K = 2
= 2
0.36
 sin φ sin ( φ − ω)   sin 35sin ( 35 − 8.53) 
cos 2 ω 1 +  cos2 8.53 1 + 
 cos ω   cos8.53 

 kh  −1  0.15 
[deg ] tan −1  =
where: ω=  tan  =  8.53
 1 − k v   1 − 0 

18.9 × 82  2 × 20 
[ kN / m ]
Therefore, PAE= 0.36 (1 − 0 ) 1 += 275.33
2  18.9 × 8 

Horizontal seismic inertia force: Fh [ kN / m ] = k h (W + QT ) = 0.15 × 821.76 = 123.26

41
Vertical seismic inertia force: Fv [ kN / m ]= k v (W + QT )= 0.0

Substituting values of various parameters, FSSL under seismic condition is obtained as:

FSSL
5 × 4.8 ) + ( 821.76 − 0 + 275.33sin 0 ) tan 35
(= 1.50 (> 1.10 safe for seismic case)
123.26 + 275.33cos0

5.4.1.3 Bearing capacity (or basal heave) failure

Since, the soil nail wall is founded in dense silty sand, basal heave or bearing capacity

failure of the soil nail wall is not likely to occur and hence, not evaluated in this example.

5.4.2 INTERNAL FAILURE MODES

5.4.2.1 Soil nail pullout failure

For any particular nail embedded at depth z, factor of safety against pullout failure FSP

can be obtained as:

R 
(FSP )z =  P 
 T z

where: RP and T are determined similarly as in Table 5.3. While determining T, value of

earth pressure coefficient K = Ka (here: Ka = 0.27) for static case and K = Kae (here: Kae =

0.36) for seismic case shall be used. Results of the soil nail pullout failure analysis for

seismic and static cases are shown in Tables 5.4a and 5.4b respectively. The minimum

recommended FSP for static case is 2.0 and for seismic case is 1.50. It may be observed

from Tables 5.4a and 5.4b that for the bottom two nails (04 and 05) FSP values are

marginally less than the minimum recommendations, though practically insignificant.

5.4.2.2 Soil nail tensile strength failure

Factor of safety against nail tensile strength failure FST for any nail embedded at depth z

can be obtained as:

42
( FST )z = 
RT 

 T z

where: RT and T are determined similarly as in Table 5.3. While determining T, value of

earth pressure coefficient K = Ka (here Ka = 0.27) for static case and K = Kae (here Kae =

0.36) for seismic case shall be used. Results of the soil nail tensile failure analysis for

seismic and static cases are shown in Table 5.5. It may be observed from Table 5.5 that

the factors of safety against nail tensile strength failure are well above the minimum

recommended FST for static case is 1.80 and for seismic case is 1.35.

Table 5.4a Factor of safety against soil nail pullout failure (static case).
Axial force Factor of safety
Depth of Effective Nail pullout
Nail No. j developed against nail
nail z pullout capacity RP
(from top) in nail T pullout failure
[m] length Lp [m] [kN]
[kN] FSP
01 0.75 1.37 55.95 20.85 2.68
02 2.25 2.08 84.95 38.14 2.22
03 3.75 2.78 113.53 55.54 2.04
04 5.25 3.49 142.53 72.73 1.95
05 6.75 4.19 171.12 90.00 1.90

Table 5.4b Factor of safety against soil nail pullout failure (seismic case).
Axial force Factor of safety
Depth of Effective Nail pullout
Nail No. j developed against nail
nail z pullout capacity RP
(from top) in nail T pullout failure
[m] length Lp [m] [kN]
[kN] FSP
01 0.75 1.37 55.95 27.68 2.02
02 2.25 2.08 84.95 50.64 1.68
03 3.75 2.78 113.53 73.61 1.54
04 5.25 3.49 142.53 96.57 1.48
05 6.75 4.19 171.12 119.53 1.43

43
Table 5.5 Factor of safety against soil nail tensile strength failure.
Axial force Axial force Factor of safety
Nail No. Depth of Nail tensile developed in developed in against nail tensile
j (from nail z capacity RT nail T [kN] nail T [kN] failure FST
top) [m] [kN]
Static Seismic Static Seismic
01 0.75 203.71 20.85 27.68 9.77 7.36
02 2.25 203.71 38.14 50.64 5.34 4.02
03 3.75 203.71 55.54 73.61 3.66 2.76
04 5.25 203.71 72.73 96.57 2.80 2.11
05 6.75 203.71 90.00 119.53 2.26 1.70

5.4.3 FACING DESIGN AND CHECKS

Step 1: Calculate design nail head tensile force at the face To

For Tmax = 90.0 kN (static case); Tmax = 119.53 kN (seismic case) and Smax =1.5 m, nail

head tensile force at the wall face To can be obtained as:

To [ kN
= ] Tmax [0.6 + 0.2(Smax − 1)=] 90 [0.6 + 0.2(1.5 − 1)=] 63.0 (for static case)
To [ kN
= ] Tmax [0.6 + 0.2(Smax −= 1)] 83.67 (for seismic case)
1)] 119.53[0.6 + 0.2(1.5 −=

Step 2: Adopt wall facing thickness

Temporary facing thickness h: 100 mm

Permanent facing thickness h: 200 mm

Step 3: Adopt appropriate facing materials

(a) Steel reinforcement: Grade Fe 415 with characteristic strength: fy = 415 MPa

(b) Concrete/shotcrete: Grade M20 with characteristic compressive strength: fck = 20 MPa

(c) Welded wire mesh (temporary facing): WMM 102 x 102–MW19 x MW19 (see Table

A.3)

(d) Horizontal and vertical waler bars (temporary facing): 2 x 10 mm diameter, (fy = 415

MPa, Avw = Ahw = 2 x 78 = 156 mm2) in both directions.

44
(e) Bearing plate (temporary facing): Grade 250 (fy =250 MPa); Shape: Square; Length:

LBP = 225 mm; Thickness: tp = 25 mm.

(f) Reinforcement bar (permanent facing): 16 mm diameter @ 300 mm both ways.

(g) Headed-studs: 4 numbers; Size: 1 × 4 1 ; LS = 100 mm; DH =25 mm; DS =13 mm; tH
2 8

= 8 mm; SHS = 150 mm (see Table A.5)

Step 4: Checks for facing reinforcement

Determine the minimum and the maximum reinforcement ratios as:

f ck [ MPa ] 20
ρmin [ % ] = 20 = 20 = 0.21
f y [ MPa ] 415
f ck [ MPa ]  600  20  600 
ρmax [ % ] 50
=  =  50 =  1.42
f y [ MPa ]  600 + f y [ MPa ]  415  600 + 415 

Therefore, the adopted reinforcement shall be ρmin ≤ ρ ≤ ρmax . In addition the ratio of the

reinforcement in the nail head and mid-span zones should be less than 2.5 to ensure

comparable ratio of flexural capacities in theses areas.

Temporary facing

Reinforcement in vertical avm and horizontal ahm directions in mid-span:

avm = ahm = 184.2 mm2/m for WWM 102 x 102 – MW19 x MW19 (see Table A.3)

Reinforcement in vertical avn and horizontal ahn directions around soil nail head:

Since same amount of reinforcement is provide in both directions (see Step 3d), avn = ahn

A vw 156
a vn =a hn =a vm + =184.2 + =288.2 mm 2 / m
Sh 1.5
Reinforcement ratio ρ at nail head and mid-span in vertical direction

a vn (288.2 /1000)
ρn [=
%] =
100 ×100
= 0.58
0.5h 0.5 ×100

a vm (184.2 /1000)
ρm [=
%] =
100 ×100
= 0.37
0.5h 0.5 ×100

45
checks:

ρn = 0.58% > ρmin = 0.21% … ok

ρn = 0.58% < ρmax = 1.42% … ok

ρm = 0.37% > ρmin = 0.21% …ok

ρm = 0.37% < ρmax = 1.42% … ok and ρn / ρm = 1.56 < 2.5 …ok

Permanent facing

Total area of 16 mm diameter @ 300 mm c/c is equal to 670 mm2/m (see Table A.4).

This area of reinforcement is provided in both vertical and horizontal directions,

therefore, avn = ahn = avm = ahm = 670 mm2/m (no waler bars are provided in permanent

facing).

Reinforcement ratio ρ at nail head and mid-span in vertical direction

a (670 /1000)
ρn [ % ] =
ρm [ % ] =vn 100 = 0.67 (satisfies both ρmin = 0.21% and
×100 =
0.5h 0.5 × 200

ρmax = 1.42% and ρn / ρm = 1.0 < 2.5)

Step 5: Verify facing flexural resistance RFF

Temporary facing

Calculate facing flexural resistance RFF as:

CF S 
R FF [ kN ] = × ( a vn + a vm )  mm 2 / m  ×  h h [ m ]  × f y [ MPa ]
265  Sv 

For temporary facing with thickness h = 100 mm (= 0.1 m), adopt CF = 2.0

( a vn + a vm ) = 288.2 + 184.2 = 472.4 mm 2 / m ; nail spacing ratio: Sh/Sv =1.0.

2
Therefore, R FF [ kN ]= × 472.4 × (1 × 0.1) × 415= 148
265

Safety factor against facing flexural failure FSFF is given by

R FF 148
FS=
FF = = 2.35 (> 1.50 safe for static case)
To 63

46
R FF 148
=
FS FF = = 1.76 (> 1.10 safe for seismic case)
To 83.67

Permanent facing

For permanent facing with thickness h = 200 mm (= 0.2 m), adopt CF = 1.0

( a vn + a vm ) = 670 + 670 = 1340 mm2 / m ; nail spacing ratio: Sh/Sv =1.0.

1
Therefore, R FF [ kN ]= ×1340 × (1 × 0.2 ) × 415= 420
265

Safety factor against facing flexural failure FSFF is given by

R FF 420
=
FS FF = = 6.67 (> 1.50 safe for static case)
To 63

R FF 420
=
FS FF = = 5.02 (> 1.10 safe for seismic case)
To 83.67

Step 6: Verify facing punching shear resistance RFP

Temporary facing: Check for bearing-plate connection

Facing punching shear capacity RFP is given by:

R FP [ kN ] 330 f ck [ MPa ] πDc' [ m ] h c [ m ]


=

Here: fck = 20 MPa; hc = h = 0.1 m; D'c = LBP + h = 225 + 100 = 325 mm = 0.325 m

Substituting values of various parameters, temporary facing punching shear capacity RFP

is calculated as

R FP [ kN
= ] 330 × 20 × π× 0.325 × 0.1
= 150

Safety factor against facing punching shear failure FSFP is given by

R FP 150
=
FS FP = = 2.38 (> 1.50 safe for static case)
To 63

R FP 150
=
FS FP = = 1.79 (> 1.10 safe for seismic case)
To 83.67

47
Permanent facing: Check for headed-stud connection

Here: fck = 20 MPa; LS = 100 mm; DH = 25 mm; DS =13 mm; tH = 8 mm; SHS = 150 mm;

tp = 25 mm.

hc = LS – tH + tP = 100 – 8 + 25 = 117 mm = 0.117 m; 2hc = 0.234 m

D'c = minimum of (SHS + hc and 2hc)

SHS + hc = 150 + 117 = 267 mm = 0.267 m

Therefore, D'c = 0.234 m (minimum of SHS + hc and 2hc)

Substituting values of various parameters, permanent facing punching shear capacity RFP

is calculated as:

R FP [ kN
= ] 330 × 20 × π× 0.234 × 0.117
= 127

Safety factor against facing punching shear failure FSFP is given by

R FP 127
FS=
FP = = 2.01 (> 1.50 safe for static case)
To 63

R FP 127
=
FS FP = = 1.52 (> 1.10 safe for seismic case)
To 83.67

Step 7: Verify headed-stud resistance RHT (only for permanent facing)

Tensile capacity of the headed-studs

πDS2 π×132
=
For: NH = 4; A SH = = 132.73mm 2 ; fy = 0.415 kN/mm2
4 4

Nail head capacity against tensile failure of the headed-studs RHT can be computed as:

R HT [ kN ] =
N H ASH f y =
4 ×132.73 × 0.415 =
220.33

Safety factor against headed-stud tensile failure FSHT is given by

R HT 220.23
=
FSHT = = 3.49 (> 1.50 safe for static case)
To 63

R HT 220.23
=
FSHT = = 2.63 (> 1.10 safe for seismic case)
To 83.67

48
Check for tolerable limits of compression on the concrete behind headed-stud

πD 2H π× 252 πDS2 π×132


=
AH = = 490.87 mm and A
2
= SH = = 132.73mm 2
4 4 4 4

To assure that the compression on the concrete behind headed-stud is within tolerable

limits, following two conditions shall be satisfied:

(a) A H ≥ 2.5ASH ⇒ 490.87 ≥ 2.5 (132.73


= ) 490.87 mm 2 ≥ 331.82 mm 2 .......ok

(b) t H ≥ 0.5 ( D H − DS ) ⇒ 8 ≥ 0.5 ( 25 − 13


= ) 8 mm ≥ 6 mm.......ok

5.4.4 GENERAL CONSIDERATIONS AND DESIGN SUMMARY

5.4.4.1 General considerations

(a) Permissible lateral deformation

Depending on the soil type, permissible lateral deformation of the soil nail wall shall be

within 0.1-0.3% of the vertical height H. For sandy soils (in present example),

permissible lateral wall deformation = 0.002H = 0.002 x 8000 = 16 mm.

(b) Drainage and corrosion resistance

Since the groundwater table at the construction site is well below the zone of influence,

internal drainage can be considered sufficient in the form of geocomposite drain strips,

weepholes and toe drains. For surface drainage, suitable provisions shall be made in

agreement with opinion of contractor and the site-engineer. Although, the ground

corrosion potential is unknown in the present example, suitable measures against

corrosion protection shall be adopted. A minimum grout thickness of 25 mm over nail

bars shall be provided.

5.4.4.2 Design summary

Tables 5.6-5.8 presents the summary of soil nail wall design example presented above.

Further, Fig. 5.2 shows the details of reinforcement in both temporary and permanent

facings.

49
Table 5.6 Summary of important design parameters.
Parameter Value
Number of nails per section 5.0
Nail inclination i [degrees] 15.0
Drillhole diameter DDH [mm] 130.0
Diameter of nails d [mm] 25.0
Length of nails LN [m] 4.8
Spacing Sh x Sv [m x m] 1.5 x 1.5
Maximum axial tensile force Tmax [kN] (seismic case) 90.0 (119.53)

Table 5.7 Summary of factors of safety attained for various failure modes.
Factor of safety
Failure mode Remarks
Seismic Static
Global FSG -- 1.77 1.54
Sliding FSSL -- 2.90 1.50
Pull-out resistance FSP Minimum 1.90 1.43
Nail bar tensile strength FST Minimum 2.26 1.70
Temporary facing 2.35 1.76
Facing flexure FSFF
Permanent facing 6.67 5.02
Temporary facing 2.38 1.79
Facing punching FSFP
Permanent facing 2.01 1.52
Headed stud tensile FSHT -- 3.49 2.63

Table 5.8 Summary of facing design (temporary and permanent).


Element Description Temporary facing Permanent facing
Thickness h 100 200
General Facing type Shotcrete CIP concrete
Concrete grade M20 M20
Type Welded wire mesh (WWM) Steel bars
Reinforcement Steel grade Fe415 Fe415
Denomination 102 x 102 – MW19 x MW19 16 φ @ 300 b/w

Other
Type Waler bars 2 - 10 φ b/w --
reinforcement
Type Square 4H-Studs 1/2 x 4 1/8
Bearing plate Steel Fe250 --
Dimensions 225 x 225 x 25 --
-- Nominal length, LS = 100
-- Head diameter, DH = 25
Headed studs Dimensions -- Shaft diameter, DS = 13
-- Head thickness, tH = 08
-- Spacing, SHS =150
All dimensions are in mm

50
(a) Side view

(b) Section A-A

Fig. 5.2 Details of facing reinforcement.

51
SECTION 6: FIELD PULLOUT TESTING OF SOIL NAILS

6.1 INTRODUCTION

Field pullout testing of soil nails shall be conducted (a) to verify that the nail design loads

can be carried without excessive movements and with an adequate safety factor for the

service life of the structure, and (b) to verify the adequacy of the contractor’s drilling,

installation, and grouting operations prior to and during construction of production soil

nails.

6.2 FIELD PULLOUT TEST APPARATUS

A center-hole hydraulic jack and hydraulic pump shall be used to apply a test load to a

nail bar. The axis of the jack and the axis of the nail must be aligned to ensure uniform

loading. Typically, a jacking frame or reaction block is installed between the shotcrete or

excavation face and the jack. The jacking frame should not react directly against the nail

grout column during testing. Once the jack is centered and aligned, an alignment load

should be applied to the jack to secure the equipment and minimize the slack in the set-

up. The alignment load should not be permitted to exceed 5 percent of the maximum test

load.

Movement of the nail head should be measured with at least one, and preferably

two, dial gauges mounted on a tripod or fixed to a rigid support that is independent of the

jacking set-up and wall. The use of two dial gauges provides: (1) an average reading in

case the loading is slightly eccentric due to imperfect alignment of the jack and the nail

bar, and (2) a backup if one gauge malfunctions. The dial gauges should be aligned within

5 degrees of the axis of the nail, and should be zeroed after the alignment load has been

applied. The dial gauges should be capable of measuring to the nearest 0.02 mm. The dial

52
gauges should be able to accommodate a minimum travel equivalent to the estimated

elastic elongation of the test nail at the maximum test load plus 25 mm, or at least 50 mm.

A hydraulic jack is used to apply load to the nail bar while, a pressure gauge is

used to measure the applied load. A center-hole load cell may be added in series with the

jack for use during creep tests. For extended load hold periods, load cells are used as a

means to monitor a constant applied load while the hydraulic jack pump is incrementally

adjusted. Over extended periods of time, any load loss in the jack will not be reflected

with sufficient accuracy using a pressure gauge. Recent calibration data for the jack,

pressure gauge, and load cell must be obtained from the contractor prior to testing. Fig.

6.1 shows schematically a test set up for field testing of soil nails.

Fig. 6.1 Field pullout test set up (FHWA 2003).

6.3 TYPES OF FIELD PULLOUT TESTS

Depending upon the type of test being performed, the maximum test load, the load

increments, and the time that each load increment is held shall be determined. To prevent

chances of explosive failure of the steel, in no case, the soil nail tendon be stressed to

53
more than 80 percent of its minimum ultimate tensile strength for grade Fe415 steel, or

more than 90 percent of the minimum yield strength for grade Fe500 steel.

6.3.1 Verification test

A verification test on soil nail is performed: (a) to determine the ultimate bond capacity

(if carried to pullout failure); (b) verify the design bond factor of safety, and (c) to

determine the soil nail load at which excessive creep occurs. In general, the maximum

verification test load shall verify 200 percent of the design bond capacity to ensure a

minimum factor of safety of 2.0 against pullout failure of soil nails. Verification tests are

generally conducted on non-production “sacrificial” nails as a first order of work prior to

construction. Verification test when conducted to failure is known as ultimate test.

Verification test nails shall have both bonded and (temporary) unbonded lengths i.e. nail

length without grout cover.

6.3.2 Proof test

A proof test is typically performed on a specified number of the total number of

production soil nails installed. Typically, successful proof tests shall be performed on 5

percent of the production nails in each row or a minimum of 1 test per row. This test is a

single cycle test in which the load is applied in increments until a maximum test load,

typically 125 to 150 percent of the design bond capacity, is achieved. Proof tests provide

information necessary to evaluate the ability of production soil nails to safely withstand

design loads without excessive structural movement or long-term creep over the

structure’s service life.

6.3.3 Creep Test

Creep tests are typically performed as part of a verification or proof test. Creep testing is

conducted at a specified, constant test load, with movement recorded at specified time

54
intervals. The deflection-versus-log-time results are plotted on a semi-log graph, and are

compared with the acceptance criteria specified in the contract documents.

6.4 TEST PROCEDURE

6.4.1 Procedure for verification test

(a) Calculate design test load as:

The design test load DTL shall be determined by the following equation:

= L B × Qall
DTL (6.1)

where:

LB = soil nail bonded length; which shall be not be less than 3 m and more than

(LB)max (so as to ensure that the nail load does not exceed 80 percent of the allowable

nail bar tensile strength during verification test) given by Eq. (6.2).

0.8 × A t × f y
( L B )max = (6.2)
Qall × ( FST )ver

where: At = nail bar cross-sectional area; fy = nail bar yield strength; Qall = allowable

pullout resistance per unit length (Qall = Qu /FSP); ( FST )ver = factor of safety against

tensile failure of nail during verification tests (generally 2.5 or 3.0).

Note: The unbonded length of test nail shall be at least 1 m.

(b) The verification test shall be conducted by measuring and recording the incremental

test load applied to the verification soil nail and the movement of the soil nail head at

each load increment. Verification test nail may be loaded to failure or a maximum test

load of 300 percent of the DTL in accordance with the loading schedule shown in

Table 6.1.

(c) Each increment of load shall be shall be held for at least 10 minutes. Monitor the

verification test nail for creep at the 1.50DTL load increment. Measure and record

55
nail movements during creep portion of the test at 1 minute, 2, 3, 4, 5, 6, 10, 20, 30,

40, 50, and 60 minutes. Maintain the load during the creep test within 2 percent of the

intended load by use of the load cell.

Table 6.1 Loading schedule for verification test.


Test load Hold time
0.05 DTL max. (AL) 1 minute
0.25 DTL 10 minutes
0.50 DTL 10 minutes
0.75 DTL 10 minutes
1.00 DTL 10 minutes
1.25 DTL 60 minutes
1.50 DTL (Creep test) 10 minutes
1.75 DTL 10 minutes
2.00 DTL 10 minutes
2.50 DTL 10 minutes max.
3.00 DTL or Failure 10 minutes max.
0.05 DTL max. (AL) 1 minute (record permanent set)
Note: The alignment load AL should not exceed 5 percent of DTL and dial gages should be set to
zero after the application of AL.

6.4.2 Procedure for proof test

(a) Calculate design test load DTL using Eqs. (6.1) and (6.2) by adopting factor of safety

against tensile failure of nail for proof production test equal to 1.5.

Note: Production proof test nails shorter than 4 m in length may be tested with bond

length less than 3 m.

(b) Perform proof tests by incrementally loading the proof test nail to 150 percent of the

DTL in accordance with the loading schedule shown in Table 6.2. Record the soil nail

movements at each load increment.

(c) The creep period shall start as soon as the maximum test load 1.50 DTL is applied and

the nail movement shall be measured and recorded at 1 minute, 2, 3, 5, 6, and 10

minutes. Where the nail movement between 1 minute and 10 minutes exceeds 1 mm,

maintain the maximum test load for an additional 50 minutes and record movements

56
at 20 minutes, 30, 50, and 60 minutes. Maintain all load increments during creep test

period within 5 percent of the intended load.

Table 6.2 Loading schedule for proof test.


Test load Hold time
0.05 DTL max. (AL) Until movement stabilizes
0.25 DTL Until movement stabilizes
0.50 DTL Until movement stabilizes
0.75 DTL Until movement stabilizes
1.00 DTL Until movement stabilizes
1.25 DTL Until movement stabilizes
1.50 DTL (max. test load) Start creep test
Note: The alignment load AL should not exceed 5 percent of DTL and dial gages should be set to
zero after the application of AL.

6.5 TEST ACCEPTANCE CRITERIA

Preliminary requirement of field pullout testing is that the both test nails and the

production nails shall have the same method of installation, the soil/rock conditions,

equipment, and the operator. Moreover, testing should be completed in each row of nails

prior to excavation and installation of the underlying row. If inadequate test results

indicate faulty construction practice or bond capacities less than required, the contractor

should be required to alter nail installation/construction methods. In the event that

required design adhesion capacities are still not achievable, redesign may be necessary.

6.5.1 Acceptance criteria for verification test

The verification test acceptance criteria require that:

(a) no pullout failure occurs at 200 percent of the design load where pullout failure is

defined as the inability to maintain constant test load without excessive movement;

and

(b) the total measured movement ∆L at the test load of 200 percent of design load must

exceed 80 percent of the theoretical elastic movement of the unbonded length UL i.e.

57
P × UL
∆L min =
0.8 (6.3)
EA

where: ∆L min = minimum acceptable movement; P = maximum applied test load; UL

= unbonded length of test nail (measure from the back of reference plate to top of the

grouted length); A = cross-sectional area of the nail bar; and E = Young’s modulus of

steel (typically 200 GPa).

This criterion ensures that load transfer from the soil nail to the soil occurs only in the

bonded length and not in the unbonded length.

6.5.2 Acceptance criteria for proof test

The acceptance criteria for proof test require that no pullout failure occurs and that the

total movement at the maximum test load of 150 percent of design load must exceed 80

percent of the theoretical elastic movement of the unbonded length. Again, the measured

movement must be ΔL ≥ ΔLmin, where ΔLmin is as defined in Eq. (6.3).

6.5.3 Acceptance criteria for creep test

For verification tests, the total creep movement should be less than 2 mm between the 6-

and 60-minute readings and the creep rate should show linear or decreasing trend

throughout the creep test load hold period. For proof tests, the total creep movement

should be less than 1 mm during the 10-minute readings or the total creep movement

should be less than 2 mm during the 60-minute readings and the creep rate should show

linear or decreasing trend throughout the creep test load hold period.

6.6 TYPICAL TEST DATA SHEET

Fig. 6.2 shows the typical format for soil nail field pullout test data sheet.

58
SOIL NAIL TEST DATA SHEET

Project…………………………………………… Length………………………………………
Project No……………………………………….. Bonded length………………………………
Station…………………………………………… Unbonded length……………………………
Nail No…………………………………………… Drillhole diameter…………………………
Date……………………………………………… Tendon diameter……………………………
Field Inspector…………………………………… Tendon grade………………………………..
Test type Verification …… Ultimate ……... Proof……..

Load Movement [mm]


Time Load Load
Dial Dial Average Remarks
[minutes] increment increment
gage 1 gage 2 dial gage
[%] [kN]

Fig. 6.2 Typical format of soil nail field pullout test data sheet.

59
APPENDIX A

GENERAL DESIGN TABLES

Table A.1
Minimum Recommended Factors of Safety for the Design of Soil Nail
Structures using the ASD Method

Factors of safety
Failure Static Seismic
Resisting component Symbol (Temporary
mode Temporary Permanent and permanent
structures structures structures)
Global stability FSG 1.35 1.5 1.1
External
Sliding stability FSSL 1.3 1.5 1.1
stability
Bearing capacity FSH 2.5(1) 3.0(1) 2.3(1)

Internal Pullout resistance FSP 2.0 2.0 1.5


stability Nail bar tensile strength FST 1.8 1.8 1.35

Facing flexure FSFF 1.35 1.5 1.1


Facing
Facing punching shear FSFP 1.35 1.5 1.1
strength
Headed-stud tensile FSHT 1.5-1.8 1.7-2.0 1.3-1.5
Source: FHWA (2003)

Note:
(1) The safety factors for bearing capacity are applicable when using standard bearing-capacity
equations. When using stability analysis programs to evaluate this failure mode, the factors of
safety for global stability apply.

60
Table A.2
Estimated Bond Strength of Soil Nails in Soil and Rock

Construction Ultimate bond


Material Soil/rock type
method strength, qu [kPa]
Marl/limestone 300 -400
Phyllite 100 -300
Chalk 500 -600
Soft dolomite 400 -600
Rock Rotary drilled Fissured dolomite 600 - 1000
Weathered sandstone 200 -300
Weathered shale 100 -150
Weathered schist 100 -175
Basalt 500 -600
Slate/hard shale 300 -400
Sand/gravel 100 - 180
Silty sand 100 - 150
Rotary drilled Silt 60 - 75
Piedmont residual 40 - 120
Fine colluvium 75 - 150
Sand/gravel
low overburden 190 - 240
Cohesionless soils Driven casing high overburden 280 - 430
Dense moraine 380 - 480
Colluvium 100 - 180
Silty sand fill 20 - 40
Augered Silty fine sand 55 - 90
Silty clayey sand 60 - 140
Sand 380
Jet grouted
Sand/gravel 700
Rotary drilled Silty clay 35 -50
Driven casing Clayey silt 90 -140
Loess 25 - 75
Fine-grained soils Soft clay 20 - 30
Augered Stiff clay 40 - 60
Stiff clayey silt 40 - 100
Calcareous sandy clay 90 - 140
Source: FHWA (2003)

61
Table A.3
Welded Wire Mesh Dimensions

Mesh designation(1), (2) Wire cross-sectional area per Weight per unit area
(mm x mm –mm2 x mm2) unit length(3) (mm2/m) (kg/m2)
102x102 - MW9xMW9 88.9 1.51
102x102 -MW13xMW13 127.0 2.15
102x102 -MW19xMW19 184.2 3.03
102x102 -MW26xMW26 254.0 4.30
152x152 - MW9xMW9 59.3 1.03
152x152 -MW13xMW13 84.7 1.46
152x152 -MW19xMW19 122.8 2.05
152x152 -MW26xMW26 169.4 2.83
Source: FHWA (2003)
Notes:
(1) The first two numbers indicate the mesh opening size, whereas the second pair of numbers
following the prefixes indicates the wire cross-sectional area.
(2) Prefix M indicates metric units. Prefix W indicates plain wire. If wires are pre-deformed, the
prefix D shall be used instead of W.
(3) This value is obtained by dividing the wire cross-sectional area by the mesh opening size.

62
Table A.4
Area of Reinforcement Bars at given Spacings
(Values in cm2 per Meter width)

Spacing Reinforcement bar diameter (mm)


(cm) 6 8 10 12 14 16 18 20 22 25 28 32

5 5.65 10.05 15.71 22.62 30.79 40.21 50.89 62.83 76.03 98.17 123.15 160.85
6 4.71 8.38 13.09 18.85 25.66 33.51 42.41 52.36 63.36 81.81 102.63 134.04

7 4.04 7.18 11.22 16.16 21.99 28.72 36.35 44.88 54.30 70.12 87.96 114.89

8 3.53 6.28 9.82 14.14 19.24 25.13 31.81 39.27 47.52 61.36 76.97 100.53

9 3.14 5.59 8.73 12.57 17.10 22.34 28.27 34.91 42.24 54.54 68.42 89.36

10 2.83 5.03 7.85 11.31 15.39 20.11 25.45 31.42 38.01 49.09 61.58 80.42

11 2.57 4.57 7.14 10.28 13.99 18.28 23.13 28.56 34.56 44.62 55.98 73.11

12 2.36 4.19 6.54 9.42 12.83 16.76 21.21 26.18 31.68 40.91 51.31 67.02

13 2.17 3.87 6.04 8.70 11.84 15.47 19.57 24.17 29.24 37.76 47.37 61.87

14 2.02 3.59 5.61 8.08 11.00 14.36 18.18 22.44 27.15 35.06 43.98 57.45

15 1.88 3.35 5.24 7.54 10.26 13.40 16.96 20.94 25.34 32.72 41.05 53.62

16 1.77 3.14 4.91 7.07 9.62 12.57 15.90 19.63 23.76 30.68 38.48 50.27

17 1.66 2.96 4.62 6.65 9.06 11.83 14.97 18.48 22.36 28.87 36.22 47.31

18 1.57 2.79 4.36 6.28 8.55 11.17 14.14 17.45 21.12 27.27 34.21 44.68

19 1.49 2.65 4.13 5.95 8.10 10.58 13.39 16.53 20.01 25.84 32.41 42.33

20 1.41 2.51 3.93 5.65 7.70 10.05 12.72 15.71 19.01 24.54 30.79 40.21

21 1.35 2.39 3.74 5.39 7.33 9.57 12.12 14.96 18.10 23.37 29.32 38.30

22 1.29 2.28 3.57 5.14 7.00 9.14 11.57 14.28 17.28 22.31 27.99 36.56

23 1.23 2.19 3.41 4.92 6.69 8.74 11.06 13.66 16.53 21.34 26.77 34.97

24 1.18 2.09 3.27 4.71 6.41 8.38 10.60 13.09 15.84 20.45 25.66 33.51

25 1.13 2.01 3.14 4.52 6.16 8.04 10.18 12.57 15.21 19.63 24.63 32.17

26 1.09 1.93 3.02 4.35 5.92 7.73 9.79 12.08 14.62 18.88 23.68 30.93

28 1.01 1.80 2.80 4.04 5.50 7.18 9.09 11.22 13.58 17.53 21.99 28.72

29 0.97 1.73 2.71 3.90 5.31 6.93 8.77 10.83 13.11 16.93 21.23 27.73

30 0.94 1.68 2.62 3.77 5.13 6.70 8.48 10.47 12.67 16.36 20.53 26.81

32 0.88 1.57 2.45 3.53 4.81 6.28 7.95 9.82 11.88 15.34 19.24 25.13

34 0.83 1.48 2.31 3.33 4.53 5.91 7.48 9.24 11.18 14.44 18.11 23.65

36 0.79 1.40 2.18 3.14 4.28 5.59 7.07 8.73 10.56 13.64 17.10 22.34

38 0.74 1.32 2.07 2.98 4.05 5.29 6.70 8.27 10.00 12.92 16.20 21.16
40 0.71 1.26 1.96 2.83 3.85 5.03 6.36 7.85 9.50 12.27 15.39 20.11

63
Table A.5
Headed-Stud Dimensions

Headed-Stud Nominal Length Head Diameter Shaft Diameter Head Thickness


Size Ls (mm) DH (mm) DS (mm) tH (mm)
1 ×4 1 105 12.7 6.4 4.7
4 8
3 ×4 1 105 19.1 9.7 7.1
8 8
3 ×6 1 156 19.1 9.7 7.1
8 8
1 ×4 1 105 25.4 12.7 7.9
2 8
1 ×5 5 135 25.4 12.7 7.9
2 16
1 ×6 1 156 25.4 12.7 7.9
2 8
5 ×69 162 31.8 15.9 7.9
8 16
3 × 311 89 31.8 19.1 9.5
4 16
3 ×43 106 31.8 19.1 9.5
4 16
3 ×5 3 132 31.8 19.1 9.5
4 16
3 ×6 3 157 31.8 19.1 9.5
4 16
7 ×43 102 34.9 22.2 9.5
8 16
7 ×5 3 127 34.9 22.2 9.5
8 16
7 ×6 3 152 34.9 22.2 9.5
8 16

Source: FHWA (2003)

Note: Locally available headed-studs may be used in-lieu of those mentioned in above table.

Reference figure for dimensions of Headed-stud

64
APPENDIX B: TYPICAL SOIL NAILING PROJECT SUMMARY

Project No………. Report Date…………………

Project Title…………………………………………………………………………………………………………………………………

Project Site………………………………………………………………………………………………………………………………….

Client………………………………………………………………. Contractor………………………………………………………......

Type of structure……………………………………………………………………………………………………………………………

Purpose of structure………………………………………………………………………………………………………………………...

Duration of project……………………………….. (Start date…………………………… Completion Date………………………...)

Attachments

Sl.No. Item Yes No Remark (s), if any

1 Soil testing report

2 Field pullout test report

3 Photographs

4 Other (if any)

5 Other (if any)

65
TABLE B.1 TECHNICAL INFORMATION
Parameter Symbol Unit Value Remark (s), if any

Wall Layout
Vertical height of soil nail wall H m

Maximum longitudinal stretch L m

Face batter (wrt vertical) α Degrees

Slope of backfill β Degrees

Surcharge (if any) qs kPa

Soil Properties (adopted from soil testing data)


Cohesion c kPa

Friction angle φ Degrees

Unit weight γ kN/m3

Nail (or Reinforcement) Properties


Type

Installation method

Steel grade

66
Reinforcement or Nail Properties Continued
Length LN m

Nail diameter D mm

Drill hole diameter DDH mm

Nail spacing (vertical x horizontal) SV x SH mxm

Nail inclination (wrt horizontal) i Degrees

Compressive strength of grout fck MPa

TABLE B.2 WALL FACING COMPONENTS


Element Description Temporary Facing Permanent Facing

Thickness (mm)

General Facing type

Concrete grade

Type

Reinforcement Steel grade

Size / Denomination

67
Type

Bearing Plate Steel

Dimensions

Other Information (s), If any

TABLE B.3 FIELD PULLOUT TEST SUMMARY


Parameter Test 1 Test 2 Test 3 Remark (s), if any

Depth of test nail (m)

Maximum pullout load (kN)

Maximum nail displacement (mm)

Prepared by

Name………………………………………………………..

Signature……………………. …………………………….. Date………………………………………….

Designation…………………………………………………. Place…………………………………………

68
Indian Geotechnical Society
TC on Ground Improvement and
Geosynthetics
Technical Note on Ground
Improvement using Deep Soil
Mixing: Theory & Practice
-by Keller Ground Engineering India Pvt. Ltd

Revision Details:

00 10-11-2016 For Circulation Vani Madan Hari


Init. Sign. Init. Sign. Init. Sign.
Rev. Date Document Details
Prepared Checked Approved
TABLE OF CONTENTS
1 Background .....................................................................................................................................1
1.1 General .....................................................................................................................................1
1.2 Technical Committee ................................................................................................................1
1.3 Brainstorming Session ..............................................................................................................1
2 Deep Soil Mixing (DSM) .................................................................................................................1
2.1 Introduction ...............................................................................................................................1
2.2 Design Philosophy ....................................................................................................................3
2.2.1 Mix Design of Grout Slurry..............................................................................................4
2.2.2 Process Design ...............................................................................................................4
2.3 Application and Process of DSM ..............................................................................................4
2.3.1 Bearing Capacity Improvement & Settlement Control (Wet DSM) .................................5
2.3.2 Bending Rigidity for Excavation Support (Gravity & CDSM- Compound DSM) .............7
2.3.3 General Soil improvement for shallow depth (Mass stabilisation) ..................................8
2.3.4 Quality control and testing of DSM .................................................................................9
3 Case studies ................................................................................................................................. 11
3.1 Improvement of Bearing Capacity & Settlement Control (Wet DSM) .................................... 11
3.1.1 Case study 1: Optimized Foundation for 6 Storey Condominium, Singapore. ............ 11
3.1.2 Case study 2: Optimised Foundation Systems for Jelutong Sewage Treatment Plant in
Penang Island, Malaysia. ....................................................................................................... 12
3.2 Bending rigidity for excavation support (Gravity wall and CDSM) ......................................... 13
3.2.1 Case Study 1: Black soil Strengthening of Slime for Deep Excavation at Bunus
Sewerage treatment plant, Kuala Lumpur, Malaysia ............................................................. 13
3.2.2 Case Study 2: Deep Excavation support for Basement of 3-storey commercial
complex with 2-level basement, Kuala Lumpur, Malaysia. .................................................... 14
3.2.3 Case Study 3: Deep Excavation support for Music Academy Poznan, Poland. ......... 15
3.2.4 Case Study 4: Excavation Support for TBM Retrieval Shaft using Deep Soil Mixing
Technique, Kuala Lumpur ...................................................................................................... 16
3.3 Ground treatment for shallow depth ...................................................................................... 18
3.3.1 Case Study 1: Mass Stabilisation of Peat and Mud for a Road Embankment
Gärtunavägen, Sodertalje, Sweden. ...................................................................................... 18
4 Bibliography ................................................................................................................................. 19

LIST OF FIGURES
Figure 1: Basic concept of Deep Soil Mixing (Source: Broms, 2004) ......................................................2
Figure 2: Single and Multi-shaft mixing ....................................................................................................2
Figure 3: Soil mixing patterns in DSM ......................................................................................................3
Figure 4: Geomechanical design philosophy for deep stabilisation .........................................................3
Figure 5: Wet DSM machine & Ejection of the binder (wet form) form the machine ...............................5
Figure 6: Mixing operations in Wet DSM process ....................................................................................6
Figure 7 Overall process in Wet DSM ......................................................................................................6
Figure 8: Typical cross section of gravity DSM wall .................................................................................7
Figure 9: Typical arrangements of DSM columns ....................................................................................8
Figure 10: Process of installing Steel H sections & Overall view of CDSM arrangements ......................8
Figure 11: Types of mixing in Shallow depth ...........................................................................................9
Figure 12 Typical computer record of the Wet DSM ............................................................................. 10
Figure 13 Typical testing methods (Unconfined compressive strength testing on backflow samples) . 10
Figure 14 Visual examination of exposed columns by means of excavation ....................................... 11
Figure 15: Pressure vs settlement curve & Completed view of building ............................................... 11
Figure 16: Plan layout and Proposed foundation system of SBR ......................................................... 12
Figure 17: Settlement monitoring & UCS test results of SBR ............................................................... 12
Figure 18: Completed SBR tanks during hydro test .............................................................................. 13
Figure 19: Typical plan & cross section of Treatment scheme and SPT results of Digester tank ........ 14
Figure 20 Typical exposed DSM and Structure after DSM installation ................................................. 14
Figure 21: Schematic of DSM gravity wall block. .................................................................................. 15

Page i
Figure 22 Completed excavation........................................................................................................... 15
Figure 23: Typical CDSM scheme and exposed CDSM columns ......................................................... 16
Figure 24: Completed structure of Music Academy Poznan ................................................................. 16
Figure 25 Typical cross section of the tunnel ........................................................................................ 17
Figure 26: DSM scheme and exposed columns after excavation ......................................................... 17
Figure 27: DSM Core Sample UCS Strength Test Results ................................................................... 17
Figure 28 Typical cross section of Mass stabilisation ........................................................................... 19

LIST OF TABLES
No table of figures entries found.
LIST OF ENCLOSURES

Annexure I Technical paper on “Soil mixing - Challenges of Application ranging


from Ground Improvement to Structural Elements” by Topolnicki M,
2006.
Annexure II Technical paper on “Quality Control of Wet Deep Soil Mixing with
reference to Polish Practices and Applications”.
Annexure III Technical paper on the project case study on Deep Soil Mixing in Mine
tailing for 8m Deep Excavation”, Y. W Yee, V.R. Raju, H K Yandamuri,
Kuala Lumpur
Annexure IV Technical paper on the project case study on “Excavation Support for
TBM Retrieval Shaft using Deep Soil Mixing Technique, Kuala Lumpur”,
Ir. Yew Weng, Yee and Ir. Yean Chin &Tan.

Page ii
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice

1 Background
1.1 General
The President Indian Geotechnical Society (IGS) has constituted several Technical
Committees (TCs) in order to contribute substantial technical innovations to serve the
geotechnical community by publishing guidelines in the field of ground engineering. In this
endeavour, IGS formed various TCs to seek support in the preparation of guidelines and
publish them on behalf of IGS. In order to form the guidelines, the modus operandi
suggested by IGS was to conduct brain-storming sessions in local chapters in each of the
selected themes and topics and further to record the proceedings. Each member of the
committee shall have to make a presentation followed by a detailed discussion. The
chairman of each TC will decide the sub-topics on which the theme paper will be presented
by a particular member of the committee, followed by a thorough discussion. The individual
TC will develop guidelines with regard to various fields of Geotechnology on behalf of IGS
who will contribute in a meaningful way to better geotechnical practices in India.

1.2 Technical Committee


With the above background, IGS has identified Ground Improvement and Geosynthetics is
one of the TC and the main objective is to prepare an implementable document for practicing
engineers covering Ground Improvement technology, limitations, codal provisions, case
histories esp. in India with their performances.

1.3 Brainstorming Session


IGS Hyderabad Chapter has taken initiative to support IGS and conducted one-day National
Workshop on Ground Improvement and Geosynthetics on 29th August 2015 in JNTU
premises. Minutes of meeting was prepared and circulated among the TC members. It was
agreed in the meeting that design and construction aspects of ground improvement using
Deep Soil Mixing (DSM) shall be addressed by Keller Ground Engineering Pvt. Ltd (Keller).
This document describes concept, theory, design & construction, performance of ground
improvement using Deep Soil Mixing for variety of projects executed in Asia.

2 Deep Soil Mixing (DSM)


2.1 Introduction
Ground improvement using deep soil mixing is accepted world-wide in order to enhance the
engineering properties such as increase strength, lower permeability and reduced
compressibility of soils. The experiences have been positive and deep soil mixing methods
are undergoing rapid development, particularly with regard to the explicabilities and cost
effectiveness.
The roots of Deep Soil Mixing (DSM) go back to the mid-1950s, when the Mixed in Place
(MIP) piling technique was developed. In this method a mechanical mixer was used to mix
cementitious grout such as lime, cement or a combination of both in different proportion into
the soil for the purpose of creating foundation elements and retaining walls. The grout was
injected from the tip of the mixing tool consisting of a drilling head and separated horizontal
blades. The technique forms columns within the treated zone whilst improving shear and
compressibility parameters of the in-situ loose/soft soils. The basic concept of Deep Soil
Mixing is illustrated in Figure 1.

1
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice

Figure 1: Basic concept of Deep Soil Mixing (Source: Broms, 2004)


In general, single or multi shaft are used for mixing process that needs primarily cement-
based slurries to create isolated elements, continuous walls or blocks. The different mixing
process are shown in Figure 2

Figure 2: Single and Multi-shaft mixing


Depending on the purpose of deep mixing works, specific condition of the site, stability
calculations and costs of treatment, different patterns of column installations are used to
achieve desired results by utilising spaced or overlapping & single or combined columns.
Typical patterns of DSM in practice are shown in Figure 3.

2
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice

Figure 3: Soil mixing patterns in DSM


DSM method is best suited in cohesive soil with high moisture contents & loose, saturated &
fine granular soils. It is also used in less cohesive soils, but not feasible in very dense/ stiff
materials/ in ground with obstructions characteristics such as cobbles, boulders etc.

2.2 Design Philosophy


The design philosophy for DSM is to produce a relatively rigid and high strength pile-like
column that mechanically interacts with the surrounding natural soil. The applied load is
partly carried by the columns and partly by the unstabilised soil between the columns.
Therefore, a too stiffly stabilised material is not necessarily the best solution since such a
material will behave like a pile. Instead, the increased stiffness and strength of the stabilised
soil should not prevent an effective interaction and load distribution between the stabilised
and natural soil. This philosophy is schematically described in Figure 4.

Figure 4: Geomechanical design philosophy for deep stabilisation


To compare various column patterns in terms of the treatment area and to evaluate
composite properties of the treated elements and surrounding untreated soil, a purposely
defined ratio of area improvement (ap) is used
Net area of soil mixing
ap =
Respective total area
For group of columns, to avoid too risky designs with low ap values and high strength
elements, it has been generally recommended that the width of the improved ground should

3
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice
be larger than the thickness of soft soil. Depending upon the application, the DSM are
frequently designed with area improvement ratios between 15% to 50 %.

2.2.1 Mix Design of Grout Slurry

The grout slurry has to be designed after evaluating the natural moisture content in the in-situ
soil, desired viscosity / flow rate of grout slurry to suit available equipment and curing period
for hydration process. Typically, water-cement (W/C) ratio ranges between 0.8 and 1.2 with
grout slurry densities in the range of 1.4 to 1.6 t/m3.

2.2.2 Process Design

The objective of process design is to achieve efficient mixing of grout slurry and in-situ soil
whilst delivering design quantities of grout slurry per cubic meter of soil. The process
operating parameters has to be designed to achieve design binder content and also to suit
available equipment (i.e. torque of base machine, maximum rotation speed of mixing tool,
type of grout pump, achievable maximum flow rates and flow pressures, diameter and
position of nozzles, no. of nozzles and mixing blades, no. of cycles etc.).

Based on the process operating parameters, the blade rotation number (T) is defined as the
total number of effective mixing/cutting blades passing during 1m of single shaft movement
through the soil (Topolnicki, 2004), and is expressed as follows,

 R p Rw 
T  M    
V 
 p Vw 
Where,
M is total number of effective mixing/cutting blades;
Rp is rotation speed during penetration;
Vp is rate of penetration;
Rw is rotation speed during withdrawal;
Vw is rate of withdrawal.

2.3 Application and Process of DSM


The Deep Soil Mixing techniques can be adopted for the following ground engineering
applications
 Enhances the bearing capacity of weak and problematic soil types such as loose
sands, soft marine clays, ultra-soft slimes, weak silty clays & sandy silts.
 Control Settlements
 Foundations of embankment fill for highways, railways & runways.
 Slope stabilisation, stabilisation of cuts and excavations.
 Excavation support walls.
 Ground treatment.
 Hydraulic cut-off walls.
 In-situ reinforcement piles and gravity walls.
 Environmental remediation.
Technical paper on “Soil mixing- the challenges of application ranging from ground
improvement to structural elements” is attached in Annexure I.

4
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice
Different methods of DSM used for the above applications are listed below
 Wet Deep Soil Mixing
 Gravity Wall/Compound Deep Soil Mixing (CDSM - I Sections)
 Mass Stabilisation

2.3.1 Bearing Capacity Improvement & Settlement Control (Wet DSM)

Wet DSM is used to enhance the bearing capacity and to reduce the settlement of the soft
soils. A typical wet DSM unit consists of an installation rig fitted with a mast and a drill motor
as shown in Figure 5. The wet binder is mixed separately at a mixing plant (consists of high
speed colloidal mixers and agitators) and transported to the rig point in a slurry form with
desired flow rate and grout pressure using custom-built pumps. The grout slurry is delivered
in the ground through nozzles located below mixing blades.

Mixing is achieved by using an auger-mixing tool connected to the drill motor by a kelly bar.
The mixing tool is drilled down to firm ground or intended depth whilst delivering and mixing
of required quantity of grout slurry with in-situ soil uniformly. Once at the required depth, the
tool is drilled out with the simultaneous injection of grout slurry. Depending on the process
design, one or two cycles of mixing operations can be executed. The rotation speed during
penetration and withdrawal, rate of penetration and withdrawal, flow rate of grout slurry and
grout pressure are adjusted such that the desired amount of grout slurry is thoroughly mixed
with the in-situ soil. The rotation speed ranges between 40 and 70 rpm; the rate of
penetration ranges between 0.6 and 1.2 m/min and rate of withdrawal ranges between 2 and
3 m/min. Typically, the flow rate of grout slurry ranges between 60 and 120 lit/min with grout
pressures of 30 to 60 bars.

Figure 5: Wet DSM machine & Ejection of the binder (wet form) form the machine

The process of efficient mixing of grout slurry with in-situ soil is shown in Figure 6. The
amount of slurry injected into the ground is continuously monitored by flow sensors to verify
whether the required amount of grout slurry has actually been utilised uniformly over the
length of the column.

5
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice

Figure 6: Mixing operations in Wet DSM process

The overall Wet DSM process is illustrated in Figure 7. The execution method of the soil
mixing process is laid down in the European Standard prEN14679 (2003).

Figure 7: Overall process in Wet DSM

The amount of cement in the grout slurry is usually in the range of 250 to 400 kg/m3 of soil.
The final result of the Wet DSM process is a relatively rigid soil mass in the shape of a
cylindrical column with improved deformation and shear strength characteristics. Typically,
undrained shear strength of the columns ranges between 500 and 2000 kPa. The rate of
strength gain is dependent on soil conditions but typically requires about 2 to 4 weeks of
minimum curing period. Typical column diameter ranges from 0.8m to 1.5m and maximum
treatment depth from 15m to 20m.

6
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice
2.3.2 Bending Rigidity for Excavation Support (Gravity & CDSM- Compound DSM)
Gravity type structures are subjected to large horizontal forces caused by earth pressure. For
this type of structures DSM techniques that utilize an array of soil/cement columns arranged
in the native soil in a manner that creates a gravity block. This block resists the sliding and
overturning loads and thus eliminated the need for bracing or tieback anchors for stability. All
the support elements are placed prior to any excavation, thus speeding the construction
schedule. Because it is a gravity block, no steel reinforcement is needed. Typical gravity wall
cross section is shown Figure 8.

Figure 8: Typical cross section of gravity DSM wall


Retention systems comprise applications associated with restraining the earth pressure
mobilised during deep excavations and vertical cuts in soft ground, with protection of
structures surrounding excavations measures against base heave and prevention of
landslides and slope failures. In these applications blocks wall column patters are mainly
used while the soil binder mix is typically engineered to have strength and stiffness. To
overcome soil and water lateral pressures the DSM columns should have adequate internal
shear.
Steel pipes or H beams can be installed in DSM columns executed with the wet method to
increase the bending resistance and create a structural wall for excavation support.
Elongated mixing time and /or full restroking are usually applied to ensure easier installation
of soldier elements immediately after mixing. Panels of mixed soil between H beam
reinforcement are designed to work in arching. Concrete facing, tieback anchors or stage
struts are typically used in combination with the DM walls. Drainage media may be required
behind the wall to prevent build up excess hydrostatic pressures. Deep circular shafts can be
constructed using 2 to 3 concentric rings of overlapping DSM columns acting together in
hoop compression.
The typical arrangement of DSM excavation support walls is illustrated in Figure 9. Steel
reinforcement is installed in every other DSM column after mixing and the accessible face of
each column is trimmed off once the excavation is complete. The wall system is supported
with at least one level of struts or anchors and walers for horizontal support. It is common for
DSM excavation systems to have multiple levels of support.

7
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice

Figure 9: Typical arrangements of DSM columns


The use of CDSM is to produce blocks of over lapping piles for nailed or self supporting
gravity retaining structures. Steel H sections are installed as structural reinforcement in
retaining walls prior to the hardening of the soil-cement mixture. The soil-cement is designed
to arch between adjacent steel H sections. The process of installing steel H sections &
overall view of CDSM arrangements is shown in Figure 10.

Figure 10: Process of installing Steel H sections & Overall view of CDSM
arrangements

2.3.3 General Soil improvement for shallow depth (Mass stabilisation)


The Complementary Shallow Mixing Method (SMM) has been specially developed to reduce
the costs of improving loose or soft superficial soils overlying substantial areas, including
land disposed dredged sediments and wet organic soils a few meters thick. It is also a
suitable method for in-situ remediation off contamination soils and sludges. In such
applications, the soils have to be thoroughly mixed in-situ with an appropriate amount to
binders to ensure stabilisation of the entire volume of treated soil. Therefore, this type of soils
mixing is often referred to as mass stabilisation. Mass stabilisation can be used to a depth of
about 5 meters.

8
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice
In this method special mixing tools are used, which are in most cases fixed to an excavator’s
rig arm. Mixing is executed vertically or horizontally, with mixing tools that resemble screw
propellers having a centrally provided nozzle for binder. The binder is fed from a separate
unit which houses the pressurised binder container, compressor, air dryer and supply control
unit.
Stabilisation is executed in phases, according to the operational range of the drilling rig,
which generally comprises an area of 8 to 10 m² and depth up to 4 m. Once the required
binder volume has been applied, mixing is continued to assure the optimum mixing
properties. The different type of mixing is shown in Figure 11.

Figure 11: Types of mixing in Shallow depth


The total soil volume is stabilised in order to create a block that can carry the load of the
embankment. The binder generally consists of cement, quick lime, slag or a mixture of the
above. After mass stabilisation the stabilised block is much stiffer than the original soil, which
will reduce the settlements and improve stability. Mass-stabilisation (MS) is usually used in
combination with lime/cement (LC) columns where the top very soft (primary organic) soils
are mass-stabilised and the underlying soft clay is stabilised with lime/cement columns. The
mass-stabilisation production rate is highly depended on the type of project and the amount
of binder. Generally, the production rate varies between 300-500 m3/16 hrs shift.
Mass stabilization can be achieved by installing vertical overlapping columns with up and
down movements of rotating mixing tools, as in the case of DSM and is most cost effective
when using large diameter mixing augers or multiple shaft arrangements.

2.3.4 Quality control and testing of DSM


Quality assessment of the DSM products is regarded as one of the pressing issues
confronting the implementations of soil mixing. Quality assessment is obtained from the
installation records of the columns and from the results of appropriate laboratory and field
verification tests.

Quality control during execution is important for DSM also to ensure uniform improvement of
the soil. The mixing units are equipped with computerized recording devices to measure real
time depth of mixing tool; amount of grout slurry used; flow rate of grout; grout pressure;
rotation speed during penetration and withdrawal; and rate of penetration and withdrawal.
Typical computer record showing relevant quality control parameters is shown in the Figure
12.

9
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice

Figure 12: Typical computer record of the Wet DSM

After allowing for sufficient curing time, the columns of DSM also can be tested using single
and group column plate load tests, unconfined compressive strength on cored and backflow
samples, visual examination on exposed columns by means of excavation etc. Typical
testing methods are represented in following Figure 13. The visual examinations of the
exposed columns are shown in Figure 14. The technical paper on “Quality control of wet
deep soil mixing with reference to polish practices and applications” are enclosed in
Annexure II.

Figure 13: Typical testing methods (Unconfined compressive strength testing on


backflow samples)

10
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice

Figure 14: Visual examination of exposed columns by means of excavation

3 Case studies
3.1 Improvement of Bearing Capacity & Settlement Control (Wet DSM)

3.1.1 Case study 1: Optimized Foundation for 6 Storey Condominium, Singapore.


M/s Prestigious W- Residences at Sentosa Cove, Singapore proposed to develop a
condominium in Singapore. The project comprises of 6 storey condominiums and the
approximate area of development is about 1120 m2.
The sub soil in the project site comprises of sand for top 2m followed by soft clayey silt up to
10m.This layer was underlying by stiff clay up to 50m.The loading intensity of the proposed
structure on the soft soil is 180kPa.
Considering the project boundary conditions, DSM technique with 25 % to 35 % area
replacement ratio (Wet DSM) up to maximum of 8 m was adopted as a viable method for
subsoil improvement and a full raft foundation supported by the treated found as an
alternative foundation system.
Keeping the importance of post performance of the structure, plate load test has been
conducted on the improved ground. The graph showing settlement from plate load test and
completed structures are shown in Figure 15.
The results of post construction are shown below
 Achieved bearing capacity : > 180kPa
 Settlement : 10mm @ max 180 kPa loading intensity

Figure 15: Pressure vs settlement curve & Completed view of building

11
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice
3.1.2 Case study 2: Optimised Foundation Systems for Jelutong Sewage Treatment
Plant in Penang Island, Malaysia.
A Jelutong Sewage Treatment plant is under construction in Penang Island and when
completed will cater for an ultimate capacity of 1.2 million populations equivalent. The project
will serve as a centralized sewage treatment facility and will include 12 nos. of Sequential
Batch Reactor (SBR) tanks and associated process tanks.
The subsoil primarily consists of 3m to 5m thick reclaimed fill / domestic waste dumps
followed by 5m to 7m thick soft marine clay. This is followed by stiff to very stiff cohesive
deposits to over 50m depth. The ground water table varied between 1m and 2m below
existing ground level.
The foundation system was designed to ensure adequate bearing capacity (to support
loading intensity of 92kPa), limit the total settlement of the structure to be less than 75mm
and differential settlement to be less than 1(V):360(H).
DSM technique with 8 % to 12 % area replacement ratio (Wet DSM) up to maximum of 14 m
was adopted. The plan layout and proposed foundation system using cement mix piles at
SBR is shown in Figure 16.

Figure 16: Plan layout and Proposed foundation system of SBR


The settlements were monitored using precise survey instruments during and after Hydro
tests and UCS results are shown in Figure 17. The completed SBR tanks during hydro test is
shown in Figure 18.
The results of post construction are shown below
 Achieved bearing capacity : > 90kPa
 Settlement : 5 to 20mm (During hydro test)
Load (Tons) Results of UCS Tests on VCC Cores - Jelutong STP (Penang)
0 5 10 15 20 25 30 35 40 45 50 55 60 49.6
0 50
44.6
45

-5 40
36.5
Settlement (mm)

35 33.5

-10
30 28.0 28.5
UCS (MPa)

27.5

25 22.7 23.4
22.5
-15 21.0
20 17.3 17.9
17.0 16.8 16.2 15.4 14.7
15 12.9
-20 11.9
10

-25 5 Design UCS = 5MPa

1st Cycle (100% - 35T) 2nd Cycle (125% - 44T) 3rd Cycle (150% - 52.5T) 0
SBR 1a SBR 1b SBR 2a SBR 2b SBR 3a SBR 3b SBR 4a SBR 4b SBR 7a SBR 7b SBR 8a SBR 8b TNB

Jelutong STP Penang : VCC : SBR 4 - Column No. 1666 Structure Reference

Inclined Cores (100mm dia.) Vertical Cores (54mm dia.)

Figure 17: Settlement monitoring & UCS test results of SBR

12
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice

Figure 18: Completed SBR tanks during hydro test

3.2 Bending rigidity for excavation support (Gravity wall and CDSM)

3.2.1 Case Study 1: Black soil Strengthening of Slime for Deep Excavation at Bunus
Sewerage treatment plant, Kuala Lumpur, Malaysia
The Sewage Services Department of the Ministry of Housing and Local Government,
Malaysia proposed to construct a Sewage Treatment Plant in Kuala Lumpur. As part of the
treatment process, four (4) numbers of 23m diameter digesters were constructed at equi-
distance from each other. Each digester is essentially a 15m high tank with a coned shaped
(45 degree) base buried 8m below ground.
Being a former tin mining area, the site is underlain by highly variable soil conditions. Very
soft slime was found at the foot-prints of 3 of the digesters while loose sand with slime layers
was found at one of the digesters. The depth of limestone bedrock typically varied between
7m and 13m below ground. The groundwater table was about 2m below ground.
To support the excavation gravity DSM has adopted to the maximum depth of 12m. The
DSM block was also designed to be sufficiently massive to overcome potential uplift forces.
Wet DSM columns of 0.87m diameter were interlocked at 0.75m spacing was adopted as an
effective solution. The columns were designed to achieve an unconfined compressive
strength of 0.3 Mpa.
The solid cored samples were tested in a laboratory and showed more than acceptable UCS
(Unconfined Compressive Strength) between 1MPa and 3MPa about 1 to 2 months after
installation. Standard Penetration Tests (SPT) were carried out for some columns and were
used to confirm increase in stiffness of the soil. Typical results are presented shear strength
of the soil was increased by more than 50 times after 1 to 2 months of curing period. The
typical plan, cross section of treatment scheme and SPT N value after treatment is shown in
Figure 19.

13
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice

Figure 19: Typical plan & cross section of Treatment scheme and SPT results of
Digester tank
The typical exposed DSM after excavation and digester after installation of DSM as shown in
Figure 20.

Figure 20: Typical exposed DSM and Structure after DSM installation
The technical paper explaining the project case study on “Deep Soil Mixing in mine Tailings
for an 8m Deep Excavation” is enclosed in Annexure III.

3.2.2 Case Study 2: Deep Excavation support for Basement of 3-storey commercial
complex with 2-level basement, Kuala Lumpur, Malaysia.
A project comprising 3-storey commercial complex with 2-level basement car park floors
(about 7m depth below existing ground level) is under construction in the middle of Kuala
Lumpur City Centre. The proposed 2-level basement construction required 7m deep
excavation with underlying limestone interface for a total perimeter length of about 690m.
The subsoil comprised of loose silty sand deposits and ex-mining soils with SPT values in
the range of 5 blows/ft to 12 blows/ft. Underlying this loose soil layers, karstic limestone
formation was found with extremely varying rock-head levels ranging between 3m and 15m
below existing ground level. The ground water table was found to be at about 1m to 2m
below existing ground level.
The gravity wall block was designed to ensure adequate resistance against lateral earth
pressure to support the intended depth of excavation, whilst reducing seepage water inflow

14
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice
and thus, minimise the possible risk of drawdown and consequent ground subsidence to the
surroundings. Wet DSM columns of 0.85m diameter were interlocked at 0.75m centres to
form the rigid gravity wall block. The columns were designed to achieve an unconfined
compressive strength of 1.0Mpa.
The cores from DSM columns were extracted and tested in a laboratory for UCS. The test
results indicated an UCS in the range of 1MPa to 3MPa. In addition, wall movement was
monitored during excavation works, which showed a maximum horizontal movement of about
30mm to 40mm.The schematic of DSM gravity wall block is shown in Figure 21.

Figure 21: Schematic of DSM gravity wall block.


The Completed excavation of Gravity wall is shown in Figure 22.

Figure 22 Completed excavation

3.2.3 Case Study 3: Deep Excavation support for Music Academy Poznan, Poland.
The Music Academy Poznan (PL) planned in 2005 new construction of a 5-storey concert
hall "I.J. Paderewski" with circular cross-section (ø=46 m). The building with a basement was
in 8 m distance to the neighbouring building built at a road junction.
The building is under a low powerful up dividend to a depth of 4m from easy to medium
densely packed sands. Among them is clay with a Basalt Stone Liner on. The water table is
about 3.5m above the deepest excavation bottom. was for the creation of 6.7m below ground
level a 0.8m thick base plate provided.

15
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice
In cooperation with the company GT project Keller Polska developed a variant in which the
excavation by 248 overlapped and reinforced TBV-pillars as well as by a ring beam (80 x 80
cm) in the head region is secured. Despite an intermediate layer of Basaltstei-NEN with 30 to
50cm diameter could the I-beam easily in the freshly prepared lagerich- TBV columns be
installed tig. Before excavation the excavation was from the ring beam.
Reinforced concrete for the ablation of the bending moment elements and transverse forces
completed. With the production of waterproof sheeting was in April 2005 started. The
columns with lengths of 8.2m and a diameter of 70cm were placed in a Centre distance of
55cm produced. The embedment in the impermeable sandy loam was 3.5m. Each second
column was reinforced with an IPE 300 profile. The typical CDSM scheme and exposed
columns after excavation is shown in Figure 23.

Figure 23: Typical CDSM scheme and exposed CDSM columns


The completed structure of Music Academy Poznan is shown in Figure 24

Figure 24: Completed structure of Music Academy Poznan

3.2.4 Case Study 4: Excavation Support for TBM Retrieval Shaft using Deep Soil
Mixing Technique, Kuala Lumpur
The Klang Valley Mass Rapid Transit (KVMRT) Project when completed will cover a distance
of 51km and comprise of 31 passenger stations. The South Portal structure at Taman
Maluri, Kuala Lumpur (KL) acted as the transition point between the elevated and
underground sections of the Sungai Buloh - Kajang line and also, as the shaft for retrieval of
the Tunnel Boring Machines (TBM) from Cochrane Station. The rail level is about 15m below
the existing ground level. The ensuing 15m deep excavation required an earth support
system.

16
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice
Seven (7) boreholes were conducted on the site during the design stage. These showed that
thickness of overburden soil varied between 7m and 10m below existing ground level. The
soil generally comprised of sandy silt with interbedded layers of soft clay. This is typical of
former tin mining soil. The typical cross section & scheme of DSM, exposed columns after
installation are shown in Figure 25 & Figure 26.

Figure 25: Typical cross section of the tunnel

Figure 26: DSM scheme and exposed columns after excavation


Core samples were collected to examine consistency of the columns and to recover sections
for unconfined compressive strength (UCS) tests. Cores were done in both the centre and
the edge of the columns (where the columns intercept). As shown in Figure 27, test results
show UCS strength consistently above 1.5 MPa after 28 days.

Figure 27: DSM core sample UCS strength test results

17
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice
Wall movement was monitored during excavation works using both inclinometers and
settlement markers. Three inclinometers were installed, one on each wall face, and twenty-
one settlement markers on the ground surface. Maximum wall movement was observed at
East wall, showing a reading of 10 to 15mm (less than 0.15% wall height).
The shape of the deflection implied that there was some sliding movement, albeit small. This
is well within widely accepted wall defection criterion of 0.5%. The maximum ground
subsidence of 2mm to 6mm was observed behind West wall, and this was probably caused
by construction load. Back-analyses imply a stiffness modulus of the block between 35 and
100 times of UCS.
In Annexure IV, the technical paper explaining the project case study on “Excavation Support
for TBM Retrieval Shaft using Deep Soil Mixing Technique, Kuala Lumpur” is enclosed.

3.3 Ground treatment for shallow depth

3.3.1 Case Study 1: Mass Stabilisation of Peat and Mud for a Road Embankment
Gärtunavägen, Sodertalje, Sweden.
The project was a general contract, where Peab in cooperation with Keller Grundläggning AB
and SWECO received the work to construct a continuation of road 225 between Södertälje
and Nynäshamn. The new road was to be built upon a peat bog with underlying clay. Keller
Grundläggning AB made the design for the mass- and deep stabilisation.
The Ground Investigation Report showed 0.5 – 3m organic soil (peat) and underneath from 3
to 15m a soft silty clay, laying upon moraine. The peat had a water content of up to 1200%.
The organic soil had a water content of 300 – 500% and a shear strength of 3 to 7kPa. The
clay had a shear strength of 10 – 25 kPa with increasing strength towards the depth.
Mass stabilisation was chosen for the organic soil and lime-cement columns for the
underlying clay. Laboratory tests and field tests were executed and the proposed binder
content was 175kg/m3 cement for the peat and 80kg/m3 lime/cement (50%/50%) for the clay.
The dimensioning shear strength in the peat was 50kPa and 100 kPa for the clay. The
achieved shear strength in the project was from 50kPa in the peat and 200kPa in the clay.
The settlements in the peat stopped after approximately one month.
Keller’s specially designed mass stabilisation machine was used in combination with the
ordinary lime column rig. The mass stabilisation work was executed in a grid pattern with
blocks of 3×4m. This was to ensure that the right amount of binder was mixed into the soil.
Shortly after stabilisation a layer of 0.8m fill and geogrid was laid over the block to preload
the peat layer. The rigs could use the stabilised area after 24 hours. The final fill-height of the
embankment was applied after 1 month. The typical cross section of mass stabilisation is
shown in Figure 28.

18
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice

Figure 28: Typical cross section of Mass stabilisation

4 Bibliography
1. Federal Highway Administration Design manual (2013): Deep mixing for embankment
and foundation support., FHWA-HRT-13-046.
2. Raju V.R & Hari Krishna Y “Deep soil mixing”, Institute of Engineers Malaysia (2005)
3. Broms, B.B. (1999) “Design of lime, lime/cement and cement columns”, Proceedings
of the International Conference on Dry Mix Methods for Deep Soil Stabilisation,
Stockholm, Sweden, pp. 125-153.
4. Broms, B.B. (2004) “Lime and Lime/cement Columns”, Ground Improvement (2nd
Edition), Edited by Moseley, M.P. and Kirsch, K, Spon Press. 2004, pp. 252-330.
5. prEN 14679 (2003) “Execution of Special Geotechnical Works – Deep Mixing”,
European Standard.
6. Euro Soil Stab (2002) “Development of Design and Construction Methods to Stabilise
Soft Organic Soils”, Design Guide Soft Soil Stabilisation, CT97-035I, European
Commission Project BE 96-3177.
7. Larsson, S. (2005) “On the Use of CPT for Quality Assessment of Lime Cement
Columns”, Proceedings of the International Conference on Deep Mixing Best Practice
and Recent Advances, Stockholm, Sweden.
8. LCM Brochure “Dry Soil Mixing”, Sweden.
9. Raju, V.R. and Abdullah, A. (2005) “Ground Treatment Using Dry Deep Soil Mixing
for a Railway Embankment in Malaysia”, Proceedings of the International Conference
on Deep Mixing Best Practice and Recent Advances, Stockholm, Sweden.
10. Swedish Geotechnical Society (1997) “Lime and Lime Cement Columns”, SGF
Report 4:95E: Guide for Project Planning, Construction and Inspection, Linkoping,
Sweden.
11. Topolnicki, M. (2004) “In-situ Soil Mixing”, Ground Improvement (2nd Edition), Edited
by Moseley, M.P. and Kirsch, K, Spon Press. 2004, pp. 331-428.
12. V.R Raju and Y Hari Krishna, (2008) “Ground improvement Techniques for
infrastructure projects in Malaysia”, Proceedings of 12th International conference of
International Association for computer methods and Advances in Geomechanics,
Goa.

19
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice

Annexure I Technical paper on “Soil mixing - Challenges of Application


ranging from Ground Improvement to Structural Elements”
by Topolnicki M, 2006.

20
Ground Improvement Techniques

Soil mixing -
challenges of
applications ranging
from ground
improvement to
structural elements
MM.. T Tooppoollnniicckkii,, pprrooff.. ddrr hhaabb.. iinnżż..,,
GGddaańńsskk U Unniivveerrssiittyy ooff TTeecchhnnoollooggyy & &K
Keelllleerr
PPoollsskkaa SSpp.. zz oo..oo..

Presented by
XXIIIIII.. D
Daannuubbee--EEuurrooppeeaann C Coonnffeerreennccee oonn
Keller Grundbau GmbH GGeeootteecchhnniiccaall EEnnggiinneeeerriinngg,, LLjjuubblljjaannaa,,
Kaiserleistraße 44 2299..--3311..0055..22000066
D-63067 Offenbach
Tel. +49 69 8051-0
Fax +49 69 8051 244
E-mail: [email protected]
T
Teecchhnniiccaall ppaappeerr 3322--5555 EE
www.KellerGrundbau.com

1
Soil mixing - challenges of
applications ranging from ground
improvement to structural elements

M. Topolnicki
Gdańsk University of Technology & Keller Polska Sp. z o.o.

ABSTRACT
The current state of development of the wet method of deep Soil Mixing (SM) is presented,
based on experience gained in Poland, where the use of wet SM is seeing continuous
growth which begun in 1999. Considered case histories refer to ground improvement,
foundation support, retention systems and cut-off walls. They demonstrate not only a
unique flexibility of the application of SM, but also prove that deep soil mixing should be
considered as a competitive solution to a number of different geotechnical problems. Chal-
lenges of SM application are connected with varying properties of the stabilised soil, as well
as with appropriate geotechnical design. Stiff, rigid SM columns interact in a different way
with the soil than the more flexible columns. It is therefore important that design engineers
are aware of these differences. Specialised soils mixing in relation to environmental projects
is not touched upon.

1.INTRODUCTION

The use of in situ Soil Mixing (SM) in Europe to improve the engineering properties of soft
or contaminated ground is increasing rapidly, indicating growing interest and acceptance of
this relatively new technology in civil engineering. The extent of applications of SM across
European countries differs considerably, as reviewed recently by Massarsch and Topolnicki
(2005). Outside Scandinavia the total number of implemented projects is still relatively small
compared with other relevant geotechnical methods, however industrial, social and envi-
ronmental developments in Europe offer major commercial opportunities for an increase in
SM applications. In Poland, for instance, the wet soil mixing method is now in regular use.
Typical applications of SM involve ground improvement and hydraulic cut-off walls. However,
structural supports, such as pad and strip foundations and retaining structures, have also
been founded on soil mixing elements installed in various patterns ranging from individual
columns to grids, walls and blocks of stabilised soil (e.g. Topolnicki, 2004). Consequently, the
strength of soil mixing elements may differ significantly within the range determined by low
XIII. Danube-European Conference on Geotechnical Engineering, Ljubljana, 29-31.05.2006

capacity columns, with a compressive strength of about 0.3 to 0.5 MPa, and high capacity in-
dividual or combined structural elements, having unconfined compressive strength in the or-
der of 2 to 5 MPa, which perform similar to piles or block foundations. The external loads
are usually transferred down to the bearing layer resulting in a fixed type improvement, but
can be also partly or wholly transferred to the foundation soil when a more interactive or
even a floating type of improvement is desired. The choice of the required strength and of
the load transfer system is dictated by the purpose of the deep mixing application, and re-
flects the mechanical capabilities and characteristics of the particular mixing method used.
In the following sections, selected examples of soil mixing applications are presented to illus-
trate the current practice, based on the experience gained in Poland, where the use of wet
soil mixing has grown rapidly since 1999. Considered case histories refer to ground im-
provement, hydraulic cut-off wall, support of slab and individual foundations and retention
systems. They demonstrate not only a unique flexibility of the application of SM, but also
prove that deep soil mixing should be considered as a competitive solution for a number of
different geotechnical problems.

2. APPLICATIONS

2.1 Ground improvement for city highway

SM columns of diameter 0.8 m were applied to support a road embankment overlying weak
soils found to the depth of 3 to 8 m below the ground surface (Fig. 1). The subsoil consisted
of loose anthropogenic fill, underlined by peat and organic clay. The organic soils were 1 to 4
m thick. The embankment height was 1.3 to 2.5 m and the equivalent live load was 30 kPa. A
triangular column spacing of 2m×2m was selected, resulting in the design compressive stress
of 480 to 676 kPa, assuming column diameter reduced to 0.7 m. The required unconfined
compressive strength (UCS) of the stabilised soil material was 1.5 MPa. Altogether, 2402
columns with a total length of 15,532 m were constructed. The final embankment was
strengthened with two layers of geogrid, resulting in the so-called Load Transfer Platform
design (cf. Topolnicki, 1999).

2.2 Hydraulic cut-off wall

A hydraulic cut-off wall is constructed by installing intercut SM columns to intercept poten-


tial seepage flow paths. Since the hydraulic conductivity and continuity of the cut-off wall are
the most important design considerations, the slurry mixes must be tailored to soil condi-
tions, and adequate control of columns’ overlapping zones and verticality are required. This
is especially important when cut-off walls are executed to a large depth with single shaft mix-
ing equipment, as shown in Fig. 2.
For SM cut-off walls, the unconfined compressive strength is typically in the range of 0.7 to 3
MPa, and even higher if steel reinforcement is installed. The permeability is normally be-
tween 10-8 to 10-9 m/s. When bentonite and/or clayey stone dust and/or fly ash are added to
the slurry mix, the permeability can be reduced to 10-9 to 10-10 m/s, with the associated de-
crease of the unconfined compressive strength usually below 1 MPa.

2.3 Support of a slab foundation

A multistory building was located in heterogeneous soil conditions. Under superficial mixed
fill, organic clay and some peat were locally present, extending from 3.5 to about 6.7 m be-

2
M. Topolnicki, Gdańsk University of Technology & Keller Polska Sp. z o.o.

low the foundation level. Organic soils were underlined by fine sand and silt layers of varying
thickness, making ordinary piling very expensive due to the necessary pile length. Early calcu-
lations also indicated, that direct placement of the foundation slab on the existing soil would
lead to large and unequal settlements, ranging from 7 to about 50 cm.
In this situation a wet SM option was investigated and finally accepted by the client. The de-
sign was based on 3D FEM calculation, allowing for slab-soil interaction and elastic behavior
of columns. The resulting arrangement of 461 SM columns is shown in Fig. 3. For the out-
lined design it was essential to make a good estimate of the expected compressive strength
of the stabilised soil since significant variation in column strength was anticipated. For this
reason, maximum factored load acting on a single column was limited to 430 kN, resulting in
design compressive stress of 0.86 MPa, and a special mixing procedure was adopted at the
construction site. With a safety factor of 2.5, applied to the maximum factored design stress,
the 28-days UCS of the stabilised soil was required to be 1.9 MPa.

2.4 Support of strip and pad foundations

Strip and pad foundations of a commercial centre Megaplex were initially designed for CFA
piles because excessive settlement differences were expected. The subsoil consisted of a thin
fill material, underlying sandy and silty clays, and a coarse sand layer in a dense state. Loads
acting on strip foundations ranged between 230 and 729 kN/m, with resulting vertical
stresses of 230 to 430 kPa. Twelve types of rectangular footings were also designed for
loads between 1170 and 5670 kN, yielding vertical stress of 310 to 677 kPa (Fig. 4).
The initial piling design was successfully converted into a competitive SM project. Pad and
strip foundations were supported on closely spaced SM columns of 0.8 m diameter, and
were consequently designed as direct foundations, resulting in cost savings since less steel
reinforcement was needed. The number of columns under the footings ranged from 3 to 14,
and was determined by taking into account the expected strength of the stabilised soil as
well as the allowable settlement difference of 5mm over 6 m span, specified by the client.
The maximum design load acting on a single column, with slightly reduced diameter due to
aggressive groundwater, was limited to 512 kN, corresponding to compressive design stress
of 1330 kPa. Exposed columns on the bottom of foundation pits are shown in Fig. 5.

km 0+708

Fill / Peat /
Organic clay

Medium sand

SM columns, Ø 80 cm, Lmean =6.5 m

Figure 1. Road embankment supported on SM columns (Trasa Zielona, Lublin, Poland).

3
XIII. Danube-European Conference on Geotechnical Engineering, Ljubljana, 29-31.05.2006

3,0 m
Bentomat
BentomatteST
ST
1:2
A 1:2

0,0 old dam


alter Deich 1: 3

1,4 1 A
cut-off wallaus
Dichtwand from SM columns
DMM-Säulen
L= 4,5 bis 8,7 m

2,9 2

100
6,9 50 Ø6
0

7,6 4
5

Figure 2. Soil mixing cut-off wall to intercept seepage flow paths under river bank embankment (The
Vistula river near Tarnobrzeg, Poland).

4
M. Topolnicki, Gdańsk University of Technology & Keller Polska Sp. z o.o.

392.4 363.3 358.2 369.3 379.0 377.5 372.5 371.0 372.3 375.0 378.8 385.3 394.0399.0 403.1 406.7 410.0 411.6 409.4 406.5 403.0 398.4 393.1 385.0 372.6 361.1 352.9 348.7 348.1 351.5 358.3 370.0
389.7
372.7
354.6 360.3 379.1 371.2 369.7 377.0 388.0 392.9 404.8405.7 396.4 404.9 417.2 400.6 398.2 402.4 391.9 359.4 345.2 346.6 407.3 A
394.6
361.6
358.9 410.2
395.9
370.7 405.4400.2 380.1 354.4 341.8 353.6
391.7 366.9 400.8 381.3
364.1 402.0 377.5 410.2
387.3 389.3 391.1 394.2 362.3 394.7 408.9
376.0 409.8397.2
386.8 352.0 378.3 379.9 388.6 398.0 405.6 410.4 403.1 387.2
2 32 393.8 3 93

413.6 370.4 382.5 370.9 364.8 405.5


365.1 392.1 406.3 395.6 391.6 410.6 405.5 403.2
415.5398.7 384.0 377.8 390.5
420.0 402.9 398.2
383.9 341.0 385.4 383.3 387.0 398.0 405.8 409.2 411.1 412.8 368.7 416.1 398.1 379.7 391.5 400.8
364.1 399.6 378.6 386.4 384.2 399.6 E
411.1394.2 399.7
380.6 384.7 387.1 386.9
407.5 395.8
387.4 355.1 379.9 384.6 395.7 404.4 405.3 403.9
403.5 363.7 388.8 364.0 389.7 393.9397.7
361.0 396.5384.2
379.8 390.2
370.8 371.4
399.1 399.5 397.1 F
371.1 373.4 403.3
368.5 387.8 358.3 379.4 365.9 388.6 384.5 393.1 G
394.8 356.9 388.1 380.4 378.7
389.9382.3 376.0 393.9
366.8 379.9 391.1 398.9 373.2
368.5 365.9
390.0 366.2
361.6 374.1 394.3 393.1 386.8
402.2 398.5392.7 387.1
369.4 375.0 376.0 366.7 366.1 378.8 394.4 397.3 377.7 370.3
370.2 379.8 395.7 411.2 403.4 366.7
383.1 393.2 390.5 380.8
398.5 380.6 386.8
407.5 386.3 385.1 394.6 385.4 364.8 365.1 388.0 404.7 394.2 390.5 398.1395.9 382.1 387.2 400.8 418.7 416.5 415.4 402.8 379.0373.3
423.3 424.1 H
392.1 411.7 414.7 414.4
373.0 397.0 381.7 383.4 388.6 383.5 387.3 398.9
409.1 419.8
420.3 383.4
359.3 379.6 370.7 375.5
I

401.3 424.1 405.2


386.3 414.4
398.2 406.0 410.4 404.2
411.6 399.2 394.6405.7 362.1 379.1
405.4 415.1 375.0 361.1
364.7 383.6 374.1
404.1 366.6
35 m

388.5 373.1 368.8 361.0 377.5


413.5
383.3
378.6
407.6 373.5 412.0 413.3 383.2 371.9 387.4 416.5 399.8 371.1 364.5 373.9
365.5 365.2
376.0
400.6 367.8 371.7384.4 354.5 372.1
405.8 369.4 1 00
377.3 380.4
413.6 369.6
409.6 397.3
361.2 414.1 389.8 408.6 384.8
403.3 369.1 402.4 386.4 361.6 379.7
403.0 356.1 402.7 387.1
397.7409.7 416.9 393.9
400.5 358.0 406.1 364.4
353.4 355.4 373.8
360.2
393.3 377.8 407.9
400.8 403.7 339.9
345.8 375.7
354.3 350.4 414.6 327.7 410.1 388.2
391.0 419.0 350.5 379.9 412.3
405.9 387.7 358.5
407.2 367.9 372.3 374.1
416.1 399.7
365.0 393.9 387.6
421.3 429.3 K
415.4 400.7 385.5
410.6 411.0413.1 384.9 383.9 410.5 366.2
366.8 391.3 308.8 383.8 360.2
392.1 399.4
423.6 397.0 409.2 392.1 373.3

401.6 394.6 387.3 394.3 371.0 352.8


423.4 390.9 379.5 392.3 M 397.5
383.7
395.4 367.5 382.9
360.2 361.0
375.2 371.7 368.9
376.5 398.4 410.0 410.0
410.2409.2 399.7
397.8 395.8402.9
10 356.3 380.4 394.6

12 11 55 m
13

Figure 3. Arrangement of SM columns under the foundation slab of multistory building (Kielce,
Poland).

type 3 4
type 1 F=1170 kN type 2 F=1610 kN F=1420 kN type F=2160 kN

type 5 F=1820 kN type 6 F=2600 kN type 7 F=2970 kN type 8 F=3500 kN

type 9 F=3980 kN type 10 F=4280 kN type 11 F=5160 kN


type 12 F=5670 kN

Figure 4. Arrangement of SM columns under pad foundations of a commercial centre (Megaplex,


Katowice, Poland).

5
XIII. Danube-European Conference on Geotechnical Engineering, Ljubljana, 29-31.05.2006

Figure 5. Exposed SM columns under pad foundation of a commercial centre (Megaplex, Katowice,
Poland).

Figure 6. Soil mixing works and exposed columns for a bridge support foundation (A2 highway near
Poznań, Poland).

Figure 7. Temporary SM wall with embedded steel H-beams for excavation support (Music Acad-
emy, Poznań, Poland).

6
M. Topolnicki, Gdańsk University of Technology & Keller Polska Sp. z o.o.

Figure 8. Permanent soil mixing retention wall with embedded steel H-beams and concrete facing
(Oświęcim, Poland).

2.5 Bridge supports

Construction work on the A2 highway in Poland led to interesting applications of wet SM.
Careful analysis revealed that certain road bridges, originally designed on large diameter
piles, can be supported on SM columns, fulfilling all stability and settlement requirements and
offering substantial savings on foundation costs.
The solution adopted for bridge WD-23 provides a good example of this application. The
subsoil consisted of boulder sandy clays, with an average CPT cone resistance of 2 MPa. The
geotechnical design included a total of 200 columns for the five bridge supports. The applied
arrangement of 46 SM columns under bridge abutment foundation is shown in Fig. 6. The al-
lowed maximum characteristic design compressive stress was 830 kPa, and the requested
compressive strength was 2.5 MPa. The observed settlement of this support was 12 mm. It is
interesting to note that since 2002 about 80 road bridges have been founded on SM columns
in Poland.

2.6 Temporary and permanent retention systems

Retention systems using SM are mostly comprised of applications associated with restraining
the earth pressure mobilised during deep excavations. In these applications wall-type column
patterns are used, while the soil-cement mix is typically engineered to have high strength and
stiffness. To overcome soil and water lateral pressures the SM columns should also have
adequate shear resistance. Other key requirements for successful construction are a high
degree of column homogeneity and maintaining verticality tolerance to achieve the minimum
required designed thickness of columns effectively in continuous contact.
For the projects shown in Figs. 7 and 8, steel H-beams were installed in fresh SM columns to
increase the bending resistance and create a structural wall for excavation support.
Elongated mixing time and full restroking were required to ensure easier installation of sol-
dier elements immediately after mixing. Panels of mixed soil between H-beam reinforcement
were designed to work in arching. Tieback anchors or stage struts are typically used in com-
bination with the SM walls. In case of circular excavations or shafts peripheral concrete cap
beams working in hoop compression can be constructed instead, as shown in Fig 7.

7
XIII. Danube-European Conference on Geotechnical Engineering, Ljubljana, 29-31.05.2006

For permanent SM walls, concrete facing is needed to prevent the soil-cement mix from
long-term deterioration and damage and to provide smooth wall surface. Concrete facing is
easily constructed, with the reinforcing bars welded to exposed H-beams, as shown in Fig. 8.

3. CONCLUSIONS

When soil mixing is applied to support shallow embankments or foundation slabs usually to
reduce excessive or differential settlement, the quality of individual columns is less important
and the overall performance depends mainly on the combined interaction between the sup-
ported structure, soil, and strengthening elements (i.e. columns). This soil-structure interac-
tion design concept is frequently applied in the case of low strength soil mixing and is often
combined with preloading to accelerate strength gain and consolidation settlement. This
concept has proved to be efficient and cost-effective. On the other hand, when soil mixing is
performed to support high embankments or heavily loaded pad or strip foundations, and
where horizontal loads, shear forces, or bending moments may appear, the quality of load-
bearing columns is essential to prevent progressive failure mechanisms. The same applies for
economically attractive low values of the area improvement ratio, and for retention systems
with steel reinforced soil mixing columns. Consequently, a good assessment of the expected
strength and deformation properties of the stabilised soil is one of the key issues in reliable
and optimum SM design. Stiff, rigid columns interact in a different way with the soil than the
more flexible columns. It is therefore important to account for these differences in the geo-
technical design of Soil Mixing.

REFERENCES
- M. Topolnicki, Low height reinforced embankments supported on piles: application, de-
sign and reliability – a panel report, Proc. XII European Conference on Soil Mechanics
and Foundation Engineering, Amsterdam, Vol. 2, s. 1191-1194, 1999.

- R. Massarsch, M. Topolnicki, Regional Report: European Practice of Soil Mixing Technol-


ogy, Proc. of Int. Conference on Deep Mixing – Best Practice and Recent Advances,
Stockholm, Vol.1, pp. R19-R45, 2005.

- M. Topolnicki, In situ Soil Mixing, Chapter 9 in Ground Improvement book, Editors M.


Moseley and K. Kirsch, Spon Press, pp. 331-428, 2004.

8
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice

Annexure II Technical paper on “Quality Control of Wet Deep Soil


Mixing with reference to Polish Practices and
Applications” by Michal Topolnicki,2002

21
Quality Control of
wet Deep Soil Mixing
with Reference to
Polish Practice and
Applications

Michal Topolnicki,
Prof. Ph. D. Civ Eng.,
Head of Marine Civil Eng. Department
at the Technical University of Gdansk;
Managing Director and Chief Designer
of Keller Polska Ltd., Gdynia, Poland
Phone: +48 58 629 7510,
Fax: +48 58 629 7470,
[email protected]

Presented by
Keller Grundbau GmbH
Kaiserleistr. 44
D-63067 Offenbach
Tel. +69 8051 - 0
Fax +69 8051 - 244 Deep Mixing Workshop, Tokyo, 15.–18.10.2002
E-mail [email protected]
www.KellerGrundbau.com Technical Paper 32-52E
Quality Control of Wet Deep Soil Mixing with
Reference to Polish Practice and Applications
Michal Topolnicki1

Abstract
With reference to Polish application of wet Deep Soil Mixing, begun in 1999, three selected
projects are presented relating to quality control issues. The first case reflects the impor-
tance of adequate soil investigation when dealing with soil improvement. The second refers
to the evaluation procedure of strength data obtained from laboratory tests on cubic sam-
ples, indicating that the standard approach used for concrete should not be automatically
applied for DSM. The last case concerns highway bridge support founded on DSM columns
and describes quality control by means of a preloading test.

Introduction
Application of wet Deep Soil Mixing started in Poland in 1999. The first project, designed
and conducted by Keller Polska Ltd., comprised the execution of intersecting DSM columns
forming a sealing wall along an old dam of the Vistula River in Krakow, as a part of a major
food protection program for this historical city. Since then more than 30 projects have been
completed in Poland so far, indicating growing importance of this technology. Despite a rela-
tively short period of application, the range of executed DSM projects already covers quite a
broad spectrum of difficult geotechnical cases, including the following: improvement of or-
ganic soils for a new city road, the foundation of several multi-story buildings on slabs sup-
ported by DSM columns, strip and pad foundations of industrial and municipal buildings,
foundation supports of highway bridges, the sealing walls and temporary protection of exca-
vation walls.
In Polish foundation practice, DSM columns with a diameter of 80 cm and a length of 3 to 10
m have been executed (in a few cases also 60 cm dia. columns were used, but mainly for
sealing walls). This limitation led to a requirement for higher internal column strength than
usually applied elsewhere, as compared for instance to Keller’s practice in the USA, where
large diameter columns are frequently executed.

In the following, three different DSM applications are briefly presented in order to illustrate
important quality control issues. All reported cases were designed and executed by Keller
Polska Ltd.

1Prof. Ph. D. Civ Eng., Head of Marine Civil Eng. Department at the Technical University of Gdansk; Narutowicza 11,
80-952 Gdansk, Poland; Phone: +48 58 347 2611; Fax: +48 58 347 1436
Managing Director and Chief Designer of Keller Polska Ltd.; Rdestowa 51a, 81-577 Gdynia, Poland;
Phone: +48 58 629 7510; Fax: +48 58 629 7470; [email protected]
Quality Control of Wet Deep Soil Mixing with Reference to Polish Practice and Applications

Pad and strip foundations with strongly varying loading

This case has been selected in order to illustrate how insufficient soil investigation data may
influence performance of foundations supported on DSM columns.

Strip foundations of varying width between 1.0 and 1.7 m were designed for loads ranging
from 230 to 729 kN/m, with a resulting unit pressure of 230 to 430 kPa. There were also 12
types of rectangular pad foundations, loaded from 1170 kN up to 5670 kN, with resulting
unit pressure of 310 to 677 kPa. This challenging geotechnical project, originally designed for
CFA piles because of expected settlement differences, has been changed to modern soil im-
provement with wet DSM. Pad and strip foundations were designed as shallow foundations
supported on DSM columns with an 80 cm diameter. Assumed column layout under the
footings is shown in Fig.1. The number of columns under pad foundations ranged from 3 to
14 and was selected taking into account expected internal strength of DSM columns and
allowable settlement difference of 5mm over 6 m span, specified by the client. The maxi-
mum design load acting on a single column was limited to 512 kN. This corresponds to
compression stress of 1330 kPa for a column with reduced diameter due to slightly aggres-
sive groundwater.

type 3 4
type 1 F=1170 kN type 2 F=1610 kN F=1420 kN type F=2160 kN

type 5 F=1820 kN type 6 F=2600 kN type 7 F=2970 kN type 8 F=3500 kN

type 9 F=3980 kN type 10 F=4280 kN type 11 type 12 F=5670 kN


F=5160 kN

Fig. 1. Arrangement of DSM columns under pad foundations

2
Deep Mixing Workshop, Tokyo, 15-18.10. 2002

Fig. 2. Trimmed DSM columns at the bottom of a foundation pit

An initial soil investigation report, ordered by the client and used at the design stage, in-
cluded soil profiles and parameters evaluated from classical borings and dynamic penetration
tests. In general, bearing strata comprised of coarse sand and stiff clay were indicated at the
depth of 5 to 6 m below the foundation level. In the case of pad foundation No 9, for in-
stance, a coarse sand layer with compression modulus of 110 MPa was found at 5.7 m depth
and the boring was terminated at 6 m (Fig.3). Consequently, for this footing 6 m long DSM
columns were assumed in the design.

qc [MPa]
0 5 10 15 20 0
s=3.04 cm Fill, E=25 MPa
1 1-
Sand, E=50 MPa
2 2-
Initial Sandy clay, E=5 MPa
3 3- column
4 4-
Clayey sand, E=80
5 5-
Silty clay, E=17
6 s=6.93 cm 6-
Coarse sand, E=110
7 7- Final col- MPa
umn
8 8
s=1.69 cm
9 9- Initial soil profile

z [m] z [m] (boring terminated


Supplementary at 6 m depth)

Fig. 3. Initial soil profile at footing No 9 (right) in relation to supplementary cone penetration test
(left) and the resulting change in column length

3
Quality Control of Wet Deep Soil Mixing with Reference to Polish Practice and Applications

During construction, additional soil investigations were conducted by the contractor in the
framework of the quality assurance plan. Close to footing no. 9 a supplementary CPT indi-
cated below 6 m depth a silty clay layer with low stiffness (Fig. 3). Thanks to the immediate
reaction of the site engineer it was possible to extend fresh DSM columns to the depth of
8.5m and the whole work was completed successfully.
The reported case was subsequently analysed to see what could actually happen if the initial
column length was maintained. Calculations have indicated that with 6 m long columns the
expected settlement could reach 7 cm, as indicated in Fig.3. This is because the DSM col-
umns are stiffer than the surrounding soil and most of the load is transferred down to the
underlying weak layer. Interestingly, the resulting settlement would be even bigger than for a
direct foundation placed on untreated subsoil (ca 3 cm). Consequently, foundation perform-
ance would be worse despite the time and money spent on soil improvement. This case un-
derlines the role of adequate soil investigation data and of on-site control of works when
dealing with this type of soil improvement.

Foundation slab on DSM columns

A new multistory building was located in difficult heterogeneous soil conditions. Under su-
perficial mixed fill, organic clay and some peat was present, extending 3.5 to about 6.7 m
below the final slab foundation level. Organic soils were underlined by fine sand and silt lay-
ers of varying thickness, making ordinary piling very expensive due to necessary pile length.
Early calculations indicated also, that direct placement of the foundation slab on the existing
soil would lead to large and unequal settlements, ranging from 7 to about 50 cm despite the
slab stiffness (Fig.4).

settlement scale in [cm]: 49


0.496

1 2 3 4 5 6

0.470 44
0.443

0.310

23 2 393
39
0.390

F
G 34
0.336

0.416

0.363
H 29
0.283
I

23
0.229
0.256 10 0

J
0.203
0.149

18
0.176

0.096
K

0.043
12
0.123

10 L
7
0.069
12 11
13
14

Fig. 4. Calculated settlement of the slab founded directly on untreated subsoil

4
Deep Mixing Workshop, Tokyo, 15-18.10. 2002

In these circumstances a wet DSM option was investigated and finally accepted by the client.
The design was based on 3D finite element calculation, allowing for slab-soil interaction and
elastic behaviour of columns. The resulting arrangement of DSM columns in the plan view
and in a cross-section is shown in Figs. 5 and 6, respectively.
392.4 363.3 358.2 369.3 379.0 377.5 372.5 371.0 372.3 375.0 378.8 385.3 394.0399.0 403.1 406.7 410.0 411.6 409.4 406.5 403.0 398.4 393.1 385.0 372.6 361.1 352.9 348.7 348.1 351.5 358.3 370.0
389.7
372.7
354.6 360.3 379.1 371.2 369.7 377.0 388.0 392.9 404.8405.7 396.4 404.9 417.2 400.6 398.2 402.4 391.9 359.4 345.2 346.6 407.3 A
394.6
361.6
358.9 410.2
395.9
370.7 405.4400.2 380.1 354.4 341.8 353.6
391.7 366.9 400.8 381.3
364.1 402.0 377.5 410.2
387.3 389.3 391.1 394.2 362.3 394.7 408.9
376.0 409.8397.2
386.8 352.0 378.3 379.9 388.6 398.0 405.6 410.4 403.1 387.2
2 32 393.8 3 93

413.6 370.4 382.5 370.9 364.8 405.5


365.1 392.1 406.3 395.6 391.6 410.6 405.5 403.2
415.5398.7 384.0 377.8 390.5
420.0 402.9 398.2
383.9 341.0 385.4 383.3 387.0 398.0 405.8 409.2 411.1 412.8 368.7 416.1 398.1 379.7 391.5 400.8
364.1 399.6 378.6 386.4 384.2
399.7 399.6 E
411.1394.2 384.7
380.6 407.5 395.8 387.1 386.9
387.4 355.1 379.9 384.6 395.7 404.4 405.3 403.9
403.5 363.7 388.8 364.0 389.7 393.9397.7
361.0 396.5384.2
379.8 390.2
370.8 371.4 399.5 397.1 F
403.3 399.1
371.1 373.4
368.5 387.8 358.3 379.4 365.9 388.6 384.5 393.1 G
394.8 356.9 388.1 380.4 378.7
389.9382.3 376.0 393.9
366.8 379.9 391.1 398.9 373.2
368.5 365.9
390.0 366.2
361.6 374.1 394.3 393.1 386.8
402.2 398.5392.7 387.1
369.4 375.0 376.0 366.7 366.1 378.8 394.4 397.3 377.7 370.3
370.2 379.8 395.7 411.2 403.4 366.7
383.1 390.5 380.8
398.5 393.2 386.8
407.5 386.3 385.1 394.6 385.4 388.0 404.7 394.2 390.5 398.1395.9 382.1 387.2 400.8 418.7 416.5 415.4 380.6
364.8 365.1
424.1 402.8 379.0373.3
411.7 414.7 414.4 423.3 H
392.1 373.0 397.0 381.7 383.4 388.6 383.5 387.3 398.9 420.3 383.4
359.3 409.1 419.8 375.5
405.2 379.6 370.7
I

401.3 424.1
386.3 414.4
398.2 406.0 410.4 404.2
411.6 399.2 394.6405.7 362.1 379.1
405.4 415.1 375.0 361.1
364.7 383.6 374.1
404.1 388.5 368.8 366.6 377.5
413.5 373.1 361.0
383.3
378.6
407.6 373.5 412.0 413.3 383.2 371.9 387.4 416.5 399.8 371.1 364.5 373.9
365.5 365.2
376.0
400.6 367.8 371.7384.4 354.5 372.1
405.8 369.4 1 00
377.3 380.4
413.6 369.6
409.6 397.3
361.2 414.1 389.8 408.6 384.8
403.3 369.1 402.4 386.4 361.6 379.7
403.0 356.1 402.7 387.1
397.7409.7 416.9 393.9
400.5 358.0 406.1 364.4
353.4 355.4 373.8
360.2
393.3 377.8 407.9
400.8 403.7 339.9
345.8 375.7
354.3 350.4 414.6 327.7 410.1 388.2
391.0 419.0 350.5 379.9 412.3
405.9 387.7 358.5
407.2 367.9 372.3 374.1
416.1 399.7
365.0 393.9 387.6
421.3 429.3 K
415.4 410.5 400.7 385.5
410.6 411.0413.1 384.9 383.9 366.2
366.8 391.3 308.8 383.8 360.2
392.1 399.4
423.6 397.0 409.2 392.1 373.3

401.6 394.6 387.3 394.3 371.0 352.8


423.4 390.9 379.5 392.3 M 397.5
383.7
395.4 367.5 382.9
360.2 361.0
375.2 371.7 368.9
376.5 398.4 410.0 410.0
410.2409.2 399.7
380.4 394.6 397.8 395.8402.9
10 356.3

12 11
13
Fig. 5. Arrangement of DSM columns under the foundation slab
(size of each circle indicates column load in kN)

Foundation slab
45 cm, B30

Lean concrete B10

Foundation slab
60 cm, B30

DSM columns
dia. 80 cm
L=5.5 to 9 m

Fig. 6. Cross-section of the foundation slab and DSM columns

5
Quality Control of Wet Deep Soil Mixing with Reference to Polish Practice and Applications

For the outlined design, comprised of 461 DSM columns with a total length of about 3280 m,
it was essential to make a good estimate of the expected compression strength of the
grouted soil, referred to as internal column strength. Due to heterogeneous soil and the
presence of organic layers, significant differences in column strength was anticipated. For this
reason, maximum factored load acting on a single column was limited to 430 kN, resulting in
design compression stress of 0.86 MPa, as well as special mixing procedure was adopted at
the construction site. The average unit density of the cement slurry was about 1700 kg/m3
and the mean consumption rate was about 180 l/m. According to the assumed design crite-
ria, a general safety factor of 2.5 had to be applied to the maximum factored design stress
acting on a single column. This led to the required strength of at least 1.9 MPa after 28 days
of curing. The actual strength was checked on 32 standard cubic samples, extracted from
fresh DSM columns and tested for uniaxial compression at an independent laboratory. The
obtained results are presented in Tab. 1.
Once the test results became available for control it turned out that three samples (no 2, 3,
and 4) had lower strength than required. Although even the minimum strength of 1.11 MPa
was higher than the design stress 0.86 MPa, discussion started about the actual margin of
safety. Later on it has been also found that the first series of samples were left unprotected
during a chilly night and got partly frozen. This observation has not been duly reported,
however, and all samples were brought to the laboratory for testing.
The outlined case is reported in order to illustrate that the classical evaluation procedure of
sample strength data, based on 95% of confidence as prescribed for ordinary concrete (e.g.
Polish Standard PN-88/B-06250), should not be mechanically applied to DSM. The calcula-
tions presented in Tab. 1 show that standard deviation for DSM material can be very large,
and generally exceeds 20% of the mean value (in case of concrete this would require addi-
tional inspection of concrete quality). Consequently, the classical evaluation approach may
lead to overly strong restrictions for DSM applications. It is also demonstrated by evalua-
tions C and D in Tab. 1 that truncation of the obtained strength to, for example, 6 MPa
yields automatically higher calculation strength Rcd. This however should not lead to the con-
clusion that a reduced amount of cement in the slurry would lead to increased safety of soil
improvement. Therefore, a new evaluation procedure for DSM strength data, carefully tai-
lored to this technology, is actually needed. One possibility is to introduce fixed safety fac-
tors in relation to mean and minimum obtained strength, like for jet grouting, or to reduce
the level of confidence.

Highway bridge supports founded on DSM columns


Recent works for the A2 highway in Poland have initiated new, interesting applications for
wet DSM. After careful analyses it turned out that certain road bridges, originally designed
on large diameter piles, could be founded on DSM columns fulfilling all technical require-
ments with respect to stability and settlement of supports and at the same time offering sub-
stantial economical savings.
For illustration the solution adopted for bride WD-105 is presented. The design included
soil improvement for 5 bridge supports using 168 DSM columns. A typical layout is shown in
Fig. 7 for support P3 with 30 columns. Allowed maximum factored load for a single column
was 458 kN, resulting in compression stress of 916 kPa. The requested compression
strength was 2.3 MPa, applying a partial safety factor of 2.5. Predicted settlement for the
whole support was 0.95 cm.

6
Deep Mixing Workshop, Tokyo, 15-18.10. 2002

Tab.1 Results of compression tests of DSM samples and four evaluation procedures

Evaluation Evaluation Evaluation Evaluation


Compression test results
A B C D
Mean strength Strength Strength Strength Strength
No. Date of sampling Strength [MPa]
[MPa] [MPa] [MPa] [MPa] [MPa]
1. 3,11 3,11 3,11
samples samples
2. 1,60 1,60 1,60
21.03.2002 1,87 partly fro- partly fro-
3. 1,11 1,11 1,11
zen zen
4. 1,64 1,64 1,64
5. 6,76 6,76 6,76 6,00 6,00
6. 3,20 3,20 3,20 3,20 3,20
27.03.2002 5,07
7. 6,36 6,36 6,36 6,00 6,00
8. 3,96 3,96 3,96 3,96 3,96
9. 3,96 3,96 3,96 3,96 3,96
10. 4,09 4,09 4,09 4,09 4,09
28.03.2002 4,34
11. 4,53 4,53 4,53 4,53 4,53
12. 4,80 4,80 4,80 4,80 4,80
13. 7,60 7,60 7,60 6,00 6,00
14. 8,27 8,27 8,27 6,00 6,00
30.03.2002 7,04
15. 8,44 8,44 8,44 6,00 6,00
16. 3,87 3,87 3,87 3,87 3,87
17. 8,18 8,18 8,18 6,00 6,00
18. 8,13 8,13 8,13 6,00 6,00
02.04.2002 7,32
19. 7,64 7,64 7,64 6,00 6,00
20. 5,33 5,33 5,33 5,33 5,33
21. 7,91 7,91 7,91 6,00 6,00
22. 8,04 8,04 8,04 6,00 6,00
05.04.2002 8,04
23. 8,44 8,44 8,44 6,00 6,00
24. 7,78 7,78 7,78 6,00 6,00
25. 6,67 6,67 6,67 6,00 6,00
26. 6,49 6,49 6,49 6,00 6,00
08.04.2002 6,08
27. 5,38 5,38 5,38 5,38 5,38
28. 5,78 5,78 5,78 5,78 5,78
29. 6,36 6,36 6,36 6,00 6,00
30. 6,00 6,00 6,00 6,00 6,00
09.04.2002 5,98
31. 6,22 6,22 6,22 6,00 6,00
32. 5,33 5,33 5,33 5,33 5,33
Calculation according to Rm = Mean value: 5,72 6,27 4,99 5,44
Polish Standard PN-88/B-06250
(confidence 95%) SD = Standard Deviation: 2,14 1,63 1,47 0,87
Rcg = Guaranteed strength Rcg = Rm-1,64*SD (95%) 2,21 3,60 2,57 4,01
Rck = Characteristic strength Rck = Rcg/1,25: 1,77 2,88 2,06 3,21
Rcd = Calculation strength Rcd = Rck/1,8: 0,98 1,60 1,14 1,78
Maximum factored design stress in DSM column: 0,86 0,86 0,86 0,86
Additional calculation Rcg = Rm-1,28*SD (90%) 2,98 4,19 3,10 4,32
for 90% of confidence: Rck = Rcg/1,25: 2,39 3,35 2,48 3,46
Rcd = Rck/1,8: 1,33 1,86 1,38 1,92
Evaluation A: All data, as measured in the laboratory

Evaluation B: Samples 1 to 4 excluded from analysis

Evaluation C: as A, with strength limited to 6.0 MPa for all samples having strength exceeding 6.0 MPa

Evaluation D: as B, with strength limited to 6.0 MPa for all samples having strength exceeding 6.0 MPa

7
Quality Control of Wet Deep Soil Mixing with Reference to Polish Practice and Applications

47 48 49 50 51 52 53 54 55 56

37 38 39 40 41 42 43 44 45 46

27 28 29 30 31 32 33 34 35 36

±0,0=78,3 m npm.
60

300 Gp/Pd
120

-1,40
-1,80 Gp/Pd
-2,20
50 100 100 50
Gp/Pd
-3,30

Gp/Pd

-4,70

-5,50
Gp zw

Gp/Pd = Sandy clay / Fine sand


Gp zw = Stiff sandy clay

Fig. 7. Typical arrangement of DSM columns (support P3, bridge WD-105)

In order to convince the client and the independent engineer of the merits of this new foun-
dation solution it was recommended that two loading tests of selected single DSM columns
were performed. The test was not intended to check the bearing capacity of a column, as it
would be required in case of piles, but rather to verify the load-settlement characteristic of
DSM column and to confirm the applied design method and the predicted settlement. Pre-
loading was conducted in 12 steps up to 572 kN, which is 150% of the value of characteristic
maximum load for a single column, equal to 458/1.2=382 kN. Each loading step was main-
tained for at least 30 minutes or until the observed settlement was less than 0.05 mm in 10
min. The obtained settlement curve is presented in Fig. 8.

8
Deep Mixing Workshop, Tokyo, 15-18.10. 2002

Load [kN]

0 100 200 300 400 500 600


0
-1
Loading
-2
Settlement [mm]

-3
-4
Unloading Max applied load
-5
Q=150% N
-6 Design load
Q=100% N
-7
-8
-9
-10

Fig. 8. Static loading test of a single DSM column (col. 156, bridge WD-105 )

As can be observed, the total settlement corresponding to the design characteristic load was
3.28 mm, and 2.11 mm after unloading. Under the maximum applied load the behaviour was
still far from an ultimate condition and the total settlement was only 8.22 mm and 5.83 mm
after full unloading.

The above test results were also reanalysed with the same calculation method in order to
check the settlement prediction. For a single column loaded with 100% load the calculated
settlement was 6.0 mm, while the observed value was 3.28 mm. This gave evidence that the
applied calculation approach is on the safe side and that the predicted settlement for the
whole support can be considered as the upper bound estimate. Further settlement observa-
tions, collected during subsequent construction stages of the bridge and afterwards, will be
also used for future back analysis.

Conclusions

Increasing applications of the wet Deep Soil Mixing in Poland have shown that this method of
soil improvement is very versatile and can be used with technical and economical success.
Quality control measures applied so far were closely related to individual project require-
ments and were based on active control during work execution, including the recording of
production parameters and additional soil testing, as well as on post execution tests, includ-
ing column inspection, sampling and even preloading, if necessary
The DSM technology still needs optimisation in relation to the equipment, mixing procedure
and energy as well as with respect to type and amount of binders applied for different soils,
in particular. Moreover, internationally accepted standards for DSM would be very helpful to
assure rational and responsible application of this technology in the future.

9
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice

Annexure III Technical paper on the project case study on “Deep Soil
Mixing in Mine tailing for 8m Deep Excavation”, Y.W Yee,
V.R.Raju, H K Yandamuri, Kuala Lumpur.

22
Ground Improvement Techniques

Deep Soil Mixing


in Mine Tailings
for a 8 m
Deep Excavation
Y.W. Yee, V.R. Raju and
H.K. Yandamuri
Keller (M) Sdn. Bhd., Kuala Lumpur

Presented by Presented at the 16th Southeast Asian


Geotechnical Conference, 2007, Kuala Lumpur,
Keller Grundbau GmbH Malaysia
Kaiserleistraße 44
D-63067 Offenbach
Tel. +49 69 8051-0
Fax +49 69 8051 244
E-mail: [email protected]
www.KellerGrundbau.com Technical paper 32-57 E

1
Deep Soil Mixing in Mine Tailings
for a 8m Deep Excavation
Y. W. Yee, V. R. Raju & H. K. Yandamuri
Keller (M) Sdn. Bhd., Kuala Lumpur, Malaysia
[email protected]

Abstract:
Soil improvement using mixing technology has advanced appreciably and can now solve a
broad spectrum of geotechnical problems. The method basically involves the introduction
and mechanical mixing of binder (cement or lime) with the in-situ soils to improve the
strength, deformation and permeability characteristics of the problematic soils. The process
procedure and operating parameters suitable for use in the delivery and mixing of cement in
the ground would depend on the ground conditions such as soil type, density, water content
and the end product required. This paper describes the design and construction methods
carried out to enable an 8m deep excavation in very soft ex-mining slime using Deep Soil
Mixing (DSM) technique for a sewage treatment plant in Kuala Lumpur. It is demonstrated
that cement mixing increased the shear strength of the soil by more than 50 times within a
short period which enable the excavation to be carried out at a steep 45 degrees whilst
excluding groundwater. This procedure proved to be more cost and time effective
compared to the original idea of using sheet piles, anchor tie-backs and grouting. The design
process is explained touching on soil investigation to derive engineering and process
parameters and subsequent computer simulations. The importance of quality control
measures during construction stage are emphasized and available proving methods for the
post construction stage are also discussed.

1 BACKGROUND

The Sewage Services Department of the Ministry of Housing and Local Government,
Malaysia proposed to construct a Sewage Treatment Plant in Kuala Lumpur. As part of the
treatment process, four (4) numbers of 23m diameter digesters were constructed at
equi-distance from each other (see Fig. 1). Each digester is essentially a 15m high tank with a
coned shaped (45 degree) base buried 8m below ground.

To construct the cone shaped base, it was necessary to firstly, excavate the soil. Being a
former tin mining land, the ground is underlain by very soft slime. Conventional excavation
method would require multiple handling of installing anchored sheet piles and subsequent
removal of soft slime before replacement with competent soil formed at 45 degrees. An
alternative method was realised using deep soil mixing technique which involved treating the
Y. W. Yee, V. R. Raju and H. K. Yandamui, Keller (M) Sdn. Bhd., Kuala Lumpur, Malaysia

soft ground beneath the tank footprint and subsequent excavation of the cone shaped base
without the need of any excavation support system.

Fig.1 Typical view of four digesters

2 SITE LOCATION

The site is located off Tasik Titiwangsa in Kuala Lumpur. The site is a former tin mining land
and situated adjacent to private dwellings. Construction works were required to cause
minimal disruption to these surrounding properties, especially with regards to ground
movement and noise pollution. The four digest-ers are located at the north-east corner of
the site. As shown on the aerial view (Fig. 2), the four digesters are arranged at equal
distance from each other and named N1, N2, S1 and S2.

2
Deep Soil Mixing in Mine Tailings for a 8m Deep Excavation

Fig.2 Arial view of the site showing location of digesters

3 SUBSOIL CONDITIONS

Being a former tin mining area, the site is underlain by highly variable soil conditions. Very
soft slime was found at the foot-prints of 3 of the digesters (N1, S1 and S2) while loose sand
with slime layers was found at one of the digesters (N2). The depth of limestone bedrock
typically varied between 7m and 13m below ground. The groundwater table was about 2m
below ground.
The slime has typically the following characteristics: unit weight 1.5t/m3; moisture content
80% - 100%; liquid limit 70% - 80%; plasticity index 30% - 40%. Geonor vane shear test
showed undrained shear strength (Cu) of between 5kPa and 10kPa. Typical result of dynamic
penetration test (DPT) showing extent of slime (essentially zero blow count) is shown in
Fig. 3.
The loose sand (beneath N2) has typically the following characteristics: unit weight 1.7t/ m3;
moisture content 50%. Typical result of dynamic penetration test (DPT) showing the mixed
soil conditions is shown in Fig. 4, generally less than 10 blows / 10cm. Limited SPT tests
done on the loose sand, which showed SPT N between 3 to 8 blows / 30cm. Table 1
summarises the typical subsoil conditions for the four digesters.

3
Y. W. Yee, V. R. Raju and H. K. Yandamui, Keller (M) Sdn. Bhd., Kuala Lumpur, Malaysia

Fig. 3 Typical result of DPT showing slime below N1, S1 and S2

Digester Typical Subsoil Conditions


Reference Top level Description
0m Fill material
Digesters 1m Very soft slime
N1, S1 and Soft to firm sandy
8m
S2 clay
7 – 13m Weathered limestone
0m Fill material
Digester N2 Loose silty sand
1m
(with slime layers)
12m Weathered limestone

Table 1 Summary of typical subsoil conditions

4
Deep Soil Mixing in Mine Tailings for a 8m Deep Excavation

Fig. 4 Typical result of DPT showing loose sand with inter-bedded slime layers below N2

4 GROUND IMPROVEMENT CONCEPT

The objective of ground improvement was to treat the soft ground by improving its strength
and stiffness characteristics such that excavation could be carried out safely. The soil was
also required to be made fairly impermeable to water to reduce risk of piping and ground
loss.
After reviewing many options, it was found that Deep Soil Mixing (DSM) presented the most
technically acceptable and cost effective solution.
The soft soil would be mechanically mixed with cement and the end product would be a stiff
stabilised soil to allow the required 8m deep excavation. A schematic drawing showing the
DSM treated block and geometry of excavation is shown in Fig.5.

5
Y. W. Yee, V. R. Raju and H. K. Yandamui, Keller (M) Sdn. Bhd., Kuala Lumpur, Malaysia

Fig. 5 Schematic of DSM treated block

5 DEEP SOIL MIXING TECHNIQUE

Deep Soil Mixing (DSM) technology was invented almost 30 years ago and is a form of
ground improvement involving the introduction and mechanical mixing of in-situ soft soils
with cementitious compound (CDIT 2002). The compound (which is of-ten referred to as
the binder) can be injected into the ground in dry or wet form. A mixing tool is drilled to
the intended depth and then withdrawn to form individual columns (diameter can range
from 0.5m to 1.5m). The tool is lowered and withdrawn at pre-determined rate of rotation,
rate of penetration and rate of withdrawal while delivering the design binder content at a
specified flow rate and pressure. The end product is an improved soil with undrained shear
strength ranging from 0.1MPa to 6.0MPa, depending on the soil type, mixing process and
binder content. Typical applications of deep soil mixing include foundations of embankment
fill for roads and highways, stabilization of excavations, foundations for structures and
subgrade improvement (Topolnicki 2004; Raju & Abdullah, 2005).
The ‘dry’ method is more suitable for soft soils with very high moisture content and hence,
it was used at digesters N1, S1 and S2. Typical picture showing ‘dry’ DSM rig is shown in
Fig. 6.

6
Deep Soil Mixing in Mine Tailings for a 8m Deep Excavation

Fig. 6.”Dry” DSM rig

The ‘wet’ method is more appropriate in mixed soil conditions which are generally stiffer,
with lower water content. Hence, the ‘wet’ method was adopted at digester N2. Typical
picture showing ‘wet’ DSM rig is shown in Fig. 7.

Fig. 7.”Wet” DSM rig

7
Y. W. Yee, V. R. Raju and H. K. Yandamui, Keller (M) Sdn. Bhd., Kuala Lumpur, Malaysia

6 ENGINEERING ANALYSES

The excavation was designed to satisfy the following conditions:

a) A 45 degree slope was to be excavated down to 8m depth below ground.

b) Water seepage into the excavation from the sides and base was required to be minimal.

c) The treated block was to be able to resist flotation forces.

The DSM design (in terms of cement content and mixing parameters) was determined using
theory established by Broms (2004) and in accordance to the design methodology developed
by Swedish Geotechnical Society (SGF Report 4:95E 1997).
Slope stability analyses were carried out employing the composite improved parameters.
The following Fig. 8 shows critical slip circles from the slope stability analyses (Bishop
modified method) to achieve factor of safety above 1.4.

Fig. 8 Typical slope stability analyses

The columns were generally designed to be contiguous (touching columns representing 87%
treatment) in the slime soils to minimise water infiltration i.e. at N1, S1 and S2. In the more
permeable mixed soils of N2, greater precautionary measure was taken by overlapping the
columns (100% treatment).
The DSM block was also designed to be sufficiently massive to overcome potential uplift
forces (see Fig. 9). Besides, the block was checked to be adequately stiff to prevent any shear
type failure. It was also ensured that no tension forces developed within the block using FEM
analyses.

8
Deep Soil Mixing in Mine Tailings for a 8m Deep Excavation

Fig. 9 Design model against flotation forces

7 CONSTRUCTION

The soil beneath the digester footprint was essentially treated to form a massive block with
conical shape geometry. Individual columns of soil were mixed with cement in contiguous
fashion (or with overlap for N2). A typical layout of the treatment scheme is shown in Fig.
10.

Fig.10 Typical Layout showing DSM treatment scheme

As explained in section 5, both the ‘dry’ and ‘wet’ methods of installation were used, due to
the variable soil conditions. Basic parameters of ‘wet’ and ‘dry’ DSM are summarized in
Table 2.

9
Y. W. Yee, V. R. Raju and H. K. Yandamui, Keller (M) Sdn. Bhd., Kuala Lumpur, Malaysia

Parameters ‘Dry’ DSM ‘Wet’ DSM


Column
0.6m 0.87m
diameter
Grid pattern Triangular Triangular
Grid spacing 0.6m 0.75m
Depth of
7 to 13m 10 to 12m
treatment
Design
undrained
150kPa 150kPa
shear
strength (Cu)
Dry Grout slurry
Binder type cement (w/c ratio =
(OPC) 1.0 to 1.5)
Volume of 150 to 200 150 to 200
binder kg/m3 kg/m3
About 3 to About 3 to 4
Curing time
4 weeks weeks

Table 2 Summary of ‘dry’ and ‘wet’ DSM parameters

Construction was carried out with one ‘dry’ rig and one ‘wet’ rig. The mixing tools of ‘dry’
and ‘wet’ DSM is shown in Fig. 11 and Fig. 12, respectively. The DSM works were
commenced in November 2004 and completed in February 2005.

Fig. 11 Process of ‘dry’DSM Fig. 12 Process of ‘wet’ DSM

8 QUALITY CONTROL AND MONITORING DATA

Numerous quality control measures (pre, during and post construction) were implemented
which included:
a) Additional soil investigation to confirm actual soil conditions prior to commencement of
soil improvement works as high lighted in section 3.

b) Trial columns on site prior to construction of working columns.

10
Deep Soil Mixing in Mine Tailings for a 8m Deep Excavation

c) Real-time computerised monitoring of operating process parameters during construction.

d) Post construction testing using Standard Penetration Test (SPT) in the field and horizontal
coring and subsequent un confined compressive strength (UCS) test in a laboratory.

e) Monitoring of lateral displacements using inclinometers during excavation works.

Prior to construction of working columns, trial columns were constructed at designated


locations to confirm operating parameters (cement content, rotation speed, withdrawal rate,
etc.). After allowing for sufficient curing time, the columns were exposed by excavation for
examination. The trial 2.5m deep excavation clearly demonstrated that the previously very
soft / loose soil has been successfully treated to allow a vertical cut (see Fig. 13). The
diameter and consistency of the columns were also proven.

After initial trials, installation of working columns was com-menced with appropriate
operating process parameters. The process parameters were closely monitored during
installation us-ing real time computer and printout. A typical real-time comput-erized
installation record during construction stage is shown in Fig. 14.

Fig. 13. Exposed trial columns and subsequent excavation

11
Y. W. Yee, V. R. Raju and H. K. Yandamui, Keller (M) Sdn. Bhd., Kuala Lumpur, Malaysia

Fig 14 Typical real-time computerised installation record

After sufficient curing period, selected working columns were cored vertically using
conventional sol investigation rig (see Fig. 15) to recover samples of 50mm to 60mm
diameter. It was ob-served that it was difficult to retrieve continuous intact cores samples,
probably because the columns were relatively low strength. Hence, such coring processes
were mostly useful only for visual examination of samples. Standard Penetration Tests (SPT)
were carried out for some columns and were used to con-firm increase in stiffness of the
soil. Typical results are presented in Fig. 15.

Fig. 15. Vertical coring and SPT result

12
Deep Soil Mixing in Mine Tailings for a 8m Deep Excavation

Horizontal coring was found to be more effective in recover-ing continuous intact samples.
After localised excavation, selected working columns were cored horizontally using
hand-held coring machine to recover 100mm diameter samples (see Fig. 16). The solid cored
samples were tested in a laboratory and showed more than acceptable UCS (Unconfined
Compressive Strength) between 1MPa and 3MPa about 1 to 2 months after installation

Fig. 16 Horizontal coring and UCS testing

Cone Penetration Tests (CPT) were attempted but generally gave misleading results. It is
noted that within each column cross section there was strength variability and the CPT
probe was too small to provide representative result for the entire cross section. Such
limitation of the testing method has been observed by others (Larsson 2005).

9 EXCAVATION

Excavation was supposed to have been carried out about one month after installation.
However, this was delayed by 3 months due to other outside factors. Typical exposed DSM
columns dur-ing excavation works are shown in Fig. 17. The intended depth of excavation
was reached safely without incident (see Fig. 18).

13
Y. W. Yee, V. R. Raju and H. K. Yandamui, Keller (M) Sdn. Bhd., Kuala Lumpur, Malaysia

Fig. 17 Typical exposed DSM columns during excavation

Fig. 18 Completed excavation

Water infiltration into the excavation was minimal for N1, S1 and S2. At digester N2,
localised compaction grouting was im-plemented to arrest water inflow where weathered
rock was found and where mixing had to be terminated prematurely.
Inclinometers were installed to monitor lateral displacement during excavation works which
showed minimal lateral movement in the range of 1mm to 3mm (see Fig. 19).

14
Deep Soil Mixing in Mine Tailings for a 8m Deep Excavation

Fig. 19 Results of inclinometer during excavation

8 CONCLUSIONS

Deep Soil Mixing technology has been shown to be effective in the treatment of very soft
mine tailings. The shear strength of the soil was increased by more than 50 times within a
curing period of about 1 to 2 months. The soil treatment enabled 8m deep excavation to be
carried out at a steep 45 degrees slope for four (4) digesters. Water infiltration, within the
high groundwater environment, was largely excluded without the need for any cut-off wall.
The DSM method was proven to be able to provide significant savings in construction cost
and time compared to the conventional method of using sheet piles, anchor tie-backs and
grouting.

ACKNOWLEDGMENTS

The authors wish to thank the Main Contractors, Road Builder Bhd for allowing new
engineering ideas to be implemented in this project. The support provided by the Project
Consultant, Minconsult Sdn. Bhd. is also gratefully acknowledged. Contribution from Keller’s
staff has been significant, notably Dr. How Y.C. & Mr. D. Schrader (design), Mr. U.P. Raju
(project management) and Mr. N. Peterzen (site superintendent).

15
Y. W. Yee, V. R. Raju and H. K. Yandamui, Keller (M) Sdn. Bhd., Kuala Lumpur, Malaysia

REFERENCES
nd
Broms, B.B. 2004. Lime and lime/cement columns. Ground Improvement (2 Edition), Edited by
Moseley, M.P. & Kirsch, K. Spon Press: 252–330.
Coastal Development Institute of Technology (CDIT). 2002. The Deep Mixing Method –
Principle, Design and Construction, A.A. Balkema Publishers, Japan.
Larsson, S. 2005. On the Use of CPT for Quality Assessment of Lime Cement Columns.
Proceedings of the International Conference on Deep Mixing Best Practice and Recent
Ad-vances, Stockholm, Sweden.
Raju, V.R. & Abdullah, A. 2005. Ground Treatment Using Dry Deep Soil Mixing for a Railway
Embankment in Malaysia.
Proceedings of the International Conference on Deep Mixing Best Practice and Recent Advances,
Stockholm, Sweden.
Swedish Geotechnical Society (SGS): SGF Report 4:95E. 1997. Lime and Lime Cement Columns
– Guide for Project Planning, Construction and Inspection, Linkoping, Sweden.
nd
Topolnicki, M. 2004. In situ soil mixing. Ground Improvement (2 Edition), Edited by Moseley,
M.P. & Kirsch, K. Spon Press: 331–428.

16
Technical Note on Ground Improvement using Deep Soil Mixing: Theory & Practice

Annexure IV Technical paper on the project case study on “Excavation


Support for TBM Retrieval Shaft using Deep Soil Mixing
Technique, Kuala Lumpur”, Ir. Yew Weng, Yee and Ir.
Yean Chin &Tan.

23
International Conference & Exhibition on Tunnelling & Underground Space (ICETUS2015) 3-5 March 2015 Kuala Lumpur

Excavation Support for TBM Retrieval Shaft using Deep Soil Mixing Technique,
Kuala Lumpur

Ir. Yew Weng, Yee1 and Ir. Yean Chin, Tan2


1
Keller (M) Sdn Bhd, Kuala Lumpur
2
G&P Professionals, Kuala Lumpur
Email: [email protected]

ABSTRACT: As part of the construction for the Klang Valley Mass Rapid Transit (KVMRT) Line 1, the South Portal structure at Taman
Maluri acted as the shaft to retrieve the Tunnel Boring Machine (TBM) from Cochrane Station. An earth retaining system was required to
retain the 15m deep shaft. Deep soil mixing (DSM) columns were used to provide excavation support instead of conventional secant bored
pile walls in limestone. The DSM columns were constructed to form a gravity block which has no steel reinforcements and allowed
excavation to be carried out without lateral supports. This enabled the TBM to be driven through and retrieved without the need to cut
through reinforcing bars or having to design complex placement of struts.

KEYWORDS: Deep Soil Mixing, Limestone, Excavation, Cement

1. INTRODUCTION bedrock is highly irregular. Adding to the natural complexity of the


ground, the site was a former tin mining area and hence, highly
The Klang Valley Mass Rapid Transit (KVMRT) Project when variable soil composition is to be expected.
completed will cover a distance of 51km and comprise of 31
passenger stations. The South Portal structure at Taman Maluri,
Kuala Lumpur (KL) acted as the transition point between the
elevated and underground sections of the Sungai Buloh - Kajang
line and also, as the shaft for retrieval of the Tunnel Boring
Machines (TBM) from Cochrane Station. The rail level is about
15m below the existing ground level. The ensuing 15m deep
excavation required an earth support system. Conventional secant
pile retaining walls in limestone have to be designed to resist
bending moments and minimise lateral deflection. This is normally
done using steel reinforcing bars and steel struts (or anchors). These
steel elements do not provide convenient exit for the TBM and
hence, an alternative retention system had to be devised.
Deep Soil Mixing (DSM) walls have been used with
increasing regularity in Kuala Lumpur (Yee & Chua, 2008),
especially over KL limestone formation. The advantages stem not
only from the omission of steel reinforcement and lateral bracings,
but when designed as a gravity structure, the DSM wall does not
need to be socketed into limestone rock. However, there is no prior
record of use of this type of wall for mass-transit station boxes and
TBM retrieval shafts. Figure 1a Location Plan (aerial view)
This paper explains the design philosophy adopted for
the DSM retention system. The ensuing construction technique,
including quality control and safeguards, are also explained.
Subsequent performance, during and after TBM break-out, is
assessed, together with suggested improvements for future
application.

2. SITE LOCATION
The site is located within a commercial hub and beside a very busy
Jalan Cheras, in Taman Maluri, Kuala Lumpur (see Figure 1). The
site was a former petrol station. 3-storey commercial buildings are
found just off the south of the site. Construction logistics issues
included limited movement of construction vehicle during traffic
peak hours and tight space constraint.

3. SUBSOIL CONDITION
Based on Geological Map of Selangor, Sheet 94 Kuala Lumpur
1976 and 1993, published by the Mineral and Geoscience
Department, Malaysia, the proposed site is located over Kuala Figure 1b Location Plan (drawing)
Lumpur limestone formation. Ground conditions in limestone areas
are known to be exceptionally challenging (Chan, 1986). Due to the Seven (7) boreholes were conducted on the site during
inherent karstic feature of limestone bedrock, depth of the limestone the design stage. These showed that thickness of overburden soil
varied between 7m and 10m below existing ground level. The soil
International Conference & Exhibition on Tunnelling & Underground Space (ICETUS2015) 3-5 March 2015 Kuala Lumpur

generally comprised of sandy silt with interbedded layers of soft comes to TBM exit point, these steel elements do not provide a
clay. This is typical of former tin mining soil. The soil was of soft to convenient passage. Hence, an alternative retention system was
stiff consistency with SPT blow-counts typically in the range from 2 devised for this retrieval shaft.
to 12 blows/ft. Deep Soil Mixing (DSM) walls are becoming more
Before commencement of site work, further information common in Kuala Lumpur (Yee & Chua, 2008), especially for
on the rock head profile was gathered by conducting a series of excavation over KL limestone formation. When designed as a
probes along the perimeter of the wall and also, perpendicular to the gravity block, steel reinforcements and lateral steel bracings are not
excavation. This enabled a profile of the rock head to be generated required; and rock socketing is also not needed. DSM walls have not
for the purpose of design (see Figure 2). been used for KVMRT station boxes mainly due to lack of case
history for such application. There was concern of medium term
durability of the unreinforced wall elements (as the excavation may
be kept opened for more than 2 years.

Figure 3 Layout Plan of Shaft

Figure 4 Cross Section of Shaft (Section A-A)

5. DESIGN CONSIDERATIONS FOR A DSM WALL


DSM technique involves the process of mixing soil with cement
slurry by using a mechanical tool, which is drilled into the ground.
Figure 2 Rock Profile along Wall Perimeter The mixing tool has cutting blades which are rotated as the tool is
pushed into the ground. Pre-mixed cement grout is pumped at high
4. PROPOSED STRUCTURE pressures through the mixing tool and injected into the soil during
penetration and withdrawal, such that the cement paste and in-situ
At the Maluri South Portal, two numbers of TBM crossed beneath soil are well blended. Through this process, the in-situ soil is
the ground (invert level about 15m below ground surface). At the improved by cement hydration hardening, bonding of soil particles
end of the tunnelling route, a retrieval shaft was required to retrieve and filling of voids by pozzolanic hardening (CDIT, 2002). The end
the two TBMs. The retrieval shaft was to be formed by vertical product will have greatly enhanced strength, low permeability and
excavation retained by an earth support system. As shown in Figure low compressibility compared to the original soil. Typical mixing
3, three faces of the wall need to be retained, the most critical being tool is shown in Figure 5.
the TBM drive exit. The retaining wall was not only required to
resist active earth (and water) pressures but also the TBM thrust
pressures at the drive face induced as the TBM daylights into the
shaft. Figure 4 shows the cross section of the TBM drive exit face.
Conventional station boxes for the KVMRT were
formed using secant bored piles and braced by horizontal struts (or
anchors where space permits). Secant piles have advantages in the
limestone geology as (i) each pile element can be terminated at
different depth (depending on the rock head) after adequate rock
socket is achieved, (ii) the interlocking pile elements minimise
groundwater ingress into the excavation shaft. However, secant pile
retaining walls have to be designed to resist bending moments and
minimise lateral deflection. This is usually achieved by means of Figure 5 DSM Mixing Tool
dense steel reinforcing bars and steel struts (or anchors). When it
International Conference & Exhibition on Tunnelling & Underground Space (ICETUS2015) 3-5 March 2015 Kuala Lumpur

For application at the Maluri Retrieval Shaft, the design minimises uneven movements; and there is less concern of the effect
requirements for the composite wall were reviewed from past of non-uniformity within the block.
projects and literature. Various column configurations were
considered.

5.1 Discrete Columns Arrangement


Individual DSM discrete columns have been shown to perform well
to support vertical load for say, railway embankments (Raju and
Abdullah, 2005) and even foundation rafts (Tolponicki, 2002).
However, such discrete columns are not suitable when required to
resists lateral forces and cases of failure have been documented.
Topolnicki (2004) reported that tensile strength may be as low as
8% of UCS and unlikely to be higher than 200 kPa. Since the
material is rather brittle, lateral shear forces, uneven movements, or
bending stresses may result in failure of the column; and if a series
of discrete columns are lined up together, progressive failure may
ensue (see Figure 6). Hence, the design has to ensure that
compressive stress acting on the columns should not be exceeded
and tensile stress avoided.

Figure 6 Typical Modes of Failure Observed in Centrifuge Tests


(Kitazume et al, 2000) Figure 7 DSM Configuration for Slope Stabilisation (Leong W.K. et
al. 2012)

5.2 Column Grids Arrangement


Topolnicki (2004) describes applications where grids of columns are
constructed and tied together, primarily, to increase rigidity and
reduce risks of progressive failure. Since the risks of subjecting the
DSM wall to bending forces and movements are increased, detailed
analyses using finite element need to be carried out. Strain
compatibility between the soil (to mobilise peak shear strength), the
structure being supported and the DSM elements would need to be
assessed carefully. For example, the failure strain of soft clay is
generally 2% to 5%, whilst the DSM column is less than 1% as
reported by Topolnicki (2004). Leong W.K. et al (2012) describe (a) Cross Section (b) Column Pattern
such detailed considerations for a slope stabilisation application in
soft soils in Singapore. 3D FE analyses were carried out to derive a
DSM wall configuration (about 50% replacement ratio) that not
only fully utilises the advantages of the composite wall but also
isolates the development of tensile stresses (Figure 7). The
constructed wall performed efficiently with only 6mm deflection. It
should be noted that the Mohr-Coulomb soil model should be used
with caution for DSM wall design. At high stress levels, large
volumetric change or strain softening may occur, which cannot be
captured in the model (Lee, 2011).

5.3 Block Arrangement


The Maluri Shaft was designed to retain a vertical cut face up to
10m high. Together with rock excavation beneath, the total
excavated shaft was 15m deep. The consequences of failure were
severe and such risks had to be minimised. As such, the design (c) Wall Layout
intent was to mix the entire block of soil (100% replacement) using
interlocking columns rather than a more economical grid pattern.
Having said that, this DSM wall type was still found to be less Figure 8 (a) Cross Section of DSM Block at the Maluri Shaft (b)
expensive than conventional secant pile wall. The DSM block was Column Pattern (c) Wall Layout
formed by 1m diameter columns overlapping each other by 0.12m
thickness in a honeycomb pattern (see Figure 8). Such a
configuration ensures that the columns experience less stress;
International Conference & Exhibition on Tunnelling & Underground Space (ICETUS2015) 3-5 March 2015 Kuala Lumpur

6. DSM WALL ANALYSES


6.5 Toe stability
The DSM block design was checked against the following failure
modes: The wall needs to be taken to a sufficient depth to prevent toe “kick
out” and basal heave. However, for this project, the rock head is
6.1 Wall overturning stability found at a higher level than the excavation level and hence, this
check was not relevant.
The wall had to be sufficiently wide to ensure that it will act like a
gravity block. No element of the wall will be subjected to tension or 6.6 Groundwater cut-off
bending forces. Although not an issue on this project site,
construction of the wall requires sufficient space (width) behind the The wall has to be effective in reducing water inflow into the
excavation face. Generally, if the depth of soil face is H, then a excavation, both across the wall and also, the interface between the
block width of 0.6H to 0.8H would be required. A factor of safety DSM and rock. Construction of the block was very effective in
against overturning of 3.0 was established (see Figure 9). excluding water, as the size of the block and also, the jagged rock
head decreased permeability manifold. The bigger concern of water
infiltration stemmed from the untreated rock below the DSM block,
with natural fissures and cavities (see Figure 11). It is common to
treat the rock by grouting to reduce water infiltration (see Raju and
Yee, 2006). Rock fissure grouting was carried out before
construction of DSM block at 4m intervals along the excavation
perimeter.

Figure 9 Check for Wall overturning

6.2 Wall sliding resistance


Once again, the wall has to be wide enough to provide resisting
surface against lateral forces. In the limestone, the uneven rock
surface provides excellent interface friction against lateral
movement. However, it should be checked that the rock head profile
does not incline toward the excavation, which may result in lower
resistance than assumed. Sliding check assumed a safety factor of
2.0 (see Figure 10). A check was also made for reduced resistance Figure 11 Rock Fissures where Water may Ingress into Excavation
should the interface be poorly mixed and portions of soils remain.
6.7 Bedrock stability
Pre-construction soil investigation would need to establish that the
rock is stable after excavation (against block failure). This is
established indirectly, by means of examining the rock quality
designation and core recovery ratio. Advice from a geologist is
normally sought. During mining of the rock, the exposed rock face
was examined at each excavation stage. Rock bolts were installed
where there were localised defects found in the rock. Contingency
measures to underpin the block using micropiles were planned but
not found to be necessary for this project.

6.8 Movements (lateral strain)


Figure 10 Check for Sliding along Rock Interface The DSM block was not designed to withstand high levels of strain.
The design had to minimize risks of uneven wall movement.
6.3 Vertical load support
6.9 Blasting force during rock excavation
The block will be carrying temporary construction load. Since the
block is supported on limestone rock, bearing capacity to support Mining of the rock within the excavation block is mainly done by
the load was not an issue. A temporary reinforced concrete slab was controlled blasting. Adjacent to the wall, less invasive mechanical
cast on the block to spread the load, and avoid any concentrated breakers were used. Past work in DSM wall has found that the
point load. Short anchor bars were drilled into the DSM block at cement-soil composite structure can tolerate peak particle velocity
certain designed grids to tie the slab onto the DSM block. A as high as 50 mm/sec without suffering damage. As a general guide,
surcharge load of 20 kPa was assumed. blasting was not carried out within 3m of the wall face.

6.4 Inter-locking bond between column elements 6.10 TBM thrust pressure
The design relies on the effective distribution of soil, water and The North DSM wall block had to be wide enough to resist the
surcharge loads from the back of the wall into the entire block. For thrust forces from the TBM. Besides this, TBM operational
this to happen, load has to be transferred from one element to requirements necessitated extended treatment. Hence, this block was
another via their contact bond (by shear). During construction, it eventually designed to be wider than the others.
was ensured that “cold joints” were avoided and each column was
constructed within 48 hours of the preceding column.
International Conference & Exhibition on Tunnelling & Underground Space (ICETUS2015) 3-5 March 2015 Kuala Lumpur

Table 1 Dimensions of the DSM Walls

Wall Location Depth to Rock Width of DSM Column Dia Column Remarks
(m) Block (m) (m) Interlock (m)

North (TBM 4.0 to 10.0 16.7 1.0 0.12 Design governed by TBM operation
drive) requirements. As-built DSM column
depth varies between 3.5 to 10.5m

East 5.0 to 7.0 9.7 1.0 0.12 Design assumed worst case of 10m soil
depth. As-built DSM column depth
varies between 3.5 to 7.5m

West 5.0 to 7.0 8.8 1.0 0.12 Design assumed worst case of 10m soil
depth. As-built DSM column depth
varies between 3.0 to 10.0m

From the above considerations, it is clear that the design of the The formation of “cold joints” had to be avoided. It was ensured
DSM wall has to consider many practical factors; has to be robust; that corresponding columns were constructed within 48 hours of
and with allowance for a fair amount of redundancies. The installed preceding columns. Where it was anticipated that this was not
DSM columns were required to have a shear strength of 0.75MPa possible or there were ground complications, jet grouting was used
(UCS = 1.5MPa). The final design dimensions of the wall are to form larger columns. For example, in one localized areas, old
summarised in Table 1. timber pile foundations were encountered during installation, which
required jet grouting to be instituted.
7. CONSTRUCTION
Construction of the wall began with site trials to determine the
required cement content and mixing parameters. Based on work by
Topolnicki (2004) and previous experience in KL, a design cement
content between 300 and 350 kg/m3 was adopted. The operating
parameters (e.g. rotation speed, rate of penetration & withdrawal,
blade rotation number, flow rate, grout pressure, binder content,
etc.) were monitored using real-time computerised recording
systems to ensure adequate and uniform mixing of the soil (see
Figure 12). Blade rotation number T, defined as the total number of
mixing blade passing during 1m of single shaft movement through
the soil, was kept above 700 [T = M x (R/V), where M = total
number of mixing blades per m depth; R = rotational speed of
mixing tool; V = penetration or withdrawal rate m/min]. Most
practitioners recommend T > 400 for adequate mixing in normal
application.

Figure 13 Exposed DSM Columns

Figure 12 Typical Operation Parameters Monitored by Computer in


Real Time during Installation

Figure 13 shows the exposed DSM column and the


interface between the column and rock head. Good contact was
achieved by keeping the tool at the deepest penetration level for at
least 0.5 min whilst jetting and rotating.
There are layers of clay within the soil mass. To avoid
formation of soil “bulb”, which impedes thorough soil-cement
mixing, a free blade was introduced (see Figure 14). Figure 14 Free Blade in Mixing Tool to Avoid Formation of Soil
It was imperative that the columns were interlocked Bulb
such that the individual columns combined to act as a single block.
International Conference & Exhibition on Tunnelling & Underground Space (ICETUS2015) 3-5 March 2015 Kuala Lumpur

The advancement of DSM technology has resulted in the behind barricades 20m away from the TBM wall face. A video of
availability of machines using multiple shafts. Recently, the advent the break-out is available for view at http://mymrt-
of cutter-soil mixing has also been met with optimism. However, underground.com.my/videos/breakthrough. Figure 17 shows the
such multiple shaft mixing is not suitable for use in limestone given wall during and after TBM exit.
the pinnacle nature of rock. “Gaps” would result in the soil-rock With regards to wall durability, the DSM wall has been
interface as the blades will be stopped at the highest peak in the standing for 2 years (at the time of writing). Visual examination has
limestone. Hence, to ensure proper mixing down to limestone rock found no signs of distress or degradation. For longer term
head level, a single shaft mixing tool is preferred. application, weathering resistance can be enhanced with higher
cement content (and higher strength of end product). Additional
8. QUALITY CONTROL surface protection measures such as applying a gunite surface (with
or without steel mesh) may also be implemented. In the permanent
The execution practice and quality control of DSM works follow the stage, structural walls and slabs will be constructed in front of the
British Standard BS EN 14679:2005. A quality plan was drawn up, wall for long term serviceability requirements.
which included the methods and frequency of checks to be made In future, for such applications, the designer could
during construction. consider economising the design by reviewing the wall width, blade
As mentioned before, the operating parameters were rotation number and other TBM-related parameters.
monitored using real-time computerised recording systems.
Verification process included daily review of these computer
records and any deviations were investigated and rectified to the
satisfaction of the supervisory team.
Core samples were collected to examine consistency of
the columns and to recover sections for unconfined compressive
strength (UCS) tests. Cores were done in both the centre and the
edge of the columns (where the columns intercept). As shown in
Figure 15, test results show UCS strength consistently above 1.5
MPa after 28 days.

Figure 16 Location of Inclinometers and Settlement Markers

Figure 15 DSM Core Sample UCS Strength Test Results

9. PERFORMANCE
Excavation was carried out after completion of the DSM block
(subsequent to a curing period exceeding 28 days). Bedrock was
mined beneath the block using hydraulic breakers. Rock blasting
was carried out nearby and the wall was monitored to ensure that the
vibrations induced did not do any damage.
Wall movement was monitored during excavation works
using both inclinometers and settlement markers. Three
inclinometers were installed, one on each wall face, and twenty-one
settlement markers on the ground surface (see Figure 16). Maximum
wall movement was observed at East wall, showing a reading of 10
to 15mm (less than 0.15% wall height). The shape of the deflection
implied that there was some sliding movement, albeit small. This is
well within widely accepted wall defection criterion of 0.5%. The
maximum ground subsidence of 2mm to 6mm was observed behind
West wall, and this was probably caused by construction load.
Back-analyses imply a stiffness modulus of the block between 35
and 100 x UCS.
Safety precautions were taken during tunnel break-out.
This included reducing the TBM slurry pressure to very low levels
from usual pressure. The speed of boring was slowed to 30% to
40% of normal speed. In addition to the above, 2 layers of
temporary soil nails were installed to strengthen the front face of the
DSM block during TBM break-out. The actual TBM break-out
occurred on April 8, 2014 (West Tunnel) and April 24, 2014 (East
Tunnel). It was reported by the TBM Operators that both break-outs
occurred smoothly without incident. There were many spectators
viewing the emerging TBMs and they were requested to stand Figure 17 DSM Wall during and after TBM exit
International Conference & Exhibition on Tunnelling & Underground Space (ICETUS2015) 3-5 March 2015 Kuala Lumpur

10. CONCLUSION
The Maluri TBM Retrieval Shaft required an excavation of 15m
deep. DSM walls comprising 1m diameter columns interlocked with
each other were constructed to form a gravity block over three faces
of the excavation pit. The DSM block was formed over soil up to
10m high and seated on limestone bedrock beneath. The design was
approached with caution given the severe consequences of failure.
Besides ensuring high safety factors in various wall stability and
sliding checks, the block was designed to be fully treated with
cement (100%) to avoid any risk of occurrence of tensile and
bending stresses. Strains were kept to a minimum. Construction was
carried out with high level of supervision and control, especially in
ensuring thorough mixing and avoidance of cold joints between
columns. Based on previous experience in similar soils in Kuala
Lumpur, the mix design ensured that cement content up to 350 kg
per m3 and blade rotation number above 700 were achieved. The
DSM block was excavated with maximum 15mm movement and
did not display any duress during the further 5m deep rock
excavation afterwards. The DSM wall performed well as the TBM
break-out events occurred without incident. The DSM wall stood for
2 years before structural walls were constructed as permanent finish.

11. ACKNOWLEDGEMENTS
The authors wish to thank colleagues from G&P and Keller, for
contributing significantly to this project. Their ideas and suggestions
during design and construction were invaluable. Grateful
recognition is given to Dr. Ooi L.H. for his advice during design
development of this project. Appreciation is accorded to PVSR
Prasad for data compilation in the preparation of this paper.

12. REFERENCES
BS EN 14679 (2005): Execution of special geotechnical works –
Deep Mixing
Chan, S. F., and Hong, L. E. (1986). Pile foundation in Limestone
areas of Malaysia, Foundation Problems in Limestone Areas
of Peninsular Malaysia, Geo. Eng. Tech. Div., IEM
Coastal Development Institute of Technology (CDIT). (2002). The
Deep Mixing Method, A.A. Balkema Publisher.
Kitazume, M. , Okano, K. and Miyajima, S. (2000) Centrifuge
model tests on failure envelope of column type DMM
improved ground, Soils and Foundations, Vol. 40, No. 4,
pp.43-55.
Lee, F.H. (2011). Cement-Soil Treatment in Underground
Construction: Some issues and Recent Findings. Seminar on
Infrastructures in Soft Ground – Challenges and Solutions.
BCA Academy, Singapore.
Leong, W.K., Soh, K.K., Yeo, L., Chew, S.H., Leong K.W. and He,
Z.W (2012). Deep Soil Mixing Columns as Retaining
Structure. 18th SEAGC, Singapore.
Raju, V.R. and Abdullah, A. (2005). Ground Treatment Using Dry
Deep Soil Mixing for a Railway Embankment in Malaysia.
Proceedings of the International Conference on Deep Mixing
Best Practice and Recent Advances, Stockholm, Sweden.
Raju, V.R. and Yee, Y.W. (2006). Grouting in Limestone for
SMART Tunnel Project in Kuala Lumpur. International
Conference and Exhibition on Tunnelling and Trenchless
Technology, Kuala Lumpur, Malaysia.
Topolnicki, M. (2002). Deep Mixing Workshop, Tokyo, 15.–
18.10.2002
Topolnicki, M. (2004). In situ Mixing. Ground Improvement. pp.
331-428.
Yee Y.W. and Chua C.G. (2008). Case Studies: Deep Soil Mixing in
Excavation. Seminar on Excavation & Retaining Walls, Kuala
Lumpur. Institution of Engineers Malaysia.
Yee Y.W., Raju, V.R. and Yandamuri, H.K. (2007). Deep Soil
Mixing in Mine Tailings for a 8m Deep Excavation. 16th
SEAGC, Kuala Lumpur. pp 573-578.
Indian Geotechnical Society
TC on Ground Improvement and
Geosynthetics
Technical Note on Ground
Improvement using Grouting
Techniques: Theory & Practice
-by Keller Ground Engineering India Pvt. Ltd

Revision Details:

00 10-11-2016 For Circulation Vani Madan Hari


Init. Sign. Init. Sign. Init. Sign.
Rev. Date Document Details
Prepared Checked Approved
TABLE OF CONTENTS
1 Background .....................................................................................................................................1
1.1 General .....................................................................................................................................1
1.2 Technical Committee ................................................................................................................1
1.3 Brainstorming Session ..............................................................................................................1
2 Grouting Technology .....................................................................................................................1
2.1.1 Types of Grout ................................................................................................................2
2.2 Methods of Grouting .................................................................................................................3
2.2.1 Rock/ Fissure Grouting ...................................................................................................3
2.2.2 Permeation Grouting .......................................................................................................4
2.2.3 Compaction Grouting ......................................................................................................5
2.2.4 Jet Grouting ....................................................................................................................5
2.3 Quality Control & Grout Test. ....................................................................................................6
3 Applications ....................................................................................................................................8
4 Case studies ....................................................................................................................................8
4.1 Case study 1: Seepage control works at Teesta Low Dam – IV, West Bengal, India
(Permeation Grouting) ......................................................................................................................8
4.2 Case study 2: Soil arching, Delhi Metro Rail Corporation, Newdelhi (Compaction grouting) ...9
4.3 Case study 3: Rock Grouting for Deep Shaft for SMART Tunnel (Rock/ Fissure Grouting) . 10
4.4 Case study 4: Settlement control for SMART tunnel (Rock/Fissure Grouting) ...................... 11
4.5 Case study 5: Soil Stabilisation for SMART tunnel (Jet grouting) ......................................... 12
5 Bibliography ................................................................................................................................. 14

LIST OF FIGURES
Figure 1: Application range of grouting methods .....................................................................................2
Figure 2: Penetrability of various grouts ...................................................................................................2
Figure 3: Typical cross section of Rock/Fissure grouting.........................................................................3
Figure 4: Process of Rock/Fissure grouting using P, S & T grouting .......................................................3
Figure 5: Improved ground after completion of Rock/Fissure grouting ....................................................4
Figure 6: Schematic showing permeation grouting methodology ............................................................4
Figure 7: Automatic Injection Containers – Computerised Grout Pumps ................................................5
Figure 8: Process of compaction grouting ................................................................................................5
Figure 9: Execution of jet grouting using T & D system ...........................................................................6
Figure 10: Process of jet grouting ............................................................................................................6
Figure 11: Marsh cone on Grout & Hydrometer test on grout and Water ................................................7
Figure 12: Slump cone test ......................................................................................................................7
Figure 13 Quality control monitoring devices ...........................................................................................7
Figure 14 Quality control records .............................................................................................................8
Figure 15 Cross section of coffer dam & Permeation grouting ................................................................8
Figure 16: Teesta Low Dam .....................................................................................................................9
Figure 17: Schematic of the NATM tunnels under an abandoned Nallah channel; the required and
existing SPT N values are plotted on the right .........................................................................................9
Figure 18 Typical layout of NATM and compaction grouting scheme................................................... 10
Figure 19 Comparison of Design SPT values with Pre and Post SPT Values & Completed structure 10
Figure 20 Typical layout of grout curtain ............................................................................................... 11
Figure 21 NVM shaft & Dry walls after curtain grouting ........................................................................ 11
Figure 22 Cross section of grout holes below rail track ........................................................................ 12
Figure 23 Drilling and grouting beneath rail track ................................................................................. 12
Figure 24 Schematic diagram of grouting schemes at cutter-head intervention locations. .................. 13
Figure 25 Plan & Cross section of Treatment scheme & Grout column after Jet grouting ................... 13

LIST OF TABLES
No table of figures entries found.

Page i
LIST OF ENCLOSURES

Annexure 1 Technical paper on “Ground Improvement Techniques for


Infrastructure Projects in Malaysia” by Dr.V. R. Raju and Y Hari
Krishna.,2008
Annexure 2 Technical paper on “Application of Ground anchors to support deep
excavation and compaction grouting for NATM tunnel construction for
Delhi Metro Rail Corporation (DMRC)” by Jitendra Tyagi, Mohan Gupta,
Ashit Shah & Y. H. Krishna and B.C. Kanth
Annexure 3 Technical paper on “Grouting in Limestone for SMART tunnel project in
Kuala Lumpur, Malaysia” by Dr.V. R. Raju and Ir.Y.W. Yee,2006

Page ii
Technical Note on Ground Improvement using Grouting techniques: Theory & Practice

1 Background
1.1 General
The President Indian Geotechnical Society (IGS) has constituted several Technical
Committees (TCs) in order to contribute substantial technical innovations to serve the
geotechnical community by publishing guidelines in the field of ground engineering. In this
endeavour, IGS formed various TCs to seek support in the preparation of guidelines and
publish them on behalf of IGS. In order to form the guidelines, the modus operandi
suggested by IGS was to conduct brain-storming sessions in local chapters in each of the
selected themes and topics and further to record the proceedings. Each member of the
committee shall have to make a presentation followed by a detailed discussion. The
chairman of each TC will decide the sub-topics on which the theme paper will be presented
by a particular member of the committee, followed by a thorough discussion. The individual
TC will develop guidelines with regard to various fields of Geotechnology on behalf of IGS
who will contribute in a meaningful way to better geotechnical practices in India.

1.2 Technical Committee


With the above background, IGS has identified Ground Improvement and Geosynthetics is
one of the TC and the main objective is to prepare an implementable document for practicing
engineers covering Ground Improvement technology, limitations, codal provisions, case
histories esp. in India with their performances.

1.3 Brainstorming Session


IGS Hyderabad Chapter has taken initiative to support IGS and conducted one-day National
Workshop on Ground Improvement and Geosynthetics on 29th August 2015 in JNTU
premises. Minutes of meeting was prepared and circulated among the TC members. It was
agreed in the meeting that design and construction aspects of ground improvement using
Grouting techniques shall be addressed by Keller Ground Engineering Pvt. Limited (Keller).
This document describes concept, theory, design & construction, performance of ground
improvement (esp. Grouting) for variety of projects executed in Asia.

2 Grouting Technology
Grouting in ground engineering can be defined as controlled injection of material, usually in a
temporary fluid phase, into soil or rock, where it stiffens to improve the physical
characteristics of the ground for geotechnical engineering reasons. “Grouting techniques
started from practice, not from theory”.
Grouting methods are effective in sealing cavities in both coarse and fine fissures in rock,
and sealing pores in granular materials typical of all soils short of clays and very silty sands.
With the development of high-pressure pumps cement grouts came to predominate, and they
were frequently associated with the sinking of mine shafts.
The in situ deep mixing of stabilizers with soft soils to form columns, walls grids or blocks in
the foundation has been developed and applied extensively in civil engineering practice since
the 1970s.
The two types of mixing methods, namely deep mechanical mixing (DMM) and high-
pressured Jet-grout mixing have been used under deep ground conditions. Both methods
rapidly spread nationwide in the 1970s. The method of application range is shown in Figure

1
Technical Note on Ground Improvement using Grouting techniques: Theory & Practice
1.The Deep Mixing Method-DMM uses slurry state or dry powder state stabilizer. During
Grouting, the grout material was ejected at a pressure of 200 bars to cut fill and soil, but it
was capable of obtaining an improved body limited to 0.5 meters in diameter. By 1990,
creating a much larger improved body became a primary concern and a variety of
approaches were tested to achieve this.

Figure 1: Application range of grouting methods

2.1.1 Types of Grout


Two major types of grouting materials are generally available in practice:
(i) suspension-type grouts
(ii) solution-type grouts.
These are used for both impermeation and strength improvement. They develop the strength
and sealing ability when the cement hydrates and cures into a system of interlocking crystals.
Water: cement ratios are in the range of 0.5:1 to 5:1. The lower the water: cement ratio, the
greater the strength of the stabilized mass. Many types of chemical (solution) grouts have
been developed for injecting sands & silts. The more familiar are silica gel, aminioplast,
phenoplast, acrylamide, chrome lignin, vinyl polymers, epoxy and polyurethane etc. The
penetrability of various grout is shown in Figure 2.

Figure 2: Penetrability of various grouts

2
Technical Note on Ground Improvement using Grouting techniques: Theory & Practice
2.2 Methods of Grouting
There are four principal grouting methods as explained below.
 Rock/Fissure grouting
 Permeation grouting
 Compaction grouting
 Jet grouting

2.2.1 Rock/ Fissure Grouting


Rock grouting has a long history of use in dam construction and rehabilitation, and can be
applicable to challenges in mining, tunnelling, rock mechanics and environmental
remediation. Rock grouting is typically performed to reduce the hydraulic conductivity of, or
more appropriately, across a rock mass by injection of grout into the rock's joints and
fissures. Its purpose is to fill up artificially created or naturally existing caves, joints and pores
with remarkable change in the structure of the void system. Rock grouting serves for sealing
and for stabilizing rock and soil. This method is often used to prepare the foundations and
abutments for dams. It usually is done using cementing grouts. The typical cross section of
intrusion grouting is shown in Figure 3.

Figure 3: Typical cross section of Rock/Fissure grouting


The process of Rock/Fissure grouting using Primary(P), Secondary(S) & Tertiary(T) grouting
process is shown in Figure 4.

Figure 4: Process of Rock/Fissure grouting using P, S & T grouting

3
Technical Note on Ground Improvement using Grouting techniques: Theory & Practice
The Improved ground after the completion of Rock/Fissure grouting is shown in Figure 5.

Figure 5: Improved ground after completion of Rock/Fissure grouting

2.2.2 Permeation Grouting


Permeation grouting is the injection of a fluid grout into granular, fissured or fractured ground
to produce a solidified mass by filling grout in voids and fissures to control water flow. Once
the grout cures, the porous soil is transformed into a near solid mass. Although this can
sometimes be done using cement grouts, the void space in most soils is too much small to
permit passage of the Portland cement particles. Hence most permeation grouting is
performed using chemical grouts. This is sometimes called chemical grouting also. The
treated soil had a much lower hydraulic conductivity and is stronger and less compressible
than before. It is often used to form groundwater barriers and to stabilise soils in advance of
making excavations or tunnels. The process of permeation grouting and automatic injection
containers is shown in Figure 6 and Figure 7.

Figure 6: Schematic showing permeation grouting methodology

4
Technical Note on Ground Improvement using Grouting techniques: Theory & Practice

Figure 7: Automatic Injection Containers – Computerised Grout Pumps

2.2.3 Compaction Grouting


Compaction grouting is the injection of very stiff, low slump (25-75mm) mortar-type grout
under relatively high pressures to displace and compact soils in place. It is most effective in
cohesionless soils but can also be effective in finer grained soils where disturbance has
occurred. Grout mix comprises of Portland cement, sand and bentonite (for workability) or
other additives. Compaction grouting is often used to repair structures that had experienced
excessive settlement, since it both improves the underlying soils and raised the structure
back into position. The typical process of compaction grouting is shown in Figure 8.

Figure 8: Process of compaction grouting

2.2.4 Jet Grouting


The jet grouting process consists of the disaggregation of soil or weak rock and its mixing
with, and partial replacement by, a cementing agent; the disaggregation is achieved by
means of a high energy jet of a fluid which can be the cementing agent itself. Because of the
high pressures, this method is usable on a wide range of soil types. This method had been
used for ground water control, underpinning, stabilisation prior to tunnelling etc.
Jet grouting can be executed using
i. T-System (0.8m to 1.2m dia.)
ii. D-System (1.5m to 2.0m dia.)

5
Technical Note on Ground Improvement using Grouting techniques: Theory & Practice
The typical cross section of T & D system and process of get grouting is shown in Figure 9 &
Figure 10

Figure 9: Execution of jet grouting using T & D system

Figure 10: Process of jet grouting

2.3 Quality Control & Grout Test.


In all ground improvement techniques quality control during execution is important to ensure
uniform improvement of the soil. In grouting techniques appropriate quality assurance and
quality control shall be adopted to ensure intended performance of grouting.
The working parameters (e.g. depth, pressure, grout volume, etc.) need to be maintained
and recorded at each stage of grouting to determine the appropriate termination criteria.
Termination of particular stage is considered, when one of the following conditions achieved:
– Pre-determined grout volume is achieved (in accordance with column diameter)
– i.e. volume of each bulb
– Pre-determined grout pressure is achieved (in accordance with depth of treatment)
– i.e. Pressure greater than 20 bars.
– Mortar is overflowing from same grout hole collar
The following test should be done to ensure a good quality of grout
– Density test using hydrometer and mud balance
– Viscosity test using Marsh cone

6
Technical Note on Ground Improvement using Grouting techniques: Theory & Practice
– Strength tests using hand held pocket penetrometer
– Sedimentation test
The marsh cone test, hydrometer test and slump cone test are shown in Figure 11 & Figure
12. The strength of the grout is also checked by unconfined compressive strength on cored
samples. The quality control monitoring devices & records used for grouting are shown in
Figure 13 & Figure 14.

Figure 11: Marsh cone on Grout & Hydrometer test on grout and Water

Figure 12: Slump cone test

Figure 13: Quality control monitoring devices

7
Technical Note on Ground Improvement using Grouting techniques: Theory & Practice

Figure 14: Quality control records

3 Applications
The grouting techniques can be adopted for the following ground engineering applications
 Seepage control (Permeation Grouting)
 Soil Arching (Compaction Grouting)
 Rock Grouting for Deep shaft (Rock/Fissure Grouting)
 Settlement Control (Rock/Fissure Grouting)
 Soil Stabilisation (Jet Grouting)
The technical paper explaining the project case studies on “Ground Improvement
Techniques for Infrastructure Projects” in Malaysia is enclosed in Annexure 1.

4 Case studies
4.1 Case study 1: Seepage control works at Teesta Low Dam – IV, West
Bengal, India (Permeation Grouting)
M/s National Hydro Power Corporation proposed the Teesta Low Dam-IV for a Hydro Power
Plant (160MW) near siliguri in West Bengal, India. A cofferdam was planned as shown in
Figure 15 to divert the river water and allow 20m deep excavation in the river bed.
The river bed alluvium is made of a highly permeable mix of sand, gravels and boulders
followed by bedrock. A grout curtain was required to reduce the permeability up to 10-6 m/s
and to allow the construction of dam foundation in a relatively dry condition. The depth of
bedrock varied between 5m and 20m from top of the cofferdam. The cross section of coffer
dam and Teesta dam under construction is shown in Figure 16.

Figure 15: Cross section of coffer dam & Permeation grouting

8
Technical Note on Ground Improvement using Grouting techniques: Theory & Practice

Figure 16: Teesta Low Dam


Maximum excavation depth : 20m
Avg Permeability achieved after : 10-6 to 10-7 m/s
grouting works

4.2 Case study 2: Soil arching, Delhi Metro Rail Corporation, Newdelhi
(Compaction grouting)
M/s Delhi Metro Rail Corporation (DMRC) is building a metro corridor connecting Central
Secretariat and Qutub Minar. New Austrian Tunneling Method (NATM) was adopted to
construct the proposed tunnels.
The soils at site generally consist of sandy silt fill to 5m depth. The abandoned Nallah
channel was excavated and filled with locally available sandy silt to level the ground. SPT N
values in the sandy silt fill were in the range of 4 to 17, indicating loose to medium dense.
This was followed by medium dense to dense Delhi Silt alluvium layer, with SPT N values
between 20 & 30 to about 26m depth. This is underlain by moderately weathered Quartzite
bedrock.
The presence of loose filled up sandy soils over a stretch of 100m near Saket station (BC
19C package) posed problems with effective soil arching which is required for NATM
construction. Compaction Grouting was adopted to enhance the densities of loose sandy
soils to form effective arching. The schematic of tunnel section with required and design SPT
N value is shown in Figure 17.

Figure 17: Schematic of the NATM tunnels under an abandoned Nallah channel; the required
and existing SPT N values are plotted on the right

9
Technical Note on Ground Improvement using Grouting techniques: Theory & Practice
Figure 18 illustrates the layout and compaction grouting scheme – 2m and 4m square grid of
the NATM tunnels. The comparison of design SPT with pre as well as post SPT and the
completed structure after grouting are shown in Figure 19.

Figure 18 Typical layout of NATM and compaction grouting scheme


Comparision of Design SPT values with Pre and Post SPT
Values

0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34
0

4
Depth (m)

10

12
SPT N values

Design SPT Idealised Pre-SPT


Idealised Post SPT on 4m grid Idealised Post SPT on 2m grid

Figure 19: Comparison of Design SPT values with Pre and Post SPT Values & Completed
structure

Total linear meters executed : 3000 Lin.m


SPT N value after improvement : 20 to 30
The technical paper explaining the project case study on Application of Ground anchors to
support deep excavation and compaction grouting for NATM tunnel construction for Delhi
Metro Rail Corporation (DMRC) is enclosed in Annexure 2.

4.3 Case study 3: Rock Grouting for Deep Shaft for SMART Tunnel (Rock/
Fissure Grouting)
The SMART tunnel was constructed in very challenging geological terrain comprising
cavernous Limestone, with highly permeable subterranean solution channels and cavities.
Dewatering activities can result in groundwater lowering which lead to ground subsidence
and in some cases, formation of sinkholes. Likewise, actions from tunneling works can cause
disturbance to the ground with similar consequence.
The Kuala Lumpur Limestone comprises Upper Silurian marble, finely crystalline grey to
cream thickly bedded, variably dollomitic rock. Karstic features are prevalent in the limestone
formed by movement of water containing carbonate acid (dissolved carbon dioxide).
The nearly 30m deep launch shaft (named North Ventilation Shaft or NVS) is located at the
corner of Chan Sow Lin and Cheras Roads. The two 13m diameter tunnel boring machines
were mobilised from this shaft, one north-bound 6km towards Ampang, while the other

10
Technical Note on Ground Improvement using Grouting techniques: Theory & Practice
headed south 4km to Taman Desa. The depths of the excavation were about 20m and 25m
deep. The schematic section of treatment area is shown in Figure 20.

Figure 20: Typical layout of grout curtain

Figure 21: NVM shaft & Dry walls after curtain grouting
Experience from more than 2 years of grouting at this site has shown that the available
technology is effective in minimizing water seepage and ground disturbance.

4.4 Case study 4: Settlement control for SMART tunnel (Rock/Fissure


Grouting)
The SMART tunnel is a project financed by the Government of Malaysia. The proposed
tunnel will be about 10km when completed (Figure 3). The launch shaft was located at the
corner of Chan Sow Lin and Cheras Roads. Two 12m diameter tunnel boring machines
(slurry shield machines) were utilized, one Tunnel Boring Machine (TBM) moving north-
bound 6km towards Ampang, while the other TBM will move southwards 4km to Taman
Desa. The TBMs will also travel beneath some settlement sensitive areas e.g. rail crossing,
bridge crossing, important highway, beside buildings, etc. Grouting works were instituted in
these areas as preventive measures to mitigate the risk of ground movement.
There are some structures along the TBM path which were deemed to be sensitive to ground
movement that may result from the tunneling works. A program of probing and grouting were
implemented with the objective of detecting and filling any natural rock cavities and large
fissures, which may otherwise, if untreated, result in collapse and large sudden movement.

11
Technical Note on Ground Improvement using Grouting techniques: Theory & Practice
One such facility is the Light Rail Transit (LRT) line near Chan Sow Lin Road. The TBM had
to pass beneath twin rail tracks some 14m below ground shown in Figure 22. The rock level
was about 3 to 8m below ground. Drilling had to be done outside the security fencing beside
the track and as such, almost all the holes were drilled at an incline shown in Figure 23. Very
strict precautionary measures were implemented during the works which included
supervision by the train staff and continuous settlement monitoring of the tracks.

Figure 22: Cross section of grout holes below rail track

Figure 23: Drilling and grouting beneath rail track


The reported incidences of sinkholes and ground subsidence reduced considerably after
grouting works were commissioned.

4.5 Case study 5: Soil Stabilisation for SMART tunnel (Jet grouting)
A tunnel project in Kuala Lumpur involved the construction of a 13m diameter bored tunnel
over approximately 10km stretch. The tunnel will function mainly as a storm water storage
and diversion channel but also incorporates a 3km motorway in a triple deck arrangement.
The geology encountered along the tunnel path was ex-mining soils and limestone formation.

12
Technical Note on Ground Improvement using Grouting techniques: Theory & Practice
The subsoil conditions consist of highly variable mixed soils, comprising mainly of loose silty
sand and sandy silt underlain by highly variable karstic limestone formation. Groundwater
was generally at about 3m to 4m below ground.
The cutter-head of the Tunnel Boring Machine (TBM) required maintenance at regular
intervals (about 150m to 200m). Due to the existence of loose sandy material, there was a
risk of ground disturbance and subsequent ground subsidence, if left untreated. Most of the
cutter-head interventions were located within limestone bedrock and rock grouting was
carried out at some locations depending on quality of the bedrock.
At locations, where cutter-head interventions are located partially in soil stratum and partially
in bedrock, combination of compaction grouting and rock grouting was utilised and shown in
Figure 24. At other locations, where cutter-head interventions are located completely in soil
stratum a capping shield made of “Jet Grout block” was designed to ensure face stability
whilst maintenance of cutter-head was carried out. The treatment scheme of Jet grouting and
grout column is shown in Figure 25.The Jet Grout block was installed from 9m to 28m below
existing ground level. The face of the soil gets stabilized after the grouting process.

Figure 24: Schematic diagram of grouting schemes at cutter-head intervention


locations.

Figure 25: Plan & Cross section of Treatment scheme & Grout column after Jet
grouting
All the cutter-head inventions which were treated using Jet Grout block performed well. The
cutter-head of TBM was parked inside the Jet Grout block and the necessary maintenance
was carried out successfully, to withstand an air pressure of 1 to 2 bars without any pressure
drop.
The technical paper explaining the project case study on “Recent Experiences with cement
grouting and mixing techniques in Kuala lumpur” is enclosed in Annexure 3.

13
Technical Note on Ground Improvement using Grouting techniques: Theory & Practice

5 Bibliography
1. Dr. A.V. Shroff. “Development in Design & Execution in Grouting Practice”, Indian
Geotechnical Society 31st Conference (2009).
2. Dr Satyendra Mittal “An Introduction to Ground Improvement Engineering”, Indian
Institute of Technology, Roorke (2013)”.
3. Dr. V. R Raju, Ir. Y.W Yee. (2006) “Grouting in Limestone for SMART Tunnel project”,
International conference and Exhibition on tunneling and trenches technology, Kuala
Lumpur, Malaysia
4. V.R Raju and Y Hari Krishna, (2008) “Ground improvement Techniques for
infrastructure projects in Malaysia”, Proceedings of 12th International conference of
International Association for computer methods and Advances in Geomechanics,
Goa.
5. Dr. V.R. Raju & Ir. Yee Yew Weng (2005) “Recent Experiences with cement grouting
and mixing techniques in Kuala lumpur” IEM-GSM Oktoberforum2005: Case Histories
in Engineering Geology and Geotechnical Engineering, 4th Oct. 2005, Petaling Jaya.
6. Y. H. Krishna and B.C. Kanth ,Jitendra Tyagi, Mohan Gupta & Ashit Shah
“Application of Ground anchors to support deep excavation and compaction grouting
for NATM tunnel construction for Delhi Metro Rail Corporation (DMRC)”.

14
Technical Note on Ground Improvement using Grouting techniques: Theory & Practice

Annexure 1 Technical paper on “Ground Improvement Techniques for


Infrastructure Projects in Malaysia” by Dr.V. R. Raju and
Y Hari Krishna.,2008.

15
The 12th International Conference of
International Association for Computer Methods and Advances in Geomechanics (IACMAG)
1-6 October, 2008
Goa, India

Ground Improvement Techniques for


Infrastructure Projects in Malaysia

V. R. Raju
Keller Far East, Singapore

Y. Hari Krishna
Keller Ground Engineering (India) Pvt. Ltd.

Keywords: Vibro methods, Deep Soil Mixing, Grouting techniques, applications in infrastructure projects

ABSTRACT: Ground improvement techniques utilising Vibro methods, Deep Soil Mixing and Grouting
technologies are finding increasing application in Malaysia to solve a broad spectrum of geotechnical problems.
This paper will describe recent applications in Malaysia for four separate projects – Jet Grouting to form stable
cutter-head interventions for a tunnel project; Deep Soil Mixing to support deep vertical basement excavation
with limestone interface for a commercial complex; Vibro Concrete Columns to found reinforced concrete tanks
in former domestic landfill for a sewage treatment plant; Vibro Stone Columns to support high reinforced soil
walls for a highway project. The importance of quality control measures are emphasized and available proving
methods are also discussed. The case histories presented demonstrate that the techniques can provide
effective solutions to challenging engineering problems.

1 Introduction
Malaysia has seen extensive growth for the past one decade with many infrastructure projects in the
construction industry. Current technology affords many ground improvement techniques to suit a variety of soil
conditions, structure types and performance criteria. These ground improvement techniques can offer alternative
foundation systems to the conventional pile foundation systems. For more details on various available ground
improvement techniques, the reader is referred to “Ground Improvement 2 nd Edition” book edited by Moseley &
Kirsch (2004).

This paper illustrates four recent case histories in Malaysia, where innovative ground improvement techniques
were employed to suit varying needs of application type and performance criteria. The chosen techniques varied
from Jet Grouting, Deep Soil Mixing, and Vibro Concrete Columns to Vibro Stone Columns as shown in the
Figure 1.

The construction methodology and quality control procedures during execution of works were in accordance with
relevant Code of Practices (e.g. BS EN 12716:2001, BS EN 14679:2005, BS EN 14731:2005, etc.). These
techniques offered reasonably environmental friendly solutions, especially in urban areas.

Figure 1. Schematic showing various ground improvement techniques.


The 12th International Conference of
International Association for Computer Methods and Advances in Geomechanics (IACMAG)
1-6 October, 2008
Goa, India

2 Recent case histories in Malaysia


This paper will describe following four different case histories from four separate projects; where ground
improvement techniques were utilised to solve challenging problems in difficult ground conditions:
a) Jet Grouting to form stable tunnel boring machine cutter-head interventions for a tunnel project.
b) Deep Soil Mixing to support vertical basement excavation over limestone for a commercial complex.
c) Vibro Concrete Columns to support reinforced concrete process tanks in a former domestic landfill for a
sewage treatment plant.
d) Vibro Stone Columns to support high reinforced soil walls for a highway project.

3 Application of Jet Grouting

3.1 Background
A tunnel project in Kuala Lumpur involved the construction of a 13m diameter bored tunnel over approximately
10km stretch. The tunnel will function mainly as a storm water storage and diversion channel but also
incorporates a 3km motorway in a triple deck arrangement. The geology encountered along the tunnel path was
ex-mining soils and limestone formation. For more details of the project, the reader is referred to Raju & Yee
(2006).

The cutter-head of the Tunnel Boring Machine (TBM) required maintenance at regular intervals (about 150m to
200m). At such TBM stops (referred as “cutter-head intervention”), the slurry pressure will be switched off and
the stability of the rock/soil face in front of the TBM relies on air pressure and inherent strength of the in-situ
rock/soil. Due to the existence of loose sandy material, there was a risk of ground disturbance and subsequent
ground subsidence, if left untreated. Most of the cutter-head interventions were located within limestone bedrock
and rock grouting was carried out at some locations depending on quality of the bedrock. At locations, where
cutter-head interventions are located partially in soil stratum and partially in bedrock, combination of compaction
grouting and rock grouting was utilised. At other locations, where cutter-head interventions are located
completely in soil stratum a capping shield made of “Jet Grout block” was designed to ensure face stability whilst
maintenance of cutter-head was carried out. Figure 2 represents the schematic of different types of grouting
schemes implemented depending on the geological conditions. The subsequent sections explain the details of
grouting scheme using large diameter Jet Grout columns to form a stable block in the soil stratum.

Figure 2. Schematic of grouting schemes at cutter-head intervention locations.

3.2 Soil conditions


In general, the subsoil conditions consist of highly variable mixed soils, comprising mainly of loose silty sand and
sandy silt underlain by highly variable karstic limestone formation (see Figure 3). Standard penetration test
(SPT) blow counts typically vary from 0 blows/0.3m (especially along “slump zones” above rock-head) to 20
blows/0.3m. Historically, mining activities took place at some of the sections which explain the varying nature of
the soil. Groundwater was generally at about 3m to 4m below ground.
The 12th International Conference of
International Association for Computer Methods and Advances in Geomechanics (IACMAG)
1-6 October, 2008
Goa, India

Figure 3. Typical geology along the tunnel path.

3.3 Solution
The capping shield made of Jet Grout columns was designed to form a stable block at the cutter-head
intervention locations as shown in Figure 4. The shield was formed using 2 rows of 2m diameter Jet Grout
columns in front of the cutter-head of TBM. The front row (Line-B) was designed to be full depth section,
whereas back row (Line-A) designed to be hollow section to ease the cutting process and also for economy
reasons. The Jet Grout block was installed from 9m to 28m below existing ground level.

Figure 4. Schematic of Jet Grout block at cutter-head intervention location.

The construction challenges of the Jet Grout block were as follows:


a) Formation of consistent 2m diameter Jet Grout columns in the highly variable soil.
b) Proper interlocking of each Jet Grout column down to 28m depth which requires the verticality of drilling to
be within 0.5% to 1%.
c) Required minimum unconfined compressive strength (UCS) of 1MPa for each Jet Grout column.
d) Existing underground utilities which required careful attention to avoid damages.
e) High power transmission towers which limited the working head-room and associated safety issues.
The 12th International Conference of
International Association for Computer Methods and Advances in Geomechanics (IACMAG)
1-6 October, 2008
Goa, India

A trial was performed prior to the commencement of working columns to confirm the erodability of in-situ soils
and adequacy of operating parameters. The jetted column was exposed and core samples were taken to verify
the as-built diameter and achieved strength. The diameter formed was proven to be more than 2m and UCS was
more than 1MPa. The site pictures showing exposed trial columns and execution of working columns using HT
400 pump and D-system are shown in Figure 5.

During construction, as part of quality control measures, the density of backflow from Jet Grouting works were
monitored which indirectly reflected the erodability of soil and the diameter of Jet Grout column formed. Based
on the laboratory test results, the required CEM soil (i.e. the remaining cement content in the Jet Grout column)
to achieve UCS of 1MPa was about 200 to 250 kg/m 3 for sandy soils and 350 to 400 kg/m 3 for clayey soils.

Figure 5. Exposed trial Jet Grout columns (left) and execution of working columns (right).

3.4 Performance
All the cutter-head inventions which were treated using Jet Grout block performed well. The cutter-head of TBM
was parked inside the Jet Grout block and the necessary maintenance was carried out successfully, to withstand
an air pressure of 1 to 2 bars without any pressure drop.

4 Application of Deep Soil Mixing

4.1 Background
A project comprising 3-storey commercial complex with 2-level basement car park floors (about 7m depth below
existing ground level) is under construction in the middle of Kuala Lumpur City Centre. The project site is
confined between a newly completed 4-storey commercial lots, light rail transit track and existing old warehouse
(see Figure 6). The distance between face of excavation and boundary setback line is in the range of 3m to 10m,
hence open sloped excavation was limited to shallow rock-head areas.

The proposed 2-level basement construction required 7m deep excavation with underlying limestone interface
for a total perimeter length of about 690m. The conventional solution using contiguous bored piles or anchored
sheet piles proved very expensive and needed prolonged construction period. As an alternative, a rigid gravity
wall retaining system using interlocked Deep Soil Mixing (DSM) columns was implemented around the perimeter
length of about 560m as shown in Figure 6.

Figure 6. Overall layout of proposed basement excavation.


The 12th International Conference of
International Association for Computer Methods and Advances in Geomechanics (IACMAG)
1-6 October, 2008
Goa, India

4.2 Soil conditions


The subsoil comprised of loose silty sand deposits and ex-mining soils with SPT values in the range of 5 blows/ft
to 12 blows/ft. Underlying this loose soil layers, karstic limestone formation was found with extremely varying
rock-head levels ranging between 3m and 15m below existing ground level. Overhanging boulders and
pinnacles are common; hence the founding level of the bedrock formation was unpredictable. The ground water
table was found to be at about 1m to 2m below existing ground level.

4.3 Solution
The gravity wall block was designed to ensure adequate resistance against lateral earth pressure to support the
intended depth of excavation, whilst reducing seepage water inflow and thus, minimise the possible risk of
drawdown and consequent ground subsidence to the surroundings. The design of gravity wall required a width of
0.7 times the depth of overburden soil above rock-head level. The gravity wall acted as a temporary retaining
structure during the basement excavation works. Wet DSM columns of 0.85m diameter were interlocked at
0.75m centres to form the rigid gravity wall block as shown in Figure 7.

Figure 7. Schematic of DSM gravity wall block.

The columns were designed to achieve an unconfined compressive strength of 1.0MPa with binder content
(Ordinary Portland Cement with water-cement ratio of 1:1) in the range of 200kg/m 3 to 250kg/m3.The columns
were installed to a maximum depth of 12m below existing ground level. For locations, where there was space
constraint, shear pins were installed to provide wall stability. A picture of the site with on-going installation works
is shown in Figure 8. The operating parameters (e.g. rotation speed, rate of penetration and withdrawal, blade
rotation number, flow rate, grout pressure and binder content, etc.) were monitored using real-time computerised
recording systems to ensure adequate and uniform mixing of the soil.

Figure 8. Execution of DSM works.


The 12th International Conference of
International Association for Computer Methods and Advances in Geomechanics (IACMAG)
1-6 October, 2008
Goa, India

4.4 Performance
Excavation works proceeded upon completion of DSM installation works and subsequent curing period of only
14 days due to tight schedule of the project. Bedrock underneath the DSM columns was excavated using the
hydraulic breaker and blasting works. The installed DSM columns were able to withstand the high vibration
induced by rock excavation works. At the time of writing this paper, approximately 60% of the excavation works
have been completed (see Figure 9). As part of quality control procedure, cores from DSM columns were
extracted and tested in a laboratory for UCS. The test results indicated an UCS in the range of 1MPa to 3MPa. In
addition, wall movement was monitored during excavation works, which showed a maximum horizontal
movement of about 30mm to 40mm.

Figure 9. Completed excavation.

5 Application of Vibro Concrete Columns

5.1 Background
A Sewage Treatment plant is under construction in Penang Island and when completed will cater for an ultimate
capacity of 1.2 million population equivalent. The project will serve as a centralized sewage treatment facility and
will include 12 nos. of Sequential Batch Reactor (SBR) tanks and associated process tanks (see Figure 10).

Figure 10. Overall plan layout of sewage treatment plant.

The SBR tanks are major process tanks in the entire plant and were designed as twin tanks made up of
reinforced concrete (total 6 nos. of twin tanks separated by very narrow gap) supported on treated ground. The
dimension of each twin tank is approximately 90m x 60m x 7m high. One of the twin tank (SBR 1&2) has
additional 2 floors on top of the tank to accommodate administration office and storage area for process
equipment. At the time of writing this paper, the building works are almost completed, whilst mechanical and
process installation works are ongoing.
The 12th International Conference of
International Association for Computer Methods and Advances in Geomechanics (IACMAG)
1-6 October, 2008
Goa, India

5.2 Soil Conditions


The site is located on the north-eastern part of Penang Island in Jelutong, about 5 km from Georgetown. The
site was reclaimed from the sea and approximately, half of the SBR tanks area was covered by former domestic
landfill (3m to 5m thick) waste dumps. The subsoil primarily consists of 3m to 5m thick reclaimed fill / domestic
waste dumps followed by 5m to 7m thick soft marine clay. This is followed by stiff to very stiff cohesive deposits
to over 50m depth. The ground water table varied between 1m and 2m below existing ground level.

5.3 Solution
The original foundation design was piled foundation to over 40m depth; but this was later found to present a few
undesirable construction limitations like noise pollution during pile driving; requirement of pre-boring and removal
of landfill material; and transportation and storage of pre-cast piles on a congested site; as well as relatively high
cost. As an alternative, ground improvement techniques (Vibro Concrete Columns and Deep Soil Mixing) were
utilised to support the SBR tanks. Vibro Concrete Columns (VCC) were constructed for 3 nos. of twin tanks
(namely; SBR 1&2, SBR 3&4 and SBR 7&8) in the former landfill area, forming concrete pile-like elements by
displacing the domestic waste dumps rather than requiring removal. DSM columns were constructed for
remaining 3 nos. of twin tanks (namely; SBR 5&6, SBR 9&10 and SBR 11&12) in the non-landfill area.

The alternative foundation system was designed to ensure adequate bearing capacity (to support loading
intensity of 92kPa), limit the total settlement of the structure to be less than 75mm and differential settlement to
be less than 1(V):360(H). The diameter of Vibro Concrete Columns varied between 0.6m and 0.75m with
working loads of 35tons and 50tons, respectively. Typical spacing of columns (0.6m diameter) ranged between
1.8m c/c and 1.6m c/c to support foundation loads of 90kPa and 130kPa, respectively. The depth of columns
varied from 8m to 14m. The design mixture of concrete as follows:
a) Cement content ~ 200kg/m3
b) Water-cement ratio (w/c) ~ 0.5
c) Grading of aggregates ~ 8mm to 20mm

The picture showing execution of Vibro Concrete Columns using custom-built machine (Vibrocat) is shown in
Figure 11. During execution works, appropriate quality control procedures (e.g. cube strength tests, concrete
consumption and adequate compaction effort, etc.) were implemented on site.

Figure 11. Picture showing execution of VCC.

5.4 Performance
After successful execution of VCC works, the columns were exposed and it was demonstrated that the domestic
waste material was displaced sideways during installation and did not contaminate the concrete. Selected
working columns were tested up to 1.5 times the working load using plate load tests in 3-cycles.

As part of quality control, coring was carried out through selected working columns to retrieve the samples of
50mm to 100mm diameter. The retrieved samples were tested for UCS and results of tests showed UCS in the
range of 10MPa to 40MPa, which is much more than design strength of 5MPa (see Figure 12).
The 12th International Conference of
International Association for Computer Methods and Advances in Geomechanics (IACMAG)
1-6 October, 2008
Goa, India

Results of UCS Tests on VCC Cores - Jelutong STP (Penang)

49.6
50
44.6
45

40
36.5

35 33.5

30 28.0 28.5
UCS (MPa)

27.5

25 22.7 23.4
22.5
21.0
20 17.3 17.9
17.0 16.8 16.2 15.4 14.7
15 12.9
11.9
10

5 Design UCS = 5MPa

0
SBR 1a SBR 1b SBR 2a SBR 2b SBR 3a SBR 3b SBR 4a SBR 4b SBR 7a SBR 7b SBR 8a SBR 8b TNB
Structure Reference

Inclined Cores (100mm dia.) Vertical Cores (54mm dia.)

Figure 12. Typical results of UCS tests.

After quality control and testing works, the concrete structures were constructed according to the specifications
incorporating a load distribution layer (150mm thick well compacted crusher run) between foundation and super
structure. A Hydro test was carried by filling the water into the twin tank with uniform water levels in each tank.
The geotechnical objective of the Hydro test was aimed to check the settlement performance of the foundation
system under full water load prior to the actual operational stage. The rate of water filling was about 0.5m per
day and the design load was maintained for minimum 2 weeks rest period after reaching to the full height. Upon
completion of Hydro test, half of the twin tank was emptied, whilst maintaining the full water load in the other half
to simulate the loading and unloading sequences during operational stage. Pictures showing completed tanks
before and during Hydro test are shown in Figure 13.

Figure 13. Completed SBR tanks – before and during Hydro test (Left: SBR 7&8 and Right: SBR 3&4 in the
foreground and SBR 1&2 with 2 floors on top in the background).

The settlements were monitored using precise survey instruments during and after Hydro tests. The Hydro tests
for 3 nos. of twin tanks (SBR 1&2, SBR 3&4 and SBR 7&8) supported on VCC foundation have been
successfully completed. The settlement monitoring data over the past 10 months period (Sept’06 – Jul’07) has
indicated good performance with maximum settlements in the range of 5mm to 20mm. The typical results of
settlement monitoring for SBR 7&8 are shown in Figure 14. As expected, a marginal elastic rebound was
observed during unloading process (see Figure 14).
The 12th International Conference of
International Association for Computer Methods and Advances in Geomechanics (IACMAG)
1-6 October, 2008
Goa, India

Avg. Water Height (SBR 7&8)


7
6
Water Height (m) 5
4
3
2
1
0
09-Nov-06
23-Nov-06

08-Nov-07
04-Jan-07
18-Jan-07

07-Jun-07
21-Jun-07
05-Jul-07
19-Jul-07
01-Mar-07
15-Mar-07
29-Mar-07
14-Sep-06
28-Sep-06

07-Dec-06
21-Dec-06

01-Feb-07
15-Feb-07

12-Apr-07
26-Apr-07

02-Aug-07
16-Aug-07
30-Aug-07
13-Sep-07
27-Sep-07
12-Oct-06
26-Oct-06

10-May-07
24-May-07

11-Oct-07
25-Oct-07
SBR 7 Water SBR 8 Water

Settlement of Top Monitoring Points (SBR 7&8)

10-May-07
24-May-07
14-Sep-06
28-Sep-06

09-Nov-06
23-Nov-06
07-Dec-06
21-Dec-06

02-Aug-07
16-Aug-07
30-Aug-07
13-Sep-07
27-Sep-07

08-Nov-07
01-Mar-07
15-Mar-07
29-Mar-07
01-Feb-07
15-Feb-07
12-Oct-06
26-Oct-06

04-Jan-07
18-Jan-07

07-Jun-07
21-Jun-07

11-Oct-07
25-Oct-07
12-Apr-07
26-Apr-07

05-Jul-07
19-Jul-07
Cum. Settlement (mm)

-5

-10

-15

TMP-10 TMP-11 TMP-12 TMP-13 TMP-14 TMP-15


TMP-16 TMP-17 TMP-18 Average

Figure 14. Typical results of settlement monitoring (SBR 7&8).

6 Application of Vibro Stone Columns

6.1 Background
The highway network in the capital city of Malaysia (Kuala Lumpur) has seen remarkable growth in the recent
years. Most of the modern expressways were constructed on a Build, Operate and Transfer (BOT) basis. One
such modern expressway with dual three-lane carriageway was opened to the traffic in April 2004. The
expressway forms the main interchange at Kampung Pasir Dalam (referred as Pantai Dalam Interchange) to
connect three distinct routes in the city (namely; Subang Jaya, Jalan Bangsar and Jalan Kuchai Lama). Due to
site constraints at the interchange, high reinforced soil walls were constructed to form the bridge approaches
and other ramps to the required design heights (maximum up to 13m). The following Figure 15 represents the
detailed plan layout of Pantai Dalam Interchange including instrumentation monitoring scheme. For more details
of the project, the reader is referred to Yandamuri & Yee (2006).

Figure 15. Plan layout of interchange.


The 12th International Conference of
International Association for Computer Methods and Advances in Geomechanics (IACMAG)
1-6 October, 2008
Goa, India

6.2 Soil Conditions


The subsoil conditions at Pantai Dalam Interchange varied from very soft silts to soft sandy silts down to a depth
between 5m and 12m followed by hard sandy silts. Typical plot showing results of cone penetration tests is
shown in Figure 16.

Figure 16. Result of typical cone penetration test at Ramp C.

6.3 Solution
Increasingly, Vibro Stone Columns are used to support reinforced soil walls. The combination has proven
economy and has intrinsic technical advantages, i.e. the stone columns ensures relatively quick consolidation as
the embankment is built; while the wall is constructed in stages (lifts) with the wall panels placed progressively
and adjusted for any movement. For details of past Vibro applications in Malaysia, the reader is referred to Yee
& Raju (2007).

For Pantai Dalam Interchange, the design scheme comprised of 1.0m diameter columns spaced at 1.5m to 1.8m
c/c under reinforced soil walls and 2.0m to 2.3m c/c under earth fill embankments (i.e. area replacement ratios in
the range of 15% to 35%). The columns were installed to a depth between 5m and 12m to treat very soft silts
and soft sandy silt deposits. The following Figure 17a shows the site conditions during installation of Vibro Stone
Columns, whereas Figure 17b shows completed reinforced soil wall (about 13m high) at the same location.

Figure 17a. During construction. Figure 17b. After completion.

In total, an area of approximately 23,000m 2 was treated with proper quality control measures to ensure design
diameter and compaction effort throughout the construction process. The installation works were successfully
carried out even adjacent to existing dwellings and very close to the constructed bridge abutments. Vibration
monitoring was carried out for such locations and the measured vibration levels in terms of peak particle velocity
were less than 20mm/s even when Keller’s Mono vibrator was working 1.0m away from the monitoring point (see
Figure 18). The British and Australian standards (BS 5228 Part 4 and AS 2187) accept vibration levels between
20mm/s and 50mm/s for normal structurally sound structures.
The 12th International Conference of
International Association for Computer Methods and Advances in Geomechanics (IACMAG)
1-6 October, 2008
Goa, India

25

Peak Particle Velocity


20

(mm/s)
15

10

0
0 2 4 6 8
Distance from Vibration Source (m)

Figure 18. Vibration monitoring during installation works using Keller’s Mono vibrator.

6.4 Performance
After completion of ground improvement works using Vibro Stone Columns in June 2003, construction of
reinforced soil walls and embankments were commenced. The long-term performance of the treated ground to
support high reinforced soil walls is evaluated based on the results of instrumentation monitoring for more than 3
years, both during construction of embankment and operational stages. The data of settlement measurements
showed that the Vibro Stone Columns has provided effective drainage paths to dissipate excess pore water
pressures under the newly placed fill loads by means of radial and vertical consolidation processes. The time
rate of consolidation was also relatively quick; 90% degree of consolidation was achieved within construction
period of embankment itself.

The embankments and reinforced soil walls were constructed with a rate of filling of about 1m per week and the
highest sections (about 13m high) were completed in about 3 months period. Most of the predicted settlements
occurred during the construction period (see Figure 19) leaving minimal residual settlements for the post
construction stage. The treated ground settled to a maximum of about 100mm only even under 13m high
reinforced soil wall.

PDI : Settlement Markers (SM-3, SM-4, SM-5, SM-6 & SM-7)


36
34
Embankment RL (m)

32
30
28
26
24
22
20
0 90 180 270 360 450 540 630 720 810 900 990 1080 1170 1260 1350 1440
Time (Days)

SM-3 SM-4 SM-5 SM-6 SM-7

Time (Days)
0 90 180 270 360 450 540 630 720 810 900 990 1080 1170 1260 1350 1440
Settlement (mm)

-50

-100

-150

SM-3 SM-4 SM-5 SM-6 SM-7

Figure 19. Summary of results of settlement markers.

Inclinometers were installed at highest reinforced soil wall locations to monitor lateral displacements below
original ground level, both during construction of the wall and post construction stages. Figure 20 below shows
the results of inclinometer measurements at two different locations (near bridge approaches), the results of
which indicated less than 30mm lateral displacement (i.e. less than 1/3 rd of vertical displacement).
The 12th International Conference of
International Association for Computer Methods and Advances in Geomechanics (IACMAG)
1-6 October, 2008
Goa, India

Inclinometer (I-1), AXIS A Inclinometer (I-2), AXIS A


Cumulative Lateral Deflection (mm) Cumulative Lateral Deflection (mm)
-50 -40 -30 -20 -10 0 10 20 30 40 50 -50 -40 -30 -20 -10 0 10 20 30 40 50
0 0
1 1
2 2
3 3

Depth (m)
Depth (m)

4 4
5 5
6 6
7 7
8 8
9 9
10 10

Figure 20. Results of inclinometers (lateral displacements).

7 Conclusions
In Malaysia, ground improvement techniques are finding increasing applications in infrastructure projects. Many
ground improvement techniques are available to suit the particular needs of soil type, structure type, application
type and performance criteria. These techniques offer cost effective solutions, whilst reducing construction
period considerably. Furthermore, these techniques also offer environmental friendly systems, which is important
for urban areas.

The case histories presented in this paper have demonstrated their effective usage. A Jet Gout block was
successfully utilised to form stable ground for a 13m diameter tunnel boring machine cutter-head interventions.
Gravity retaining wall formed using Deep Soil Mixing was utilised for a 7m deep basement excavation support
(strut free) over pinnacled limestone. Vibro Concrete Columns supported concrete process tanks in a former
landfill area without need for removal of domestic waste dumps. Last but not least, Vibro Stone Columns
supported reinforced soil walls up to 13m height over formed tin-mined soils. The techniques enabled innovative
solutions to be applied, which relied on design according to methods recommended in relevant Code of
Practices; proper quality control measures during construction, suitable post construction testing methods and
long-term instrumentation monitoring.

8 Acknowledgements
The author wish to acknowledge the management and staff of Main Contractors and Consults for their valuable
contribution in the implementation of the ground improvement works. The author also wishes to acknowledge
colleagues in Keller who contributed immensely in the design and construction, and not forgetting: Mr. Saw
Hong Seik, Mr. Chua Chai Guan, Mr. P. Sreenivas and Mr. Joe Chang for their valuable input in the preparation
of this paper. The review of this paper by Mr. Yee Yew Weng is also greatly acknowledged.

9 References
Australian Standard: AS2187:1993. Explosive Code.
British Standard: BS 5228-4:1992. Noise and Vibration Control on Construction and Open Sites. Code of Practice for Noise
and Vibration Control applicable to Piling Operations.
British Standard: BS EN 12716:2001. Execution of special geotechnical works – Jet grouting.
British Standard: BS EN 14679:2005. Execution of special geotechnical works – Deep mixing.
British Standard: BS EN 14731:2005. Execution of special geotechnical works – Ground treatment by deep vibration.
Moseley M.P., Kirsch K. 2004. Ground Improvement (2nd Edition). Published by Spon Press.
Raju V.R., Yee Y.W. 2006. Grouting in Limestone for SMART Tunnel Project in Kuala Lumpur. International Conference and
Exhibition on Tunneling and Trenchless Technology, Kuala Lumpur, Malaysia.
Yandamuri H.K., Yee Y.W. 2006. Performance of A High Reinforced Soil Wall Supported on Vibro Stone Columns. GSM-IEM
Oktoberforum 2006 on Engineering Geology and Geotechnical Engineering, Petaling Jaya, Malaysia.
Yee Y.W., Raju V.R. 2007. Ground Improvement Using Vibro Replacement (Vibro Stone Columns) – Historical Development,
Advancements and Case Histories in Malaysia. 16th Southeast Asian Geotechnical Conference, Kuala Lumpur, Malaysia.
Technical Note on Ground Improvement using Grouting techniques: Theory & Practice

Annexure 2 Technical paper on “Application of Ground anchors to


support deep excavation and Compaction Grouting for
NATM tunnel construction for Delhi Metro Rail Corporation
(DMRC)” by Jitendra Tyagi, Mohan Gupta, Ashit Shah & Y.
H. Krishna and B.C. Kanth.

16
APPLICATION OF GROUND ANCHORS TO SUPPORT DEEP EXCAVATION
AND COMPACTION GROUTING FOR NATM TUNNEL CONSTRUCTION FOR
DELHI METRO RAIL CORPORATION (DMRC)

Authors: Jitendra Tyagi, CPM, Delhi Metro Rail Corporation,


Mohan Gupta, Divisional Director, Mott MacDonald Pvt Ltd,
Ashit Shah, Project Director, Mott MacDonald Pvt Ltd,
Y. H. Krishna and B.C. Kanth, Keller Ground Engineering India Pvt. Ltd.

ABSTRACT

Metro rail construction is planned and is underway in several cities in India, including New
Delhi, Mumbai, Chennai and Bengaluru. In New Delhi, the Delhi Metro Rail Corporation
(DMRC) has successfully completed and commissioned the 1st phase of the metro network
covering 65 km. As part of its 2nd phase, construction of about 121 km of metro network is
almost complete. This includes an exclusive “Airport Metro Express Line”, which is under
final phase of commissioning. As part of the Airport Metro Express Line, an underground
metro station is planned next to the existing New Delhi Railway Station, which required deep
excavations in range of 11 m to 19 m. A retaining wall system comprising of soldier pile walls
and multi-level ground anchors was adopted to support the deep vertical excavation., A
different geotechnical challenge was faced at one of the underground metro corridors near
Saket Station. Here the presence of loose sandy silts along an abandoned Nallah channel
posed problems with regard to effective soil arching, which is necessary for tunnel
construction using the proposed NATM method. Compaction Grouting was used to increase
the stiffness in the in-situ soils and to enable effective soil arching above the tunnel crown.
This paper presents the construction methodology, QA/QC measures and performance
testing results related to Ground Anchors and Compaction Grouting.
1. INTRODUCTION
Delhi Metro Rail Corporation (DMRC) has successfully completed the 1st 65 km phase of
metro network in New Delhi. As part of its 2nd phase, construction of about 121 km of metro
network is almost complete. This includes an exclusive link, namely the “Airport Metro
Express Line”, which is under final phase of commissioning.

As a part of this project, an underground metro station, a multi-level car park and a cut-and-
cover tunnel have been constructed, which required 11 m to 19 m deep excavation. The
designed retaining wall system includes soldier pile walls and multi-level ground anchors to
support the vertical deep excavation. Fig. 1 shows part of the DMRC network and the
location of the Airport Metro Express Line.

Fig. 1: DMRC network showing Airport Metro Express Line and Saket Station

On the link connecting Central Secretariat Station with Gurgaon, at one of the underground
tunnel section (near Saket Station, Fig. 1) connecting Central Secretariat and Qutub Minar,
the New Austrian Tunnelling Method (NATM) was adopted for tunnel construction. The
presence of loose sandy silts along an abandoned Nallah channel posed problems with
regard to effective soil arching, which is required for safe tunnel construction. To address
this problem, compaction grouting was chosen to increase the stiffness of the in-situ soils to
allow effective soil arching above the tunnel crown.

2. CASE STUDY 1: GROUND ANCHORS AT NEW DELHI METRO STATION

2.1 Introduction to Ground Anchors


A ground anchor is a structural element installed in soil or rock that is used to transmit an
applied tensile load (as a result of horizontal earth pressure) into the ground. The basic
components of the anchor include (a) the anchorage, (b) the free (or un-bonded) length and
(c) the fixed (or bonded or grouted) length. Depending on the application, the anchors may
be classified as (a) permanent anchors, (b) temporary non-retrievable anchors or (c)
temporary retrievable anchors.
Fig. 2 illustrates the schematic of the types of anchors according to method of installation
(BS 8081, 1989).

Fig. 2: Type of Anchors according to method of installation (BS 8081, 1989). (a) Straight shaft gravity-
grouted anchors, (b) Straight shaft pressure grouted anchors, (c) Post grouted anchors (d) Under-
reamed anchors.

In general anchor capacity and performance are influenced by four main factors, namely (a)
the number of strands to achieve the desired structural capacity, (b) ground characteristics,
especially shear strength, to achieve the desired geotechnical capacity, (c) installation
techniques and (d) workmanship attained in the field.

2.2 Soil conditions


In general site consists of silty clay (Delhi Silt Alluvium) with depth of bedrock varying from
as low as 5m to as deep as 18m. The rock can be described as highly to moderately
weathered Quartzite. The following profiles (Fig. 3) describe the general stratigraphy with
respect to varying depth of bedrock.

(a) (b)
Fig. 3: a) Typical soil profile at New Delhi Station b) Typical soil profile at cut & cover tunnel site
2.3 Geotechnical problem
For the construction of the underground metro station, multi-level car park and cut-and-cover
tunnel on Airport Metro Express Line stretch, an 11 m to 19 m deep excavation was
required. The site is next to the existing New Delhi Railway Station and is surrounded by
other structures like hotels and hospitals. Therefore, deep vertical excavations were
necessary. Fig. 4 shows the layout of station building & car park location.

Fig. 4: Layout showing station building and multi-level car park

To retain the soil of the 11 m to 19 m deep vertical excavation, a retention system


comprising of soldier pile walls in combination with multi-level soil and rock anchors was
proposed. Two to three levels of ground anchors (60 tons and 80 tons) were installed
depending upon the depth of excavation. Where the rock level was high, only a single level
of anchors was installed.

Fig. 5 shows the schematic of the three levels of soil anchors and rock anchors.

(a) (b)

Fig. 5: a) Schematic showing three levels of soil and rock anchors at New Delhi Station site
b) Schematic showing one level of strut followed by rock anchors at cut & cover tunnel site
2.4 Structural and Geotechnical Capacity of Anchors
2.4.1 Structural Capacity
To achieve the desired structural capacity of the anchors i.e., say 80 tons, the anchors are
fabricated using 6 Nos. of each 12.7mm diameter steel strands (7 ply) LRPC confirming to
IS: 14268-1995, clause-II, were used as per the following calculations:

Design capacity of the anchor = 80 T

Capacity of each strand of 12.7mm dia. = 18.74 T


(As per IS 14268:1995 For 7 ply, 12.7 mm nominal dia., LRPC strands, clause II)

No. of strands = 6nos (say)

Total Structural capacity of the anchor = 18.74T x 6nos = 112.44T

Factor of safety against STRUCTURAL capacity of the anchor


= Theoretical capacity / design capacity
= 112.4 T / 80T
= 1.41 > 1.4 (as per BS 8081: 1989)

2.4.2 Geotechnical Capacity


The main components of the geotechnical capacity of the anchor are free and fixed length,
which are arrived at based on the following calculations:

Design Capacity of the anchor = 80T

Length of anchor in the active wedge zone = 10.5m (as per to failure wedge analysis)

Free length of anchor = 12.5m (incl. 2m additional buffer length)

Fixed length of the anchor, L = 9.5m (say)


Geotechnical capacity of the anchor = x D x L x f (Sandy silt)

D, Dia of drill hole = 0.152m


f (Sandy silt) is theoretical skin friction (> 400 kN/sq.m) for Sandy Silts having Consistency
Index, Ic=1.25, according to BS 8081: 1989, clause 6.2.5.3, Fixed Length in Type C
Anchorages
Considering, theoretical skin friction f (Sandy silt) = 400 kN/sq.m

The ultimate geotechnical capacity of anchor = ( x 0.152 x 9.5) x 400


= 1,815 kN ~ 182T

Factor of safety against geotechnical capacity of the anchor


= Theoretical capacity / design capacity
= 182 T / 80T
= 2.3 > 2 (as per BS 8081: 1989)

Total length of the anchor = Free length + Fixed length


= 12.5m + 9.5m = 22m
2.5 Installation method
Installation of ground anchors consists primarily of drilling, installation of fabricated anchor,
cement grouting and finally by pre-stressing after a curing period of 7 to 10 days. Fig. 6
shows anchor installation works in progress and construction of underground New Delhi
metro station in full swing after the successful excavation to the desired depth.

(a) (b)
Fig. 6: a) Installation of ground anchors using Casagrande C6 Hydraulic drill rig
b) Construction of underground New Delhi metro station after successful excavation to the desired
depth (New Delhi railway station is also seen in the back ground)

2.6 Testing results


After a curing period of 7 to 10 days, each and every anchor was tested / pre-stressed using
a 100 ton multi-strand pre-stressing jack. The anchors are stressed to a test load of 1.1
times of the working load. The working loads were 60 tons at station building and 80 tons at
cut-and-cover tunnel location. Every anchor was tested to confirm the respective design
capacities. Fig. 7 shows the stressing activity in progress.

Fig. 7: Pre-stressing using 100T capacity multi strand jack


2.7 Quality Assurance and Control
State-of-the-art of anchor installation includes appropriate QA-QC procedures throughout the
construction process. The following QA-QC parameters were monitored and recorded on
site, during the installation of anchor:
 Drilling
- Drilling logs consisting of type of soil encountered with depth were kept
- (It must be ensured that the soil / rock encountered is not significantly different
from the assumptions made during the design. Any significant deviations would
trigger a design review.)
 Anchor fabrication parameters
- Components such as fixed length, free length, length of grout pipes, etc. were
checked before the anchor was installed in the drill hole
 Primary Grouting
- Volume of the grout pumped in and the flow rate was recorded
 Secondary Grouting
- Volume of the grout pumped in and the flow rate was recorded
- Grout pressures (> 20kg/sq.cm) at which the grout is pumped was recorded
 Pre-stressing
- All the ground anchors were pre-stressed (100% frequency)
- All anchors were tested to 1.1 times the design load
- Elongation of steel were recorded and checked to be under acceptable limits
- Staged loading and deformations were recorded

2.8 Performance of Ground Anchors


For the New Delhi Railway station excavation, multi-level carpark and cut-and-cover tunnel,
a total of 600 ground anchors were installed. Excavation was completed successfully in
before the middle of 2010.

3. CASE STUDY 2: COMPACTION GROUTING FOR NATM TUNNEL AT SAKET

3.1 Introduction to Compaction Grouting


The compaction grouting technique uses displacement and compaction to improve ground
conditions. A very viscous (low-mobility), aggregate grout is pumped in stages, forming grout
bulbs, which displace and densify the surrounding soils. Significant improvement can be
achieved by correctly sequencing the grouting work from primary to secondary to tertiary
grids.

The compaction grouting method may be used for the improvement of non-cohesive soils,
especially in cases, where soils of loose to medium density are encountered. This method is
also used in fine-grained soils in order to install elements of higher strength and bearing
capacity, thus improving the load bearing behaviour of the soil.

3.2 Soil conditions


The soils at site generally consist of sandy silt fill to 5m depth. The abandoned Nallah
channel was excavated and filled with locally available sandy silt to level the ground. SPT N
values in the sandy silt fill were in the range of 4 to 17, indicating loose to medium dense.

This was followed by medium dense to dense Delhi Silt alluvium layer, with SPT N values
between 20 & 30 to about 26 m depth. This is underlain by moderately weathered Quartzite
bedrock.
3.3 Geotechnical problem
Tunnel excavation by NATM was proposed at a depth of about 9 m below existing ground
level. The soil above the tunnel crown is fill material (along the Nallah alignment) consisting
of sandy silt/silty sand in the top 5 to 6 m. This was followed by Delhi silt alluvium down to
the tunnel crown. Fig. 8 illustrates the layout of the NATM tunnels and the alignment of
Nallah channel.

Fig. 8: Layout of the NATM Tunnels and abandoned Nallah channel


NATM is a method where the surrounding rock or soil formations of a tunnel are integrated
into an overall ring-like support structure, thus the supporting formations will themselves be
part of this supporting structure. But the pre-improvement soil conditions (loose to medium
dense sandy silt/silty sand) was not expected to allow effective arching.

3.4 Geotechnical Requirement


Hence, in order to permit safe and stable NATM tunnel excavation and primary lining
construction, it was necessary to carry out a combination of shallow and deep ground
treatment by compaction grouting. An theoretical SPT ‘N’ value profile between 10 and 18
with respect to depth was proposed by using the correlation, SPT N = 10 + 1.75Z, where, Z
is depth, to form effective soil arching during tunnel construction.

Fig. 9 illustrates the proposed NATM tunnels under a filled up soil strata at abandoned
Nallah channel location along with the existing and required SPT N value profile.

Fig. 9: Schematic of the NATM tunnels under an abandoned Nallah channel; the required and
existing SPT N values are plotted on the right
3.5 Installation method
Generally construction consists of drilling, installation of stinger rods and pumping the low
slump grout mix from the bottom of the treatment depth to the working platform in steps. For
compaction grouting, a low slump cement with a mix proportion of 1:3, water-cement ratio of
0.5 and admixtures like Bentonite and Glenium are used as a plasticizer to increase the
workability of grout mix. The slump value of grout mix is about 120 to 150 mm. A truck
mounted hydraulic drill rig was used to drill a nominal diameter hole of 90 mm to a depth of
about 8 m through the over burden soils. After drilling, the grout mix was pumped through
the stinger rods, to form a bulb like element in the loose soils, in stages (0.5 m each) from
bottom to the top of the working platform. Fig. 10 shows the compaction grouting works at
site.

(a) (b)
Fig. 10: a) Picture illustrating the progress of compaction grouting works at site b) Measurement of
slump as QA-QC procedures
3.6 Testing results
Field trials were carried out to establish a suitable grid pattern to achieve the intended post
compaction grouting SPT ‘N’ values. Trials were carried using 2 m and 4 m square grids. Pre
and post compaction grouting SPT ‘N’ values were recorded and analysed.

Fig. 11 illustrates the typical layout of the compaction grouting – 2m and 4m square grid:

Fig. 11: Layout illustrating the compaction grouting grid – 2 m and 4 m


Pre and post treatment analysis are also done to find the strength of the improved ground.
Post treatment SPT ‘N’ values in the filled up soil increased and ranged between 20 & 30.
Both 2 m and 4 m grids were generally able to achieve the required design SPT N values.
Fig. 12 shows the Comparison between Design SPT values with Pre and Post-improvement
SPT values.

Comparision of Design SPT values with Pre and Post SPT


Values

0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34
0

4
Depth (m)

10

12
SPT N values

Design SPT Idealised Pre-SPT


Idealised Post SPT on 4m grid Idealised Post SPT on 2m grid

Fig. 12: Comparison of Design SPT values with Pre and Post SPT Values

3.7 Quality Assessment and Control


As with other ground improvement techniques, proper quality assurance and quality control
(QA-QC) procedures were adopted. The working parameters (e.g depth, pressure, grout
volume, heave etc) were maintained and recorded at each stage of compaction grouting
process to determine the appropriate termination point. Termination of particular grouting
stage was considered when one of the following conditions achieved:

– Pre-determined grout volume is achieved (in accordance with bulb diameter i.e., 0.5 m)
– Pre-determined grout pressure is achieved (in accordance with depth of treatment i.e.,
12 kg/sq.cm to 18 kg/sq.cm)
– Mortar is overflowing from same grout hole collar
– Excessive ground heave is measured i.e. greater than or equals to 15 mm

Working parameters (grout volume, pressure, depth, etc.) were monitored using automated
quality control systems, which are recorded and printed real-time during the installation of
the compaction grouting columns.

3.8 Performance of improved ground

A total of 296 grout points were drilled with over 420 m3 of grout pumped. 19 pre-
improvement and 17 post-improvement SPT boreholes were drilled. The NATM tunnel
excavation was successfully completed in the middle of 2009.
4. CONCLUSIONS
Soldier pile walls in combination with ground anchors as a retention system was successfully
carried out to support the 11m to 19m deep excavations for the underground station and
tunnelling works Delhi Metro Rail Project sites. This retention system facilitated the space for
construction activities of the underground station and cut & cover tunnel. The construction of
these underground structures is now complete (Fig. 13).

(a) (b)
Fig. 13: Pictures illustrating the regular traffic movement over completed underground a) New Delhi
Metro Station Building b) Cut & Cover Tunnel at Ram Manohar Lohia Hospital

Similarly, Compaction Grouting proved to be effective in densifying the loose silty sandy
deposits above the tunnel crown. This facilitated the construction of tunnel by NATM method
as the loose silty sandy soils densified after the compaction grouting there by forming self
arching which is required for NATM method of construction. This was for the first time Delhi
Metro Rail Corporation has constructed a tunnel in loose deposits using NATM method.

References

BS 8081: 1999, “British Standard Code of Practice for Ground Anchorages”, British
Standards Institute, London.

IS 14268: 1995, “Specification for uncoated stress relieved low relaxation seven ply strand
for prestressed concrete”, Bureau of Indian Standards, New Delhi.
Technical Note on Ground Improvement using Grouting techniques: Theory & Practice

Annexure 3 Technical paper on “Grouting in Limestone for SMART


tunnel project in Kuala Lumpur, Malaysia” by Dr.V. R. Raju
and Ir.Y.W. Yee,2006.

17
Grouting in
Limestone for
SMART Tunnel
Project in
Kuala Lumpur,
Malaysia
Dr. V.R. Raju and
Ir. Y. W. Yee
Keller (M) Sdn. Bhd.,
Malaysia

Presented by International Conference and


Exhibition on Tunneling and Trenchless
Keller Grundbau GmbH
Kaiserleistraße 44 Technology,
D-63067 Offenbach March 2006, Kuala Lumpur, Malaysia
Tel. +49 69 8051-0
Fax +49 69 8051 244
E-mail: [email protected]
www.KellerGrundbau.com Technical paper 60-53 E
Grouting in Limestone for SMART
Tunnel Project in Kuala Lumpur
Dr. V.R. Raju and Ir. Y. W. Yee
Keller (M) Sdn. Bhd.
Kuala Lumpur, Malaysia
([email protected])

ABSTRACT
The SMART tunnel was constructed in very challenging geological terrain comprising cavernous
Limestone, with highly permeable subterranean solution channels and cavities. Dewatering
activities can result in groundwater lowering leading to ground subsidence and in some cases,
formation of sinkholes. Likewise, actions from tunneling works can cause disturbance to the
ground with similar consequence. This paper describes grouting works that were carried out to
reduce water inflow into open excavations and to minimize ground disturbance for the SMART
Tunnel project. It also details the program of preventive grouting that was implemented,
comprising mainly of rock fissure grouting and compaction grouting, and also, jet grouting. The
planning and design of the grouting activities, which were carried out to suit specific site
conditions are explained. Demanding aspects of working in public areas, beneath sensitive
structures like rail track, mitigating heavy water flow and under tight time schedules are
discussed.

1.0 INTRODUCTION

The SMART tunnel was constructed through Kuala Lumpur Limestone, home to cavernous and
karstic features with highly permeable subterranean solution channels. This presented one of
the most challenging geological terrains for the construction team. Construction activities
during initial construction of the launch shaft led to higher than expected groundwater inflow
into the excavation through these solution channels. Systematic grouting was carried out which
mitigated the situation. Subsequently, the Main Contractor, MMC-Gamuda JV implemented a
grouting program primarily aimed at minimising water inflow into open excavations and also
involved ground treatment along the tunnel corridor at areas identified to be prone to ground
subsidence.

The following pages will describe the variety of grouting methods selected depending on ground
conditions and site constraints, including fissure grouting, compaction grouting and jet grouting.
The systematic approach adopted in the design, execution and monitoring was crucial to
achieve the desired end result. This required close working relation between engineers and
technicians from the Main Contractor and Specialist Contractor to implement the appropriate
solution. This paper demonstrates that the grouting program was successful in reducing ground
disturbance due to the tunneling activities.
Dr. V.R. Raju and Ir.Y. W. Yee, Keller (M) Sdn. Bhd. , Kuala Lumpur, Malaysia

2.0 GEOLOGICAL FEATURES OF KUALA LUMPUR LIMESTONE

2.1 General
According to Gobbett & Hutchinson (1973), the Kuala Lumpur Limestone comprises Upper
Silurian marble, finely crystalline grey to cream thickly bedded, variably dollomitic rock. Karstic
features are prevalent in the limestone formed by movement of water containing carbonate
acid (dissolved carbon dioxide). As water flows downwards, the bedrock profile near the
surface is eroded to form sharply varying pinnacles, cliffs and ravines. Cavities in the rock (infill
or empty) seldom exist in isolation but as part of a complex matrix of solution channels. Over
time the roof of some cavities may dissolve or collapse which may trigger sinkholes or
depressions on the ground surface. Under some conditions, the soil overburden may arch
around the cavity (slump zone) and a quasi-stable condition may persist for years (see Figure 1).
The occurrences of ground subsidence and formation of sinkhole in limestone are frequently
associated with construction activities i.e. when the ground water is lowered, rock/ soil is
removed or triggered by vibrations (Tan, 2005).

Figure 1: Limestone Rock Profile

2.2 The Tunnel Path


The tunnel is located approximately between 10m and 16m below existing ground level. In
most areas along the tunnel corridor, the TBM bored almost entirely within the rock mass.
Tunneling activities were hence, largely shielded from disturbing the ground surface by the
relatively strong rock mass. However, where cavities or rock fissures are of significant
proportions or where the rockhead is at greater depths, drilling activities could trigger ground
movement. Significant ground displacement could lead to the formation of sinkholes at the
surface.

Various interesting geological features were exposed during the SMART Tunnel construction
works including steep cliffs, potholes and cavities (see Figure 2). They underline the complex
ground conditions in which the tunnel had to be built.

2
Grouting in limestone for SMART tunnel project in Kuala Lumpur

Figure 2: Variable Features in Limestone

3.0 OBJECTIVES OF DRILLING AND GROUTING WORKS

Grouting works were initially commissioned to reduce water inflow into the excavated launch
shaft (described in section 6.2). Thereafter, a program of “preventive” grouting works was
designed and purposefully implemented by the Main Contractor.

The objectives of preventive grouting works were mainly to reduce groundwater lowering and
minimize disturbance to the overburden soil, which were achieved by the following means:
(i) fill solution channels or cavities in rock with slurry and mortar grout
(ii) grout behind any gaps of retaining walls
(iii) compact loose overburden using thick mortar grout

In addition, the Main Contractor set up “emergency grouting teams” which were on-call to
quickly carry out any remedial works due to ground movement.

4.0 PARTICULAR AREAS REQUIRING GROUTING TREATMENT

4.1 Open Excavated Deep Shafts


(i) The nearly 30m deep launch shaft (named North Ventilation Shaft or NVS) is located at the
corner of Chan Sow Lin and Cheras Roads. The two 13m diameter tunnel boring machines
were mobilised from this shaft, one north-bound 6km towards Ampang, while the other
headed south 4km to Taman Desa.
(ii) Besides the NVS, open excavations were dug at the North Junction Box (NJB), South
Ventilation Shaft (SVS) and South Junction Box (SJB), all typically down to between 25m and
30m below ground.
(iii) As part of the route will incorporate a traffic tunnel, open excavations were also carried
out to form the north and south ingress/ egress traffic entry/ exit points (NIE and SIE).
The depths of the excavation were about 20m and 25m deep.

3
Dr. V.R. Raju and Ir.Y. W. Yee, Keller (M) Sdn. Bhd. , Kuala Lumpur, Malaysia

The rockhead level was generally found to be about 5m to 10m below ground. Hence, grout
treatment for these open excavation pits mainly comprised of rock fissure grouting to form
“grout curtain” around the excavation shafts (Figure 3). The spacing between grout holes was
typically 4m, deemed to be reasonable having taken practical constraints into considerations.
The depth of drill holes were normally taken to 2m below the depth of the excavation. Where
fracture rock was encountered or where the grout take was high, depth of drilling was
extended further down and drill holes were added. The grout holes were drilled at least 6m
behind the excavation face to be clear of rock bolts.

Figure 3: Typical Grout Holes Around Excavation Shaft

4.2 Subterranean Excavation


The tunnel also required the construction of nine Cross Passages as safety exit points along the
3km stretch of traffic tunnel. All Cross Passages were formed by excavation in rock, dug
underground from within the tunnel shaft.

Ground treatment consisted of rock fissure grouting to form “grout curtain” around the sides
and roof of the proposed box-like underground excavation (Figure 4). The spacing between
grout holes was typically 4m. Where fracture rock was encountered or where grout-take was
high, drill holes were added in between holes.

4
Grouting in limestone for SMART tunnel project in Kuala Lumpur

Figure 4: Typical “Grout Curtain” Around Cross Passage

4.3 Settlement Sensitive Areas


The TBMs also traveled beneath some settlement sensitive areas e.g. rail crossing, bridge
crossing, important highway, beside buildings, etc. These sites were firstly investigated by
“exploratory drilling” to ascertain the depth and quality of rock. Depending on the findings,
grouting works were carried out, primarily to fill cavities and seal large solution channels.

4.4 Cutter-head Intervention Locations


The TBM cutter-head required maintenance at regular intervals. At such TBM stops, there was
a risk of ground disturbance. These locations were carefully selected by the Main Contractor
based on known soil data. Grouting works were usually specified as a precautionary measure to
form a “grout block” where the cutter-head could be parked whilst maintenance was carried
out (Figure 5).

Figure 5: Typical Grout Treatment of Cutter-head Intervention Locations

5
Dr. V.R. Raju and Ir.Y. W. Yee, Keller (M) Sdn. Bhd. , Kuala Lumpur, Malaysia

4.5 Areas of Deep and Cavernous Rock


The Main Contractor carried out detailed geophysical survey and soil investigation ahead of the
TBM, along the proposed tunnel path. Particular features which were deemed to pose certain
risks to the tunnel mining activities were identified. These features included:
(i) deep rockhead (or thick soil overburden), especially where soft/ loose soil is found within
the tunnel path
(ii) high density of fractured rock
(iii) sizeable caverns in the rock mass

Rock fissure grouting and compaction grouting were then instituted (Figure 6).

Figure 6: Artist Impression of Cavity Treatment

4.6 Grouting Behind Retaining Walls


The soil encountered on the site was highly variable including loose sand, very soft slime, soft
peaty clay and man-made dumping ground. The groundwater was generally at about 3m to 4m
below ground. Typically, such mixed soils were highly permeable. The construction of some of
the deep excavation shafts (described in section 4.1) required retaining piled wall (usually
contiguous bored piles).

Jet grouting was instituted behind some of these walls to minimize water seepage and also to
reduce lateral active earth pressures. Figure 7 shows typical detail of the treatment scheme.

Figure 7: Typical Jet Grout Treatment Behind Contiguous Bored Pile

6
Grouting in limestone for SMART tunnel project in Kuala Lumpur

5.0 GROUTING TECHNIQUES

Various grouting techniques were applied for different site conditions on this project including
the following:
(i) Rock fissure grouting
(ii) Compaction grouting
(iii) Jet Grouting

The techniques are described below.

5.1 Rock Fissure Grouting


Rock fissure grouting was used widely across the site. It was primarily used to form “grout
curtain” around excavations to minimize water seepage and was applied at the NJB, SVS, SJB,
NIE, SIE and Cross Passages. Besides these, the sites for all cutter-head interventions and other
settlement sensitive locations were grouted to minimize potential ground movement.

Description of the methodology involved is given below. In general, they are in line with
guidelines given in BS_EN 12715 (2000).

5.1.1 Drilling
The overlying soil strata was drilled and retained with casing with adequate diameter for the
following rock drilling. Drilling in rock was mostly carried out using down-the-hole hammer
(DTHH) with drill hole diameter of about 95 to 115mm. The drilling air pressure was
controlled to ensure that excessive pressure is not introduced into the ground.

5.1.2 Materials and Properties


Grout comprised Ordinary Portland Cement with some bentonite. Washed sand is added into
the mix when the grout take in the hole exceeded 10m3. Bentonite was added mainly to reduce
shrinkage and aid pumpability. In some cases, additives were included to reduce shrinkage. In
general, water cement ratio was kept below 1.5, as thinner mixes were found to segregate
under the conditions of use.

5.1.3 Grouting Parameters


Typical grout pressures are as shown in Table 1. The selection of appropriate pressure is
crucial to ensure adequate distance of grout flow without causing hydrofracture.

Depth Min. Pressure Max. Pressure


(m) ( bar ) ( bar )
0–5 1 3
5 – 10 3 6
10 – 20 6 10
20 - 30 10 20

Table 1 – Typical Grouting Pressure

Grout mix was prepared by adding pre-determined volume of water and bentonite to cement
and mixing using high speed colloidal mixer. Since much of the work was carried out in public
areas or within tight space constraints where a silo batching system was not suited, the process
was relatively labour intensive. The bentonite used was allowed to soak 24hr before utilization.

7
Dr. V.R. Raju and Ir.Y. W. Yee, Keller (M) Sdn. Bhd. , Kuala Lumpur, Malaysia

5.1.4 Grout Delivery


Delivery of the grout was generally carried out using hydrostatic pumps with pumping
pressures of up to 15 bars and flowrates up to 60 lit/min. Where high grout take was
anticipated, mono pumps with flowrates exceeding 100 lit/min were used. When grouting large
cavities or where there was a need to compact the in-fill material, concrete pumps with
capacity of 200 lit/min were used.

Grout was delivered down the grout hole through high pressure grout hoses. In general,
pneumatic packers were used to inject grout following an ascending staged grouting process at
5m lift intervals. Where cavities or large fissures were encountered, these are normally grouted
immediately and left for 24hrs. The affected hole was then re-drilled and grouting was
continued for the next stage (Figure 8).

Mechanical packers were used only where grouting was required near the ground or
excavation wall surface.

Figure 8: Depiction of Fissure Grouting and End Product

5.1.5 Closure Method


The following closure grouting technique was performed as part of the rock grouting works
(Figure 9):
i) Primary exploration holes were drilled generally at 4.0m centers around the perimeter of the
excavation box.
ii) In stable boreholes, grouting was carried out in stages, ascending from the base of the
grout hole.
iii) When the grout take in the primary holes exceeded certain volume, secondary grout holes
were drilled midway between the primary holes and grout injected.
iv) The next injection sequence (if required) involves drilling tertiary holes, located midway
between the previous grout holes, to depths indicated by previous local grout takes and
grout injected.
v) The above process is repeated, with grout injections from grout holes at gradually reduced
spacing, resulting in overlapping of grout from different injection phases, which enabled the
grout to fill progressively finer fissures and discontinuities.

8
Grouting in limestone for SMART tunnel project in Kuala Lumpur

Figure 9: Grouting Using Closure Method

5.1.6 Grout Termination Criteria


For grouting in the fissured rock mass one of the following termination criteria had to be
satisfied:
i) Final pressure (max. pressure) to be maintained for a predetermined period (about 5 to 10
minutes) with a reduced flow rate of 2 to 3 lit./min.
ii) Suspend the grouting if the grout volume reaches a pre-determine amount (normally 10m3)
for at least 24 hours in order to allow setting of the grout. After this time lapse, grouting
was resumed. (Where suspension of grouting occurred, secondary grout holes were usually
added thereafter).

5.1.7 Quality Control


i) Material Suitability
The quality and consistency of the grout slurry were monitored on site by way of the following
means:
• density test using hydrometer and mud balance
• viscosity test using Marsh cone
• strength tests using hand held pocket penetrometer
• sedimentation test

ii) Monitoring during Execution


The volume of grout used and flowrate of grout delivery were recorded using grout flowmeter
connected to a computer. A hardcopy of the data was printed in real time and was inspected
frequently by the supervision personnel.

iii) As-built Records


As-built records were necessary not only for recording what has been done but were
important to aid the design engineer determine whether additional works were required (e.g.
need for secondary holes or deeper grouting depths). These included:
• drilling logs showing duration of drilling and depths where fractured rock were
encountered
• grouting logs indicating grout volume consumed versus depth and grouting pressures
• plan layout drawing showing grout take at each point
• other data such as ground monitoring data during grouting was also kept

9
Dr. V.R. Raju and Ir.Y. W. Yee, Keller (M) Sdn. Bhd. , Kuala Lumpur, Malaysia

Figure 10 shows typical as-built records of the grouting procedure.

Figure 10 – Typical As-built Grouting Record

iv) Water Tests


Where required by the Main Contractor, water tests were carried out on selected grout holes
in accordance to BS5930:1999. The grouted zone was required to demonstrate water loss of
less than 5 Lugeon.

5.2 Compaction Grouting


Compaction grouting was implemented mainly to fill cavities and compact very loose soil
overburden on site. Compaction grouting process was first applied in the USA in the 1950s and
has developed to include a wide range of applications (Rubright and Bandimere, 2004).
Compaction grouting is now the preferred method for soil improvement in the USA (replacing
slurry injection), primarily because the grouting process can be better controlled within the
localized treatment area. Traditional slurry grouting tends to result in extensive grout travel,
often to a distance far beyond the treatment zone and normally wastes large volumes of
expensive grout. Whether the targeted soil zone is repaired or not, is uncertain.

The execution method of the compaction grouting process is governed by the European
Standard of BS_EN12715 (2000).

10
Grouting in limestone for SMART tunnel project in Kuala Lumpur

5.2.1 Drilling
The technique involves the installation of a grouting pipe (“stinger rod”) to the required depth.
The spacing between grout points was varied depending on area of treatment. For overburden
compaction, this was typically at 3m centres. Where grout take was high, treatment points
were added at appropriate distance.

5.2.2 Materials
The grout material comprised of stiff mortar grout (cement, water and sand mix) which was
mixed at batching plant and delivered to site. To aid flowability, bentonite and lime were
sometimes added. Typical slump suitable for soil compaction is less than 100mm. More flowable
mix was used for cavity grouting or treatment at soil/ rock interface to allow greater grout
travel.

5.2.3 Compaction Process


The stiff mix is pumped into the soil under high pressure until a pre-determined termination
criteria is met. The overburden soil was treated as the grout pipe was withdrawn in steps
upwards. The end product is a homogeneous grout bulb or series of linked bulbs, formed near
the tip of the grout pipe as the pipe is withdrawn in steps (Figure 11). The grout bulbs formed
compacts the surrounding ground by displacing loose soil and closing voids existing within the
soil (without causing hydrofracture).

Figure 11: Schematic Drawing of Compaction Grouting Process

The displacement ability of the compaction bulbs also raised the subsided ground surface,
thereby remedying any previous ground settlement. For such treatment, several grouting points
were normally required in each subsided zone. The method is most effective in non-cohesive
soils. When using this technique in saturated fine-grained soils, care should be exercise as
temporary increase of pore pressure can result which may lead to temporary reduction in soil
shear strength.

11
Dr. V.R. Raju and Ir.Y. W. Yee, Keller (M) Sdn. Bhd. , Kuala Lumpur, Malaysia

5.2.4 Grouting Parameters


Typical compaction grouting parameters are listed in Table 2.

Description Parameter Remarks


Depending on soil conditions, displacement ratio and
Grout Point Spacing 2m to 3m grid
compaction effort (improvement) required
Sequence of Grout Downward grouting used only when grouting near
0.5m steps upwards
Injection surface or where sensitive structures nearby
Grout pressure must be high enough to overcome
Injection Pressure 10 to 20 bar
line losses
Slower rate applied for slow draining soil; Higher
Injection Rate 50 to 100 lit/min rate acceptable for treating “slump” ground or
cavities
Depending on soil conditions, displacement ratio and
Injection Volume Varies
compaction effort required.

Table 2: Typical Parameters for Compaction Grouting

5.2.5 Compaction Grouting Termination Criteria


Grouting at each step (depth) was terminated when one of the following criteria was achieved:
(i) surface ground heave was observed exceeding prescribed limit
(ii) refusal of further grout flow at pre-determined pressure (e.g. overburden stress + line
losses + 5-10 bars)
(iii) volume of grout exceeded pre-determined volume

5.2.6 Cavity Filling


Almost all the so-called “cavities” encountered on site were actually infilled with very loose soil.
The compaction grouting process involved pumping thick mortar which remained localized
around the grout pipe and cavity being targeted. The loose soil within the cavity was hence
displaced and compacted which helped stabilize the cavity.

One treatment feature to be noted is that high slump compaction grout was also injected at the
interface of rock and loose soil, which formed a grouted mass which flowed to fill voids on the
rock surface and created a grout cover to minimize downward migration of soil.

5.2.7 Overburden Soil Compaction


Compaction grouting was used widely to treat soil for sinkhole prevention. This was carried
out by drilling and injection at pre-determined grids of grouting points at various locations along
the project corridor. The insitu loose soil was thereby compacted. The treatment was similar
to that shown in Figure 5.

5.2.8 Quality Control


During the installation process, the drilling and grouting parameters were monitored carefully.
The pressure and grout volume were measured at every step of grout placement. Ground
heave was also monitored throughout the grouting process to ensure that the ground was not
over-remedied which could otherwise lead to uplift damages. Post-treatment settlement
monitoring was also carried out to ascertain that ground movement has been arrested.

12
Grouting in limestone for SMART tunnel project in Kuala Lumpur

Tests can be done in the ground to verify that the soil has actually achieved the desired level of
compaction. Figure 12 shows a typical plot of pre and post treatment SPT results conducted at
the project site for compaction grouting points carried out at 3m centres grid. Generally, in the
silty and sandy soil encountered, the soil was compacted appreciably from a loose state to a
medium dense soil with SPT N values improving by 5 to 10 blows/ft. Naturally, closer spaced
grouting points would achieve greater compaction.

Figure 12: Typical Plot of Pre and Post Compaction Grouting Treatment SPT

5.3 Jet Grouting


Jet Grouting is a well-established type of cement soil stabilization. Publications on its usage are
numerous (Essler and Yoshida, 2004) and include foundation underpinning, excavation support
and props, water sealing slabs, etc. It is a versatile method that can be used for a wide range of
soils. For the SMART project, the method was applied to seal gaps between retaining wall piles
and reduce lateral earth pressures acting on the walls (see Figure 7).

5.3.1 Drilling and Jetting


The triple tube jetting method was used for this project. This involves jets of water and cement
grout being introduced into the soil at high pressures exceeding 400 bar and flowrates above
100 lit/min to erode the soil around the drill hole. The high pressure water is shrouded in a
cone of air to concentrate and increase the erosion capability. The eroded soil is rearranged
and mixed with the cement grout (Figure 13). Some of the soil will be flushed out to the top of
the drill hole through the annular space around the drill rod. The erosion distance (diameter of
the grout column formed) varies according to the design and the soil type being treated. This
was typically 1.0 to 1.4m diameter.

13
Dr. V.R. Raju and Ir.Y. W. Yee, Keller (M) Sdn. Bhd. , Kuala Lumpur, Malaysia

Factors like the size of the columns, proportion of interlocking columns, as well as construction
verticality tolerance were taken into consideration in the design. The execution method of the
jet grouting process was in compliance to the European Standard of BS_EN12716 (2001).

Figure 13: Schematic Drawing of Jet Grouting Process

5.3.2 Materials
The sealing effect of jet grout columns to minimize water ingress was primarily determined by
the grout composition (with addition of bentonite if necessary) and the quantity injected into
the ground. Compressive strength was generally specified by the Main Contractor to be
between 0.5 and 1.0 MPa, which was determined by the cement content and the remaining
portion of soil in the treated mass.

5.3.3 Jetting Parameters


The triple tube system was employed which enabled formation of 1.0m to 1.4m diameter
columns required. The typical parameters used are shown in Table 3.

Ref. Description Values


1 Grout Pressure >1.5 MPa
2 Grout Flow Rate 50-200 ltr/min
3 Water Pressure 30-40 MPa
4 Water Flow Rate 80 – 150 ltr/min
5 Water/Cement Ratio 0.67 – 1.0

Table 3 Typical Working Parameters for Triple-tube Jet Grout Columns

5.3.4 Quality Control


During the installation process, drilling and grouting parameters were measured in real time.
The lifting speed and the rotation of the jetting string, the depth, the pressure and the volume
of the erosive and placement jets were automatically recorded in a customized computer and
printed out simultaneously. Figure 14 shows the typical print-out from the computer.

14
Grouting in limestone for SMART tunnel project in Kuala Lumpur

Figure 14: Typical Jetting Parameters Computer Printout

Another important parameter monitored during jet grout columns installation was the spoil
return. It gave direct indication of the quality of the treatment in terms of soil erodibility,
mixability and column diameter. Low spoil return generally means that the column is not being
properly formed. The density of the spoil at the said site was typically 1.6 t/m3. Cube samples
were retrieved, cured and subjected to compression tests to verify strength. In general, the
cube test showed strength of 1.0 to 2.0 MPa after 28 days.

6.0 EXAMPLES OF CHALLENGING ASPECTS OF THE WORK

6.1 “Emergency Grouting”


Throughout the tunneling works, there were some occurrences of ground depression and
sinkholes. Where these features formed over public areas, quick action was required to
minimise propagation of the disturbed zone. This necessitated almost immediate mobilization of
the required personnel and equipment (see Figure 15).

Where the slump ground “daylighted” on the ground surface (i.e. sinkhole), the area was
normally cordoned off from the public. The Main Contractor would then fill the hole with

15
Dr. V.R. Raju and Ir.Y. W. Yee, Keller (M) Sdn. Bhd. , Kuala Lumpur, Malaysia

aggregates. Subsequently, compaction grouting was carried out to infill remaining cavities
through a grout pipe drilled to the base of the slump zone. Depending on the extent of the
damage, grout holes were added until the ground recovered (usually observable through
ground heave).

Figure 15: Treatment of Sinkhole and Ground Depression

6.2 Grouting to Stop Heavy Water Flow


The launch shaft (North Ventilation Shaft) for the TBM was the first shaft excavated for the
SMART project. The shaft was located at the corner of Chan Sow Lin Road and Cheras Road.
The depth of excavation was about 27m below ground and the width of the shaft about 25m.
The plan layout of the shaft is shown in Figure 16. Limestone bedrock was generally found 5m
below ground but deep ravines were found on the south-east and south-west corners of the
shaft.

Water inflow into the excavation was mainly found to be coming through the SE corner of the
shaft and was measured to exceed 120 m3/hr. The heavy water inflow made it almost
impossible to construct the base slab for the TBM launch structure. At that stage, the concrete
floor of the excavation had been constructed over the entire pit except for this 360m2 (20m x
18m) area of exposed rock.

Initially, the Main Contractor tried chemical grouting. However, the flow of water was too high
which did not permit time for the grout to set. Subsequently, the Main Contractor decided to
use systematic grouting to arrest the problem. This was successfully carried out after working

16
Grouting in limestone for SMART tunnel project in Kuala Lumpur

over an intense 2 week period over 24 hrs. The step-by-step method employed is described
below.

Figure 16: Location of Grouting Works at Launch Shaft

(i) Identify Problem Area


The area of highest water inflow on the shaft floor was identified. 3 holes were drilled in this
region down to a depth about 3m to 6m deep. Water exiting these holes was then channeled
through steel pipes into a water sump at a corner of the shaft.

(ii) Cast Concrete Slab


A 600mm thick concrete slab was cast to cover the entire exposed rock area (with the
exception of the sump). This allowed water flow to be channeled only through openings cored
in the slab. The slab also would act as reaction slab for subsequent pressure grouting.

(iii) Drill and Grout Perimeter “Curtain”


Grout holes were determined based on an arbitrary spacing of 4m centres. Using closure
method, the initial grout holes furthest away from the sump were drilled down to 12m and
grouted. These perimeter holes were grouted to form a “grout curtain” to limit lateral flows
from outside the area to be grouted. Generally, grouting was carried out using the ascending
method, grouting with pneumatic packer placed at 3m intervals.

(iv) Grouting at Slab/ Rock Interface


The gap between the slab and the rock was grouted with the aid of mechanical packers.

(v) Rock Grouting


Where fractured rock was encountered, grouting was carried out using descending method at
3m intervals, with a 24 hr time lapse between each step to allow for the grout to cure. Through
closure method, groundwater inflow was eventually confined to exit from 3 holes at one

17
Dr. V.R. Raju and Ir.Y. W. Yee, Keller (M) Sdn. Bhd. , Kuala Lumpur, Malaysia

corner of the site. Water was flowing at high pressures and when confined to flow through a
140mm casing, the resultant water fountain reached 4 to 5m high (see Figure 17).

Figure 17: Confined Water Inflow into Shaft During Grouting

(vi) Compaction Grouting


Compaction grouting was used to treat the final 3 holes. Ready-mix Grade 30 concrete was
utilized and was delivered using a grout pump primed at up to 10 bar. After delivering almost
75m3 of concrete, the water flow was brought to a manageable level.

The Main Contractor decided not to completely stop water inflow due to concern of water
uplift which may damage the base slab. The groundwater level rose significantly after grouting,
to a level which was thought to be sustainable and not detrimental to surrounding areas. With
the grouting works completed, subsequent construction works were able to proceed
accordingly.

6.3 Grouting Beneath Rail Tracks


The TBM had to pass about 14m beneath twin rail tracks near Chan Sow Lin Road (see Figure
18). The rock level was about 3 to 8m below ground but there were concerns that cavities and
solution features in the rock may be unstable. Hence, ground treatment was implemented by
the Main Contractor.

18
Grouting in limestone for SMART tunnel project in Kuala Lumpur

Figure 18: Schematic Section of Treatment Area

Drilling had to be done outside the security fencing beside the track and as such, almost all the
holes were drilled at an incline (Figure 19). Very strict precautionary measures were
implemented during the works which included continuous supervision by the train staff and
settlement monitoring of the tracks. The methodology employed was briefly:
(i) Investigative probing using continuous rotary coring to recover rock samples to understand
geological conditions beneath the track.
(ii) Perimeter holes were drilled and grouted to form a “grout curtain” to limit lateral flows
from beyond the area to be grouted.
(iii) Inclined holes at various angles were drilled and grouted to cover almost the entire
footprint beneath the tracks. Grouting was carried out at relatively low pressures (2 to 5
bars) basically to fill fissures and cavities, not to compact the ground.
(iii) The grout was allowed 1 to 2 weeks to cure before the TBM passed beneath the tracks
without incident.

Figure 19 Grouting Beneath the “Live” Rail Track

19
Dr. V.R. Raju and Ir.Y. W. Yee, Keller (M) Sdn. Bhd. , Kuala Lumpur, Malaysia

7.0 PERSONNEL AND EQUIPMENT

Experienced engineers from the Main Contractor were involved from the onset of the project
to identify the problems and potential pitfalls ahead of TBM arrival. The Specialist Contractor
aided in suggesting the techniques available to carry out the required works. It was important
that engineers and technicians from both Main Contractor and Specialist Contractor worked
closely together to implement the appropriate solution.

It was ensured that the correct “tools” were employed for the job. The following lists some of
the equipment used on the Project:
• Drilling Rigs – equipped with computer measuring depth, rotation speed and drilling rate
• Drilling Tools – down-the-hole hammer, percussion hammer, rotary bits
• Mixing Plants – colloidal Mixer (>1400 rpm), bentonite mixer, agitator, water tank
• Grouting Plants – concrete pump, hydrostatic pump, mono pump, pneumatic &
mechanical packers
• Monitoring Devices – grout flowmeter, grout density measurement devices, testing
tools, survey equipment

8.0 SAFETY

Many precautions were taken for working in public area which included:
(i) Provision of safety gear for personnel such as safety helmet, harness and reflective garment
(ii) Placement of safety barricade, lighting and warning signages
(iii) Identification of location of buried services in the area to avoid damaging them during the
works
(iv) Monitoring of ground or property movement to ensure no sudden downward movement
or excessive ground heave

9.0 PERFORMANCE OF GROUTING WORKS

9.1 Systematic Design and Implementation


Grouting works implemented at the SMART project have demonstrated that they can be
successfully carried out provided that they are properly engineered. This would include the
following steps:
(i) understand the problem and define clearly the desired result
(ii) apply the appropriate grouting method by using the proper equipment
(iii) program the sequence of work systematically with constant review of findings on site
(iv) monitor the work with the aid of experienced eye and quality control tools

9.2 General performance


The “success” of the grouting program could be inferred from the following:
(i) The heavy water inflow at the launch shaft (North Ventilation Shaft) was reduced from a
high level of 120 m3/hr to a manageable level.
(ii) Such large inflow of ground water was not observed at excavation pits at the NJB, SVS,
SJB, SIE and NIE shafts, where pre-excavation grouting programs were implemented. All
these shafts have been excavated without water inflow problems (Figure 20).

20
Grouting in limestone for SMART tunnel project in Kuala Lumpur

(iii) The reported incidences of sinkholes and ground subsidence were uncommon in areas
where grouting works were commissioned.

Figure 20: Dry Walls of Excavated Shaft

There were a few incidences of water seepage reported in certain areas where grouting was
done. This is to be expected since it would be impossible to totally stop water flow given such
highly variable fissured rock. Water seepage in these treated areas was mainly observed in the
following situations:
(a) excavation floor where the depth of grouting was not taken to sufficient depth or where
there was design change in excavation level (deeper)
(b) excavation wall face where secondary grout holes were not carried out
(c) excavation wall where rock bolts/ anchors have been drilled through or where blasting
works have had a detrimental effect
(d) ceiling of cross passages where grout material became dislodged during excavation
Remedy action, in most cases, has been fairly straightforward.

21
Dr. V.R. Raju and Ir.Y. W. Yee, Keller (M) Sdn. Bhd. , Kuala Lumpur, Malaysia

10.0 CONCLUSION

Like most ground improvement works, grouting works are mainly carried out by specialists and
involve more than just simply drilling a hole in the ground and injecting grout slurry. The design
of the treatment (e.g. spacing between grout holes, pumping pressure and appropriate mix for
the given soil, etc.) needs to be given sufficient consideration. The requirements of existing
standards should be adhered to, especially with regards to quality control procedures
(e.g. BS_EN12715 (2000)).

The SMART project afforded ground engineers with the unique challenge of mitigating ground
disturbance associated with construction work in Limestone, using grouting technology.
Experience from more than 2 years of grouting at this site has shown that the available
technology is effective in minimizing water seepage and ground disturbance. It has
demonstrated that the success of grouting depended on proper identification of the problem by
experienced engineers and subsequent implementation of appropriate mitigating measures
(using suitable type of grout, grout parameters, etc.). Proper equipment and tools have to be
used and such specialised works required close supervision by experienced personnel. Since
grouting works cannot be assessed visually, a strict quality control system was essential to
ensure the desired end result was achieved.

ACKNOWLEDGEMENTS

The authors would like to record their gratitude to Jabatan Pengairan dan Pembetungan for
permission to publish this paper. The opportunity granted by the Main Contractor, MMC-
Gamuda JV to work on this prestigious project is gratefully acknowledged; in particular Mr.
Paramsivalingam, Mr. S.K. Ho, Cik Aswana, Mr. Koo K.S., Mr. Tioh, Mr. Gus Klados, Mr. Yeo
H.K., Mr. Andrew Yeow, Dr Ooi L.H. and many others. Special thanks are afforded to Mr.
Satkunaseelan and Mr. Doraisingam for sharing their knowledge in dealing with the challenging
environment. The support provided by the Consultants, SSP and Mott MacDonald is also
valued.

The authors wish to thank Keller’s SMART Project Team, especially the Project Manager Mr.
Sreenivas P. and Project Engineer, Mr. Gobiramanan. Mr.Toth, Mr. Schrader and Mr Saw H.S.
also provided significant assistance. It is noted that some of the grouting works were carried
out by other sub-contractors to the JV and their contributions to the understanding of the
ground conditions are acknowledged.

22
Grouting in limestone for SMART tunnel project in Kuala Lumpur

REFERENCES

BS 5930. Code of Practice for Site Investigations. 1999

BS EN 12715. Execution of Special Geotechnical Work – Grouting. 2000

BS EN 12716. Execution of Special Geotechnical Work – Jet Grouting. 2001

Essler, R. and Yoshida, H. Jet Grouting. Ground Improvement (2nd Edition).


Edited by Moseley & Kirsch. Spon Press. 2004

Gobbett, D.J. and Hutchinson, C.S. Geology of the Malay Peninsular,


New York: Wiley Interscience. 1973

Rubright, R. and Bandimere, S. Compaction Grouting. Ground Improvement (2nd Edition).


Edited by Moseley & Kirsch. Spon Press. 2004

Tan S.M. Karstic Features of Kuala Lumpur Limestone. Bulletin of the Institution of Engineers,
Malaysia. June 2005

23
GEOSYNTHETICS – Glossary & Description
The word Geosynthetics is a generic term encompassing all the polymeric (synthetic or
natural) materials used in contact with soil/rock and/or any other geotechnical material in
civil engineering applications. These products are used for several purposes in civil
engineering applications like reinforcement, drainage, filtration, erosion control, etc. The
geosynthetics includes a broad range of products which are as follows:

 Geotextiles
 Geogrids
 Geonets
 Geomembranes
 Geofoams
 Geocomposites
 Geocells
 Geopipes
 Geotubes
 Geobar
 Geomat
 Geomesh
 Geofabric
 Geonatural
 Geostrip
 Geomatress
 Electrokinetic geosynthetic
 Geosynthetic clay liner.
TERMINOLOGY AND DESCRIPTION of different geosynthetics:
Geotextile is a sheet like material made up of natural or synthetic fibres. These sheets are
flexible and permeable. It has an appearance similar to a regular fabric. Classification of
geotextiles based on the manufacturing process is as follows:
 Woven geotextiles- these are made from yarns by a conventional weaving process with a
regular textile structure.
 Non-woven geotextiles- these are made from directionally or randomly oriented fibres in a
loose web by bonding with partial melting, needle punching or chemical bonding agents
like glue, rubber, latex, cellulose derivative, etc.
 Knitted geotextiles- these are produced by interloping one or more yarns together.
 Stitch-bonded geotextiles- these are formed by the stitching together of fibres or yarns.
Geogrids are polymeric, mesh like planar products formed by ribs which are joined at the
junction in the same plane. The opening between the longitudinal and transverse ribs are
called apertures, are large enough to create interlocking with the surrounding soil particles.
The ribs are linked by extrusion, bonding or interlacing.
The extruded geogrids (also called as stretched grids) are classified into two categories based
on the direction of stretching during the manufacturing process.
 Uniaxial geogrids- these are made by stretching the punched polymer sheets in
longitudinal direction, the tensile strength in the longitudinal direction is higher than in
the transverse direction.
 Biaxial geogrids- these are made by stretching the punched polymer sheets in both the
longitudinal and the transverse direction, and, therefore possess considerable strength in
both the principal directions.

Geonets are thick planar products consisting of ribs in different direction at two different
planes. The apertures are in the shape of diamond. Geonets are also termed as geospacers.
Geomembrane is an impermeable sheet made up of one or more synthetic materials. They are
flexible in nature. Due to its extremely high impermeability properties, it is used as a fluid
barrier.
Geocomposites are materials, which are made of two or more combinations of geosynthetics
to meet advantages of both materials. Examples: Geotextile-geonet,geotextile-geogrid,
prefabricated vertical drains(PVD) ,Geosynthetic clay liners, pavement overlays, etc.
 Geosynthetic clay liner is a geocomposite. In this the bentonite clay layer is sanwitched
between thick non-woven geotextiles.Geotextile-encased GCLs are stitched or needle
punched through the bentonite core to give higher shear resistance. When hydrated they
are effective as a barrier. Often used in conjunction with geomembranes.
 Prefabricated vertical drains are made of corrugated plastic sheets surrounded by
geotextile filters used as a drain in soft clays to accelerate the consolidation process.
Geocell is a three dimensional honeycombed structure formed by joining the polymeric sheet
strips in a cellular manner. The geocells can be collapsed like an accordion during transport
and stretched during the installation. The pockets of the geocells are filled with granular
materials to form a semi-rigid base to support load bearing elements like flexible roads,
container yards, etc. The geocells help in spreading the applied loads over a large area and
provide excellent support even under cyclic loads.
Geopipes are perforated or solid-wall polymeric pipes placed beneath the ground surface and
backfilled.
Geofoam are low density network of closed, gas filled cells made by expansion of
polystyrene foam.
Geotube is a geotextile fabric with an oval cross section filled with sediment used for
shoreline protection and dewatering process.
Geobar is a polymeric material in the form of bar.
Geoblanket is a permeable blanket which is biodegradable in nature. It is used in slopes
where vegetation is possible thereby protecting the slopes.
Electrokinetic geosynthetic is a mesh made from a metal wire stringer coated in a conductive
polymer; it resembles a reinforcing geomesh, and is available also in the form of sheets, strips
or tubes. In addition to electrical conduction it also provides drainage, filtration, and
reinforcement functions.
Geomat is a three dimensional, polymeric structured made of bonded filaments which are
permeable. Used as reinforcement to roots of grass and small plants which provides in turn
permanent erosion control.
Geomatress is a three dimensional, permeable geosynthetic structure which is filled with soil
or concrete after placing over a soil layer to prevent erosion.
Geostrip is a strip of polymeric material.

FUNCTIONS AND APPLICATIONS


Basic functions of Geosynthetics are,
 Barrier - some of the Geosynthetics are impermeable in nature thus it acts as a barrier
to fluids or gases. For example, geomembrane, geosynthetic clay liner, thin film
geotextile composites and field coated geotextiles are used as fluid barrier to restrict
flow of liquids or gases. This function is also used in encapsulation of swelling soils,
asphalt pavement overlays and waste containment.
 Drainage-some geosynthetics allow in plane flow of fluid which serves the function
as drains. For example Prefabricated vertical drains (PVDs) have been used to
accelerate the rate of consolidation of soft clays under foundations. Due to this
function geosynthetics are used as pavement edge drains, slope interceptor drains, and
abutment and retaining wall drains.
 Surficial erosion control- soil erosion caused by rainfall and runoff on slopes are
reduced by providing either temporary geosynthetics or permanent light weight
geomat.
 Filtration – geosynthetics allow only fluids to pass across them this function helps in
filtration. For example geotextiles prevents migration of soil particles in drains.
 Protection- geosynthetic is used as a localised stress reduction layer to prevent
damage to a given surface or layer this termed as protection function.
 Reinforcement – the major function of geosynthetics is to increase the strength of
soil mass by its inclusion and thus it maintains the stability of soil mass, which is
called reinforcement. As a reinforcement geosynthetics takes the tensile load.
Geosynthetics as reinforcement enables the embankments to be constructed over the
soft clays and to build embankment with steeper slopes.
 Separation -The geosynthetic are used to separate two layers of soil that have
different particle size distributions. For example, geotextiles are used to prevent road
base materials from penetrating into soft underlying soft subgrade soils, thus
maintaining design thickness and roadway integrity. Separators also help to prevent
fine grained subgrade soils from being pumped into permeable granular road bases.
Table 1: Functions served by different geosynthetics
Functions served by the geosynthetic
Type of
Erosion
geosynthetic Barrier Drainage Filtration Protection Reinforcement Separation
control
Woven
√ √ √
Geotextile
Non-woven
√ √ √ √ √
Geotextile
Geogrid √
Geonet √ √
Geomembrane √
Geocell √ √
Geocomposite √ √ √ √ √ √ √
Geopipes √
Geofoam √ √

Some Common Types of polymers used for geosynthetics


Polyethylene (PE) – used for manufacture of geotextiles, geomembranes, geogrids, geopipe,
geonets, geocomposites
Polyvinyl chloride (PVC) – geomembranes, geopipes, geocomposites
Polyester (PET) – Geotextiles & geogrids
Polypropylene (PP) – geotextiles, geomembranes, geogrids, geocomposites
Polystyrene (PS) – geocomposites, geofoam

Standard Graphical Illustrations for Different Geosynthetics:


NAME SYMBOL GRAPHICAL
REPRESENTAITION
Geotextile GTX
Geomembrane GMB
Geobar GBA
Geoblanket GBL
Geocomposite drain with geotextile GCD
on both sides
Geocell GCE
Geocomposite clay liner GCL
Surficial geosynthetic erosion control GEC
Electro-kinetic geosynthetic GEK
Geogrid GGR
Geomat GMA
Geomattress GMT
Geonet GNT
Geospacer GSP
Geostrip GST
DESIGN MONOGRAPH FOR DESIGN OF SHALLOW FOUNDATION
WITH GEOSYNTHETICS

By
Satyendra Mittal,

Department of Civil Engineering, IIT Roorkee, Email: [email protected]

1. INTRODUCTION

Foundation is the integral part of any superstructure. Foundations are required to be dependable,
intact, firm and sound. Two types of foundations are well known that is shallow foundations and
deep foundation. Sometimes shallow foundations have to be provided due to site constrains,
limited budgets etc. Size of shallow foundation has to be in commensurate to the loads of the
superstructure. There may be such situations where it may be difficult to provide that much space
on ground as per design loads. The reason could be the neighbors’ property, some water body,
historical monuments etc. In such cases, if the size of foundation can be restrained suiting to site
conditions but its bearing capacity be increased by use of reinforced element, the purpose gets
solved. These days many reinforcing material made of polymers, textiles are available which
have made the foundation designs quite innovative, economical and effective. The construction
of shallow footings supported on geosynthetic reinforced foundation soils has considerable
potential as a cost-effective alternative to conventional deep foundations. In this technique, one
or more layers of geosynthetic reinforcement (geotextile, geogrid, geocell or geocomposite) are
placed inside a controlled granular fill beneath the footings, to create a composite material with
improved performance characteristics. The geosynthetic-reinforced foundation soils provide
improved bearing capacity and reduced settlements by distributing the imposed loads over a
wider area of weak subsoil. In the conventional construction techniques without the use of any
reinforcement, a thick granular layer is needed which may be costly or may not be possible,
especially in the sites that have a limited availability of good-quality granular materials.
Moreover, the simplicity of the basic principles and the economic benefits over the conventional
approaches make the geosynthetic-reinforced foundation soil very attractive to the designers.

1
Geosynthetics is a generic term for all synthetic materials used in conjunction with soil, rock
and/or any other civil-engineering-related material as an integral part of a man-made project,
structure or system. A geosynthetic shows its reinforcement function by increasing the strength
of a soil mass as a result of its inclusion, thus it maintains the stability of the soil mass. It
includes a broad range of synthetic products; the most common ones, used for reinforcement are:
• Geotextiles
• Geogrids
• Geocells
These products are almost exclusively polymeric, and those based on natural fibres (jute, cotton,
wool, silk, etc.) are generally not included.

2. GEOSYNTHETICS PRODUCTS USED FOR REINFORCEMENT


Following are popular geosynthetics products:
2.1 Geotextiles are permeable, polymeric textile products in the form of flexible sheets.
Currently available geotextiles are classified into the following categories based on the
manufacturing process:

• Woven geotextiles - They are made from yarns (made of one or several fibers) by
conventional weaving process with regular textile structure.

• Non-woven geotextiles - They are made from directionally or randomly oriented fibres
into a loose web by bonding with partial melting, needle punching or chemical binding
agents (glue, rubber, latex, cellulose derivative, etc.).

• Knitted geotextiles - They are produced by interlooping one or more yarns together.

• Stitch-bonded geotextiles - They are formed by the stitching together of fibres or yarns.

2.2 Geogrid is a polymeric, mesh-like planar product formed by intersecting elements, called
ribs, joined at the junctions. The ribs can be linked by extrusion, bonding or interlacing,
and the resulting geogrids are called extruded geogrid, bonded geogrid and woven geogrid,

2
respectively. Extruded geogrids are classified into the following two categories based on
the direction of stretching during their manufacture:

• Uniaxial geogrids - They are made by the longitudinal stretching of regularly punched
polymer sheets and, therefore, possess a much higher tensile strength in the longitudinal
direction than in the transverse direction.

 Biaxial geogrids - They are made by both the longitudinal and the transverse stretching
of regularly punched polymer sheets and, therefore, possess equal tensile strength in both
the longitudinal and the transverse directions.

2.3 Geocell is a three-dimensional honeycomb-like cellular confinement system, which acts as


a foundation reinforcement mat for improvement of load bearing capacities of weak soils
and as an erosion control barrier for unstable slope surfaces.

3. SOIL-GEOSYNTHETIC INTERACTION MECHANISMS

Soil- geosynthetic interaction is of the utmost importance in several applications of


geosynthetics, especially when they act as soil reinforcement. Soil reinforcement consists of the
placement of elements duly oriented in the soil, which, by their character, improve the
mechanical properties of the new material (reinforced soil) when compared with that of the
unreinforced soil. The main target of reinforcement is to inhibit the development of tensile
strains in the soil and, consequently, to support the tensile stresses that the soil cannot withstand.
The tensile stress imparted by reinforcement improves the soil mechanical properties by reducing
the shear stress that has to be carried by the soil and by increasing its available shearing
resistance, as the normal stress acting on potential shear surfaces increases. The effectiveness of
reinforcement depends on its alignment, it being most effective when aligned in a direction of
tensile strain in the soil, so that tensile reinforcement stress develops (McGown et at., 1978;
Jewell and Wroth, 1987; Jewell, 1996). The behavior of the reinforced soil depends on several
factors, such as:
• Soil and reinforcement mechanical characteristics
• Soil-reinforcement interaction mechanism and properties

3
• Geometry of the reinforced system
• Shape, number, location and alignment of reinforcements
• Process of construction, etc.

Three mechanisms of interaction can be identified in reinforced systems:


• Skin friction along the reinforcement
• Soil-soil friction
• Passive thrust on the bearing members of the reinforcement.

Skin friction is the only mechanism with geotextiles and strips. In the case of geogrids, the
passive thrust on the bearing members of the grids must also be considered as soil- soil friction,
if relative movement occurs in the soil along the grids' apertures. Shear strength mobilization
between granular soils and geotextiles is a two-dimensional phenomenon, where soil dilatance is
allowed, strongly affected by the extensibility of geotextiles. In the case of strips, the
phenomenon is three-dimensional and greatly dependent on the characteristics of soil dilatance
and on the roughness of the reinforcement surfaces. In fact, the volume of soil shearing around
the reinforcement is influenced by its geometry and roughness. With regard to geogrids, the
phenomenon can also be considered three-dimensional, mobilizing skin friction for small
displacements and progressively mobilizing the passive thrust on the bearing members of the
grid as displacement increases. Figure 1 shows the stress distribution in the cases of free soil
dilatance (two-dimensional phenomenon) and restricted soil dilatance (three dimensional
phenomenon).
Since geogrids are less extensible than geotextiles, the improvement in soil strength and the
mobilization of shear resistance along the interface with the soil increase when the reinforcement
used is a geogrid.

4
Fig. 1 Stress conditions in reinforced soil: (a) free dilatance; and (b) restricted dilatance
(after Hayashi et al., 1994)

The geosynthetics, in conjunction with foundation soils, may be considered to have five different
roles in improving their load-carrying capacity and settlement characteristics.

(a) Geosynthetics reduce the outward shear stresses transmitted from the overlying soil/fill to the
top of the underlying foundation soil. This action of geosynthetics is known as the shear stress
reduction effect. This effect results in a general-shear failure, rather than a local-shear failure
(Fig. 2a), thereby causing an increase in the load-bearing capacity of the foundation soil
(Bourdeau et al. , 1982; Guido et al., 1985; Love et al. , 1987; Espinoza, 1994; Espinoza and
Bray, 1995; Adams and Collin, 1997). The reduction in shear stress and the change in the failure
mechanism is the primary benefit of the geosynthetic layer at small deformations.

(b) Geosynthetic redistributes the applied surface load by providing restraint of the granular fill
if embedded in it, or by providing restraint of the granular fill and the soft foundation soil if

5
placed at their interface, resulting in reduction of applied stress (Fig. 2b). This is referred to as
the slab effect or confinement effect of geosynthetics (Bourdeau et al., 1982; Giroud et al. , 1984;
Madhav and Poorooshasb, 1989; Sellmeijer, 1990; Hausmann, 1990). The friction mobilized
between the soil and the geosynthetic layer plays an important role in confining the soil.

(c) The deformed geosynthetic, sustaining normal and shear stresses, has a membrane force with
a vertical component that resists applied loads, i.e. deformed geosynthetics provide a vertical
support to the overlying soil mass subjected to loading. This action of geosynthetics is popularly
known as its membrane effect (Fig. 2c) (Giroud and Noiray, 1981; Bourdeau et al. , 1982;
Sellmeijer et al. , 1982; Love et al. , 1987; Madhav and Poorooshasb, 1988; Bourdeau, 1989;
Sellmeijer, 1990; Shukla and Chandra, 1994a). Depending on the type of stresses - normal stress
and shear stress - sustained by the geosynthetics during their action, the membrane support may
be classified as 'normal stress membrane support' and 'interfacial shear stress membrane support',
respectively (Espinoza and Bray, 1995). Edges of the geosynthetic layer need to be anchored in
order to develop the membrane support contribution that results from normal stresses, whereas
the membrane support contribution resulting from mobilized interfacial membrane shear stresses
does not require any anchorage. The membrane effect of geosynthetics causes an increase in the
load-bearing capacity of the foundation soil below the loaded area, with a downward loading on
its surface either side of the loaded area, thus reducing its heave potential. It is to be noted that
both the geotextile and geogrid can be effective in membrane action in case of high deformation
of the reinforced foundation soils (Hass et al., 1988).

(d) The use of geogrids has another benefit due to the interlocking of the soil through the
apertures of the grid, which is known as the anchoring effect (Guido et al., 1986). The transfer of
stress from the soil to the geogrid reinforcement is made through bearing at the soil to the grid
cross-bar interface.

6
Fig. 2 Influence of geotextile inclusion on a two-layer soil system: (a) change of failure mode;
(b) redistribution of the applied surface load; and (c) membrane effect (after Bourdeau et al.,
1982; Espinoza, 1994)

(e) Geosynthetics (particularly, geotextiles, but perhaps also geogrids) improve the performance
of the reinforced soil system by acting as a separator between the soft foundation soil and the

7
granular fill. This influence is known as the separation effect of geosynthetics (Guido et al.,
1986; Nishida and Nishigata, 1994). The separation can be an important function compared to
the above functions (which may collectively be called the reinforcement function) when the ratio
of the applied stress (σ) on the subgrade soil to the shear strength (cu) of the subgrade soil has a
low value (less than 8), and it is basically independent of the settlement of the reinforced soil
system (Fig. 3).

In general, the improved performance of a geosynthetic-reinforced foundation soil can be


attributed to an increase in shear strength of the foundation soil from the inclusion of the
geosynthetic layer. The soil geosynthetic system forms a composite material that inhibits
development of the soil-failure wedge beneath shallow spread footings .

Fig. 3 Relationship between the separation and the reinforcement functions (after Nishida and
Nishigata , 1994)

4. MODES OF FAILURE
There are four possible modes of failure for geosynthetic-reinforced shallow foundations. They
are as follows.
(a) Bearing capacity failure of soil above the uppermost geosynthetic layer (Fig. 4a) - this type of
failure is likely to occur if the depth of the uppermost layer of reinforcement (u) is greater than
about 2/ 3 of the width of the footing (B), i.e. u/B > 0·67, and if the reinforcement concentration
in this layer is sufficiently large to form an effective lower boundary into which the shear zone

8
will not penetrate. This class of bearing-capacity problem corresponds to the bearing capacity of
a footing on shallow soil overlying a strong rigid boundary.

(b) Pullout of geosynthetic layer (Fig. 4b) - this type of failure is likely to occur for a shallow and
light reinforcement (u/B < 0·67, and the number of reinforcement layers, N < 3).

(c) Breaking of geosynthetic layer (Fig. 4c) - this type of failure is likely to occur with long,
shallow and heavy reinforcement (u/B < 0·67, N > 3 or 4). The reinforcement layers always
break approximately under the edge or towards the centre of the footing. The uppermost layer is
most likely to break first, followed by the next deep layer, and so forth.

(d) Creep failure of the geosynthetic layer (Fig. 4d) - this failure may occur due to long-term
settlement caused by sustained surface loads and subsequent geosynthetic stress relaxation. The
first three modes of failure were first reported by Binquet and Lee (1975a, 1975b) on the basis of
the observations made during laboratory model tests (on a footing resting on a sand layer
reinforced by metallic reinforcements). The fourth mode of failure, i.e. creep failure, was
explained by Koerner (1990).

9
Fig. 4 Possible modes of failure of geosynthetic reinforced shallow foundations (after Binquet
and Lee, 1975b; Koerner, 1990)

5. STATE OF ART IN BRIEF


A large number of model tests have been conducted in order to evaluate the beneficial effects of
reinforcing the soils with geosynthetics, as related to the load-carrying capacity and the
settlement characteristics of shallow foundations .

5.1 Reinforced granular soil


Guido et al. (1985) conducted laboratory model tests to study the bearing capacity of a
square footing (side B = 0.31 m) resting on loose sand (relative density = 50%) reinforced
with geotextiles of strength varying from 0·67 to 2·16 kN/m. The tests were performed in a
square stiffened plexiglass box of dimensions shown in Fig. 5(a). The square sheets of

10
Fig. 5 (a) Geometry of model; (b) load-settlement curves (u/B = 0·5, h/B = 0·25, b/B = 2); (c)
BGR variation with N (u/B = 0·5, h/B = 0·25, b/B = 2); (d) BGR variation with width ratio (u/B
= 0·5, h/B = 0·25, N = 2); and (e) BGR variation with tensile strength (u/B = 0·5, h/B = 0·25, b/
B = 3) (after Guido et al., 1985)

geotextile were placed concentrically under the square footing. For these tests, several
parameters were varied -- the depth below the footing of the first layer of geotextile, u; the

11
vertical spacing of the layer of geotextile, h; the number of layers of geotextile, N; the width of
the square sheet of geotextile, b; and the tensile strength of geotextile. For convenience in
expressing and comparing test data, the results were presented in terms of a bearing capacity
ratio (BCR), a term introduced by Binquet and Lee (I 975a). This term is defined as follows:
𝑞
BCR = 𝑞𝑅
𝑢

where qu is the ultimate bearing capacity of the unreinforced soil, and qR is the bearing capacity
of the geotextile reinforced soil at a settlement corresponding to the settlement S u at the ultimate
bearing capacity qu for the unreinforced soil.

Based on the test results, the following generalized conclusions can be made.

(a) All the parameters stated above have a substantial effect on the load-bearing capacity of the
geotextile-reinforced foundation.
(b) When the geotextile layers are placed within a depth equal to the width of the foundation,
they increase the load bearing capacity of the foundation - but only after a measurable settlement
has occurred.
(c) The presence of the geotextile layers changes the failure mode from one of local shear to one
of general shear. The trends of variation for BCR have been reported to be independent of the
soil type.

Small-scale laboratory model test results for the ultimate bearing capacity of strip and square
footings supported by sand reinforced with geogrid layers, as shown in Fig. 6, have been
presented by Omar el al. (1993). The general conclusions from the test observations are as
follows.

(a) For development of maximum bearing capacity ratio (BCRmax), the effective depth of geogrid
layer z is about 2B for strip footings and l.4B for square footings .
(b) The maximum width of geogrid layers bmax required for mobilization of maximum bearing
capacity ratio is about 8B for strip footings and 4·5B for square footings .
(c) The maximum depth of placement umax of the first layer of geogrid should be less than about
B for the geogrid to be effective.

12
Fig. 6 Strip and square footings supported by sand reinforced with layers of geogrid (q = load per
unit area) (after Omar et al., 1993)

Yetimoglu et at. (1994) investigated the bearing capacity of rectangular footings on geogrid-
reinforced sand by performing laboratory model tests. From the test results, the following
generalized conclusions can be drawn.

(a) The bearing capacity of rectangular footings can be increased significantly by incorporating
geogrid layers at strategic elevations in the foundation soil. However, the settlement at failure
may not be affected significantly by the geogrid layer.
(b) For single-layer reinforced sand, the optimum embedment depth (the depth of the
reinforcement layer at which the bearing capacity is highest) is approximately 0·3 times the
footing width B. For multi-layer reinforced sand, the highest bearing capacity occurs at an
embedment depth (for the first layer of reinforcement) of approximately 0·25B. The optimum
vertical spacing of the reinforcement layer is between 0·2B and 0.4B.
(c) The bearing capacity of reinforced sand increases significantly with the size of the geogrid
reinforcement and the number of reinforcement layers within a certain effective zone. The extent
of the effective zone lies approximately within 1·5B from both the base and edges of the footing.

13
Ju et al. (1996) performed a series of bearing capacity tests on reinforced sand with strip
footings. The sand was reinforced with a geonet of relatively weak tensile strength. The types of
reinforcement used were one layer, multi-layer, and mattress. Of the three reinforcing methods,
the greatest ultimate bearing capacity was obtained from the multilayer type, the optimum layer
number was 4, and the ultimate bearing capacity ratio was 3·65.

A total of 34 large model load tests were conducted by Adams and Collin (1997) in order to
evaluate the potential benefits of reinforcing the sand with goesynthetic layers below the shallow
spread footings. The tests were performed in a reinforced concrete box 5.4 m wide by 6·9 m long
by 6 m deep. One to three layers of the geogrid reinforcement, or one layer of geocell, were
placed beneath the 0·30, 0.46, 0·61 and 0·91 m square footings. The depth of the reinforcement
layers varied between 0·25 and 1·5 m. In the tests, precast, steel reinforced, concrete footings
were loaded with a hydraulic ram jacked against a reaction frame. The generalized conclusions
from the tests are as follows.
(a) The use of geosynthetic-reinforced soil foundations may increase the ultimate bearing
capacity of shallow spread footings by a factor of 2·5.
(b) The maximum improvement in bearing capacity at low strains (s/B = 0.5%; s is settlement,
and B is footing width) occurs when the top layer of reinforcement is within a depth of 0.25B
from the bottom of the footing.
(c) For one layer of reinforcement, improvement in the bearing capacity occurs if the sand within
the reinforced zone is compacted to a high relative density so that stress transfer to the
reinforcement takes place before large soil strains occur.
(d) The spread footings on the reinforced soil foundation are likely to experience a general-shear
plunging failure, if the first layer of reinforcement is placed 0.4 B beneath the base of the
footing.

Small-scale laboratory model test results of the ultimate bearing capacity of a strip footing
supported by sand reinforced with multiple layers of geogrid were presented by Shin and Das
(2000). The tests were conducted with one type of sand compacted at two relative densities and
only one type of geogrid. The foundation depth was varied from zero to 0.75B (B is the footing
width). The test results indicated that the BCR value determined from the surface footing tests

14
would provide conservative estimates of the ultimate bearing capacity for footings at depths
greater than zero.

5.2 Reinforced clay


One of the possibilities for increasing the ultimate bearing capacity of a shallow footing
supported by a saturated clay foundation under undrained conditions is by reinforcing it by
means of geosynthetic layers. Ingold and Miller (1982) reported model test results conducted on
geogrid-reinforced clay. The apparatus consisted of a rigid steel box 150mm wide, 150mm deep
and 710 mm long, in which the clay was, loaded under undrained plane strain conditions using a
rigid strip footing 50 mm wide. Figure 7 shows some model footing test results. It is noted from
Fig. 7(a) that the bearing capacity ratio, in general, increases with the number of reinforcing
layers (N); however, at low settlement ratios (namely s/B = 5%; s is the footing settlement, B is
the width of footing) and for a number of reinforcement layers less than 5, the reinforcement
appears to weaken the foundation as indicated by bearing capacity ratios less than unity. This
tendency is repeated in Fig. 7(b), which shows that BCR is < 1 for depth ratio, u/B > 0·65 (u is
the the depth below the footing to the top of the reinforcement layer), and settlement ratio s/B =
5%.

Sakti and Das (1987) reported some model test results on the bearing capacity of a strip footing
on saturated clay. They used a heat-bonded non-woven geotextile as reinforcement. From their
tests, the following general conclusions can be drawn.

(a) Beneficial effects of geotextile reinforcement are realized when reinforcement is placed
within a depth equal to the width of the footing.
(b) For maximum benefit, the first layer of geotextile should be placed at a depth of about 0·35
times the width of the footing.

15
Fig. 7 Model footing test results: (a) BCR versus N; and (b) BCR versus u/ B (after Ingold and
Miller, 1982)

(c) The minimum length of the reinforcing geotextile layer for maximum benefit is about four
times the width of the footing.
(d) Geotextile reinforcements do not have much influence on the foundation settlement at
ultimate load.

Koerner (1990) reported the results of model tests conducted at Drexel University's Geosynthetic
Research Institute. The loading tests were carried out on 6 inch round footings resting on soft

16
saturated clay silt, at saturation above the plastic limit and reinforced with woven slit-film
geotextile layers at 1.5 in. spacings (Fig. 8). Some improvement in the load-bearing capacity is
noted throughout, but the improvement is noteworthy only at large deformations.

Bearing capacity tests on model footings resting on clay subgrades reinforced with horizontal
layers of geogrids were conducted by Mandal and Sah (1992). The test results show that the
geogrid reinforcement increases the bearing capacity of subgrades, with improvements being
observed at nearly all levels of deformation. The maximum percentage reduction in settlement
with the use of geogrid reinforcement below the compacted and saturated clay is about 45% and
it occurs for the geogrid layer at a depth of 0·25B (B is the footing width) from the base of the
square foundation.

Fig. 8 Model footing test results (after Koerner, 1990)

17
6. DESIGN METHODOLOGY

6.1 Reinforced Granular fill


Binquet and Lee (1975a; 1975b) performed a pioneering study on the load-bearing capacity of
footings resting on sandy ground reinforced with aluminium foil strips, and proposed a design
method based on an assumed failure mechanism as shown in Fig. 9. According to this
mechanism, the tensile force, developed in the vertically bending part of the reinforcement
across the assumed shear band, increases the bearing capacity of the reinforced sandy ground.
When the length of the reinforcement is short, e.g. equal to the width of footing (B), as shown in
Fig. 10 (Huang and Tatsuoka, 1990), the model proposed by Binquet and Lee (1975b) is invalid.
Schlosser et at. (1983) proposed a failure mechanism, shown in Fig. 11, for the reinforced
ground. Based on this failure mechanism, the bearing capacity of a strip footing resting on
reinforced ground can be expressed as:

qu(reinforced) = γ x DR x Nq sq dq + O·5(B + ΔB) x γ x Nf x sf …..(1)

Fig. 9 Failure mechanism for reinforced sandy ground assumed by Binquet and Lee (1975b)

18
Fig. 10 Failure surface observed by Huang and Tatsuoka (1988; 1990)

qu(reinforced)

Fig. 11 Failure mechanism of reinforced ground proposed by Schlosser et al. (1983)

where qu(reinforced) is the ultimate bearing capacity of footing resting on reinforced ground, γ is the
unit weight of sand, Nq , Ny are bearing capacity factors, DR is the depth of the reinforced zone
from the ground surface, sq,sy are shape factors, dq is the depth factor, B is the width of surface
footing, ΔB is the increase of footing width at the depth of DR due to the wide slab effect
expressed by 2DR tan α, and α is the load-spreading angle as described in Fig. 11. According to
equation (1), the following two mechanisms account for the increase in the bearing capacity of
footings resting on densely reinforced sandy ground:
• deep-footing mechanism
• wide-slab mechanism.
The deep-footing mechanism is applicable when a quasi-rigid zone is developed beneath the
footing (Huang and Tatsuoka, 1988; 1990). The wide-slab mechanism is applicable only when a
quasi-rigid earth slab below the footing extends beyond the width of the footing. For densely
reinforced conditions (for either short or long strips), shear bands starting from the edges of the
footing extend straight down approximately to the depth DR, then form a wedge beneath the

19
reinforced zone (Fig. 12a). In this case, the bearing capacity of the reinforced ground is
controlled by the strength of the zone, including the wedge denoted by B in Fig. 12(a).
For lightly reinforced conditions, the shear bands that start from the edges of the footing form a
wedge within the reinforced zone, but the apex of the wedge is deeper than that for the
unreinforced ground (Fig. 12b). In this case, the bearing capacity of the reinforced

Fig. 12 Failure modes of reinforced sand: (a) densely reinforcing; and (b) lightly reinforcing
(after Huang and Tatsuoka , 1988)

ground is controlled by the strength of the block A immediately beneath the footing. In this
situation, the failure may occur because of one of the following factors:
(a) Bond failure between the sand and reinforcement.
(b) An insufficient CR (covering ratio, which is the width of the reinforcing strip/centre-to-
centre horizontal spacing of the reinforcing strips) of reinforcement.
(c) Rupture failure of reinforcement (Huang and Tatsuoka, 1990).
For estimating the ultimate bearing capacity of a deep footing (0 < Df/B < 2·5; B = width of
footing; Df = depth of footing) placed on a homogeneous dry sand, the following equation
has been suggested by Terzaghi (1943):

qu(unreinforced ,Df> 0) = Ƞ x B x γ x Nγ + γ x Df x Nq ….(2)

where qu(unreinforced ,Df > 0) is the ultimate bearing capacity for unreinforced deep footing,
Ƞ= 0·5 for strip footing and Ƞ = 0.4 for square footing.
Equation (2) can be extended for the reinforced ground based on the deep-footing and wide-
slab mechanisms as:
qu(reinforced) = Ƞ x (B + ΔB) x γ x Nγ + γ x Df x Nq ....(3)

20
The tangent of the load-spreading angle from the vertical, namely tan α, can be obtained as
follows: tan α= ΔB/2DR ….(4)
Based on comparisons of measured and multiple-variable data regression for several model
test results, the following relationship between the load spreading angle, α, and the factors
that control the scheme for reinforcement were presented by Huang and Menq:
tan α = 0·680 - 2·071d/B + 0·743CR + 0·030L/B + 0·076 …(5)
where d is the vertical spacing between two reinforcing layers, B is the footing width, L is
the length of reinforcing layers, and n is the total number of reinforcing layers. This
relationship is valid under the conditions: tan α > 0; 0·25 <=df/B<=0·5; 0·02 <=CR <=1.0; 1
<= L/B<=10; I<= N <=5.

7. DESIGN EXAMPLE 1 (For cohesionless soils)


A design example is being given below for cohesionless soil.

A shallow square foundation of width 2m, with its base at a depth of 1 m, is resting on a dry
sand stratum having properties as:
γd = 17 kN/m3 , ɸ = 280 & c=0
Geogrid sheets are used as means of reinforcement. Compare the results of bearing capacity of
shallow foundation in reinforced and unreinforced case.

Solution:
7.1 Unreinforced case
The equation 2 in modified form can be written as below:
qultimate = γ Df Nq + 0.5 γ B Nγ .... (2a)
Where,
Df = depth of foundation from ground surface = 1m
Nq, Nγ = Bearing capacity factors
Nq = 15.30 Nγ= 17.79 (Mittal & Shukla, 2014)
γ = 17 kN/m3 B= 2m

from Equation (2a),


qultimate = 17 X 1 X 15.30 + 0.5 X 17 X 2 X 17.79 = 260.1 + 302.43

21
= 562.63 kN/m2
Applying a factor of safety = 3
562.63
q allowable = qultimate/ F.O.S = = 187.54 kN/m2 (Say 18 t/m2 )
3

7.2 Reinforced case


Equation for ultimate bearing capacity for the reinforced ground based on the deep footing and
wide slab mechanism is given as below in modified form of equation 1:
qu(reinforced) = η x (B+ΔB) x γ x Nγ + γ x DR x Nq .… (1a)
η = 0.4 (for square footing)
B= width of footing= 2m
DR = depth of the reinforced zone from the ground surface = 3m (PLATE ‘A’)
ΔB = increase of footing width at the depth of DR due to the wide slab effect.
= 2DR tanα [α= load spreading angle, given by Eq. 5]
Nq, Nγ = Bearing capacity factors.
tanα= 0.680-2.071d/B + 0.743 CR + 0.030 L/B + 0.076 N ….(5)
[after Huang & Menq, 1997]
Equation (5) is valid under the conditions:
tanα > 0 ; 0.25 ≤ df/B ≤ 0.5 ; 0.02 ≤ CR ≤ 1.0 ; 1< L/B ≤ 10; 1 ≤ N ≤ 5
d = vertical spacing between two Reinforcing layer = 1 m.
L= length of Reinforcing layer
N = total no. of Reinforcing layer, =2
𝑤𝑖𝑑𝑡ℎ 𝑜𝑓 𝑅𝑒𝑖𝑛𝑓𝑜𝑟𝑐𝑖𝑛𝑔 𝑠𝑡𝑟𝑖𝑝
CR = Covering Ratio = 𝑐𝑒𝑛𝑡𝑟𝑒 𝑡𝑜 𝑐𝑒𝑛𝑡𝑟𝑒 ℎ𝑜𝑟𝑖𝑧𝑜𝑛𝑡𝑎𝑙 𝑠𝑝𝑎𝑐𝑒 𝑜𝑓 𝑅𝑒𝑖𝑛𝑓𝑜𝑟𝑐𝑒𝑚𝑒𝑛𝑡 𝑠𝑡𝑟𝑖𝑝𝑠

GL
1m B

DR=3m 1m 2m
2m
d=1m L

PLATE 'A'

22
from equation (5)
1
tanα= 0.680-2.071 x 2 + 0.743 CR + 0.030 x 6 + 0.076 x 2 = 0.01365.

[say, CR = Covering ratio = 0.05 , L/B = 6]


ΔB = 2 DR tanα = 2 x 3 x 0.01365
= 0.0819 m.
for ɸ = 280 , Nq =15.30 Nγ= 17.79 (Mittal & Shukla, 2014)

Therefore, from equation (1a)


qu(reinforced) = η x (B+ΔB) x γ x Nγ + γ x DR x Nq
= 0.4 x ( 2 + 0.0819) x 17 x 17.79 + 17 x 3 x 15.30
= 1032.38 kN/m2
Applying a factor of safety = 3,
𝑞𝑢𝑙𝑡𝑖𝑚𝑎𝑡𝑒 1032.38
qallowable = = = 344.13 kN/m2 (Say 34 t/m2 )
𝐹.𝑂.𝑆. 3

Note 1: The available literature reveals that practically there is no inclusion of tensile strength
factor (related to reinforcing element) in the equation pertaining to cohesionless soil. However,
this factor is included in case of cohesive soil (discussed later in Para 7.3).

Note 2: The above computations clearly show that when soil is reinforced with geogrid, the
bearing capacity increases from 18t/m2 to 34 t/m2 (which is almost doubled ) for just two layers
of reinforcement. It may further increase as the number of reinforcing layers increase ( within
the influence zone ). Research work (Mittal, 2013) indicates that when the geogrid is included in
soil, not only the angle of internal friction is increased, but there is development of pseudo
cohesion also, developed within the composite soil mass. However, that part is ignored in above
design example. The users are advised to conduct the tests in the laboratory to determine revised
value of ɸ and c' (apparent cohesion ) developed between soil at site and the reinforcement
material, being used for shallow foundations (Mandal & Mhaiskar, 1994).

For clayey soils, the method for load bearing capacity analysis is defined as below:

23
7.3 Reinforced granular fill-soft foundation soil system
A bearing capacity analysis, presented by Espinoza and Bray (l995) for a single layer geotextile-
reinforced granular fill - soft foundation soil, is described here. The bearing capacity equation
derived, satisfies both vertical force and horizontal force equilibrium along the geotextile
reinforcement and incorporates two important membrane support contributions, namely normal
stress membrane support and interfacial shear stress membrane support. The subgrade shear
stress reduction effect of geotextile is also included in the equation.
By considering the vertical force equilibrium of a differential geotextile element of unit area as
shown in Fig. 13(a), one gets a general equilibrium equation as:

qapp(x) = qs(x) + qg(x) ……… ..... (6)


where qapp(x) is the force per unit area above the geotextile, qs(x) is the vertical soil reaction per
unit area, qg(x) is the membrane support constribution per unit area, and x is the horizontal
coordinate.
Assuming plane-strain conditions and considering the vertical and horizontal force equilibrium
of the deformed geotextile (Fig. 13(b)), it can be shown that (Espinoza, 1994):

𝑑2 𝑦(𝑥)
qg(x) = Th(x) 𝑑𝑥 2

with:
Th= T(x)cos β(x)
T(x) = Jε(x)
𝑑𝑦
tanβ(x) = 𝑑𝑥
………………………… …. (7)
where y(x) is the vertical deflection of the geotextile, β(x) is the angle that the deformed
geotextile makes with the horizontal line at a distance x

24
Fig. 13 Forces on a geotextile: (a) membrane contribution provided by geotextile; and (b)
vertical and horizontal force equilibrium of the deformed geotextile (after Espinoza and Bray,
1995)

from the centreline, T(x) is the geotextile tensile force, J is the geotextile stiffness modulus, Ɛ(x)
is the geotextile strain, and Th (X) is the horizontal component of the tensile force T(x).

Espinoza (1994) defined the average membrane support contribution, qg' as:
2
1 𝐿/2 1 𝐿/2 𝑑 𝑦(𝑥)
qg = ∫ 𝑞𝑔 (𝑥)𝑑𝑥 = ∫ 𝑇ℎ (𝑥) 𝑑𝑥 .....(8)
𝐿 −𝐿/2 𝐿 −𝐿/2 𝑑𝑥2

where L is the effective horizontal length of geotextile (defined by the segment joining the
stationary points B and D as shown in Fig. 15). This equation satisfies global vertical and
horizontal force equilibriums.
The geotextile located outside the effective length (i.e. AB and DE in Fig. 14) exerts a vertical
pressure, qlat, due to membrane support, thus reducing the heave potential of the subgrade soil.

25
Considering an average surcharge lateral load (qlat + , h), the subgrade bearing capacity is given
by:
qs= cNc +γh + qlat .....(9)
where:
1 𝐿𝐶 +𝐿/2
qlat = ∫ 𝑞𝑔 (𝑥)𝑑𝑥 …. (10)
𝐿 𝐿/2

Fig. 14 Failure mechanism (after Espinoza and Bray, 1995)

Table 1 Load spreading angle (note, h is expressed in cm)

and :
𝜋
Nc= 1+ 2 + 𝛼 + 𝑠𝑖𝑛𝛼 …. (11)

where α = cos-1 (τc/cu), τc is the shear applied on the clay surface, cu is the undrained shear
strength of clay, Nc is the bearing capacity factor, h is the thickness of the granular fill, γ is the
unit weight of the fill, and Lc is the length of geotextile preventing heave (Fig. 14).
Equation (11) is based on the lower bound plasticity theory for undrained loading on a semi-
infinite saturated clay layer (Bolton, 1979). If the shear above the clay surface is zero (smooth

26
footing), then α = π/2 and Nc becomes (π+ 2), which is the classical bearing capacity factor for
vertical loads on rigid-perfectly plastic material. An Nc factor larger than (π + 2) may be used for
rough footings that transmit inward shear to the clay.
The average vertical stress within the fill can be estimated using a load spreading angle, B. The
average pressure applied to the geotextile is given by:
qap = γh + αbp .… (12)
where αb = b/L, width factor, and L = b + 2h tan θ. Table 1 shows different empirical values of
the load spreading angle, θ, as reported in literature.
Combining equations (6), (8), (9) and (12), an average equilibrium equation is obtained as:
αbp = cuNc + qt .… (13)
where:
2 𝐿/2 𝐶 1 𝐿 +𝐿/2
qt = ∫0
𝑞𝑔 (𝑥)𝑑𝑥 + ∫𝐿/2 𝑞𝑔 (𝑥)𝑑𝑥 ....(14)
𝐿 𝐿

where qt is the total membrane support contribution, which includes both normal stress
membrane support (membrane contribution obtained from outside the effective length) and
interfacial shear stress membrane support (membrane contribution obtained from within the
effective length). Normal stress membrane support depends on proper anchorage outside the
effective length. Interfacial shear stress membrane support depends upon the applied load and the
mobilized interface friction.
Assumptions regarding the geotextile strain distribution and deformation are nepessary to
numerically evaluate the integral expression given by equation (14). An equation for the
admissible surface pressure, P adm, can be estimated as:

𝛽
𝑐𝑢 𝑁𝑐 +𝑇0 𝑠𝑖𝑛 0⁄𝐿 + 2𝛼𝑟 𝛾 ℎ 𝑡𝑎𝑛𝜓𝑚
padm = ….. (15)
𝛼𝑏 (1−2𝛼𝑟 𝑡𝑎𝑛𝜓𝑚 )

Fig. 15 Mobilized shear (after Espinoza and Bray, 1995)

27
where αr = r/L , rutting factor, r is the rutting depth (Fig. 14), To is the tensile force in the
geotextile layer at point D, β0 is the inclination of geotextile layer at point D, and ψm is the
mobilized interface friction angle. The normal stress membrane support is reflected in the tensile
force T0 , and the angle of deflection β0 developed at the stationary points B and D in Fig. 14. In
many practical field situations, proper anchorage cannot be ensured at all times during
construction (i.e. there is not enough anchorage length, La, or surcharge load, γh, or a
combination of both). In such cases, T0 = 0 should be used to estimate the admissible pressure.
Even in cases where proper anhorage is provided (i.e. T0 > 0), its effect will not be felt until large
deformations are induced (i.e. (β0 » 0).
An expression can also be derived for the mobilized interface friction angle based on the strict
equilibrium between the membrane and sliding block above it. An expression for this, valid for
the situation shown in Fig. 15, is:

[𝛼ℎ (𝐾−𝐾𝑝𝑚 ) + 𝑀𝑐 (𝜂𝐾−𝑡𝑎𝑛𝛿𝑚 )]


tan ψm = ….. (16)
[1+ 𝑀𝑐 +2𝛼𝑟 {𝛼ℎ (𝐾− 𝐾𝑝𝑚 )−𝜂𝐾+𝑡𝑎𝑛𝛿𝑚 }]

where K is an earth pressure coefficient, Kpm = tan2(π/4 + ɸm/ 2), the mobilized passive earth
pressure coefficient, ɸm is the mobilized soil friction angle, δm is the mobilized interface friction
angle at the footing base, and αh = h/Land Mc = (cuNc + T0 sin β0/L) γh are dimensionless
parameters.

Equations (15) and (16) have been used to predict admissible pressures for a small-scale model
test setup and the results are compared with the footing pressures measured by Love et al. (1987)
and Milligan et at. (1989) for a series of model tests with various granular fill thicknesses and
subgrade strengths. Overall, the computed values of the admissible pressures compare
favourably with those measured, and this finding provides support to the validity of the proposed
equations (15) and (16).

Ochiai et al. (1994) described a conventional approach for the assessment of the improvement of
the bearing capacity due to placement of the geogrid-mattress foundation . In this approach, a
vertical load of intensity p and width B, applied on the mattress, is transmitted widely to the
supporting foundation soil with the corresponding intensity pm and width Bm (Fig. 16).

28
Fig. 16 Effects of the use of a geogrid mattress (after Ochiai et al., 1994)

The ultimate bearing capacity q without the use of the mattress may be given by Terzaghi's
equation, as follows:

q = cNc + 1/2 γBNγ ….. (17)

where c is cohesion and γ is the unit weight of the supporting foundation soil. On the other hand,
the ultimate bearing capacity, qm, with the use of a mattress, may be given as follows (assuming
that the placement of the geogrid mattress has a surcharge effect on the bearing capacity of the
supporting foundation)

qm = cNc + γmHNq + 1/2 γBmNγ …. (18)

where γm is the unit weight of the mattress, and H is the thickness of the mattress. Therefore, the
increase in the bearing capacity Δq due to the placement of the mattress can be given as follows:

Δq = γmHNq + 1/2 γ(Bm-B)Nγ .… (19)

It is therefore found that the evaluation of the bearing capacity improvement requires the
estimation of the width Bm. The experimental studies have revealed that the width of the
supporting foundation soil over which the vertical stress is distributed becomes larger as the
thickness of the geogrid mattress becomes greater, and as the vertical stiffness of the supporting
foundation soil becomes lower. It was suggested, from a design point of view, that the width of
the geogrid mattress should be at least large enough to accommodate the vertical stress
distribution which takes place under the mattress.

Several authors analysed the geosynthetic-reinforced granular fill - soft soil system by finite
element method (Love et aI. , 1987; Koga et al. , 1988; Poran et al., 1989; Floss and Gold, 1994;
Otani et al. , 1998). The advantage of such an analysis is that displacement distribution, and

29
stress distribution, can both be obtained in the subsoil as well as in the soil- geosynthetic layer
system. Nevertheless, it should be realized that the accuracy of the finite element results depends
on the appropriate material properties used and the type of modelling adopted for the analysis. In
the finite element analysis, the complete soil- geosynthetic layer system can be modelled using
individual elements, such as bar elements for the geosynthetic layer, continuum elements for the
soil and joint elements for the interface behaviour, or by using composite elements that comprise
the soil- geosynthetic system as whole. In the latter case, the properties of the composite element
can be evaluated either experimentally or by a separate numerical analysis.

The bearing capacity analysis of a geosynthetic-reinforced cohesive foundation loaded by a


flexible uniform strip footing was carried out by Otani et al. (1998) using a rigid plastic finite
element formulation . This method is based on the upper bound theorem of the theory of
plasticity, and the bearing capacity is obtained as a load factor at the ultimate limit state. The
geosynthetic reinforcement and the surrounding sand layer (constructed around the geosynthetics
in the cohesive ground for the purpose of increasing the friction between the geosynthetics and
the adjacent soil) are modelled as a single composite material with an equivalent cohesion. The
underlying soft ground is also assumed to be purely cohesive and, hence, both the reinforced soil
and soft ground are modelled using the von-Mises failure criterion. The method of analysis
proposed was checked against the field measurements or the model test results. The analysis
indicated that the bearing capacity of the ground of the geosynthetic-reinforced foundation is
increased as the depth and the length of the reinforcement are increased, but there is an optimum
depth for which the maximum reinforcing effect is obtained. There is also an optimum number of
geosynthetic layers. Figure 17 shows a simple design chart for the estimation of the bearing
capacity of geosynthetic-reinforced foundations on soft ground. In this chart, L is the half length
of geosynthetic layer, B is the half width of footing, D is the depth of geosynthetic layer, T is the
tensile strength of geosynthetic layer, qu is the ultimate bearing capacity of unreinforced
foundation soil and quR is the ultimate bearing capacity of reinforced foundation soil.

30
Fig. 17 Effects of the geosynthetics on the bearing capacity of the foundation: (a) T = 80 kN/m;
(b) T = 55 kN/m; (c) T = 35 kN/m; and (d) T = 15 kN/m (after Otani et al., 1998)

DESIGN EXAMPLE 2 (For Cohesive soils)

Perform the bearing capacity analysis of a geosynthetic reinforced foundation loaded by a


flexible uniform strip footing of width 1.5m .The geotextile layer having tensile strength as 55
kN/m is placed at a depth of 0.40 m from the base of footing and at the interface of granular fill
and soft soil.

Solution: The failure mechanism in such cases is illustrated in Fig. 14 which is reproduced

below:

31
l = b + 2h tanθ

Failure mechanism (after Espinoza and Bray, 1995)

The geosynthetic layer located outside the effective length (i.e. AB and DE in Fig. 15) exerts a
vertical pressure, qlat, due to membrane support, thus reducing the heave potential of the
subgrade soil.

Where, l = Effective horizontal length of geotextile (defined by the segment joining the
stationary points B and D as shown in Fig. 15).

l = b + 2h tanθ

Where,

b = width of footing = 1.5 m

h = Thickness of granular fill = 0.40 m

θ = Load spreading angle (degrees) with geotextile = 26.6 to 31.0 (Love et al. 1987)

= 270 (say)

Now, l=1.5 + 2 x 0.40 x tan 270 = 1.91 m

The bearing capacity analysis was carried out by Otani et al. (1988).This method is based on :

1. Upper bound theorem of the theory of plasticity, and the bearing capacity is obtained as a
load factor at the ultimate limit state.
2. The geosynthetic reinforcement and the surrounding sand layer (constructed around the
geosynthetics in the cohesive ground for the purpose of increasing the friction between

32
the geosynthetics and the adjacent soil) are modeled as a single composite material with
an equivalent cohesion.
3. The underlying soft ground is also assumed to be purely cohesive and, hence, both the
reinforced soil and soft ground are modelled using the von-Mises failure criterion.
4. The method of analysis proposed was checked against the field measurements or the
model test results.

Design charts (Fig. 17) for the estimation of the bearing capacity of geosynthetic
reinforced foundation on soft ground are reproduced as below for ready reference:

T = 80 kN/m T = 55 kN/m

T = 35 kN/m T = 15 kN/m

Effects of the geosynthetics on the bearing capacity of the foundation (after Otani et al., 1998)

In these charts,
L = half length of geosynthetic layer = l/2 = 0.95 m
B = half width of footing = 0.75 m

33
D = the depth of geosynthetic layer = 0.40 m
T = Tensile strength of geosynthetic layer = 55 kN/m (say)
qu = ultimate bearing capacity of unreinforced foundation soil.
quR = ultimate bearing capacity of reinforced foundation soil.

From above design charts,

For D/B= 0.53 m , L/B = 1.26 and T= 55 kN/m (Fig. 18 b)

(quR/qu) -1 = 0.43

quR/qu = 1.43

The above computations clearly show that the bearing capacity increases by 1.4 times with the
use of only one geosynthetic layer at the interface of granular fill and soft soil.

8. CONCLUSION

In this chapter the results give an idea for the use of geosynthetics in field applications for
shallow foundations based on small scale model footing tests. In most practical situations, the
improvement in the load bearing
capacity will be due to membrane shear effects (both the interfacial shear stress membrane
support and the subgrade shear stress reduction effect) without the need of full anchorage. Users
may draw sufficient guidelines and directions from above discussions for design of shallow
foundations by use of geosynthetics.

ACKNOWLEDGEMENT

Ms Disha Joshi, PG student, Geotechnical engg, IIT Roorkee helped the author in compilation of
material and design examples. Author expresses his sincere thanks to her.

34
REFERENCES

1. Adams, M. T. and Collin, J. G. (1997). Large model spread footing load tests on
geosynthetic reinforced soil foundations. Journal of Geotechnical and Geoenvironmental
Engineering, 123, No. I, 66- 72.

2. Adams, M. T. and Collin, J. G. (1997). Large model spread footing load tests on
geosynthetic reinforced soil foundations. Journal of Geotechnical and Geoenvironmental
Engineering, 123, No. I, 66- 72.

3. Binquet, J. and Lee, K. L. (1975a). Bearing capacity tests on reinforced earth slabs.
Journal of Geotechnical Engineering Division, ASCE, 101, No. 12, 1241 -1255.

4. Binquet, J. and Lee, K. L. (l975b). Bearing capacity analysis of reinforced earth slabs.
Journal of the Geotechnical Engineering Division, ASCE, 101, No. 12, 1257- 1276.

5. Bolton, M. D. (1979). A guide to soil mechanics. Macmillan Press, London.

6. Bourdeau, P. L. (1989). Modeling of membrane action in a two-layer reinforced soil


system. Computers and Geotechnics, 7, 19- 36.

7. Bourdeau, P. L., Harr, M. E. and Holtz, R. D. (1982). Soil-fabric interaction - an


analytical model. Proceedings of the 2nd International Conference on Geotextiles. Las
Vegas, USA, Nevada, pp. 387- 391.

8. Espinoza, R. D. (1994). Soil-geotextile interaction: evaluation of membrane support.


Geotextiles and Geomembranes, 13, 281 - 293.

9. Espinoza, R. D. and Bray, J. D. (1995). An integrated approach to evaluating single-layer


reinforced soils. Geosynthetics International, 2, No.4, 723- 739.

10. Espinoza, R. D. and Bray, J. D. (1995). An integrated approach to evaluating single-layer


reinforced soils. Geosynthetics International, 2, No.4, 723- 739.

11. Floss, R. and Gold, G. (1994). Causes for the improved bearing behaviour of the
reinforced two-layer system. Proceedings of the 5th International Conference on
Geotextiles, Geomembranes and Related Products. Singapore, pp. 147- 150.

12. Giroud, J. P. and Noiray, L. (1981). Geotextile-reinforced unpaved road design. Journal
of the Geotechnical Division, ASCE, 107, 1233- 54.

35
13. Giroud, J. P., Ah-Line, C. and Bonaparte, R. (1984). Design of unpaved roads and
trafficked areas with geogrids. Proceedings of the Symposium on Polymer Grid
Reinforcement in Civil Engineering, Paper No. 4.1, London.

14. Guido, V. A., Biesiadecki, G. L. and Sullivan, M. J. (1985). Bearing capacity of a


geotextile-reinforced foundation. Proceedings of the 11th International Conference on
Soil Mechanics and Foundation Engineering. San Francisco, California, USA, pp. 1777-
1780.

15. Guido, V. A., Dong, K. G. and Sweeny, A. (1986). Comparison of geogrid and geotextile
reinforced earth slabs. Canadian Geotechnical Journal, 23, No. I , 435- 440.

16. Haas, R., Walls, J. and Carroll, R. G. (1988). Geogrid reinforcement of granular bases in
flexible pavements. Transportation Research Record, No. 1188, 19- 27.

17. Hausmann, M. R. (1990). Engineering principles of ground modification. McGraw-Hill,


New York.

18. Huang, C. C. and Menq, F.Y. (1997). Deep-footing and wide-slab effects in reinforced
sandy ground. Journal of Geotechnical and Geoenvironmental Engineering, 123, No. I,
30- 36.

19. Huang, C. C. and Tatsuoka, F. (1988). Prediction of bearing capacity in level sandy
ground reinforced with strip reinforcement. Proceedings of the International Geotechnical
Symposium on Theory and Practice of Earth Reinforcement. Fukuoka, Japan, pp. 191 -
196.

20. Huang, C. C. and Tatsuoka, F. (1990). Bearing capacity of reinforced horizontal sandy
ground. Geotextiles and Geomembranes, 9, 51 - 82.

21. Ingold, T. S. and Miller, K. S. (1982). Analytical and laboratory investigations of


reinforced clay. Proceedings of the 2nd International Conference on Geotextiles. Las
Vegas, Nevada, USA, pp. 587- 592.

22. Jewell, R. A. (1996). Soil reinforcement with geotextiles, CIRIA and Thomas Telford
Publishing, London.

23. Jewell, R. A. and Wroth, C. P. (1987). Direct shear tests on reinforced sand.
Geotechnique, 37, No. I, 53- 68.

24. Ju, J. W., Son, S. J., Kim, J. Y. and Jung, I. G. (1996). Bearing capacity of sand
foundation reinforced by geonet. Proceedings of the International Symposium on Earth
Reinforcement. Fukuoka, Japan, pp. 603- 608.

25. Koerner, R. M. (1990). Designing with geosynthetics, second edition, Prentice Hall, New
Jersey, USA. Omar, M. T., Das B. M., Puri, V. K. and Yen, S. C. (1993). Ultimate

36
bearing capacity of shallow foundations on sand with geogrid reinforcement. Canadian
Geotechnical Engineering, 30, 545- 549.

26. Koga, K., Aramaki, G. and Valliappan, S. (1988). Finite element analysis of grid
reinforcement. Proceedings of the International Geotechnical Symposium on Theory and
Practice of Earth Reinforcement. Fukuoka, Japan, pp. 407- 411.

27. Love, J. P., Burd, H. J. , Milljgan, G. W. E. and Houlsby, G. T. (1987). Analytical and
model studies of reinforcement of a layer of granular fill on soft clay subgrade. Canadian
Geotechnical Journal, 24, 611 - 622.

28. Love, J. P., Burd, H. J. , Milljgan, G. W. E. and Houlsby, G. T. (1987). Analytical and
model studies of reinforcement of a layer of granular fill on soft clay subgrade. Canadian
Geotechnical Journal, 24, 611 - 622.

29. Madhav, M. R. and Poorooshasb, H. B. (1988). A new model for geosyntheticreinforced


soil. Computers and Geotechnics, 6, 277- 290.

30. Madhav, M. R. and Poorooshasb, H. B. (1989). Modified Paternak model for reinforced
soil. Mathematical and Computational Modelling, an International Journal, 12, 1505-
1509.

31. Mandal, J.N and Mhaiskar, S.Y. (1994), "An Overview of Pavement Design Methods
With Geosynthetics ", Geosynthetic world, Wiley Eastern limited , Mumbai

32. MandaI, J. N. and Sah, H. S. (1992). Bearing capacity tests on geogrid-reinforced clay.
Geotextiles and Geomembranes, 11, 327- 333.

33. McGown, A. , Andrawes, K. Z. and AI-Hasani, M. M. (1978). Effect of inclusion


properties on the behaviour of sand. Geotechnique, 28, No.3, 327- 346.

34. Milligan, G. W. E., Jewel, R. A. , Houlsby, G. T. and Burd, H. J. (1989). A new approach
to the design of unpaved roads - Part I. Ground Engineering, 25- 29

35. Mittal (2013), " An Introduction to Ground Improvement Engineering", SIPL


Publication, Delhi.

36. Mittal, S. and Shukla, J.P. (2014), "Soil Testing for Engineers", Khanna Publishers,
Delhi.

37. Nishida, K. and Nishigata, T. (1994). The evaluation of separation function for
geotextiles. Proceedings of the 5th International Conference on Geotextiles,
Geomembranes and Related Products. Singapore, 1994.

38. Ochiai, H., Tsukamoto, Y., Hayashi, S., Otani, J. and Ju, J. W. (1994). Supporting
capability of geogrid-mattress foundation . Proceedings of the 5th International

37
Conference on Geotextiles, Geomembranes and Related Products. Singapore, pp. 321 -
326.

39. Otani, J., Ochiai, H. and Yamamoto, K. (1998). Bearing capacity analysis of reinforced
foundations on cohesive soil. Geotextiles and Geomembranes, 16, 195- 206

40. Poran, C. J., Herrmann, L. R. and Romstad, K. M. (1989). Finite element analysis of
footings on geogrid-reinforced soil. Proceedings of the Geosynthetics '89 Conference.
San Diego, USA, pp. 231 - 242.

41. Sakti, J. P. and Das, B. M. (1987). Model tests for strip foundation on clay reinforced
with geotextile layers. Transportation Research Record, No. 1153, 40- 45.

42. Schlosser, F., Jacobsen, H. M. and Juran, I. (1983). Soil reinforcement. General Rep.,
Proceedings of the 8th European Conference on Soil Mechechanics and Foundation
Engineering. Helsinki, pp. 83- 103.

43. Sellmeijer, J. B. (1990). Design of geotextile reinforced unpaved roads and parking areas.
Proceedings of the 4th International Conference on Geotextiles, Geomembranes and
Related Products. The Hague, Netherlands, pp. 177- 182.

44. Sellmeijer, J. B. , Kenter, C. J. and Van den Berg, C. (1982). Calculation method for
fabric reinforced road. Proceedings of the 2nd International Conference on Geotextiles.
Las Vegas, Nevada, USA, pp. 393- 398.

45. Shin, E. C. and Das, B. M. (2000). Experimental study of bearing capacity of a strip
foundation on geogrid-reinforced sand. Geosynthetics International, 7, No. I, 59- 7l.

46. Shukla, S. K. and Chandra, S. (1994a). A generalized mechanical model for geosynthetic-
reinforced foundation soil. Geotextiles and Geomembranes, 13, 813- 825.

47. Shukla S. K. (2002), “ Geosynthetics and their applications”, Thomas Telford


publications, 1 Heron Quay, London E14 4JD.

48. Yetimoglu, T. , Wu, J. T. H. and Saglamer, A. (1994). Bearing capacity of rectangular


footings on geogrid-reinforced sand. Journal of Geotechnical Engineering, 120, No. 12,
2083- 2099.

38
Cellular Confinement Systems
Gali Madhavi Latha
Professor, Department of Civil Engineering, Indian Institute of Science, Bangalore - 560 012,
E-mail: [email protected]

1. INTRODUCTION

Synthetic material usage in civil engineering has, after years of research and successful
installations, gained a level of confidence with the engineering community. The generic term
‘geocell’ refers to three-dimensional, polymeric, honeycomb like cellular material. A
structure of these cells interconnected by joints to form cellular network could be used for the
confinement of soil. These geocells completely encase the soil and provide all-round
confinement, thus preventing the lateral spreading of the soil. Because of this, the soil-geocell
layer acts as a stiff mat, distributing the load over much larger area of the subgrade soil. This
helps in reducing vertical and lateral deformations of the foundation soil to a large extent
besides increasing the overall bearing capacity of the foundation soil.

The concept of cellular confinement was first developed by US Army Corps of Engineers in
late seventies (Rea and Mitchell 1978). The primary application was surface stabilization of
granular soils under vehicular loading. Now these geocells have found a wide range of
applications, which include:

• Embankment base reinforcement

• Foundation support

• Subgrade stabilization

• Erosion control and slope protection

• Channel protection

• Multi-layer reinforcement for retaining walls

• Reinforcing soil covers over flexible conduits

The planar geosynthetics like geotextiles and geogrids interact with the soil through surface
friction and interlocking with soil particles. They prevent the lateral flow of soil only through
these two mechanisms. Hence, these forms of reinforcement cannot be applied when lateral

1
flow tendency is severe such as under heavy loads or in flowing water conditions. For such
applications, three dimensional confinement of soils is preferred. The three-dimensional
confinement of soils is provided either through pre-fabricated geocells or geocells made at
the site using geogrids/geotextiles of various grades. The schematic of the three-dimensional
confinement of geocells is illustrated in Figure 1.

Figure 1. Schematic of geocell confinement system

Geocells can be either prefabricated or constructed on site using geogrids. The structure of
prefabricated geocell layer is shown in Figure 2. Figure shows both the collapsed and
expanded forms of the geocell layer. The collapsed form of the geocell layer, in which the
cells are closed, allows the geocell layer to occupy very less space for transportation and
handling. Geocell layer is spread on the foundation in expanded form.

Figure 2. Collapsed and expanded forms of geocell layer

2
Few studies are available on the improvement in strength and modulus of soil due to geocell
confinement (Bathurst and Karpurapu 1993; and Rajagopal et al. 1999). Performance of
geocell reinforced earth structures is investigated and reported in literature by several
researchers (Bathurst and Jarrett 1988; Jenner et al. 1988; Bush et al. 1990; Cowland and
Wong 1993; Bathurst and Knight 1998; Krishnaswamy et al. 2000; Madhavi Latha 2000;
Dash et al. 2001; Madhavi Latha and Rajagopal 2007; Madhavi Latha et al. 2008, Madhavi
Latha et al. 2006; Madhavi Latha and Murthy 2007; Madhavi Latha and Rajagopal 2007;
Madhavi Latha et al. 2010; Han et al. 2011).

Construction Procedure
Initially the level where the first layer of geocell is to be placed is marked and a geotextile
layer of high strength is laid at the proposed level to act as a separator between the geocells
and the foundation soil. The dimensions and position of geocell layer are marked and stakes
are put at the four corners. Above this, the geocell layer is expanded and positioned and
anchored over the embedded stakes. A typical geocell layer with anchors looks as shown in
Figure 3. Then the first row of cells is filled with a dump truck and the fill sand is pushed into
cells using shovels and all the rows are filled subsequently. No cell should be filled
completely until the adjacent cell is at least half-filled. No traffic is allowed to move over the
unfilled cells. The cells should be overfilled slightly to allow for consolidation. Next, the
infill sand inside the cells is compacted through multiple passes by the tracked equipment
used to spread the infill (Figure 4). A vibrating roller and/or water may be required to achieve
the specified level of compaction. Once the cells are filled and the system is compacted, the
geocell layer is ready to withstand moving construction traffic.

Figure 3. Expanded geocell layer with anchors

3
Figure 4. Filling of geocells

This document describes and compiles the laboratory and small scale field studies carried out
on geocell reinforced soil structures. Embankments supported on geocell layers, foundations
resting on geocell layer and road subgrades stabilized with geocell layer are considered and
the beneficial role of geocells in these structures is investigated and explained.

2. GEOCELL SUPPORTED EMBANKMENTS

The embankments constructed on soft clays are prone to excessive settlements and shear
failure due to high compressibility and low shear strength of foundation soil. There is also a
tendency for lateral spreading because of the horizontal earth pressures acting within the
embankment. In view of these factors, construction of embankments over soft soils poses
interesting challenges to the geotechnical engineer. Soil reinforcement is accepted as one of
the attractive solutions to support the embankments constructed on soft soils due to the
savings on time and cost apart from flexibility in space requirements. The advantages of
geocell reinforcement, which make it most appealing for the construction of embankments
over soft foundation soils, are:

• It acts as an immediate working platform for the movement of construction traffic

• It allows construction of embankments of greater heights and steeper slopes

• It promotes uniform settlements

• It minimizes construction time and required space

• It increases bearing capacity and reduces settlements to a great extent

• It provides short and long term global stability to the embankment

4
Figure 5 shows the schematic diagram of the basal geocell layer used to support embankment
construction over soft clay foundation. Application of geocells for constructing a rail road
embankment on soft clay is shown in Figure 6.

Geocell layer

Soft clay foundation

Figure 5. Geocell supported embankment

Figure 6. Geocells for constructing a rail road embankment on soft clay

A series of load tests on laboratory models of embankments constructed on soft clay

foundation were carried out by Madhavi Latha (2000). Steel tank of plan dimensions 1800

mm × 800 mm and 1200 mm depth was fabricated for conducting the model tests on

embankments. The tank was fitted with perspex sheet on one side to visualize the failure of

5
the embankment. The other three sides of the tank were made smooth and rigid to create

plane strain conditions in the tank. Soft clay bed of 600 mm depth was prepared in this test

tank. For this purpose, clay was mixed with excessive amount of water and consolidated

under a surcharge pressure of 10 kPa. The bed was cured for one full week to achieve

uniform properties of void ratio 0f 0.9 and density 1.7Mg/m3, which resulted in CBR value

0.5 and vane shear strength of 20 kPa.

After leveling the clay bed, a layer of geocells was formed on top of the clay bed. This was

done by cutting the geogrids to required length and breadth from full rolls and placing them

in transverse and diagonal directions with bodkin joints inserted at the connections. The

tensile strength properties for various geogrids used in the model tests are determined from

wide width tensile strength tests and presented in Table 2. The order of geogrids in the order

of increase in their tensile strength/stiffness is NP-1, NP-2, BX and UX. After the formation

of geocell layer, pockets of geocells were filled with soil and this soil was compacted using a

steel rod. The unit weight of the infill soil was maintained at 17 kN/m3 for all the tests. The

compaction quality of this layer was verified by testing undisturbed core samples collected

from at least six individual cells. Above the geocell layer, symmetrical half of the

embankment was constructed using clayey sand in lifts. Each layer was compacted with

calculated number of blows to achieve an average density of 1.9 Mg/m3. The properties of the

soil in the constructed embankment, determined by taking undisturbed samples are given in

Table 3. The embankments were subjected to uniform surcharge pressure on the crest until

the failure. The physical dimensions of the embankment and the set up used to apply uniform

surcharge on its crest are shown in Figure 7. The slip surfaces were observed to pass through

the soft foundation soil with deeper slip circles for embankments with stiffer geocell

reinforcement as shown in Figure 8. The soil beyond the embankment was observed to heave

up as the embankment settled into the soft clay soil.

6
The vertical and horizontal deformations and the strains developed within the geocell layer

were measured during the test. The influence of various parameters such as tensile stiffness

of geogrids used to fabricate the geocell layer, height and pocket-size of geocell layer and the

type of fill material inside the geocell on the behaviour of the embankments were studied in

detail. Results from the tests on unreinforced and geocell supported model embankments

indicated that the geocell reinforced embankments exhibited improved load carrying capacity

and reduced deformations. Geocell reinforcement was found to be beneficial in pushing the

failure envelope deeper, helping in mobilizing higher shear strength compared to the

unreinforced embankment. The efficacy of the geocell layer mainly depended on the tensile

modulus of the geocell material (M) and the aspect ratio (height to diameter ratio) of geocells.

Even clay-filled geocells provided moderate support to the embankment. As long as the

dimensions and tensile modulus of the geocell material and infill soil remained the same, the

pattern of geocell formation did not affect the performance of the embankment. Important

findings from these studies are presented in Figure 9.

7
hydraulic jack

proving ring
V1  
 V3

steel channels
expanded polystyrene sheet
700
H1 1
H2 1
400 V4
clayey sand embankment
H3 

h 100-250

geocell layer 100

600 soft clay bed

(a) Sectional Elevation

dp

800

1800

(b) Plan view


all dimensions are in mm

Figure 7. Test set-up of model embankment

8
700
mm
1
1
400 mm clayey sand
embankment geocell layer

100 mm

NP-1 geogrid
soft clay
foundation
600 mm
NP-2 geogrid
UX geogrid

1800 mm

Figure 8. Failure surfaces observed in model embankments

surcharge pressure (kPa) surcharge pressure (kPa)


0 20 40 60 80 100 120 140 160 0 20 40 60 80 100 120 140 160
0 0
h = 100 mm

2 5

4
lateral deformation (mm)
lateral deformation (mm)

10 250mm

6 200mm
150mm
15
Aspect ratio of 100mm
8 geocells (h/D)

20 1.1

10 0.89
height of geocell layer
UX (M= 200 kN/m)
BX (2nd trial) BX (M= 160 kN/m)
0.66
unreinforced 25 0.44 unreinforced
12 NP-2 (M= 70 kN/m)
(2nd trial)
NP-1 (M= 70 kN/m) unreinforced
unreinforced Diameter of cells (D) = 226 mm
14 30

surcharge pressure (kPa)


0 20 40 60 80 100 120 140 160 surcharge pressure (kPa)
0 0 20 40 60 80 100
0
h = 100 mm
5 2

10 4
lateral deformation (mm)
heave (mm)

6
15
Type of geogrid
8
20 UX (M= 200 kN/m)
BX (M= 160 kN/m) 10

NP-2 (M= 70 kN/m) clayey sand filled geocells


25
12 clay filled geocells
NP-1 (M= 70 kN/m)
unreinforced
Unreinforced
30 14

Figure 9. Results from the load tests on model embankments constructed on soft clay bed

9
Guidelines for the Construction of Geocell Supported Embankments

Based on the laboratory experiments and finite element simulations on geocell supported

embankments, the following guidelines are suggested for the construction of these

embankments on soft foundation soils (Madhavi Latha, 2000).

 A layer of planar geogrid has to be spread over the soft foundation soil layer to act as
working platform for the formation of geocells and also to avoid penetration of cells into
the soft soil.

 Granular soils are preferred for fill inside the geocells because the confinement effect is
more pronounced in these soils, leading to greater reduction in overall deformations.
However, in the absence of granular fill, locally available soils may also be used if found
suitable from other considerations like drainage.

 Even the geogrids having moderate secant modulus (not less than 200 kN/m) may be used
for forming the geocells as the influence of the modulus was found to be marginal beyond
a limit of about 200 kN/m.

 Geogrids with large aperture openings offer lesser confinement to the soil. Very small
aperture makes the geogrid unsuitable for the insertion of joints. The aperture size should
be medium for achieving significant confinement effect and better interfacial friction
along with ease in construction.

 A height to diameter ratio of 1.0 is recommended for the geocells. The height of the
geocell layer could be determined from trial finite element analysis or other simple
analyses such as slope stability analysis. In both the cases, the geocell layer could be
treated as an equivalent composite soil layer.

 The geocells can be formed in either diamond or in the chevron patterns, both of which
were observed to give similar performance.

 The geocell layer can be truncated at the toe of the embankment.

10
DESIGN METHODS FOR GEOCELL SUPPORTED EMBANKMENTS

The methods of design available for geocell-supported embankments are very few. Two of
them are discussed in detail in this document. The first method is the slip line method
proposed by Jenner et al. (1988). The second method is based on slope stability analysis,
proposed by Madhavi Latha et al (2006).

Design Based on Slip Lines

Jenner et al. (1988) suggested a method for designing geocells for supporting embankments.
In this design, plastic bearing failure of the soil was assumed instead of slip circle failure.
This type of failure was expected for embankments, whose width is more than four times the
depth of the foundation soil. The methodology developed by Johnson and Mellor (1983) for
the compression of a block between two rough, rigid plates was used for determining the
bearing capacity of the soft foundation soil. The soft soil, which was analogous to the block,
was assumed to get compressed between the geocell mattress at the top and the hard stratum
at the bottom. This analogy was used for developing a non-symmetric slip line field in the
soft foundation soil.

The concept of this design is that the geocell mattress exerts a degree of restraining influence
on the deformation mechanism of the soft soil, thus rotating the direction of principal
stresses. The direction of maximum shear stress also rotates correspondingly, pushing the
failure surface deep in to the foundation soil. A 15° slip line field was used to determine the
bearing resistance of the soft soil. Figure 10 shows the slip line field used in the design and
corresponding bearing pressure diagram. The bearing capacity diagram was developed by
working from the outer edge of the slip line field inwards to the boundary of the ‘rigid head’,
defined on the slip line field by the ratio of geocell width to the depth of the soft layer. The
‘rigid head’ is term used to denote the soil zone, which remains in the active condition and so
does not experience plasticity. Thus the slip line field is used to define the maximum
allowable pressure distribution within a zone of limiting plasticity.

11
Figure 10. Method of design of geocell mattress for supported embankments
(Jenner et al. 1988)

12
The stress distribution across the ‘rigid head’ can be determined by considering the rotations
of each of the chords of the stress field bounding the rigid field. An average pressure across
the rigid head can be calculated as

P' 22I  0.5d  P


  (1)
Cu 2X Cu
where P/Cu is the value read from the stress field at the extreme end of the rigid head
P’ is the average stress over the rigid head
I = (horizontal chord lengths  rotation)
X = (horizontal chord lengths)
d = depth of soft soil layer

A typical calculation of average stress across the rigid head for the slip line filed shown in
Figure 11is given in Table 1.

Figure 11. Conditions across rigid head boundary

Table 1. Calculation of average stress over ‘rigid head’


Slip line
A-B B-C C-D A-D
Chord X1 = 1.6 X2 = 2.2 X3 = 2.6 X = 6.4
Rotation (deg.) a1 = 37.5 a2 = 22.5 a3 = 7.5
Rotation (radians) 0.654 0.395 0.431
Chord  rotation 1.05 0.86 0.34 I=2.25

13
From Table 1,

P' 22  2.25  0.55  P P


   1.094 (2)
Cu 2  6.4 Cu Cu
Hence the additional resistance due to the average pressure across the rigid zone can be taken
as Cu. This value can be applied for all ratios of width of the geocell mattress to a depth of
soft layer greater than four. For values less than four, the additional resistance over the rigid
head will be greater and should be calculated for each case. The bearing capacity diagram can
be drawn as shown in Figure 12.

Figure 12. Bearing capacity diagram

The allowable bearing capacity is now checked against the overburden stresses and the factor
of safety against bearing capacity failure is calculated. If the factor of safety is less than one,
the following options can be considered (Bush et al., 1990).

 Adding a steep sided berm to the outer edge of the original embankment to cause
additional downward pressure on the outer passive wedge thus increasing the overall
bearing capacity of the foundation
 Increasing the strength of soft foundation soil due to consolidation can be used in the
analysis as the geocell layer filled with granular soil acts as excellent drainage blanket
and allow for the quick strength gain in soft soil during construction.
 Constructing embankment in stages, the height in each stage is a limit equilibrium
height corresponding to the strength of soft soil at that stage.

14
DESIGN EXAMPLE 1

It is proposed to construct a 6 m high embankment over 6 m thick layer of soft cohesive soil
with undrained shear strength of 15 kN/m2. The cross section of embankment is shown in
Figure 13. A surcharge of 20 kN/m2 will be applied. Design suitable geocell mattress for this
case.

36 m
2.5
1  = 19 kN/m3 4m
Geocell mattress
 = 25 6m

Cu = 15 kPa Marine Clay 6m

Figure 13. Cross section of the embankment in design example 1

To design a geocell mattress for the above problem, slip line method can be adopted as
follows:
Embankment base width = 66 m
Width of geocell mattress = 66 m – 4 m (leaving 2 m offset either side) = 62 m
Width of geocell/depth of soft soil layer = 62/6 = 10.33
From stress field diagram shown in Figure 1, P/Cu = 12
Average pressure across rigid head P’ = 12 Cu + Cu = 13 Cu
Hence the bearing capacity diagram can be drawn as shown in Figure 14.

15
5.71 Cu

12 Cu

13Cu
6m 19.5 m 7.5 m

d 1.25 d

Figure 14. Bearing capacity diagram for the design example

Load from the embankment allowing a surcharge pressure of 20 kN/m2:


(18+28)/2  4  19 + (18  20) + 33  2  19 = 3362 kN/m2
Bearing capacity from the pressure diagram :
6  5.71 Cu + (5.71+12)/2  Cu  5.71 + 7.5  13 Cu = 304.43 Cu
Cu required for equilibrium = 3362/304.43 = 11.04 kN/m2
Actual Cu = 15 kN/m2
Factor of safety against bearing capacity failure = 15/11.04 = 1.35 (against 1.25 required).
Hence safe.
For designing the geocell mattress, consider an element of soil within the granular cellular
mattress, but interfacing with the soft layer. The stress condition in element can be obtained
from a Mohr-circle construction as shown by Jenner et al (1988).

The horizontal stress on the element:


h = n – 2 x (3)

Where n is the vertical stress on the element.

  
1
2 n sin 2  '  4 2 sin 4  '  4 sin 2  '  1  n2 sin 2  '   2 
2
 n 
x

2 sin 2  '  1  (4)

16

 = shear stress at the interface = Cu in limiting condition = 11.04 kN/m2
n under highest part of the embankment = 6  19 +20 = 134 kN/m2
 = 40 for the geocell fill material
 x = 51.36 kN/m2+
h =n – 2 x = 134 – 2  51.36 = 31.28 kN/m2

The rotation of principal stress occurs within the mattress depth. Therefore mattress strength
required = 31.28 kN/m. Hence a geocell mattress with long term tensile strength more than
31.28 kN/m should be used to support the embankment, with a geogrid base.

Design Based on Slope Stability Analysis

This method uses a general-purpose slope stability program to design the geocell mattress of
required strength for embankment. The computer program developed for conducting slope
stability analysis of geocell supported embankments reads the slope parameters, height of
geocell layer, depth of foundation soil, shear strength parameters of embankment soil and
geocell layer, properties of foundation soil, pore pressure co-efficient and the value of
uniform surcharge pressure on the crest. The program uses Bishop’s method of slices for
calculating the factor of safety. The program automatically searches different trial slip circles
and gives the minimum factor of safety and coordinates of the center of the critical slip circle.
The reliability of the computer program was ensured by running some example problems.
The factor of safety obtained from the program was in agreement with the minimum factor of
safety obtained from graphical construction.

For designing geocell mattress below an embankment, geocell layer is treated as a layer of
soil with cohesive strength greater than the encased soil and angle of internal friction same as
the encased soil. This is because; geocells provide all-round confinement to the soil due to the
membrane stresses in the walls of geocells, because of which apparent cohesion is developed
in the soil. Using the rubber membrane theory proposed by Henkel and Gilbert (1952),
Bathurst and Karpurapu (1993) analyzed the cohesive strength of soil encased in a single
geocell in triaxial compression. The same analysis was extended for multiple geocells and
also for geocells made of geogrids by Rajagopal et al. (1999) and Madhavi Latha (2000).

17
Later Latha and Murthy (2007) applied the same analysis to quantify the strength and
stiffness of geocell reinforced sand. Equations developed from the above analyses can be
used for estimating the cohesive strength of a layer of geocells. In case of geocells made of
geogrids, if we consider individual cells, the soil is not fully confined as in case of geocells
made of geotextile, because of apertures in geogrids. However, during loading, the soil in
each geocell is subjected to lateral confinement due to interaction mechanism between cells.
The validity of equations for cohesive strength based on rubber membrane theory for geocells
made of geogrids was verified by Rajagopal et al. (1999) and Madhavi Latha (2000) by
testing soil encased in geocell made of open mesh.

The additional confining pressure due to the membrane stresses can be written as (Henkel and
Gilbert 1952),

2M c 1 2 M 1  1   a 
 3     (5)
D (1   a ) Do  1   a 

where a is the axial strain at failure, c is the circumferential strain at failure, Do is the initial
diameter of sample, D is the diameter of the sample at an axial strain of a and M is the
modulus of the membrane.

The above equation was used to calculate the additional confining pressure due to geocell
reinforcement, using the parameters as follows. Do was taken as the initial diameter of
geocell. The geocell pockets are not circular but are triangular in shape. The equivalent
diameter for the triangular shaped geocells can be obtained by equating the area of the
triangle to a circle of equivalent area. M is the modulus of the geocell material at axial strain
a, determined from the load-strain curves obtained from wide width tensile strength test on
geogrids.

The relation between the induced apparent cohesive strength and the additional confining
stress due to the geocell can be derived by drawing Mohr circles for the unreinforced and
reinforced soil samples as shown in Figure 15.

18
reinforced

unreinforced

composite

with
confinement

cr unreinforced

3 3 1u 1

Figure 15. Mohr circles for calculating the strength improvement due to geocell
reinforcement

From Mohr-Coulomb failure theory, the ultimate stress on soil sample can be calculated by
considering the soil as a composite as (large circle)

 1  k p 3  2cr k p (6)

In which kp is the coefficient of passive earth pressure. If the same is considered as an


unreinforced soil with an additional confining stress of 3, the failure stress can be
calculated as

 1  k p ( 3   3 ) (7)

Equating (6) and (7), the additional cohesive strength due to geocell layer can be obtained as
 3
cr  kp (8)
2

Substituting the value of 3 obtained from equation (5) in equation (8), we will get the
apparent cohesion induced to soil due to geocell confinement. This additional cohesive
strength is added to the original cohesive strength of soil encased in geocells to get the
cohesive strength of geocell layer (cg).

19
For preliminary design problems, if the geometry of the embankment, properties of
foundation and embankment soils are given, we can perform slope stability analysis with trial
values of height of geocell layer and determine the cohesive strength of geocell layer required
to get a design value of factor of safety. From this cohesive strength, we can back calculate
the modulus of geocell required for assumed values of pocket-size of geocell and axial strain
in the walls of geocell.

This design method has been verified for the case of geocell supported model embankments
constructed in laboratory with varying pocket sizes of cells, varying height of geocell
mattress, for geocell layers made of different geogrids and for sand and clay infill materials
by Madhavi Latha et al. (2006). It was observed that the maximum surcharge pressure at
which the embankments failed in the model tests was agreeing well with the surcharge
pressure at which the factor of safety was obtained as one in the slope stability analysis.

DESIGN EXAMPLE 2

It is proposed to construct a 4 m high embankment over a 6 m thick layer of soft cohesive soil
having undrained shear strength of 15 kPa. A surcharge of 55 kPa will be applied. The
embankment soil has got cohesion of 12 kPa and angle of internal friction of 35. Find out the
type and configuration of geocells needed to achieve a desired factor of safety of 3. Cross-
section of the embankment in given problem is shown in Figure 16. From slope stability
analysis of unreinforced embankment, the minimum factor of safety was obtained as 0.633.
The center of critical slip circle was obtained as (7.975, 15.983).

55 kPa
1.5
c=12 kPa, =35 1
4m
=20 kN/m3

cu=15 kPa, =15 kN/m3 6m

Figure 16. Cross-section of the embankment in design example 2

Assuming that the embankment soil itself will be used as fill material inside the geocells and
the height of geocell layer is 2 m, geocell layer will have angle of internal friction of 35. By

20
conducting slope stability analysis with trial values of cohesion of geocell layer, for a factor
of safety of one, the cohesive strength of geocell layer (cg) was obtained as 30 kPa. As the fill
soil has original cohesive strength of 12 kPa, additional cohesive strength to be derived from
geocell reinforcement (cr) is 18 kPa. For a  value of 35, kp is 3.69. Substituting the values
of cr and kp as 18 kPa and 3.69 in equation (8), 3 is obtained as 18.7 kPa.

Assuming that the axial strain in geocell wall is 5% and the pocket-size of the geocell layer is
1 m, equivalent diameter of cells is calculate as 0.564 m. Substituting the values of 3, D
and a in equation (5), M is obtained as 200 kN/m. Thus a geocell layer of 2 m height and
pocket size of 1 m with geocells made of geogrids having secant modulus at 5% strain (M) as
200 kN/m could be provided at the base of the given embankment to achieve a factor of
safety of three against bearing capacity failure.

3. GEOCELL RETAINING WALLS AND SLOPES

Stacking of geocell layers to create retaining walls and slopes has solved several issues like
space constraints and complicated designs by allowing flexible design patterns with multifold
increase in the load carrying capacity of these structures. Geocell walls are extremely flexible
and hence the deformations are independent in each layer to an extent, thus avoiding
cumulative deformations piled up at the top of the wall as seen in case of rigid retaining
walls. Schematic diagram of a typical geocell retaining walls are shown in Figure 17 and a
finished geocell retaining wall is shown in Figure 18.

Figure 17. Schematic diagram of geocell retaining wall

21
Figure 18. Finished geocell retaining wall

Geocell walls are found to be extremely stable against seismic loads because of their
flexibility and wider facia that can render stability against sliding and overturning. Systematic
shaking table studies are carried out at Indian Institute of Science, on the seismic response of
geocell retaining walls, especially to study their acceleration and displacement response
affected by various levels of ground motion parameters. Tensile strength of geocell material
is scaled down to suit the similitude requirements of the model tests, keeping in view of the
range of tensile strength of geocells typically used in field. Figure 19 shows the photograph
of a typical geocell wall tested in shaking table (Latha and Manju, 2016).

Figure 19. Typical geocell wall constructed in a laminar box

22
The stability of the geocell wall increases with the increase in the number of facia cells.
Figure 20 shows the reduction in displacements of the wall with the normalized height, when
subjected to ground motion of acceleration amplitude 0.3g at different frequencies 1,2,3 and 7
Hz in different tests (S1A3F1, S1A3F2, S1A3F3, S1A3F7, S2A3F1, S2A3F2, S2A3F3,
S2A3F7), S1 representing 4 facia geocells, S2 representing 2 facia geocells, F1-F7
representing frequency range.

Figure 20. Effect of geocell configuration on wall deformations a) 1 Hz b) 2 Hz c) 3 Hz d) 7


Hz (After Latha and Manju, 2016)

4. GEOCELL SUPPORTED FOUNDATIONS

It is the pioneering work of Binquet and Lee (1975) that marked the beginning of systematic
research in the field of reinforced earth beds. Subsequently many researchers have reported

23
the beneficial effects of using soil reinforcement on the performance improvement of shallow
foundations. Geocell reinforcement for foundation strengthening has gained lot of attention in
recent times. Geocell mattresses provided below the foundation are proved to be effective in
distributing the load over larger area compared to other forms of reinforcement and it is
established that the bearing capacity can be improved as much as five times by using geocell
reinforcement. Increase in lead bearing capacity of sand beds with geocell reinforcement is
studied by Latha and Somwanshi (2009). Figure 21 shows the schematic diagram of geocell
reinforced foundation bed.

Figure 21. Schematic of the model footing on geocell reinforced earth bed

It is well established through laboratory and field studies that among the different forms of
geosynthetics, geocell is usually the compact form, which provides better bearing capacity for a
foundation bed by providing allround confinement and distributing loads over a larger are. A
geosynthetic material, when used in different forms like planar layers, geocells or discrete
elements, gives different strength improvements though the quantity of material is same. The
mechanism by which the strength is improved varies in different forms. In planar layers and
discrete elements, strength improvement is mainly due to friction. Interlocking also adds up to the
strength if the material has apertures to hold the soil grains. Randomly oriented reinforcing
elements coil around the soil particles which will be additional advantage in some cases. In case
of geocells, in addition to the friction and interlocking, allround confinement effect imparts
additional strength to the encased soil. A study is undertaken by Latha and Somwanshi (2009) to
compare the performance of different forms of geosynthetic reinforcement (i.e. geocell, planar
layers and randomly distributed mesh elements) in improving the bearing capacity of square
footings and reducing the deformations for exactly same quantity of material. A photograph of
polypropylene made geosynthetic material used in different forms for supporting foundation
loads is shown in Figure 22.

24
Figure 22. Geosynthetics in different forms used for reinforcement: (a) planar layers,
(b) randomly distributed mesh elements and (c) geocell layer

Results of model tests carried out on unreinforced sand beds and sand beds reinforced with
planar geosynthetics, random fibres and geocells are compared in Figure 23. Sand bed with
geocell reinforcement did not show a clear failure even at a large settlement equal to about
35% of the footing width as shown in Figure 23. The response is almost linear up to much
larger settlements of about 13% of the footing width. The footing with geocell reinforcement
carried load as high as five times the ultimate capacity of footings on unreinforced soil. The
footing settlement of 20-30% does not truly represent the possible field situation. However,
the tests were continued beyond this settlement so as to show that even at this higher
settlement, the load applied is well below the bearing capacity of the footing. At settlements
lower than 10% of the footing width, the geocell layer performed almost on par with the
planar reinforcing layers.

25
Figure 23. Variation of bearing pressure with footing settlement for different forms of biaxial
grid reinforcement

The extremely high load bearing capacity exhibited by the footing on geocell layer is
probably the result of three factors. First, the geocell mattress due to its cellular structure
contains and confines the sand more effectively. As a result a better composite material is
formed, which helps to redistribute the footing load over a wider area. Second, geocell
reinforcement system acts as an interconnected cage and derives anchorage from both sides
of loading area, due to friction and passive resistance developed at the soil/geocell interfaces.
Further, because of shear and bending rigidity of the geocell layer, the footing load is carried
even after shear failure of the sand inside the geocell pockets beneath footing. Third, the
planar layer below geocell mattress resists the downward movement of soil due to footing
penetration. In contrast, planar reinforcement layers underwent pullout failure leading to
abrupt failure while much of its tensile strength remains immobilized.

5. GEOCELL REINFORCED SUBGRADES

Geosynthetics can be effectively used to reinforce road subgrades in soft soils. In general, the
geosynthetic layer is placed at the interface of the subgrade and the aggregate base.
Sometimes, additional layers are placed within the base course and above the base course to
provide extra support to the wheel loads. Geosynthetics can have one or more of the
following functions when used for in the construction of roads: separation, filtration, drainage

26
and reinforcement. Compared to the unreinforced unpaved road, the presence of geosynthetic
reinforcement can provide the following benefits:

 Reduction of fill thickness


 Separates aggregate from soft soil if a geotextile is used
 Increases soft soil bearing capacity
 Reduces fill lateral deformation
 Generates a more favorable stress distribution
 Widens the spreading of vertical stress increments
 Reduces vertical deformation due to membrane effect
 Increases the lifetime of the road
 Requires less periodical maintenance
 Reduces construction and operational costs of the road

The following sections briefly describe various laboratory triaxial tests, model plate load tests
and field tests carried out on geocell reinforced unpaved road sections and the important
observations from these studies.

Large Diameter Cyclic Triaxial Tests

Cyclic loading resistance of geocell reinforced aggregate systems was investigated by Nair
and Latha (2014) through large diameter cyclic triaxial tests. Aggregates of different size
ranges were mixed in calculated proportions by weight to obtain the gradation specified for
rural roads. Triaxial samples of 300 mm diameter and 600 mm height were prepared using
this sampled aggregate. The strength and stiffness characteristics of this aggregate inside a
geocell made of woven geotextile at different elevations were determined from static and
cyclic triaxial tests. The results were compared with the tests using planar geogrid
reinforcement. Fig. 24 shows the failure patterns observed in these tests.

Aggregate reinforced with planar geogrids bulged between the layers but the geocell
encasement could arrest the bulging. Even a geocell of low seam strength of 7. 5kN/m could
provide an increase in confining pressure of 7 kPa for a large diameter triaxial sample.

27
Figure 24. Failure modes for samples reinforced with
(a) 4 layers of biaxial geogrid and (b, c) geocell

Model Plate Load Tests

Nair and Latha (2015) carried out plate load tests on model pavement sections reinforced
with planar and cellular geosynthetic systems. Granular sub-base was constructed over clayey
subgrade in a steel tank to simulate field condition. Geogrid and geocell reinforcements were
used in the studies. Repeated loading was applied on these unreinforced and reinforced
sections to understand the resilient behaviour of these systems. The effect of type, form and
position of reinforcement in reducing plastic settlements was also investigated through these
experimental studies. The influence of aspect ratio of geocell reinforcement on elastic and
plastic strains is also studied. These model studies were carried out in a steel tank of 750 mm
× 750 mm cross section and 620 mm height. Load is applied through a circular steel plate of
150 mm diameter and 10 mm thickness. The size of the plate was selected such that there
will be no interference with the boundary. A manually operated hydraulic jack of 100 kN
capacity was used to push the loading plate to the fill and the applied load was measured
using a load cell of 10 kN capacity. Schematic sketch of the experimental set-up is shown in
Fig. 25.

Locally available red soil was used, classified as clay of low plasticity (CL) is used as
subgrade in the experiments. The red soil used showed maximum dry unit weight of 18.24
kN/m3 at an optimum moisture content of 15.5% in a standard Proctor test. The subgrade soil
had an unsoaked CBR value of 19% at optimum moisture content and maximum dry unit
weight corresponding to standard Proctor effort. Granular material of various size ranges was

28
collected and sampled such that it conformed to Grading III of granular sub-base design as
given by IRC (2004). In all the reinforced tests, to prevent the intermixing of granular sub-
base with the subgrade a geotextile layer was placed at the interface. The various types of
reinforcing materials used in the experiments are strong and weak geogrids and geocells.

Figure 25. Schematic sketch of the experimental set-up of Nair and Latha (2015)

Commercially available geocells were used to reinforce the granular base. Five different
heights of geocell reinforcement were used in the experiments viz., 25 mm, 50 mm, 75 mm,
100 mm and 150 mm. Out of this 75, 100 and 150 mm heights were commercially available.
Higher height geocell samples were cut to have 25 and 50 mm height geocell samples. Based
on the height of geocell used to reinforced granular sub-base they are designated as GC 25,
GC 50, GC 75, GC 100 and GC 150. For granular sub-base reinforced with 75 mm high
geocell, the position of the geocell was also varied in three of the tests. Photograph of
aggregate being filled in commercially available geocell sample is shown in Fig. 26. The
total weight of granular sub-base used for filling the 200 mm height was 186 kg in both
unreinforced and reinforced cases.

29
Figure 26. Photograph of aggregate being filled in geocell pockets

Repeated load tests were carried out on models of geocell reinforced pavement sections. All
the reinforced systems showed punching failure and the pressure-settlement response
corresponding to the first loading stage of unreinforced and geocell reinforced sections is
shown in Fig. 27.

Figure 27. Pressure versus settlement for unreinforced and geocell reinforced sections

Fig. 28 shows the variation of percentage reduction in settlement (PRS) with height of
geocell reinforcement. From the figure it is seen that when the height of geocell
reinforcement is increased from 25 mm to 150 mm initially the PRS increased and the
maximum reduction in settlement is observed for geocell of 75 mm height. On increasing the
height of geocell beyond 75 mm PRS decreased which implies that those sections developed
enormous settlements.

30
Figure 28. Percentage reduction in settlement with height of geocell

The effectiveness of geocell reinforcement is compared in terms of elastic and cumulative


plastic settlements developed under repeated loading and are shown in Fig.29. It is evident
that geocell reinforced sections reduced the cumulative plastic settlements and the maximum
reduction is observed for 75 mm height geocell reinforced section and least reduction for GC
150 mm section. Geocell of 100 mm height developed less plastic settlements initially but
increased with number of repetitions and at the end of 100 cycles. GC 100 and GC 50
developed almost same cumulative plastic settlements. The ascending order of performance
improvement (in terms of reducing cumulative plastic settlement) for various geocell
reinforced sections was GC 150, GC 25, GC 50, GC 100 and GC 75.

Figure 29. Cumulative elastic and plastic settlements in unreinforced and


geocell reinforced sections

The elastic settlements developed in the geocell reinforced sections depend upon the stiffness
of the geocell mattress. Higher the height of geocell section, higher is the elastic settlement.

31
Though 150 mm height geocell reinforced section developed high elastic settlements initially,
it decreased drastically after 10 cycles. As the height of geocell increases the cumulative
plastic settlement decreases up to 75 mm height geocell and beyond that it again increases.
Similarly when the height of geocell increases, the elastic settlement also increases. But for
25 mm, 50 mm and 75 mm sections the elastic settlement is more or less the same at the end
of 100 cycles because when the height of geocell gets reduced, it behaves more or less like a
planar reinforcement. The surface profile for unreinforced and geocell reinforced sections of
various heights at the end of 50 and 100 cycles is shown in Fig. 30. From the figure it is
seen that unreinforced section and 150 mm high geocell reinforced section developed heave
compared to any other section. In all other geocell reinforcement sections, an overall
settlement of the granular sub-base was observed. This infers that the load applied on the
geocell mattress was distributed to a larger area and thus increases the overall stiffness of the
sub-base.

Figure 30. Surface profile at the end of 50 and 100 cycles


for unreinforced and geocell reinforced sections

32
The elastic and plastic settlements developed in the 75 mm height geocell reinforced sections
at various placement positions are compared in to understand the stiffness of the systems.
Fig. 31 compares the improvement factors corresponding to the loading stage of first cycle
for the unreinforced and 75 mm height geocell reinforced sections. It was observed that the
order of performance of various sections in reducing cumulative plastic settlement is
dependent on the height of the granular fill above i.e., higher the height of fill, lesser is the
cumulative plastic settlement. On comparing the elastic settlements it can be seen that
section with reinforcement placed at top had less elastic settlement compared to other two
sections considered. This could be due to the loose packing of the granular material within
the geocell pocket that too at a shallow depth from the surface making that section weak and
flexible. Section with geocell placed at middle exhibited a stiff behaviour compared to
section with geocell placed at interface and hence needed more elastic and plastic settlements
to mobilize the tensile force in it.

Figure 31. Variation of improvement factor for 75 mm height geocell sections at different
positions with settlement of plate for loading stage of first cycle, s/B (%)

Field Tests

Trafficability tests were conducted on geosynthetic reinforced unpaved roads in field, using
prototype materials and vehicle, thus avoiding some of the limitations of small-scale

33
experiments. Planar geogrids layers and geocell layers fabricated on site using geogrids were
used in different tests and the performance of geogrids in these two forms is compared. The
site where the experiments were carried out is situated in Indian Institute of Science,
Bangalore. The site chosen for constructing model road section was measured 2 m  1 m. The
soil at the location is classified as Sandy Clay with an undrained cohesion of 40 kPa and CBR
value of 22%. The original soil at the location was mixed with excess amount of water and
made slushy for a depth of 10 cm and leveled. The soil beneath this 10 cm depth will be dry
or wet depending on summer or rainy season respectively. This bed was left as such for at
least 24 hours so that the soil attained homogeneous consistency. The water content and unit
weight of the subgrade were maintained as 30% and 17 kN/m3 respectively. The prepared
subgrade has an undrained cohesion of 12 kPa and CBR value of 1%. The aggregate used for
the road section is of average size 12 mm. A biaxial geogrid having ultimate tensile strength
of 40 kN/m in both the directions and a uniaxial geogrid having ultimate tensile strength of
60 kN/m in longitudinal direction and 20 kN/m in transverse direction were used as
reinforcement.

In case of geotextile and geogrids, a geosynthetic layer was cut from the rolls and placed over
the test section, covering the entire test section. The longitudinal direction of geosynthetic
layer was coinciding with the length direction of the road for all the tests to achieve
maximum benefit. In case of geocell reinforcement, initially a geotextile layer was placed
over the subgrade. A layer of geocells was constructed in diamond pattern at the site to a size
of 2 m  1 m using biaxial geogrid and anchor pins of 6 mm diameter and 10 cm effective
height and placed above the geotextile as shown in Figure 32. Geotextile layer was needed
for this case to separate the subgrade and base course and to avoid mixing of layers during
vehicle passage.

Tests were done with geocell layers of two different geometries, with the aspect ratio of cells
as one and 0.5. The test with geocells of aspect ratio of 0.5 is compared with the test with
planar reinforcement because both the tests use geogrid of 2 m2 total area, comparing the
cellular and planar form of geogrid reinforcement with the usage of same amount of
reinforcement, as the area of geogrid used in tests with planar geogrid was 2 m2. The
aggregate was placed over this bed directly (in case of unreinforced tests) or over the
geosynthetic layer placed on top of the leveled soil subgrade (in case of reinforced tests).
Total quantity of aggregate required to obtain the desired unit weight of 13.05 kN/m 3 for 10

34
cm thickness was divided into three portions and after spreading each portion, it was
compacted using a hand roller and leveled.

Figure 32. Geocell layer prepared at the site

In case of tests with geocell reinforcement, aggregate was filled in geocells itself at the
required density. The in-situ dry soil was mixed with 10% water and placed over the
aggregate layer to prepare a comfortable riding surface. The thickness of this layer was
maintained as 5 cm and it was leveled using a drop hammer. A scooter weighing 106 kg was
driven by a person weighing 55 kg at the centre of the finished roadbed. The speed of the
vehicle was maintained as 18 to 20 kmph and the vehicle was passed in one direction only.
The rut depths were measured at marked grid points after every 20 passes until 200 passes
were completed. Then it was passed continuously for 50 times more and the final rut depths
were noted. If the vehicle started skidding in any point of time, the test was stopped at that
particular stage and the corresponding number of passes and rut depths were noted.

Figure 33 compares the behaviour of geotextile, biaxial geogrid and geocell layer prepared
using 5.85 m2 area of biaxial geogrid (aspect ratio of 1) with the control section in terms of
rut formation at different sections. Sections 1, 2 and 3 are spaced at equal distance along the
road section, dividing the road into . The aspect ratio of cells was 1, as maintained in most of
the field cases. From this plot, it is evident that this geocell layer is the most efficient form of
reinforcement compared to all other types tested. Also reduction in heave of adjacent road

35
surface and more uniform settlements were observed in case of road section reinforced with
geocell layer. However, the cost involved in preparing the geocell layer and the construction
time are to be considered while assessing the relative beneficial effects of these reinforcing
materials.

Figure 33. Comparison of performance of geocell layer with biaxial geogrid and geotextile at
different sections
Note - UR: Unreinforced; BG: Biaxial Geogrid reinforced; GT: Geotextile reinforced; GC: Geocell reinforced

6. GEOCELLS FOR FLOOD PROTECTION

Geocells are being successfully employed for building flood protection systems. In
catastrophic rain and emergency flood situations, these geocell flood walls offer stronger and
effective control of holding back water. geocell flood wall offers quick set-up in comparison
to sandbags and other traditional flood control techniques. The cellular system for these walls
is made up of a strong geotextile and the cells are expanded and filled with sand or other
suitable ballast material to provide structural support. These systems could be made use to
constructed flood barriers of any shape and size in very less time. One more advantage of
these systems compared to the traditional sand bags is that these systems function as a

36
cohesive singular structure as opposed to separate sandbag units that are susceptible to
structural failure. Specific advantages of these systems are: Easy installation, quick removal,
workability with tough terrains, durability and strength and economic pricing. Figure 34
shows geocell flood protection systems.

Figure 34. Geocell flood protection systems


These geocell flood protection systems have all the essential components such as speed,
structural strength, low seepage rates, and reuse, each factor contributing equally to overall
performance of a flood barrier. These flexible barrier systems can be designed to be used in
various circumstances common in flood fight arenas. Applications include road building over
unstable soils, levee seepage and boil control, mudslide diversion and control, and beach
erosion. If these flood protection systems can be deployed in regional and strategic locations
nationwide, we can ensure enhanced preparedness measures for floods in various parts of the
country. These systems are a cheaper replacement to the traditional sand bags, but are
extremely robust and suitable for rapid and easy installation. A highly successful case study
of these geocell flood protection systems was reported in Smithland, Kentucky, USA in May
2011. Within 48 hours after receiving the flood warning, the installation teams were formed

37
from volunteers including local citizens, city employees, National Guard support, and even
inmates from nearby correctional facilities. They could install more than 10,500 linear feet of
Flood Walls, stretching over one mile in length stacked two units high, almost four feet of
additional flood protection height to a key stretch of the levee to help raise the town levee to
meet the pending flood projections

7. GEOCELLS FOR EROSION CONTROL IN SLOPES AND CHANNELS

Soil slopes are prone to erosion due to wind or water forces. These forces form rills in the
exposed soil. Over the time, these forces get concentrated within the rills, which accelerate
the erosion process. Geocells could be successfully substituted for more costly conventional
erosion control systems such as riprap, revetment mats, armour stones and gabions. To
protect a slope from erosion, a layer of geocells is placed over the slope and anchored to the
slope at specified intervals. These geocells confine the fill material and protect it from being
moved by wind or water. Each cell acts as a dam that allows wind or water to pass over the
top while holding the fill in place. The cell wall inhibits the formation of rills, thus preventing
the erosive process. Grass can be grown in the pockets of geocell, making the slope more
stable and attractive. On vegetated slopes, geocell system increases erosion resistance by
encapsulating and protecting the vegetated root zone. As construction budgets tighten and
environmental concerns rise, synthetic materials used to prevent soil transport have seen a
rapid gain in popularity. Since natural surfaces are susceptible to large soil loss due to the
kinetic energy generated by precipitation impact and flowing water, the magnitude of the
erosion damage is a function of the surface's resistance to transport. Geocell erosion control
systems have been developed specifically to strengthen the soil surface for these types of
applications. These materials vary in size, shape and composition, but are all designed to
decrease soil disturbance and increase soil moisture.

Since any increase in the tensile strength and/or density of the soil results in a greater
resistance to applied forces, a dimensionally stable containment system is an attractive way of
protecting a slope. Geocells are three-dimensional polyethylene structures that physically
contain the infill material desired and resist the soils' natural weakness to detach and move
downslope. These products are economical, aesthetically pleasing and quite easy to design
and work with when involved in erosion control and channel lining projects. A variety of

38
materials can be used to build these erosion control systems into a three dimensional cellular
network and they can be filled with choice of infill materials ranging from sand to gravel.
Geocell layers provide protection for open channels and hydraulic structures. This type of
protection is ideal for channels exposed to severe erosive conditions as well as channels with
continuous flows. The hydraulic performance of conventional protection materials such as
concrete, gravel, riprap and vegetation is greatly improved by confining them within the
cellular structure.
A geocell channel protection system can be designed for a particular site based upon the
factors like compatibility with local environment, ecological and aesthetic requirements,
maximum anticipated flow conditions, associated hydraulic stresses and surface roughness.
Geocell layer with a nonwoven geotextile under-layer combined with custom outlet ports
assures effective subgrade drainage and subsoil protection. Vegetated soil can be used as
infill material in geocell pockets for swales, ditches and on upper slopes of large channels,
where low to moderate, intermittent flows occur. Geocell walls, which contain the topsoil
infill, form a series of check-dams, extending throughout the channel protection system. Rill
and gully development is restricted since the flow is continuously redirected to the surface.
Figure 35 shows the schematic and photograph of geocell erosion control systems.

Figure 35. Schematic diagram and photograph of geocell erosion control system

8. GEOCELLS FOR DEFENSE APPLICATIONS

The geocell barrier system can be utilized to provide personnel and infrastructure protection
in military and security applications. These systems are extremely light in weight, man
portable and non-metallic with a small logistical footprint. They are commercially available

39
in various flexible and modular configurations and can be installed rapidly. These systems
can be installed as defense barriers to provide protection from fire, blasts and bullets. They
offer significant logistical advantages over sandbags and other barriers systems. Each unit is
man portable, with section length close to 5m and weighing less than 10 kg. These cells can
be filled with locally available materials, the system is modular in height and width allowing
construction to meet the differing threat requirements. All parts are man portable and air
droppable, facilitating deployment in hostile environments. Geocell security barriers
(DefencellTM) have been installed in many locations in UK and abroad to provide effective
but discreet protection to infrastructure. Construction can be tailored in height, width and
configuration to meet operational force protection requirements. Cellular structure provides
considerable strength combined with built-in redundancy so if one cell is damaged the one
behind will continue to provide protection and stability. The completely non-metallic
structure has been extensively tested and protects against vehicle attack as well as blast and
ballistic threats yet is easy to install and maintain. These systems are 5-10 times lighter than
gabions and their all-textile construction minimizes risk of secondary fragmentation and RF
interference sometimes caused by wire mesh gabions. These geocell barriers are tested for a
wide range of ballistic threats and different explosive charges, and meet the protection
requirements.

Figure 36. Applications of cellular confinement systems in military and security operations
(from DefencellTM Website)

40
9 SUMMARY

The beneficial use of geocell reinforcement for embankment basal reinforcement, reinforcing
earth beds to support foundations, subgrade reinforcement for unpaved roads, retaining walls,
flood protection and erosion control systems is reviewed. Laboratory and field experiments
on geocell reinforced soil structures are presented and results showed that the geocell
reinforcement is effective in improving the load carrying capacity of these structures and in
reducing the deformations. Compared to the planar geosynthetic reinforcement, cellular
reinforcement is many times effective in supporting the loads because of the all-round
confinement effect. Major factors that influence the efficacy of the geocell layer are the
aspect ratio of cells, tensile modulus of the geocell material and the infill material. In case of
embankments, geoocell layer helped in pushing the failure surface deep into the foundation
bed, thus mobilizing more shear strength. In case of foundation beds, geocell reinforcement
redistributed the load over wider area and provided resistance to the downward movement of
the soil. In case of subgrades, geocell layer substantially improved the traffic benefit ratio.
Other applications of cellular confinement systems, including slope erosion control, channel
protection, flood protection and military barrier systems are briefly discussed.

REFERENCES

Bathurst, R.J. and Jarrett, P.M. (1988). “Large scale model tests of geocomposite mattresses
over peat subgrades.” Transportation Research Record, 1188, 28-36.
Bathurst, R.J. and Karpurapu, R. (1993). “Large scale triaxial tests on geocell reinforced
granular soils.” Geotechnical Testing Journal, 16(3), 296-303.
Bathurst, R.J. and Knight, M.A. (1998), “Analysis of geocell reinforced soil covers over large
span conduits.” Computers and Geotechnics, 22, 205-219.
Binquet, J. and Lee, K.L. (1975). “Bearing capacity analysis on reinforced earth slabs.”
Journal of Geotechnical Engineering Division, ASCE, 101(12), 1257-1276.
Bush, D.I., Jenner, C.G. and Bassett, R.H. (1990). “The design and construction of geocell
foundation mattress supporting embankments over soft ground.” Geotextiles and
Geomembranes, 9, 83-98.
Cowland, J.W. and Wong, S.C.K. (1993). “Performance of a road embankment on soft clay
supported on a geocell mattress foundation.” Geotextiles and Geomembranes, 12, 687-705.

41
Dash, S. K., Krishnaswamy, N. R. and Rajagopal, K. (2001). “Bearing capacity of strip
footings supported on geocell-reinforced sand.” Geotextiles and Geomembranes, 19, 235–
256.
Han, J., Pokharel, S. K., Yang, X., Manandhar, C., Leshchinsky, D., Halahmi. I., and Parsons.
R.L. (2011). “Performance of geocell-reinforced RAP bases over weak subgrade under full
scale moving wheel loads”, ASCE Journal of Materials in Civil Engineering, 23(11), 1525-
1535.
Henkel, D.J. and Gilbert, G.D. (1952). “The effect of the rubber membrane on the measured
triaxial compression strength of clay samples”. Geotechnique, 3, 20-29.
Jenner, C.G., Bush, D.I. and Bassett, R.H. (1988). “The use of slip line fields to assess the
improvement in bearing capacity of soft ground given by a cellular foundation mattress
installed at the base of an embankment.” Proc. Int. Geotech. Symp. Theory and Practice of
Earth Reinforcement, Balkema, Rotterdam, 209-214.
Johnson, W. and P.B. Mellor (1983). Engineering Plasticity. Ellis Marwood Ltd., Chichester.
Van Nostrand Reinhold, UK.
Krishnaswamy, N.R., Rajagopal, K. and Madhavi Latha, G. (2000), “Model studies on
geocell supported embankments constructed over a soft clay foundation.” Geotech. Testing
J., ASTM, 23(2), 45-54.
Latha, G.M. and Manju, G.S. (2016). “Seismic Response of Geocell Retaining Walls
Through Shaking Table Tests”. International Journal of Geosynthetics and Ground
Engineering, 2(7), 1-15.
Latha, G. M. and Somwanshi, A. (2009). “Effect of reinforcement form on the bearing
capacity of square footings on sand”, Geotextiles and Geomembranes, 27(6), 409-422
Madhavi Latha G., Rajagopal K. and Krishnaswamy N.R. (2006). “Experimental and
theoretical investigations on geocell supported embankments.” Int. J. of Geomechanics,
6(1), 30-35.
Madhavi Latha, G. (2000). “Investigations on the behavior of Geocell supported
embankments.” Ph.D. Thesis, Indian Institute of Technology Madras, Chennai, India.
Madhavi Latha, G. and Rajagopal, K. (2007). "Parametric finite element analyses of geocell
supported embankments." Canadian Geotechnical Journal, 44(8), 917-927.
Madhavi Latha, G. and Vidya S. Murthy. (2007). "Effects of reinforcement form on the
behaviour of geosynthetic reinforced sand." Geotextiles and Geomembranes, 25, 23-32.

42
Madhavi Latha, G., Asha M Nair and Hemalatha, M.S. (2010). "Performance of
geosynthetics in unpaved roads." International Journal of Geotechnical Engineering, 4(2),
151-164.
Madhavi Latha, G., Dash, S. K. and Rajagopal, K. (2008). “Equivalent continuum
simulations of geocell reinforced sand beds supporting strip footings.” Geotechnical and
Geological Engineering, 26(4), 387-398.
Madhavi Latha, G. Asha M Nair and Hemalatha, M.S. (2010) "Performance of geosynthetics
in unpaved roads", International Journal of Geotechnical Engineering. 4(2), 151-164
Nair, A.M. and Latha, G.M. (2014). “Large diameter triaxial tests on geosynthetic-reinforced
granular sub-bases”, Journal of Materials in Civil Engineering, ASCE, 04014148:1-8.
Nair, A.M. and Latha, G.M. (2016). “Repeated load tests on geosynthetic reinforced unpaved
road sections”, Geomechanics and Geoengineering, 11(2), 95-103.
Rajagopal, K., Krishnaswamy, N.R. and Madhavi Latha, G. (1999), “Behavior of sand
confined in single and multiple geocells.” Geotextiles and Geomembranes, 17, 171-184.
Rea, C., and Mitchell, J.K., (1978). “Sand reinforcement using paper grid cells”, Preprint
3130, ASCE spring convention and exhibit, Pittsburgh, PA, 644-663.

43
Ground improvement Case studies
Minimol Korulla
Chief Technical Officer at Maccaferri Environmental Solutions Pvt. Ltd.

Case Study -1: Road Over Bridge connectivity between Mundra and NH8A near
PMC Building, Port Road, Mundra, Gujarat.
Mundra Special Economic Zone is located in Kutch district, Gujarat and is one of the largest
SEZ in India. An ROB was proposed over the railway line that cuts across the connectivity road
between Mundra and NH8A. The approaches of ROB were proposed to be retained by Reinforced
soil wall system (granular fill soil unit weight 20KN/m2 and angle of internal friction 32).
Maximum height of reinforced soil wall was 9m. However, foundation soil comprising primarily
of sandy silt with clay was found to have inadequate load bearing capacity to bear the load of
retaining walls. The soil upto 3m depth was clayey silt followed by silty sand upto 4.5m depth.
And sandy silt with traces of clay till 9m depth.
Keeping in mind the high water table and high consolidation settlements, such a ground
improvement technique was to be proposed which would improve bearing capacity of soil, reduce
post construction settlement and also facilitate process of implementation. By providing
conventional solutions of soil replacement other such techniques was not warranted and would
have had made the structure commercially unviable. Thus, considering all these factors, stone
column technique with a geosynthetic raft i.e. basal reinforcement was adopted.
High strength geogrids having mono axial array of geosynthetic strips, which has planar structure
were used as basal reinforcement to improve the strength of underlying soil with a drainage layer
and geotextile in between. The uni directional ultimate strength of mono axial geogrid was
200KN/m. Stone columns were used to reduce the settlement of approach road at higher heights.
The high strength geogrids placed were effectively able to distribute the stress uniformly to
foundation soil, thereby decreasing the differential settlement. Maximum tensile load was
calculated as sum of loads needed to transfer the vertical embankment lading on the stone columns
and load needed to resist lateral sliding. Since the load was transmitted to stone columns,
settlement of the soil in between the columns was also reduced considerably.
Figure 1. Schematic representation of Ground improvement- Pile (stone columns) supported Basal
Reinforcement.

The complete Pile (stone columns) supported Basal Reinforcement was designed as per BS 8006.
Ultimate limit state i.e. Stone column group capacity, extent, vertical load shedding onto stone
columns caps, lateral sliding stability of embankment fill and over all stability was considered.
Serviceability limit state for Excessive strains in reinforcement and settlement of stone column
foundation were also checked. In this project the stone columns were designed as per 15284 (Part
1).

Figure 2. Schematic representation of global stability check output


Figure3. Variables used in analysis of overall stability of basal reinforced piled embankments
Note:
1 Slip circle centre, 2 Slice i, 3 Embankment, 4 Reinforcement, 5 Pile caps, 6 Piles, 7 Most critical slip surface

The area is recommended to be excavated till the founding level where the stone columns are
installed.
Stone columns:
• South side - entire zone including the reinforced and unreinforced section was improved
by stone column technique.
• North side - owing to relatively better quality of soil, stone columns were provided only
under reinforced soil.
High strength geogrid was laid throughout the entire zone both in the North and South side of
the structure. The surface of the stone column was covered with a free draining granular fill
compacted to 95% of Modified Proctor density.
Table Summary of ground improvement
Wall Heights(m) 8.8 8 7.2 6.4 5.2 & down
Ultimate tensile strength of 200 200 200 200 200
geogrid (KN/m)
Stone column Dia(m) 0.6 0.6 0.6 0.6 0.6
Stone column Spacing(m) 1.55 2 2 2.2 2.5

Figure 4. Typical Cross section of Ground Improvement and Reinforced soil wall- North Side

Figure 5. Typical cross section of ground improvement at Cross wall.


Photo 1 Installation of stone column Photo 2 Installation of drainage layer

Photo 3 Laying of High strength uniaxial Geogrid


Case study 2: Ground improvement for ramp at Calcutta riverside, Bhatnagar

River Bank Developers Pvt. Ltd. was widening the approach ramp that leads to a ROB at
the entrance of the project. Existing ramp is an open embankment, however due to scarcity of
space Reinforced Soil Wall is to be constructed for the proposed ramp. The proposed road/ ramp
will move parallel to the existing road and then will merge with the existing road before the
location of the ROB. Length of the ramp was about 160m and the height varies from 1m at the
start to about 6.4m near the ROB.
Objective
The subsoil at location of Reinforced Soil wall upto a depth of 10 m is soft to very soft
silty clay having S.P.T. value ranging from 0 - 4. Besides this the water table was at a depth of
0.5m. Area to be consolidated was 1691.2 m2. Construction without some sort of soil treatment
was impractical due to unpredictable long-term settlement. Although surcharging increases pore
water pressure, yet settlement can take considerable time, often years, as the water lacks an easy
path to leave the soil. Hence the requirement of ground improvement here was such that it is cost
effective and time saving.
Solution Proposed
Prefabricated Vertical Drains (PVD) for accelerated consolidation of soft soil was adopted
to accelerate settlements, to reduce time for consolidation & to avoid Post Construction
settlements. As a surcharge, an embankment of height equivalent to the height of ramp was
constructed and was kept for duration of 45 days. During this time consolidation of soft soil has
occurred. Consolidation of soft cohesive soils using prefabricated vertical drains reduced
settlement times from years to months. Most settlement has occured during construction, thus
keeping post-construction settlement to a minimum. The Prefabricated Vertical Drains are less
expensive, are installed more easily and quickly and gave better drainage by providing shortened
drainage paths for the water to exit the soil. The spacing of PVD's adopted was 1.2m c/c in a square
configuration to achieve 90% consolidation in 45 days. Installation of 13,860 m length of PVD's
was completed in 15 days after the sand bed was laid. Above PVD's a drainage blanket of 0.3 –
0.5 m. was placed for drainage purpose
Design Considerations
The following design parameters were considered to find out the required spacing of PVD's
to achieve 90% consolidation.
Co-efficient of vertical consolidation (Cv) = 2.31 * 10-3 cm2 / sec (Calculated from soil
investigation report by referring BH. No. 1)
Co-efficient of radial consolidation (Ch) = Cv = Ch (Assumed)
Average degree of consolidation U= 90% (Assumed)
Undrained cohesion of clay = 31 kN/m2 (As per soil investigation report)
Plasticity index of soil = 22% (As per soil investigation report)
Unit weight of clay = 17.8 kN/m3 (bulk) (As per soil investigation report)
Depth of clay layer = 10 m (As per soil investigation report)
Consolidation Period = Preferred period 1.5 – 2.0 Month (Assumed)
Area to be consolidated = 151 X 11.2 (10+0.6+0.6) = 1691.2 m2
Conclusions
Based on the design,
1) The spacing of PVD's is calculated 1.2 meter center to center (if Ch = Cv) to achieve
a ~90% consolidation within 1.5 – 2.0 months. The detail calculations are shown
in Annexure-II.
2) PVD's are provided upto 1.5 m. height of RS wall i.e. up to 151 m. length of wall.
For remaining length of a wall (approx. 29 m.) a well granular soil having angle of
internal friction 32 degree can be placed up to 1 m. below levelling pad to take care
of safe bearing capacity as well as settlement.
3) Above PVD's a drainage blanket of 0.3 – 0.5 m. should be placed for drainage
purpose.
4) The total required length of PVD's is 13860 (including 10% wastage).
Fig-1: Installed PVD’s

Fig-2: Installation of PVD


Case study 3: South kasheli creek bridge Thane Bhiwandi Vadapa road, Maharashtra

Kasheli Bridge built since British time lies on the Old Agra Road over the Thane Creek.
This bridge is 460 meters long and connects the Thane mainland to Bhiwandi. For a major bridge
across Thane Bhiwandi Vadapa Road, the solid approaches were required to be retained using
reinforced soil walls. There was an embankment existing for many years. The road had to be
widened to the increased width of the bridge. The subsurface comprised of top 4 to 6 m of very
soft to soft dark grey clay. From 7.5 m to 10.0 m soil constituted silty clay. This layer was followed
by medium dense dark grey medium sand. As the structures were near the ground water table was
at existing ground level. The construction of approaches had to be completed quickly with
minimum post construction settlement.
Solution
In order to achieve the required global and bearing stability, basal reinforcement over piles
was implemented for the new embankment. The piled embankment technique allowed
embankments to be constructed to the required heights without any restraint on construction rate
with control on post construction settlements. Basal reinforcement was used to form a geosynthetic
raft over piles and transfer the load to the piles, and thus enabling to maximize the economic
benefits of the piles installed in soft foundations. The reinforcement also helped in counteracting
the horizontal thrust of the embankment fill and the need for raking piles along the extremities of
the foundation could be eliminated. Fill soil properties were considered as: cohesion-0 kN/m2,
angle of friction-32°, unit weight-20 kN/m2. The maximum height of the embankment was 9.6m.
Fig-1: Typical Cross sectional drawing
In the direction along the length of the embankment, the maximum tensile load should be
there which needed to transfer the vertical embankment loading onto the pile caps. In the direction
along the width of the embankment the maximum tensile load should be the sum of the load which
needed to transfer the vertical embankment loading onto the pile caps and the load needed to resist
lateral sliding. Basal reinforcement proposed here was high strength geogrid which has planar
structure consisting of a uni-axial array of composite geosynthetics strips.
Each single longitudinal strip had a core of high tenacity polyester yarns tendons encased
in a polyethylene sheath; the single strip was connected by cross laid polyethylene strip which
gave a grid like shape to the composite. Two layers of geogrid having uni-axial strength 400 kN/m
each along and across the road were given. The design was carried out according to BS: 8006
(1995). The design of piled embankments was not included in the scope of the present document.
Fig-2: Laying of Geotextile over Pile caps
Fig-3: Laying of basal reinforcement over pile caps

Fig-4: Completed Reinforced soil wall structure


Case study 4 Road Over Bridge near Dibrugarh (ROB 15) Assam, India

Objective
A Road Over Bridge was to be constructed in Dibrugarh, district Assam by Northeast
Frontier Railway. The approaches of the ROB were to be retained by Reinforced Soil walls.
Moreover, the soil at the site was cohesive (CI) for the top 3m, followed by loose to medium dense
fine silty sand. After investigation it was found that the shear properties of the soil were weak and
hence a major problem of bearing and global instability could be expected.
Solution
For retaining the bridge approaches, mechanically stabilized concrete panel wall using
geosynthetic strip as reinforcement was constructed as the vertical retaining wall. In order to avoid
deep excavation and replacing the soil, panel wall system with basal reinforcement as ground
improvement was constructed.
Basal reinforcement was provided for construction of embankments on soft soils, which is
a very efficient technique to improve the bearing capacity and global stability of the foundation as
all the stresses on the foundation is taken care by the reinforcement that is provided. Basal
reinforcement prevents collapse and limit vertical movement of the embankment surface following
the formation of a void in the foundation.
Ground improvement was done for heights ranging from 6m to 5m and 5m to 1m for
Seismic and Static conditions present at the site. High strength uni-axial geogrids as basal
reinforcement with a tensile strength of 50kN/ m was used for ground improvement.
Fig-1: Cross section of Reinforced soil wall for 6m height near abutment

Fig-2: Installation of Geotextile


Fig-3: Spreading of fill material over installed Geotextile

Fig-4: Compaction of fill material over installed Geotextile


Fig-5: Installed high strength uni-axial Geogrid (basal reinforcement)

Fig-6: Installation of RS wall system over improved ground

You might also like